11.02.2017 Views

DESCRIPTIONS OF MEDICAL FUNGI

fungus3-book

fungus3-book

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>DESCRIPTIONS</strong> <strong>OF</strong> <strong>MEDICAL</strong> <strong>FUNGI</strong><br />

THIRD EDITION<br />

(revised November 2016)<br />

SARAH KIDD 1,3 , CATRIONA HALLIDAY 2 ,<br />

HELEN ALEXIOU 1 and DAVID ELLIS 1,3<br />

1 National Mycology Reference Centre<br />

SA Pathology, Adelaide, SOUTH AUSTRALIA<br />

2 Clinical Mycology Reference Laboratory<br />

Centre for Infectious Diseases and Microbiology<br />

Laboratory Services, Pathology West, ICPMR,<br />

Westmead Hospital, Westmead, NEW SOUTH WALES<br />

3<br />

DEPARTMENT <strong>OF</strong> MOLECULAR & CELLULAR BIOLOGY<br />

SCHOOL <strong>OF</strong> BIOLOGICAL SCIENCES<br />

UNIVERSITY <strong>OF</strong> ADELAIDE, ADELAIDE<br />

AUSTRALIA<br />

2016<br />

We thank Pfizer Australia for an unrestricted educational grant to the Australian<br />

and New Zealand Mycology Interest Group to cover the cost of the printing.


Published by the Authors<br />

Contact: Dr. Sarah E. Kidd<br />

Head, National Mycology Reference Centre<br />

Microbiology & Infectious Diseases<br />

SA Pathology<br />

Frome Rd, Adelaide, SA 5000<br />

Email: sarah.kidd@sa.gov.au<br />

Phone: (08) 8222 3571<br />

Fax: (08) 8222 3543<br />

www.mycology.adelaide.edu.au<br />

© Copyright 2016<br />

The National Library of Australia Cataloguing-in-Publication entry:<br />

Creator:<br />

Title:<br />

Edition:<br />

ISBN:<br />

Notes:<br />

Kidd, Sarah, author.<br />

Descriptions of medical fungi / Sarah Kidd, Catriona Halliday, Helen Alexiou,<br />

David Ellis.<br />

Third edition.<br />

9780646951294 (paperback).<br />

Includes bibliographical references and index.<br />

Subjects: Fungi--Indexes. Mycology--Indexes.<br />

Other Creators/Contributors:<br />

Halliday, Catriona L., author.<br />

Alexiou, Helen, author.<br />

Ellis, David (David H.), author.<br />

Dewey Number: 579.5<br />

Printed in Adelaide by<br />

Newstyle Printing<br />

41 Manchester Street<br />

Mile End, South Australia 5031<br />

Front cover: Cryptococcus neoformans, and montages including Syncephalastrum,<br />

Scedosporium, Aspergillus, Rhizopus, Microsporum, Purpureocillium, Paecilomyces<br />

and Trichophyton. Back cover: the colours of Trichophyton spp.


Descriptions of Medical Fungi<br />

iii<br />

PREFACE<br />

The first edition of this book entitled Descriptions of Medical QAP Fungi was published<br />

in 1992 by David Ellis, Steve Davis, Helen Alexiou, Tania Pfeiffer and Zabeta<br />

Manatakis. The original concept was to provide all laboratories in the Royal College of<br />

Pathologists of Australasia (RCPA) Mycology Quality Assurance Program (QAP) with<br />

a set of description sheets covering medically important fungi. A second edition entitled<br />

Descriptions of Medical Fungi was released in 2007 by David Ellis, Steve Davis, Helen<br />

Alexiou, Rosemary Handke and Robyn Bartley. We now provide an updated third<br />

edition which includes new and revised descriptions. We have endeavoured to reconcile<br />

current morphological descriptions with more recent phylogenetic studies, however<br />

nomenclature changes in mycology are ongoing. To search for current accepted<br />

fungal names go to Index Fungorum (www.indexfungorum.org) and Mycobank (www.<br />

mycobank.org).<br />

Morphological Descriptions: These descriptions have by necessity been kept brief<br />

and many have been based on descriptions by other authors. For further information<br />

regarding any of the mycoses or pathogenic fungi mentioned, the reader is referred to<br />

the citations provided. For the precise definitions of the mycological terminology used,<br />

the reader is referred to Ainsworth and Bisby’s Dictionary of the Fungi (Kirk et al. 2008).<br />

Classification of the Fungi<br />

Kingdom Fungal Phyla Examples<br />

Protozoa Myxomycota Slime moulds<br />

Chromista Oomycota Pythium<br />

Eumycota<br />

Ascomycota<br />

Basidiomycota<br />

Chytridiomycota<br />

Glomeromycota<br />

Microsporidia<br />

Zygomycota<br />

Candida, Aspergillus, Scedosporium,<br />

Fusarium, Paecilomyces, Penicillium,<br />

Cladophialophora, Bipolaris, and other<br />

hyphomycetes, including the dimorphic<br />

fungi, dermatophytes, and Pneumocystis<br />

(Taphrinomycotina).<br />

Cryptococcus, Trichosporon, Malassezia.<br />

Chytrids<br />

Endomycorrhizal on plants<br />

170 genera, 1300 species<br />

Apophysomyces, Lichtheimia, Mucor,<br />

Saksenaea, Rhizomucor, Rhizopus.<br />

Fungi are now classified across three Kingdoms. Descriptions in this book are<br />

limited to the Eumycota and include medically important representatives from the<br />

Ascomycota, Basidiomycota and Zygomycota.


iv<br />

Descriptions of Medical Fungi<br />

PREFACE<br />

Key Morphological Characters<br />

Culture Characteristics:<br />

• Surface texture [glabrous, suede-like, powdery, granular, fluffy, downy, cottony]<br />

• Surface topography [flat, raised, heaped, folded, domed, radial grooved]<br />

• Surface pigmentation [white, cream, yellow, brown, pink, grey, black etc]<br />

• Reverse pigmentation [none, yellow, brown, red, black, etc]<br />

• Growth rate [colony diameter 5 cm in 15 days]<br />

• Growth at 37 O C, 40 O C, 45 O C.<br />

Zygomycota. Sporangia characteristics:<br />

• Arrangement of sporangiospores [multispored, sporangiola, merosporangium]<br />

• Arrangement of sporangiophores [unbranched often in groups or frequently branched]<br />

• Sporangium shape [pyriform, spherical, flask-shaped etc]<br />

• Sporangium size [100 μm diam.]<br />

• Columella [Present or Absent]<br />

• Apophyses [Present or Absent]<br />

• Sporangiophore height [1 mm]<br />

• Rhizoids [Present or Absent] (look in the agar)<br />

• Sporangiospore size [6 μm]<br />

Hyphomycetes - Conidial Moulds<br />

1. Conidial characteristics:<br />

• Septation [one-celled, two-celled, multicelled with transverse septa only, or multicelled with<br />

both transverse and longitudinal septa]<br />

• Shape [spherical, sub-spherical, pyriform, clavate, ellipsoidal, etc]<br />

• Size [need a graduated eyepiece, length 10 μm]<br />

• Colour [hyaline or darkly pigmented]<br />

• Wall texture [smooth, rough, verrucose, echinulate]<br />

• How many conidial types present? [i.e. micro and macro]<br />

2. Arrangement of conidia as they are borne on the conidiogenous cells:<br />

• Solitary [single or in balls]<br />

• Catenulate (in chains) [acropetal (youngest conidium at the tip) or basipetal (youngest<br />

conidium at the base]<br />

3. Growth of the conidiogenous cell:<br />

• Determinant (no growth of the conidiophore after the formation of conidia)<br />

• Sympodial (a mode of conidiogenous cell growth which results in the development of<br />

conidia on a geniculate or zig-zag rachis)<br />

4. Type of conidiogenous cell present:<br />

• Non-specialised<br />

• Phialide (specialised conidiogenous cells that produces conidia in basipetal succession<br />

without increasing in length)<br />

• Annellide (specialised conidiogenous cell producing conidia in basipetal succession by a<br />

series of short percurrent proliferations (annellations). The tip of an annellide increases in<br />

length and becomes narrower as each subsequent conidium is formed)<br />

5. Any additional features present:<br />

• Hyphal structures [clamps, spirals, nodular organs, etc]<br />

• Synnemata, Sporodochia, Chlamydoconidia, Pycnidia<br />

• Confirmatory tests for dermatophytes


Descriptions of Medical Fungi<br />

v<br />

PREFACE<br />

Molecular and/or MALDI-T<strong>OF</strong> MS Identification: The use of PCR-based assays, DNA<br />

sequencing, and other molecular methods, including those incorporating proteomic<br />

approaches such as matrix assisted laser desorption ionization time of flight mass<br />

spectroscopy (MALDI-T<strong>OF</strong> MS) have shown promising results to aid in accurate species<br />

identification of fungal cultures. These are used mainly to complement conventional<br />

methods since they require standardisation before widespread implementation can be<br />

recommended (Halliday et al. 2015). Molecular-based fungal identification is particularly<br />

helpful for fungi that lack distinguishing morphological features, e.g. Apophysomyces<br />

elegans, or to distinguish between species of the Aspergillus fumigatus complex.<br />

Comparative sequence analysis is now the ‘gold standard’ for identification of fungi.<br />

Methods are referenced where available and in many instances are recommended for<br />

more definitive identifications.<br />

Schematic diagram of the fungal rDNA gene cluster (adapted from CLSI MM18-A<br />

and Halliday et al. 2015). The 18S, 5.8S and 28S rDNA genes are separated by the<br />

two internal transcribed spacers. The 28S and 5S rDNA genes are separated by the<br />

intergenic spacer 1 (IGS1). The intergenic spacer 2 (IGS2) separates the rDNA repeat<br />

units from each other.<br />

Regardless of the genetic locus selected, accurate sequence-based identification is<br />

dependent upon database accuracy and adequate species representation. GenBank<br />

is well known to contain numerous errors in sequences and the species names<br />

attributed to the sequences, which are rarely corrected. Therefore caution must be<br />

used when interpreting sequencing comparisons against this database, and the use of<br />

multiple sequence databases is encouraged. Well-curated databases that are helpful<br />

for species identification include:<br />

1. International Society for Human and Animal Mycoses (ISHAM) ITS database (http://<br />

its.mycologylab.org/).<br />

2. CBS-KNAW Fungal Biodiversity Centre database (http://www.cbs.knaw.nl).


vi<br />

Descriptions of Medical Fungi<br />

PREFACE<br />

Frequently used molecular targets for species identification are outlined below:<br />

ITS<br />

D1/D2<br />

β-tubulin<br />

Molecular Target<br />

Internal transcribed spacer<br />

regions (ITS1-5.8S-ITS2)<br />

D1/D2 variable domains of<br />

the 28S rDNA gene<br />

Beta tubulin II<br />

Application<br />

Species level identification of wide<br />

range of fungi<br />

Species identification of many of the<br />

Mucorales<br />

Accurate species resolution of<br />

Aspergillus.<br />

Cal Calmodulin Species discrimination of Alternaria.<br />

EF-1α<br />

RPB1<br />

RPB2<br />

ACT<br />

GPDH<br />

Elongation factor alpha<br />

subunit<br />

RNA polymerase I subunit<br />

RNA polymerase II subunit<br />

Actin<br />

Glycerol-3-phosphate<br />

dehydrogenase<br />

Species complex identification of<br />

Fusarium.<br />

Species complex identification within<br />

genera of Fusarium, Penicillium and<br />

Talaromyces.<br />

Species discrimination of Aspergillus,<br />

Cladosporium, Coniochaeta, Verticillium,<br />

Verruconis.<br />

Species discrimination of Bipolaris,<br />

Curvularia, Verticillium.<br />

CHS Chitin synthase Species discrimination of Sporothrix<br />

Chi18-5 Chitinase 18-5 Species discrimination of Trichoderma<br />

Antifungal Susceptibility: For many species, antifungal susceptibility data has also<br />

been provided. This has been derived from both the literature and data from Australian<br />

clinical isolates generated by using the CLSI M27-A Standard for yeasts and the CLSI<br />

M38-A Standard for moulds. This composite data is provided as a guide only. In many<br />

cases the clinical relevance of in vitro antifungal susceptibility results remains difficult<br />

to interpret, and expert advice from a consulting microbiologist or infectious disease<br />

specialist may be required.<br />

CLSI M27-S4 clinical breakpoints are marked where available (green for susceptible,<br />

yellow for susceptible dose dependant or intermediate, red for resistant).<br />

Abbreviations: Amphotericin B (AmB), Fluconazole (FLU), Itraconazole (ITRA),<br />

Posaconazole (POSA), Voriconazole (VORI), Anidulafungin (ANID), Caspofungin<br />

(CAS), Micafungin (MICA), 5-Fluorocytosine (5FC), Terbinafine (TERB).<br />

Risk group (RG) recommendations are based on published data and on current<br />

definitions in accordance with the Australian/New Zealand Standard AS/NZS<br />

2243.3:2010. Safety in laboratories Part 3: Microbiological safety and containment.<br />

Note: International biosafety guidelines vary in their RG ratings of fungal species.


Descriptions of Medical Fungi<br />

vii<br />

CONTENTS<br />

Acremonium 1<br />

Acrophialophora fusispora 2<br />

Alternaria 3<br />

Aphanoascus fulvescens 5<br />

Apophysomyces complex 6<br />

Arthroderma insingulare (formerly Trichophyton terrestre) 9<br />

Arthroderma uncinatum (formerly Trichophyton ajelloi) 10<br />

Arthrographis kalrae 11<br />

Aspergillus 12<br />

Aspergillus flavus complex 14<br />

Aspergillus fumigatus complex 16<br />

Aspergillus felis 16<br />

Aspergillus fumigatus 17<br />

Aspergillus lentulus 19<br />

Neosartorya fischeri 20<br />

Aspergillus nidulans complex 21<br />

Aspergillus niger complex 23<br />

Aspergillus terreus complex 25<br />

Aureobasidium pullulans 27<br />

Basidiobolus ranarum 28<br />

Beauveria 30<br />

Bipolaris 31<br />

Blastomyces dermatitidis 32<br />

Candida 34<br />

Candida albicans 37<br />

Candida catenulata 38<br />

Candida dubliniensis 39<br />

Candida glabrata complex 40<br />

Candida bracarensis 40<br />

Candida glabrata 41<br />

Candida nivariensis 42<br />

Candida haemulonii 43<br />

Candida inconspicua 44<br />

Candida parapsilosis complex 45<br />

Candida metapsilosis 45<br />

Candida orthopsilosis 46<br />

Candida parapsilosis 47<br />

Lodderomyces elongisporus 48<br />

Candida rugosa 49<br />

Candida tropicalis 50<br />

Chaetomium 51<br />

Chrysosporium tropicum 52<br />

Cladophialophora 53<br />

Cladophialophora bantiana 53<br />

Cladophialophora carrionii 55<br />

Cladosporium 57<br />

Clavispora lusitaniae (formerly Candida lusitaniae) 59<br />

Coccidioides immitis/posadasii complex 60


viii<br />

Descriptions of Medical Fungi<br />

CONTENTS<br />

Colletotrichum coccodes 62<br />

Conidiobolus coronatus 63<br />

Coniochaeta hoffmannii (formerly Lecythophora hoffmannii) 64<br />

Cryptococcus 66<br />

Cryptococcus albidus 67<br />

Cryptococcus gattii 68<br />

Cryptococcus laurentii 70<br />

Cryptococcus neoformans 71<br />

Cunninghamella bertholletiae 73<br />

Curvularia 75<br />

Cyberlindnera fabianii (formerly Candida fabianii) 79<br />

Cylindrocarpon 80<br />

Debaryomyces hansenii (formerly Candida famata) 81<br />

Drechslera 82<br />

Epicoccum nigrum 83<br />

Epidermophyton floccosum 84<br />

Exophiala 85<br />

Exophiala dermatitidis 85<br />

Exophiala jeanselmei complex 86<br />

Exophiala spinifera complex 89<br />

Exserohilum rostratum 91<br />

Fonsecaea complex 93<br />

Fusarium 94<br />

Fusarium chlamydosporum complex 96<br />

Fusarium dimerum complex 97<br />

Fusarium fujikuroi complex 98<br />

Fusarium incarnatum-equiseti complex 98<br />

Fusarium oxysporum complex 99<br />

Fusarium solani complex 100<br />

Geotrichum candidum 102<br />

Gliocladium 104<br />

Graphium 105<br />

Histoplasma capsulatum 107<br />

Hortaea werneckii 108<br />

Kluyveromyces marxianus (formerly Candida kefyr) 109<br />

Lasiodiplodia theobromae 110<br />

Lichtheimia corymbifera (formerly Absidia corymbifera) 111<br />

Lomentospora prolificans (formerly Scedosporium prolificans) 113<br />

Lophophyton gallinae (formerly Microsporum gallinae) 115<br />

Madurella complex 116<br />

Madurella mycetomatis 116<br />

Trematosphaeria grisea (formerly Madurella grisea) 117<br />

Magnusiomyces capitatus (formerly Geotrichum capitatum) 118<br />

Malassezia 119<br />

Malbranchea pulchella 121<br />

Meyerozyma guilliermondii (formerly Candida guilliermondii) 122<br />

Microsphaeropsis arundinis 123


Descriptions of Medical Fungi<br />

ix<br />

CONTENTS<br />

Microsporum 124<br />

Microsporum audouinii 125<br />

Microsporum canis 126<br />

Microsporum ferrugineum 129<br />

Mortierella wolfii 130<br />

Mucor 131<br />

Mucor amphibiorum 132<br />

Mucor circinelloides 133<br />

Mucor indicus 134<br />

Mucor irregularis 135<br />

Mucor ramosissimus 135<br />

Myrmecridium schulzeri (formerly Ramichloridium schulzeri) 136<br />

Nannizzia 137<br />

Nannizzia fulva (formerly Microsporum fulvum) 137<br />

Nannizzia gypsea (formerly Microsporum gypseum) 138<br />

Nannizzia nana (formerly Microsporum nanum) 139<br />

Nannizzia persicolor (formerly Microsporum persicolor) 140<br />

Neoscytalidium dimidiatum (formerly Hendersonula toruloidea) 141<br />

Ochroconis 143<br />

Onychocola canadensis 144<br />

Paecilomyces 145<br />

Paecilomyces marquandii 146<br />

Paecilomyces variotii 147<br />

Paracoccidioides brasiliensis 148<br />

Paraphyton cookei (formerly Microsporum cookei) 149<br />

Penicillium 150<br />

Phaeoacremonium parasiticum 152<br />

Phialophora verrucosa 154<br />

Phoma 155<br />

Pichia 156<br />

Pichia kudriavzevii (formerly Candida krusei) 156<br />

Pichia norvegensis (formerly Candida norvegensis) 157<br />

Pithomyces chartarum 158<br />

Pleurostomophora richardsiae (formerly Phialophora richardsiae) 159<br />

Prototheca 160<br />

Purpureocillium lilacinum (formerly Paecilomyces lilacinus) 161<br />

Quambalaria cyanescens 163<br />

Rhinocladiella 164<br />

Rhinocladiella atrovirens 164<br />

Rhinocladiella mackenziei (formerly Ramichloridium mackenziei) 165<br />

Rhizomucor 166<br />

Rhizomucor miehei 166<br />

Rhizomucor pusillus 167<br />

Rhizopus 168<br />

Rhizopus arrhizus (formerly Rhizopus oryzae) 169<br />

Rhizopus microsporus 170<br />

Rhodotorula 171<br />

Rhodotorula glutinis 172<br />

Rhodotorula mucilaginosa 173


x<br />

Descriptions of Medical Fungi<br />

CONTENTS<br />

Saccharomyces cerevisiae 174<br />

Saksenaea vasiformis complex 175<br />

Saprochaete clavata (formerly Geotrichum clavatum) 177<br />

Sarocladium (formerly Acremonium) 178<br />

Scedosporium 179<br />

Scedosporium apiospermum 179<br />

Scedosporium aurantiacum 181<br />

Scedosporium boydii (formerly Pseudallescheria boydii) 182<br />

Schizophyllum commune 183<br />

Scopulariopsis 184<br />

Sepedonium 185<br />

Sporothrix schenckii complex 186<br />

Stemphylium 188<br />

Syncephalastrum racemosum 189<br />

Talaromyces marneffei (formerly Penicillium marneffei) 190<br />

Torulaspora delbrueckii (formerly Candida colliculosa) 192<br />

Trichoderma 193<br />

Trichophyton 194<br />

Trichophyton concentricum 195<br />

Trichophyton equinum 196<br />

Trichophyton eriotrephon (formerlyTrichophyton erinacei) 198<br />

Trichophyton interdigitale 200<br />

Trichophyton mentagrophytes 203<br />

Trichophyton quinckeanum 205<br />

Trichophyton rubrum 207<br />

Trichophyton schoenleinii 211<br />

Trichophyton soudanense 212<br />

Trichophyton tonsurans 213<br />

Trichophyton verrucosum 215<br />

Trichophyton violaceum 217<br />

Trichosporon 218<br />

Trichosporon asahii 219<br />

Trichosporon asteroides 221<br />

Trichosporon cutaneum 221<br />

Trichosporon inkin 222<br />

Trichosporon mucoides 222<br />

Trichosporon ovoides 222<br />

Trichothecium roseum 223<br />

Ulocladium 224<br />

Veronaea botryosa 225<br />

Verruconis gallopava (formerly Ochroconis gallopava) 226<br />

Verticillium 228<br />

Wickerhamomyces anomalus (formerly Candida pelliculosa) 229<br />

Yarrowia lipolytica (formerly Candida lipolytica) 230


Descriptions of Medical Fungi<br />

xi<br />

CONTENTS<br />

Microscopy Stains & Techniques 231<br />

KOH with Calcofluor White 231<br />

KOH with Chlorazol Black 231<br />

India Ink mounts 231<br />

Lactophenol cotton blue (LPCB) 232<br />

Direct microscopic preparations 232<br />

Cellotape flag preparations 232<br />

Slide culture preparations 233<br />

Specialised Culture Media 234<br />

Bird seed agar 234<br />

Bromocresol purple milk solids agar 234<br />

Creatinine dextrose bromothymol blue thymine (CDBT) media 235<br />

Canavanine glucose bromothymol blue (CGB) media 235<br />

Cornmeal agar 236<br />

Cornmeal glucose sucrose agar 236<br />

Czapek Dox agar 236<br />

Modified Dixon’s agar 236<br />

Hair perforation test 237<br />

Lactritmel agar 237<br />

Littman oxgall agar 237<br />

Malt extract agar 238<br />

1% Peptone agar 238<br />

Potato dextrose agar 238<br />

Rice grain slopes 238<br />

Sabouraud’s dextrose agar (SDA) + cycloheximide and antibiotics 239<br />

Sabouraud’s dextrose agar (SDA) + antibiotics 239<br />

Sabouraud’s dextrose agar (SDA) 5% salt 239<br />

Tap water agar 240<br />

Urea agar with 0.5% glucose 240<br />

Vitamin free agar 240<br />

References 241


xii<br />

Descriptions of Medical Fungi<br />

Schematic for the identification of medically important fungi.<br />

Acremonium Link ex Fries<br />

Key Features: Hyphomycete with solitary, erect, hyaline, awl-shaped phialides<br />

producing single-celled, globose to cylindrical conidia, mostly in slimy heads.<br />

Antifungal Susceptibility: Acremonium spp. data from about 60 isolates (Perdomo<br />

et al. 2011 and Australian National data); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 0.25-16 16 VORI 0.06-8 8<br />

ITRA 0.25-16 16 POSA 0.125-8 16<br />

References: Gams (1971), Domsch et al. (2007), Samson et al. (1995), de Hoog et<br />

al. (2000, 2015), Glenn et al. (1996), Perdomo et al. (2011a), Summerbell et al. (2011).


Descriptions of Medical Fungi 1<br />

The genus Acremonium contains many species; most are saprophytic being isolated<br />

from dead plant material and soil. Several species including A. recifei and A.<br />

alabamense are recognised as opportunistic pathogens of man and animals, causing<br />

mycetoma, mycotic keratitis and onychomycosis. Recently, several Acremoniumlike<br />

species recognised as opportunistic pathogens have been transferred to other<br />

genera; Fusarium falciforme (formerly A. falciforme), Sarocladium kiliense (formerly<br />

A. kiliense), Gliomastic roseogriseum (formerly A. roseogriseum) and Sarocladium<br />

strictum (formerly A. strictum) (Glenn et al. 1996, Summerbell et al. 2011).<br />

RG-2 for species isolated from humans.<br />

Acremonium Link ex Fries<br />

Morphological Description: Colonies are usually slow growing, often compact and<br />

moist at first, becoming powdery, suede-like or floccose with age, and may be white,<br />

grey, pink, rose or orange in colour. Hyphae are fine and hyaline and produce mostly<br />

simple awl-shaped erect phialides with inconspicuous collarettes. Conidia are usually<br />

one-celled, hyaline or rarely pigmented, globose to cylindrical, and mostly aggregated<br />

in slimy heads at the apex of each phialide. Chlamydospores may be present.<br />

Comments: Microconidial Fusarium isolates may be confused with Acremonium, but<br />

they usually grow faster and have colonies with a characteristic fluffy appearance.<br />

Phialemonium species differ by having short, tapering phialides, mostly lacking a basal<br />

septum. Coniochaeta is characterised by having sessile phialidic collarettes that are<br />

formed directly on the hyphae.<br />

Molecular Identification: Summerbell et al. (2011) revised the genus on the basis of<br />

18S and D1/D2 sequence phylogeny. Sequence based identification may be performed<br />

using the D1/D2 or the ITS region. Caution must be exercised in the interpretation of<br />

database sequence comparisons due to the scarcity of database sequences from wellcharacterised<br />

strains, and some sequences may have been attributed to species that<br />

have been reclassified (Perdomo et al. 2011a).<br />

10 µm<br />

Acremonium spp. showing long awl-shaped phialides producing cylindrical, onecelled<br />

conidia mostly aggregated in slimy heads at the apex of each phialide.


2<br />

Descriptions of Medical Fungi<br />

The genus Acrophialophora contains 16 species that are most commonly associated<br />

with soil, especially from India (Zhang et al. 2015a). Two species have been reported<br />

as human pathogens, A. fusispora and A. levis (Sandoval-Denis et al. 2015). ITS<br />

sequencing is recommended for species identification.<br />

RG-1 organism.<br />

Acrophialophora fusispora (S.B. Saksena) Samson<br />

Morphological Description: Colonies fast growing, greyish-brown with a black<br />

reverse. Conidiophores arising singly, terminally and laterally from the hyphae, erect,<br />

straight or slightly flexuose, tapering towards the apex, pale brown, rough-walled, up<br />

to 15 μm long, 2-5 μm wide, with whorls of phialides on the upper part. Phialides<br />

flask-shaped with a swollen base and a long, narrow neck, hyaline, smooth-walled or<br />

echinulate, 9-15 x 3-4.5 μm in the broadest part. Conidia in long chains, limoniform,<br />

one-celled, pale brown 5-12 x 3-6 μm, smooth to finely echinulate with indistinct spiral<br />

bands. Temperature: optimum 40 O C; maximum 50 O C.<br />

Key Features: Hyphomycete with flask-shaped phialides producing long chains of<br />

one-celled, limoniform, pale brown conidia, with indistinct spiral bands. Note: In A.<br />

levis the conidia lack spiral bands and the phialides have verruculose walls.<br />

Molecular Identification: D1/D2, ITS, 18S and β-tubulin sequences have been<br />

reported by Sandoval-Denis et al. (2015) and Zhang et al (2015a).<br />

References: Domsch et al. (2007), de Hoog et al. (2000, 2015), Al-Mohsen et al.<br />

(2000), Guarro et al. (2007), Sandoval-Denis et al. (2015), Zhang et al. (2015a).<br />

b<br />

10 μm<br />

a<br />

Acrophialophora fusispora (a) culture and (b)<br />

phialides and conidia with spiral striations (arrows).<br />

Antifungal Susceptibility: A. fusispora data from about 40 isolates (Sandoval-<br />

Denis et al. 2015 and Australian National data); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 1-32 16 VORI 0.06-0.5 0.25<br />

ITRA 0.125-4 1 POSA 0.25-1 1


Descriptions of Medical Fungi 3<br />

A ubiquitous genus containing common saprophytes in soil and air, and plant pathogens.<br />

A. infectoria is the most common clinical species (Pastor and Guarro, 2008). Although<br />

usually seen as saprophytic contaminants, Alternaria species in particular A. alternata<br />

and A. infectoria are recognised causative agents of subcutaneous phaeohyphomycosis<br />

and mycotic keratitis. They are a rare cause of onychomycosis, usually following trauma<br />

to the nail.<br />

RG-1 organisms.<br />

Alternaria Nees ex Fries<br />

Morphological Description: Colonies are fast growing, black to olivaceous-black or<br />

greyish, and are suede-like to floccose. Microscopically, branched acropetal chains<br />

(blastocatenate) of multicellular conidia (dictyoconidia) are produced sympodially from<br />

simple, sometimes branched, short or elongate conidiophores. Conidia are obclavate,<br />

obpyriform, sometimes ovoid or ellipsoidal, often with a short conical or cylindrical beak,<br />

pale brown, smooth-walled or verrucose. Temperature: optimum 25-28 O C; maximum<br />

31-32 O C.<br />

Molecular Identification: Multilocus genotype studies have shown the Alternaria<br />

complex currently comprises nine genera and eight Alternaria sections (Woudenbert et<br />

al. 2013). ITS sequencing is sufficient for genus and usually species level identification<br />

and can clearly differentiate A. alternata and A. infectoria (Pastor and Guarro, 2008).<br />

However, it is estimated that >14% of GenBank sequences of Alternaria species<br />

are misclassified, so unknown sequences should be compared to those of wellcharacterised<br />

reference strains (Woudenberg et al. 2013).<br />

Comments: Alternaria species soon lose their ability to sporulate in culture. Potato<br />

dextrose agar and cornmeal agar are the most suitable media to use, and incubation<br />

under ultra-violet light is recommended to maintain sporulation.<br />

Key Features: Dematiaceous hyphomycete producing chains of darkly pigmented,<br />

ovoid to obclavate dictyoconidia, often with short conical or cylindrical beaks.<br />

References: Simmons (1967, 2007), Ellis (1971), Domsch et al. (2007), Samson et al.<br />

(1995), de Hoog et al. (2000, 2015), Pryor and Gilbertson (2000), de Hoog and Horre<br />

(2002), Pastor and Guarro (2008), Woudenberg et al. (2013).<br />

Antifungal Susceptibility: Alternaria spp. (Australian National data); MIC µg/mL.<br />

No 8<br />

AmB 10 3 1 4 2<br />

VORI 10 1 3 5 1<br />

POSA 9 2 1 5 1<br />

ITRA 10 1 1 2 5 1


4<br />

Descriptions of Medical Fungi<br />

Alternaria Nees ex Fries<br />

Alternaria alternata colonies are black to olivaceous-black or<br />

greyish, and are suede-like to floccose.<br />

20 μm<br />

Alternaria alternata showing branched acropetal chains and multicelled,<br />

obclavate to obpyriform conidia with short conical beaks.


Descriptions of Medical Fungi 5<br />

Aphanoascus fulvescens is a soil-borne keratinolytic ascomycete that occasionally<br />

causes dermatomycosis in humans and animals.<br />

RG-2 organism.<br />

Aphanoascus fulvescens (Cooke) Apinis<br />

Morphological Description: Colonies are moderately fast growing, white to tan with<br />

the production of numerous spherical, pseudoparenchymatous, buff to light brown<br />

cleistothecia (non-ostiolate ascocarps). Asci are subspherical to ellipsoidal and eightspored.<br />

Ascospores light brown, yellowish to pale brown in mass, irregularly reticulate,<br />

lens-shaped, 3.5-4.7 x 2.5-3.5 µm. Aphanoascus fulvescens has a Chrysosporium<br />

anamorph showing typical pyriform to clavate-shaped conidia with truncated bases,<br />

15-17.5 x 3.7-6 µm, which are formed either intercalary, laterally or terminally.<br />

Molecular Identification: ITS sequencing will differentiate most species. The<br />

calmodulin gene may also be useful (Cano et al. 2002, Halliday et al. 2015).<br />

Key Features: Keratinolytic, cleistothecia, and a Chrysosporium anamorph.<br />

References: Domsch et al. (2007), McGinnis (1980), de Hoog et al. (2000, 2015),<br />

Cano and Guarro (1990), Cano et al. (2002).<br />

a<br />

b<br />

100 μm<br />

c<br />

10 μm<br />

Aphanoascus flavescens (a) culture, (b) cleistothecium and (c) conidia.


6<br />

Descriptions of Medical Fungi<br />

Apophysomyces complex<br />

Historically the genus Apophysomyces was considered to be monotypic and A. elegans<br />

was reported to be an important human pathogen in immunocompetent patients<br />

following traumatic implantation. A phylogenetic revision of the genus has identified<br />

three additional species, A. ossiformis, A. trapeziformis and A. variabilis. Many isolates<br />

previously identified as A. elegans, now appear to be A. variabilis (Alvarez et al. 2010a).<br />

Molecular identification is required to accurately differentiate these species.<br />

Morphological characteristics overlap so identification and reporting of Apophysomyces<br />

complex is recommended for most diagnostic laboratories.<br />

Molecular Identification: The ITS region and D1/D2 domain may provide for accurate<br />

species identification (Halliday et al. 2015). ITS restriction fragment length polymorphism<br />

analysis has also been described (Chakrabarti et al. 2003).<br />

RG-2 organism.<br />

Apophysomyces elegans Misra, Srivastava & Lata<br />

Morphological Description: Colonies are fast growing, white, becoming brownish<br />

grey with age, downy with no reverse pigment, and are composed of broad, sparsely<br />

septate (coenocytic) hyphae. Sporangiophores are unbranched, straight or curved,<br />

slightly tapering towards the apex, up to 540 µm long, 3-6 µm in width near the<br />

apophysis, and hyaline when young but developing a light to dark brown pigmentation<br />

and a conspicuous subapical thickening 10-16 µm below the apophysis with age.<br />

Sporangiophores arise at right angles from the aerial hyphae and often have a septate<br />

basal segment resembling the “foot cell” commonly seen in Aspergillus. Rhizoids are<br />

thin-walled, subhyaline and predominantly unbranched. Sporangia are multispored,<br />

small (20-58 µm diameter), typically pyriform in shape, hyaline at first, sepia-coloured<br />

when mature, with distinct apophyses and columellae. Columellae are hemispherical<br />

in shape and the apophyses are distinctly funnel or bell-shaped. Sporangiospores are<br />

smooth-walled, subspherical to cylindrical, (5-8 x 4-6 µm), subhyaline to sepia in mass.<br />

Good growth at 26 O C, 37 O C and 42 O C.<br />

RG-2 organism.<br />

Apophysomyces variabilis Alvarez et al.<br />

Morphological Description: Colonies are fast growing, whitish with scarce aerial<br />

mycelium and no reverse pigment. Sporangiophores are erect, generally arising singly,<br />

unbranched, slightly tapering towards the apex, up to 100-400 µm long, 2-3.5 µm in<br />

width near the apophysis, hyaline when young but developing a light greyish brown<br />

pigmentation with age. Sporangia are multispored, small (15-50 µm diameter), typically<br />

pyriform in shape, hyaline at first, sepia-coloured when mature, with distinct apophyses<br />

and columellae. Columellae are hemispherical in shape and the apophyses are short<br />

and distinctly funnel-shaped. Sporangiospores are smooth-walled, variable in shape,<br />

trapezoid, ellipsoidal, sub-triangular or claviform, (5-14 x 3-6 µm), subhyaline to sepia<br />

in mass. Good growth at 26 O C, 37 O C and 42 O C.


Descriptions of Medical Fungi 7<br />

Apophysomyces complex<br />

10 μm<br />

a<br />

10 μm<br />

b<br />

Apophysomyces elegans/variabilis (a) young, multispored, pyriform shaped sporangium<br />

showing a typical funnel-shaped apophysis but without the subapical thickening of<br />

a more mature sporangiophore, and (b) mature sporangium showing distinct funnelshaped<br />

apophyses, columellae, and a conspicuous pigmented subapical thickening<br />

which constricts the lumen of the sporangiophore below the apophysis (arrow).<br />

Sporangiospores are smooth-walled, oblong and subhyaline.


8<br />

Descriptions of Medical Fungi<br />

Apophysomyces complex<br />

Comment: Apophysomyces complex is readily distinguished from other zygomycetes,<br />

especially the morphologically similar, strongly apophysate pathogen Lichtheimia<br />

corymbifera, by having sporangiophores with distinctive funnel or bell-shaped<br />

apophyses and hemispherical-shaped columellae. In addition, there is a conspicuous<br />

pigmented subapical thickening, which constricts the lumen of the sporangiophore<br />

below the apophysis, and distinctive foot cells.<br />

Laboratory identification of this fungus may be difficult or delayed because of the mould’s<br />

failure to sporulate on primary isolation media or on subsequent subculture onto potato<br />

dextrose agar. Sporulation may be stimulated by the use of nutrient deficient media,<br />

like cornmeal-glucose-sucrose-yeast extract agar, Czapek Dox agar, or by using the<br />

agar block method described by Ellis and Ajello (1982) and Ellis and Kaminski (1985).<br />

Molecular-based identification is particularly helpful for the definitive identification of<br />

poorly sporulating cultures.<br />

Key Features: Soil fungus with a tropical to subtropical distribution. Characteristic<br />

“cocktail glass” apophysate sporangial morphology with conspicuous subapical<br />

thickening of the sporangiophore. Resistance to cycloheximide. Rapid growth at 42 O C,<br />

no growth at 50 O C.<br />

References: Misra et al. (1979), Ellis and Ajello (1982), Padhye and Ajello (1988),<br />

Wieden et al. (1985), Lawrence et al. (1986), Cooter et al. (1990), Holland (1997), de<br />

Hoog et al. (2000, 2015), Ellis (2005b), Alvarez et al. (2010a), Chakrabarti et al. (2003,<br />

2010), Guarro et al. (2011).<br />

Antifungal Susceptibility: A. variabilis limited data (Espinel-Ingroff et al. 2015a,<br />

and Australian National data); MIC µg/mL.<br />

No 64<br />

AmB 10 1 1 3 5<br />

POSA 10 1 1 7 1<br />

A. variabilis data from 20 isolates (Alvarez et al. 2010 and Chakrabarti et al. 2010);<br />

MIC µg/mL.<br />

AmB Range 0.5-4; MIC 90<br />

= 2 VORI Range 8-16; MIC 90<br />

= 16<br />

ITRA Range 0.25-2; MIC 90<br />

= 2 POSA Range 0.5-2; MIC 90<br />

= 1


Descriptions of Medical Fungi 9<br />

Arthroderma insingulare Padhye and Carmichael<br />

Synonymy: Trichophyton terrestre Durie and Frey.<br />

Arthroderma insingulare is a geophilic fungus of worldwide distribution which may<br />

occur as a saprophytic contaminant on humans and animals. Durie and Frey (1957)<br />

first described this soil fungus as Trichophyton terrestre from New South Wales,<br />

Australia. Since then T. terrestre has been described as an anamorph of three different<br />

species of Arthroderma; A. insingulare, A. lenticulare and A. quadrifidum (Padhye<br />

and Carmichael, 1972). However, ITS and D1/D2 sequencing of the original isolates<br />

obtained from the Mycology Laboratory at Royal North Shore Hospital, Sydney has<br />

now identified this fungus as Arthroderma insingulare. RG-1 organism.<br />

Morphological Description: Colonies are usually flat to downy with a suede-like to<br />

granular texture resembling T. mentagrophytes. The surface colour may range from<br />

white to cream, buff to yellow, or greenish-yellow. Reverse pigmentation is usually<br />

yellowish-brown although some variants have a deep rose red reverse. Microconidia<br />

are large, clavate or pedicellate, usually exhibiting transition forms to more or less<br />

abundant lateral macroconidia. Macroconidia are clavate to cylindrical with rounded<br />

ends, smooth and thin-walled, and are two to six-celled. Chlamydospores, hyphal<br />

spirals, racquet mycelium and antler hyphae may also be present. No growth at 37 O C.<br />

Molecular Identification: ITS and D1/D2 sequencing is recommended for definitive<br />

identification of isolates.<br />

a<br />

b<br />

20 µm<br />

Arthroderma insingulare (a) culture and (b) macroconidia.


10<br />

Descriptions of Medical Fungi<br />

Synonymy: Trichophyton ajelloi (Vanbreuseghem) Ajello.<br />

Arthroderma uncinatum is a geophilic fungus with a worldwide distribution which may<br />

occur as a saprophytic contaminant on humans and animals but infections are doubtful.<br />

Not known to invade hair in vivo, but produces hair perforations in vitro.<br />

RG-1 organism.<br />

Arthroderma uncinatum Dawson & Gentles<br />

Morphological Description: Colonies are usually flat, powdery, cream, tan to orangetan<br />

in colour, with a blackish-purple submerged fringe and reverse. Macroconidia<br />

are numerous, smooth, thick-walled, elongate, cigar-shaped, 29-65 x 5-10 µm, and<br />

multiseptate with up to nine or ten septa. Microconidia are usually absent, but when<br />

present are ovate to pyriform in shape.<br />

Key Features: Culture characteristics, macroconidial morphology, urease positive and<br />

good growth on Sabouraud’s 5% salt agar.<br />

Molecular Identification: ITS sequencing recommended (Gräser et al. 2008).<br />

References: Rebell and Taplin (1970), Rippon (1988), de Hoog et al. (2015, 2016).<br />

a<br />

b<br />

20 µm<br />

Arthroderma uncinatum (a) culture and (b) macroconidia.


Descriptions of Medical Fungi 11<br />

Arthrographis kalrae (Tewari & Macpherson) Sigler & Carmichael<br />

Arthrographis is an arthroconidial mould comprising four species: A. kalrae, A. lignicola,<br />

A. pinicola and A. alba. These fungi are commonly found in environmental samples<br />

(soil, wood, air and water), but are isolated rarely from clinical specimens (Sandoval-<br />

Denis et al. 2014a).<br />

RG-1 organism.<br />

Morphological Description: Colonies with slow to moderate growth, creamy white<br />

to tan-coloured. Initially, yeast-like then developing hyphal growth and conidiophores<br />

subhyaline, narrow, branched, often in bundles, occasionally forming whitish, large 0.5<br />

cm, linear synnemata. Arthroconidia are one-celled, hyaline, smooth-walled, oblong to<br />

cylindrical, with truncate ends, 2.5-9x 1-2 µm. Spherical blastoconidia 2-4 x 2-3 µm,<br />

may also be formed laterally and sessile on undifferentiated hyphae. Chlamydospores<br />

may also be present. Very rarely immature ascomata submerged in the agar are<br />

produced. Growth at 42 O C, and on media containing cycloheximide.<br />

Molecular Identification: ITS and D1/D2 sequencing may be used for accurate<br />

species identification (Sugiura and Hironaga 2010, Halliday et al. 2015).<br />

Key Features: Keratinolytic, in vitro hair perforation positive, growth at 37 O C and<br />

tolerance to cycloheximide.<br />

References: Sugiura and Hironaga (2010), Giraldo et al. (2014), Sandoval-Denis et<br />

al. (2014a), de Hoog et al. (2015).<br />

10 μm<br />

10 μm<br />

a b b<br />

Arthrographis kalrae (a) culture and (b) arthroconidia.<br />

Antifungal Susceptibility: A. kalrae (Australian National data); MIC µg/mL.<br />

No 8<br />

AmB 7 1 5 1<br />

VORI 7 1 3 2 1<br />

POSA 7 1 2 2 2<br />

ITRA 7 1 4 2


12<br />

Descriptions of Medical Fungi<br />

Aspergillus Micheli ex Link<br />

Aspergillus is a very large genus containing about 250 species, which are currently<br />

classified into seven subgenera that are in turn subdivided into several sections<br />

comprised of related species (Raper and Fennell 1965, Gams et al. 1985, Geiser<br />

et al. 2007). Traditionally, clinical microbiology laboratories have relied heavily<br />

on morphology-based identification methods to differentiate Aspergillus species.<br />

However many species, especially members of the section Fumigati have overlapping<br />

morphological characteristics, which has allowed several genetically distinct species<br />

to be misidentified (Balajee et al. 2005, 2007). This has led to the clustering of species<br />

with overlapping morphologies into “species complexes”, so that laboratories may<br />

report more accurately morphology-based identifications.<br />

Identification of clinical isolates of Aspergillus to species level may be important given<br />

that different species have variable susceptibilities to multiple antifungal drugs. For<br />

example, in vitro and in vivo studies have demonstrated that A. terreus isolates are<br />

largely resistant to the antifungal drug amphotericin B, A. ustus isolates appear to<br />

be refractory to azoles, and A. lentulus and Petromyces alliaceus have low in vitro<br />

susceptibilities to a wide range of antifungals including amphotericin B, azoles, and<br />

echinocandins (Balajee et al. 2005, 2007).<br />

Molecular Identification: Recommended barcoding gene: β-tubulin. General criteria<br />

for identification were outlined by Balajee et al. (2007). Phylogenetic relationships of<br />

the entire genus were presented by Wang et al. (1999) and Peterson (2000, 2008).<br />

MALDI-T<strong>OF</strong> MS: A comprehensive ‘in-house’ database of reference spectra allows<br />

accurate identification of species of Aspergillus even within complexes e.g. A. fumigatus<br />

sensu stricto and A. lentulus (Lau et al. 2013, Sleiman et al. 2015).<br />

a<br />

b<br />

c<br />

d<br />

Four species in the Aspergillus fumigatus complex showing overlapping morphological<br />

characteristics; (a) Aspergillus fumigatus, (b) Aspergillus lentulus, (c) Neosartorya<br />

fischeri and (d) Aspergillus felis.


Descriptions of Medical Fungi 13<br />

Aspergillus Micheli ex Link<br />

Morphological Description: Colonies are usually fast growing, white, yellow, yellowbrown,<br />

brown to black or shades of green, mostly consisting of a dense felt of erect<br />

conidiophores. Conidiophores terminate in a vesicle covered with either a single<br />

palisade-like layer of phialides (uniseriate) or a layer of subtending cells (metulae)<br />

which bear small whorls of phialides (the biseriate structure). The vesicle, phialides,<br />

metulae (if present) and conidia form the conidial head. Conidia are one-celled, smooth<br />

or rough-walled, hyaline or pigmented, are produced in long dry chains which may be<br />

divergent (radiate) or aggregated in compact columns (columnar). Some species may<br />

produce Hülle cells or sclerotia.<br />

For morphological identification, isolates are usually inoculated at three points on<br />

Czapek Dox agar and 2% malt extract agar and incubated at 25 O C. Most species<br />

sporulate within 7 days. Descriptions are primarily based on colony pigmentation and<br />

morphology of the conidial head. Microscopic mounts are best made using cellotape<br />

flag or slide culture preparations mounted in lactophenol cotton blue. A drop of alcohol<br />

is usually needed to remove bubbles and excess conidia.<br />

Key Features: Hyaline hyphomycete showing distinctive conidial heads with flaskshaped<br />

phialides arranged in whorls on a vesicle.<br />

References: Raper and Fennell (1965), Domsch et al. (1980), McGinnis (1980),<br />

Onions et al. (1981), Samson and Pitt (1990, 2000), Samson et al. (1995), Samson<br />

(1979), Vanden Bossche et al. (1988), Klich (2002), Steinbach et al. (2005), Samson<br />

et al. (2011a, 2014), de Hoog et al. (2000, 2015).<br />

conidia<br />

phialides<br />

vesicle<br />

metulae<br />

stipe<br />

a<br />

b<br />

Conidial head morphology in Aspergillus (a) uniseriate, (b) biseriate.


14<br />

Descriptions of Medical Fungi<br />

Aspergillus flavus complex<br />

Aspergillus section Flavi historically includes species with conidial heads in shades<br />

of yellow-green to brown and dark sclerotia. Hedayati et al. (2007) reviewed the A.<br />

flavus complex and included 23 species or varieties, including two sexual species,<br />

Petromyces alliaceus and P. albertensis. Several species of section Flavi produce<br />

aflatoxins, among which aflatoxin B1 is the most toxic of the many naturally occurring<br />

secondary metabolites produced by fungi. Aflatoxins are mainly produced by A. flavus<br />

and A. parasiticus, which coexist and grow on almost any crop or food (Varga et al.<br />

2011). Within the complex, A. flavus is the principle medically important pathogen of<br />

both humans and animals. However, some other species in the A. flavus complex,<br />

notably A. oryzae, A. avenaceus, A. tamari, A. alliaceus and A. nomius, may cause<br />

rare mostly superficial infections (Hedayati et al. 2007, de Hoog et al. 2015).<br />

Note: Accurate species identification within A. flavus complex remains difficult due<br />

to overlapping morphological and biochemical characteristics. For morphological<br />

identifications, it is recommended to report as Aspergillus flavus complex.<br />

Molecular Identification: ITS sequence analysis is sufficient to identify to species<br />

complex level only. Definitive identification requires analysis of β-tubulin, calmodulin<br />

and actin genes (Samson et al. 2007, Balajee et al. 2005a).<br />

Aspergillus flavus Link ex Grey<br />

Aspergillus flavus has a worldwide distribution and normally occurs as a saprophyte in soil<br />

and on many kinds of decaying organic matter, however, it is also a recognised pathogen<br />

of humans and animals. It is a causative agent of otitis, keratitis, acute and chronic<br />

invasive sinusitis, and pulmonary and systemic infections in immunocompromised<br />

patients. A. flavus is second only to A. fumigatus as the cause of human invasive<br />

aspergillosis (Hedayati et al. 2007).<br />

RG-2 organism.<br />

Morphological Description: On Czapek Dox agar, colonies are granular, flat, often<br />

with radial grooves, yellow at first but quickly becoming bright to dark yellow-green with<br />

age. Conidial heads are typically radiate, later splitting to form loose columns (mostly<br />

300-400 µm in diameter), biseriate but having some heads with phialides borne directly<br />

on the vesicle (uniseriate). Conidiophore stipes are hyaline and coarsely roughened,<br />

often more noticeable near the vesicle. Conidia are globose to subglobose (3-6 µm in<br />

diameter), pale green and conspicuously echinulate. Some strains produce brownish<br />

sclerotia.<br />

Key Features: Spreading yellow-green colonies, rough-walled stipes, mature vesicles<br />

bearing phialides over their entire surface and conspicuously echinulate conidia.<br />

Antifungal Susceptibility: A. flavus complex (Australian National data); MIC µg/<br />

mL.<br />

No 16<br />

AmB 68 1 5 7 30 22 3<br />

VORI 68 1 1 6 25 24 11<br />

POSA 57 2 1 5 16 26 7<br />

ITRA 68 1 3 11 43 10


Descriptions of Medical Fungi 15<br />

Aspergillus flavus Link ex Grey<br />

a<br />

b<br />

10 μm<br />

b<br />

10 μm<br />

b<br />

10 μm<br />

Aspergillus flavus (a) culture and (b) conidial heads.<br />

Note: Rough-walled stipe near vesicle (arrow) and both<br />

uniseriate and biseriate conidial heads may be present.


16<br />

Descriptions of Medical Fungi<br />

Aspergillus fumigatus complex<br />

Aspergillus section Fumigati includes species characterised by uniseriate aspergilla<br />

with columnar conidial heads in shades of blue-green and flask-shaped vesicles<br />

(Raper and Fennell, 1965). Teleomorphic species belonging to the “Aspergillus fischeri<br />

series” of the A. fumigatus group (Raper and Fennell, 1965) were placed in the genus<br />

Neosartorya (family Trichocomaceae) by Malloch and Cain (1972). Section Fumigati<br />

includes more than 23 Neosartorya species and 10 anamorphic species (Samson et<br />

al. 2007).<br />

Although A. fumigatus is recognised as the major human pathogen within the complex,<br />

recent phylogenetic studies have demonstrated that some human and animal<br />

infections may be caused by A. lentulus, A. fumigatiaffinis, A. fumisynnematus, A. felis,<br />

Neosartorya fischeri, N. pseudofischeri, N. udagawae, N. hiratsukae and N. spinosa<br />

(Coriglione et al. 1990; Summerbell et al. 1992; Padhye et al. 1994a; Lonial et al. 1997;<br />

Jarv et al. 2004; Balajee et al. 2005, 2006; Barrs et al. 2013).<br />

Aspergillus felis Barrs, van Doorn, Varga & Samson<br />

Aspergillus felis has been reported as a causative agent of invasive aspergillosis and<br />

rhinosinusitis in humans, dogs and cats. Disease in all host species is often refractory<br />

to aggressive antifungal therapeutic regimens.<br />

RG-1 organism.<br />

Morphological Description: Colonies of A. felis are suede-like to floccose, white with<br />

interspersed grey green patches of conidia (conidiation is slow to poor). Conidial heads<br />

of A. felis are short, columnar and uniseriate. Conidiophore stipes are smooth-walled<br />

and vesicles are usually subglobose in shape. Conidia globose (2-3 µm in diameter),<br />

smooth to finely roughened.<br />

Molecular Identification: A. felis can be distinguished from other members of the<br />

section Fumigati by sequence analysis of β-tubulin, calmodulin and actin genes (Barrs<br />

et al. 2013). ITS sequencing is not recommended.<br />

Comment: A. felis is phenotypically similar to Aspergillus viridinutans, but differs by<br />

its ability to grow at 45°C. This species is phylogenetically related to Neosartorya<br />

aureola and N. udagawae and differs to N. aureola in having a heterothallic mode of<br />

reproduction.<br />

Antifungal Susceptibility: A. felis (Barrs et al. 2013); MIC µg/mL.<br />

No 16<br />

AmB 13 1 11 1<br />

VORI 13 1 3 1 5 3<br />

POSA 13 4 3 2 1 2 1<br />

ITRA 13 1 2 2 4 1 3


Descriptions of Medical Fungi 17<br />

Aspergillus felis Barrs, van Doorn, Varga & Samson<br />

a<br />

b<br />

10 μm<br />

b<br />

10 μm<br />

RG-2 organism.<br />

Aspergillus felis (a) culture and (b) conidial head morphology.<br />

Aspergillus fumigatus Fresenius<br />

Morphological Description: On Czapek Dox agar, colonies are typically blue-green<br />

with a suede-like surface consisting of a dense felt of conidiophores. Conidial heads<br />

are typically columnar (up to 400 x 50 µm but often much shorter and smaller) and<br />

uniseriate. Conidiophore stipes are short, smooth-walled and have conical-shaped<br />

terminal vesicles which support a single row of phialides on the upper two thirds of<br />

the vesicle. Conidia are produced in basipetal succession forming long chains and are<br />

globose to subglobose (2.5-3.0 µm in diameter), green and finely roughened. Note:<br />

This species is thermotolerant with a maximum growth temperature of 55 O C.<br />

Key Features: Uniseriate and columnar conidial heads with the phialides limited to the<br />

upper two thirds of the vesicle and curving to be roughly parallel to each other.<br />

Molecular Identification: Sequence analysis of ITS is sufficient to identify to species<br />

complex level only. For definitive identification analysis, β-tubulin, calmodulin and actin<br />

genes is required (Samson et al. 2007; Balajee et al. 2005).


18<br />

Descriptions of Medical Fungi<br />

Aspergillus fumigatus Fresenius<br />

a<br />

b<br />

10 μm<br />

b<br />

10 μm b<br />

10 μm<br />

Aspergillus fumigatus (a) culture and (b) conidial head morphology.<br />

Note: Uniseriate row of phialides on the upper two thirds of the vesicle.<br />

Antifungal Susceptibility: A. fumigatus complex (Australian National data); MIC<br />

µg/mL.<br />

No 16<br />

AmB 523 2 40 122 167 130 58 3 1<br />

VORI 486 1 1 27 112 259 54 11 13 5 3<br />

POSA 415 7 27 50 69 162 79 13 7 0 0 1<br />

ITRA 523 1 6 17 41 74 244 115 15 5 1 4<br />

ANID 249 5 170 52 18 1 1 2<br />

MICA 249 91 95 48 12 2 1<br />

CAS 264 2 22 91 106 24 12 1 1 5


Descriptions of Medical Fungi 19<br />

Aspergillus lentulus Balajee & Marr<br />

Aspergillus lentulus appears to be widely distributed in soil and is now well documented<br />

as a causative agent of invasive aspergillosis in immunosuppressed patients. It is part<br />

of the A. fumigatus complex.<br />

RG-2 organism.<br />

Morphological Description: Colonies of A. lentulus are suede-like to floccose, white<br />

with interspersed grey-green patches of conidia (conidiation is slow to poor in most<br />

strains). Conidial heads are short, columnar and uniseriate. Conidiophore stipes are<br />

smooth-walled, sometimes sinuous and are often constricted at the neck. Vesicles<br />

are usually subglobose in shape. Conidia globose to broadly ellipsoidal (2-3.2 µm in<br />

diameter), smooth to finely roughened.<br />

Molecular Identification: A. lentulus can be distinguished from other members of<br />

the section Fumigati by sequence analysis of β-tubulin, calmodulin and actin genes<br />

(Samson et al. 2007, Balajee et al. 2005b). ITS sequencing is not recommended.<br />

a<br />

b<br />

10 μm<br />

b<br />

10 μm<br />

Aspergillus lentulus (a) culture and (b) conidial head morphology.


20<br />

Descriptions of Medical Fungi<br />

Aspergillus lentulus Balajee & Marr<br />

Antifungal Susceptibility: A. lentulus (Australian National data); MIC µg/mL.<br />

No 16<br />

AmB 5 1 1 3<br />

VORI 5 3 2<br />

POSA 5 1 1 3<br />

ITRA 5 2 1 1 1<br />

Neosartorya fischeri (Wehmer) Malloch & Cain<br />

Neosartorya fischeri is mostly found in canned foodstuffs and is now documented as a<br />

causative agent of invasive aspergillosis in immunosuppressed patients.<br />

RG-1 organism.<br />

Morphological Description: Colonies of N. fischeri are suede-like to floccose, white<br />

to pale yellow with slow to poor conidiation. Conidial heads are short, columnar and<br />

uniseriate. Conidiophore stipes are smooth-walled and vesicles are usually subglobose<br />

to flask-shaped. Conidia globose to subglobose (2-2.5 µm in diameter), smooth to<br />

finely roughened. Good growth at 37 O C.<br />

a<br />

b<br />

b<br />

10 μm<br />

10 μm<br />

Neosartorya fischeri (a) culture and<br />

(b) conidial head morphology.


Descriptions of Medical Fungi 21<br />

Neosartorya fischeri (Wehmer) Malloch & Cain<br />

Molecular Identification: N. fischeri can be distinguished from other members of<br />

the section Fumigati by sequence analysis of β-tubulin, calmodulin and actin genes<br />

(Samson et al. 2007; Balajee et al. 2005b). ITS sequencing is not recommended.<br />

Antifungal Susceptibility: N. fischeri (Australian National data); MIC µg/mL.<br />

No 16<br />

AmB 8 1 5 2<br />

VORI 8 1 3 3 1<br />

POSA 8 4 1 1 2<br />

ITRA 8 1 2 2 3<br />

Aspergillus nidulans complex<br />

Aspergillus subgenus Nidulantes; Gams et al. (1985) includes species with biseriate<br />

conidial heads, brown pigmented often short stipes, and green conidia. Cleistothecia<br />

are soft-walled, surrounded by Hülle cells, and ascospores are red to purple in colour.<br />

Section Nidulantes is one of the largest subgenera of the genus Aspergillus, and<br />

includes about 80 species. Several species have been reported as medical pathogens<br />

principally Aspergillus nidulans, but also A. sydowii, A. unguis, A. rugulovalvus and A.<br />

tetrazonus.<br />

Molecular Identification: ITS sequencing is sufficient to identify to species complex<br />

only. A. nidulans can be distinguished from other members of the section Nidulantes<br />

by sequence analysis of β-tubulin, calmodulin and actin genes.<br />

Aspergillus nidulans (Eidam) Wint.<br />

Aspergillus nidulans is a typical soil fungus with a worldwide distribution, it has also<br />

been reported to cause disease in human and animals.<br />

RG-1 organism.<br />

Morphological Description: On Czapek Dox agar, colonies are typically plain green<br />

in colour with dark red-brown cleistothecia developing within and upon the conidial<br />

layer. Reverse may be olive to drab-grey or purple-brown. Conidial heads are short,<br />

columnar (up to 70 x 30 µm in diameter) and biseriate. Conidiophore stipes are usually<br />

short, brownish and smooth-walled. Conidia are globose (3-3.5 µm in diameter) and<br />

rough-walled.<br />

Key Features: Conidial heads are short, columnar and biseriate. Stipes are usually<br />

short, brownish and smooth-walled. Conidia are globose and rough-walled.


22<br />

Descriptions of Medical Fungi<br />

Aspergillus nidulans (Eidam) Wint.<br />

a<br />

b<br />

10 μm<br />

b<br />

20μm<br />

c<br />

20 μm<br />

d<br />

10 μm<br />

Aspergillus nidulans (a) culture and (b) conidial head morphology, (c) cleistothecium<br />

of Emericella nidulans (anamorph A. nidulans) showing numerous reddish-brown<br />

ascospores and (d) thick-walled Hülle cells.<br />

Antifungal Susceptibility: A. nidulans (Australian National data); MIC µg/mL.<br />

No 16<br />

AmB 24 5 10 1 6 2<br />

VORI 23 3 7 13<br />

POSA 20 1 6 6 5 2<br />

ITRA 24 1 3 6 10 4


Descriptions of Medical Fungi 23<br />

The black aspergilli, Aspergillus section Nigri (Gams et al. 1985) includes species<br />

with uniseriate or biseriate conidial heads, spherical to pyriform vesicles, smoothwalled<br />

stipes and black or near black-coloured conidia. This group contains about 26<br />

species with Aspergillus niger being the most common species isolated. A. niger can<br />

be isolated from all continents and is not very selective with respect to environmental<br />

conditions. Other species within this group that have been linked to human and animal<br />

infection include A. acidus, A. aculeatus, A. brasiliensis and A. tubingensis.<br />

Molecular Identification: In Aspergillus section Nigri, all species can be distinguished<br />

from each other using calmodulin sequence data, and all except one can be<br />

distinguished using β-tubulin sequence data. ITS sequencing can only be used for a<br />

rough classification of the uni- and biseriate species (Samson et al. 2007).<br />

Aspergillus niger van Tieghem<br />

Aspergillus niger is one of the most common and easily identifiable species of the genus<br />

Aspergillus, with its white to yellow mycelial culture surface later bearing black conidia.<br />

This species is very commonly found in aspergillomas and is the most frequently<br />

encountered agent of otomycosis. It is also a common laboratory contaminant.<br />

RG-1 organism.<br />

Aspergillus niger complex<br />

Morphological Identification: On Czapek Dox agar, colonies consist of a compact<br />

white or yellow basal felt covered by a dense layer of dark-brown to black conidial<br />

heads. Conidial heads are large (up to 3 mm by 15 to 20 µm in diameter), globose,<br />

dark brown, becoming radiate and tending to split into several loose columns with age.<br />

Conidiophore stipes are smooth-walled, hyaline or turning dark towards the vesicle.<br />

Conidial heads are biseriate with the phialides borne on brown, often septate metulae.<br />

Conidia are globose to subglobose (3.5-5 µm in diameter), dark brown to black and<br />

rough-walled.<br />

Key Features: Conidial heads are dark brown to black, radiate and biseriate with<br />

metulae twice as long as the phialides. Conidia brown and rough-walled.<br />

Antifungal Susceptibility: A. niger (Australian National data); MIC µg/mL.<br />

No 16<br />

AmB 75 16 27 16 13 3<br />

VORI 71 3 5 8 16 28 10 1<br />

POSA 60 1 7 7 15 18 12<br />

ITRA 75 1 3 8 13 34 14 1 1


24<br />

Descriptions of Medical Fungi<br />

Aspergillus niger van Tieghem<br />

a<br />

b<br />

10 μm<br />

Aspergillus niger (a) Culture and (b) conidial head morphology.<br />

Note: Conidial heads are biseriate, large, globose, dark brown,<br />

becoming radiate with the phialides borne on metulae.


Descriptions of Medical Fungi 25<br />

Aspergillus terreus complex<br />

Aspergillus section Terrei (Gams et al. 1985); Aspergillus terreus complex includes<br />

species with biseriate, columnar conidial heads in shades of buff to brown (Raper<br />

and Fennell 1965). The most important species of this section is A. terreus, which is<br />

ubiquitous in the environment (Samson et al. 2011). Two other species have been<br />

reported as medical pathogens, A. alabamensis and A. niveus.<br />

Molecular Identification: A. terreus can be distinguished from other members of<br />

the section Terrei by sequence analysis of β-tubulin, calmodulin and actin genes. ITS<br />

sequencing is sufficient to identify to species complex level only.<br />

Aspergillus terreus Thom<br />

Aspergillus terreus occurs commonly in soil and is occasionally reported as a pathogen<br />

of humans and animals.<br />

RG-2 organism.<br />

Morphological Identification: On Czapek Dox agar, colonies are typically suedelike<br />

and cinnamon-buff to sand-brown in colour with a yellow to deep dirty brown<br />

reverse. Conidial heads are compact, columnar (up to 500 x 30-50 µm in diameter)<br />

and biseriate. Metulae are as long as the phialides. Conidiophore stipes are hyaline<br />

and smooth-walled. Conidia are globose to ellipsoidal (1.5-2.5 µm in diameter), hyaline<br />

to slightly yellow and smooth-walled.<br />

Key Features: Cinnamon-brown cultures, conidial heads biseriate with metulae as<br />

long as the phialides.<br />

References: Raper and Fennell (1965), Domsch et al. (1980), McGinnis (1980),<br />

Onions et al. (1981), Samson and Pitt (1990), Samson et al. (1995), de Hoog et al.<br />

(2000) and Klich (2002).<br />

Antifungal Susceptibility: A. terreus (Australian National data); MIC µg/mL.<br />

No 16<br />

AmB 72 1 3 11 29 24 3 1<br />

VORI 69 4 13 32 16 1 2 1<br />

POSA 59 1 10 8 23 13 2 2<br />

ITRA 72 4 3 9 17 35 4


26<br />

Descriptions of Medical Fungi<br />

Aspergillus terreus Thom<br />

a<br />

b<br />

10 μm<br />

Aspergillus terreus (a) culture and (b) conidial head morphology<br />

Note: Conidial heads are biseriate.


Descriptions of Medical Fungi 27<br />

Aureobasidium pullulans has a worldwide distribution and is usually isolated as<br />

a saprophyte, occasionally from skin and nails. It has also been reported as a rare<br />

causative agent of phaeohyphomycosis, mycotic keratitis and peritonitis in patients on<br />

continuous ambulatory peritoneal dialysis (CAPD).<br />

RG-1 organism.<br />

Aureobasidium pullulans (de Bary) Arnaud<br />

Morphological Description: Colonies are fast growing, smooth, soon covered with<br />

slimy masses of conidia, cream or pink, later becoming brown or black. Hyphae are<br />

hyaline and septate, frequently becoming dark-brown with age and forming chains of<br />

one to two-celled, thick-walled, darkly pigmented arthroconidia. These arthroconidia<br />

actually represent the Scytalidium anamorph of Aureobasidium and are only of<br />

secondary importance in recognising members of this genus. Conidia are produced<br />

synchronously in dense groups from indistinct scars or from short denticles on<br />

undifferentiated, hyaline to subhyaline hyphae. Conidia are hyaline, smooth-walled,<br />

single-celled, ellipsoidal but of variable shape and size (8-12 x 4-6 µm), often with an<br />

indistinct hilum (i.e. a mark or scar at the point of attachment). Temperature: optimum<br />

25 O C; maximum 35-37 O C.<br />

Molecular Identification: Recommended barcoding genes are ITS, EF-1α and D1/D2<br />

(de Hoog et al. 2015).<br />

Key Features: Hyphomycete (so-called black yeast) producing hyaline blastoconidia<br />

simultaneously from the vegetative hyphae, which may also form chains of darkly<br />

pigmented, thick-walled arthroconidia.<br />

References: Hermanides-Nijhof (1977), Domsch et al. (2007), McGinnis (1980), de<br />

Hoog et al. (2000, 2015), Najafzadeh et al. (2014).<br />

20 μm<br />

Aureobasidium pullulans showing one to two-celled, darkly pigmented arthroconidia<br />

and hyaline, single-celled, ovoid-shaped conidia which are produced on short denticles.<br />

Antifungal Susceptibility: A. pullulans data from 108 isolates (Najafzadeh et al.<br />

2014 and Australian National data); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 0.008-16 1 VORI 0.008-16 2<br />

ITRA 0.008-16 0.5 POSA 0.008-4 0.5


28<br />

Descriptions of Medical Fungi<br />

Synonymy: Basidiobolus meristosporus Drechsler.<br />

Basidiobolus heterosporus Srinivasan & Thirumalachar.<br />

Basidiobolus haptosporus Drechsler.<br />

Basidiobolus ranarum is commonly present in decaying fruit and vegetable matter, and<br />

as a commensal in the intestinal tract of frogs, toads and lizards. It has been reported<br />

from tropical regions of Africa and Asia including India, Indonesia and Australia.<br />

RG-2 organism.<br />

Basidiobolus ranarum Eidem<br />

Morphological Description: Colonies are moderately fast growing at 30 O C, flat,<br />

yellowish-grey to creamy-grey, glabrous, becoming radially folded and covered by a fine,<br />

powdery, white surface mycelium. Satellite colonies are often formed by germinating<br />

conidia ejected from the primary colony. Microscopic examination usually shows the<br />

presence of large vegetative hyphae (8-20 µm in diameter) forming numerous round (20-<br />

50 µm in diameter), smooth, thick-walled zygospores that have two closely appressed<br />

beak-like appendages. The production of “beaked” zygospores is characteristic of the<br />

genus. Two types of asexual conidia are formed, although isolates often lose their ability<br />

to sporulate with subculture. Special media incorporating glucosamine hydrochloride<br />

and casein hydrolsate may be needed to stimulate sporulation (Shipton and Zahari,<br />

1987). Primary conidia are globose, one-celled, solitary and are forcibly discharged<br />

from a sporophore. The sporophore has a distinct swollen area just below the conidium<br />

that actively participates in the discharge of the conidium. Secondary (replicative)<br />

conidia are clavate, one-celled and are passively released from a sporophore. These<br />

sporophores are not swollen at their bases. The apex of the passively released spore<br />

has a knob-like adhesive tip. These spores may function as sporangia, producing<br />

several sporangiospores.<br />

References: Strinivasan and Thirumalachar (1965), Greer and Friedman (1966),<br />

Dworzack et al. (1978), McGinnis (1980), King (1983), Rippon (1988), Davis et al.<br />

(1994), Jong and Dugan (2003), de Hoog et al. (2000, 2015) and Ellis (2005a).<br />

20 μm<br />

Basidiobolus ranarum showing thick-walled zygospores.


Descriptions of Medical Fungi 29<br />

Basidiobolus ranarum Eidem<br />

Basidiobolus ranarum culture showing satellite colonies formed<br />

by germinating conidia ejected from the primary colony.<br />

20 μm<br />

Basidiobolus ranarum showing conidia and a sporophore with<br />

a distinct swollen area just below the conidium (arrow).


30<br />

Descriptions of Medical Fungi<br />

Three species are recognised, two of which are well known pathogens of insects.<br />

Beauvaria bassiana is the most common species and is best known as the causal agent<br />

of muscardine disease in silkworms. Beauveria species are occasionally isolated in the<br />

clinical laboratory as saprophytic contaminants. Infections in humans are extremely<br />

rare.<br />

RG-1 organism.<br />

Beauveria Vuillemin<br />

Morphological Description: Colonies are usually slow growing, usually not exceeding<br />

2 cm in ten days at 20 O C, downy, at first white, but later often becoming yellow to<br />

pinkish. The genus Beauveria is characterised by the sympodial development of singlecelled<br />

conidia (ameroconidia) on a geniculate or zig-zag rachis. Conidiogenous cells<br />

are flask-shaped, rachiform, proliferating sympodially and are often aggregated into<br />

sporodochia or synnemata. Conidia are hyaline and globose or ovoid in shape.<br />

Key Features: Hyphomycete showing sympodial development of single-celled conidia<br />

on a geniculate or zig-zag rachis emanating from a flask-shaped conidiophore.<br />

Molecular Identification: Specific primers were developed by Hegedus and<br />

Khachatourians (1996). Full phylogeny of the genus was provided by Rehner and<br />

Buckley (2005). Biogeography of molecular types was characterised by Ghikas et al.<br />

(2010).<br />

MALDI-T<strong>OF</strong> MS: Cassagne et al. (2011) published a standardised procedure for mould<br />

identification in the clinical laboratory.<br />

References: de Hoog (1972), Domsch et al. (2007), McGinnis (1980), de Hoog et al.<br />

(2000, 2015).<br />

20 μm<br />

Beauveria bassiana showing sympodial development of conidia on a geniculate or zigzag<br />

rachis. Conidiogenous cells are flask-shaped, rachiform, proliferating sympodially<br />

and are often aggregated into sporodochia or synnemata. Conidia are hyaline and<br />

globose or ovoid in shape, 2-3 µm diameter (phase contrast image).


Descriptions of Medical Fungi 31<br />

The genus Bipolaris contains about 45 species, which are mostly subtropical and tropical<br />

plant parasites. Recent molecular studies have recognised Bipolaris cynodontis, B.<br />

micropus, and B. setariae as species isolated from clinical samples (da Cunha et al.<br />

2014). However recent phylogenetic studies have transferred several well-documented<br />

human pathogens, notably B. australiensis, B. hawaiiensis and B. spicifera to the genus<br />

Curvularia (Manamgoda et al. 2012)<br />

RG-1 organisms.<br />

Bipolaris Shoemaker<br />

Morphological Description: Colonies are moderately fast growing, effuse, grey to<br />

blackish brown, suede-like to floccose with a black reverse. Microscopic morphology<br />

shows sympodial development of hyaline to deep olivaceous pigmented, pseudoseptate<br />

conidia on a geniculate or zig-zag rachis. Conidia mostly curved, canoe-shaped, fusoid<br />

or obclavate, rarely straight, 2–14 pseudoseptate (usually more than 6), germinating<br />

only from the ends (bipolar).<br />

Key Features: Dematiaceous hyphomycete producing sympodial, pseudoseptate,<br />

pale brown, long slender, gently curving conidia, which are rounded at both ends.<br />

Comment: The genera Drechslera, Bipolaris, Curvularia and Exserohilum are all<br />

closely related. In the past, morphological differentiation of the genera relied upon<br />

a combination of characters including conidial shape, the presence or absence of a<br />

protruding hilum, the contour of the basal portion of the conidium and its hilum, the<br />

point at which the germ tube originates from the basal cell and, to a lesser degree, the<br />

sequence and location of the first three conidial septa.<br />

However, Manamgoda et al. (2012) have found that there is no clear morphological<br />

boundary between genera Bipolaris and Curvularia and some species show<br />

intermediate morphology. These authors recommend using a combined ITS and GPDH<br />

gene analysis for definitive identification of species (Manamgoda et al. 2012).<br />

Molecular Identification: ITS sequencing may be used to identify clinical species (da<br />

Cunha et al. 2012a). GPDH has been determined to be the best single phylogenetic<br />

marker of Bipolaris species (Manamgoda et al. 2012, 2014).<br />

References: Ellis (1971, 1976), Luttrell (1978), Domsch et al. (2007), Alcorn (1983),<br />

McGinnis et al. (1986b), Sivanesan (1987), Rippon (1988), de Hoog et al. (2000, 2015),<br />

Manamgoda et al. (2012, 2014), da Cunha et al. (2012a).


32<br />

Descriptions of Medical Fungi<br />

Blastomyces dermatitidis Gilchrist & Stokes<br />

At present the genus Blastomyces contains two species, Blastomyces dermatitidis<br />

and Blastomyces gilchristi, which are morphologically identical but distinguishable<br />

by sequence analysis of the ITS region (Brown et al. 2013). B. dermatitidis lives in<br />

soil and in association with decaying organic matter such as leaves and wood. It is<br />

the causal agent of blastomycosis a chronic granulomatous and suppurative disease,<br />

having a primary pulmonary stage that is frequently followed by dissemination to other<br />

body sites, typically the skin and bone. Although the disease was long thought to be<br />

restricted to the North American continent, in recent years autochthonous cases have<br />

been diagnosed in Africa, Asia and Europe.<br />

WARNING: RG-3 organism. Cultures of B. dermatitidis represent a biohazard to<br />

laboratory personnel and must be handled in a Class II Biological Safety Cabinet<br />

(BSCII).<br />

Morphological Description: Colonies at 25 O C have variable morphology and growth<br />

rate. They may grow rapidly, producing a fluffy white mycelium or slowly as glabrous, tan,<br />

nonsporulating colonies. Growth and sporulation may be enhanced by yeast extract.<br />

Most strains become pleomorphic with age. Microscopically, hyaline, ovoid to pyriform,<br />

one-celled, smooth-walled conidia (2-10 µm in diameter) of the Chrysosporium type,<br />

are borne on short lateral or terminal hyphal branches.<br />

Colonies on blood agar at 37 O C are wrinkled and folded, glabrous and yeast-like.<br />

Microscopically, the organism produces the characteristic yeast phase seen in tissue<br />

pathology; ie. B. dermatitidis is a dimorphic fungus.<br />

Comment: In the past, conversion from the mould form to the yeast form was<br />

necessary to positively identify this dimorphic pathogen from species of Chrysosporium<br />

or Sepedonium. However, culture identification by exoantigen test and/or molecular<br />

methods is now preferred to minimise manipulation of the fungus.<br />

Key Features: Clinical history, tissue pathology, culture identification by positive<br />

exoantigen test and/or by molecular methods.<br />

a<br />

b<br />

Blastomyces dermatitidis (a) culture and (b) one-celled, smooth-walled<br />

conidia borne on short lateral or terminal hyphal branches.


Descriptions of Medical Fungi 33<br />

Blastomyces dermatitidis Gilchrist & Stokes<br />

10 μm<br />

Histopathology: Blastomyces dermatitidis tissue sections show large, broad-based,<br />

unipolar budding yeast-like cells, which may vary in size from 8-15 µm, with some<br />

larger forms up to 30 µm in diameter. Tissue sections need to be stained by Grocott’s<br />

methenamine silver method to clearly see the yeast-like cells, which are often difficult<br />

to observe in H&E preparations.<br />

Molecular Diagnostics: A DNA probe assay (AccuProbe, Gen-Probe, Inc., San Diego,<br />

CA) for identification of B. dermatitidis in clinical isolates is available (Scalarone et al.<br />

1992 and Padhye et al. 1994b). However this has limited application as it can be used<br />

only with pure cultures of B. dermatitidis (yeast or mould) (Sidamonidze et al. 2012).<br />

Several conventional PCR assays have been developed for the identification of B.<br />

dermatitidis from clinical specimens (Bialek et al. 2003) and soil (Burgess et al. 2006).<br />

Sidamonidze et al. (2012) developed a real-time PCR targeting the BAD1 (formerly<br />

known as WI-1) gene for the identification of B. dermatitidis in culture and tissue and<br />

Morjaria et al. (2015) used rDNA sequencing for identification from paraffin embedded<br />

tissue.<br />

References: McGinnis (1980), Chandler et al. (1980), Kaufman and Standard (1987),<br />

Rippon (1988), Brown et al. (2013).<br />

Antifungal Susceptibility: B. dermatitidis limited data available (Sugar and Liu<br />

1996, Espinel-Ingroff et al. 2001, Espinel-Ingroff 2003, Gonzales et al. 2005 and<br />

Sabatelli et al. 2006). Antifungal susceptibility testing not recommended.<br />

For treatment options see Clinical Practice Guidelines for the Management of<br />

Blastomycosis (Chapman et al. 2008); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

FLU 0.125-64 4-16 AmB 0.03-1 0.5<br />

ITRA 0.03->16 0.125-2 VORI 0.03-16 0.25<br />

POSA 0.03-2 0.125 CAS 0.5-8 2


34<br />

Descriptions of Medical Fungi<br />

Candida Berkhout<br />

The genus Candida is characterised by globose to elongate yeast-like cells or<br />

blastoconidia that reproduce by narrow-based multilateral budding. Pseudohyphae and<br />

occasionally true hyphae may also be present. Colony pigmentation is usually absent.<br />

Ballistoconidia are not formed. Arthroconidia may be formed, but not extensively.<br />

Sexual reproduction is absent. Glucose may be fermented. Nitrate may be assimilated.<br />

Starch-like compounds are not produced. The diazonium blue B reaction is negative.<br />

The genus is highly polyphyletic, as it comprises mitosporic species that are devoid of<br />

special distinguishing features (Lachance et al. 2011).<br />

Recently, several taxonomic rearrangements have been made and many well-known<br />

Candida species have been renamed and moved to other genera, notably Pichia<br />

kudriavzevii (formerly Candida krusei), Meyerozyma guilliermondii (formerly Candida<br />

guilliermondii), Clavispora lusitaniae (formerly Candida lusitaniae), Kluyveromyces<br />

marxianus (formerly Candida kefyr) and Wickerhamomyces anomalus (formerly<br />

Candida pelliculosa). C. glabrata and C. parapsilosis are now recognised as species<br />

complexes (Tavanti 2005; Correia 2006; Alcoba-Florez 2005).<br />

Several species may be aetiological agents, most commonly Candida albicans, followed<br />

by C. parapsilosis, C. glabrata, C. tropicalis and Pichia kudriavzevii. Altogether, these<br />

five species account for >95% of human infections. However a number of other species<br />

may also be isolated. All are ubiquitous and occur naturally on humans.<br />

a<br />

b<br />

10 μm<br />

Candida albicans showing (a) typical cream-coloured, smooth surfaced, waxy<br />

colonies and (b) narrow based budding spherical to ovoid blastoconidia.


Descriptions of Medical Fungi 35<br />

Candida Berkhout<br />

Identification: see Kurtzman, Fell and Boekhout. 2011. The Yeasts, a Taxonomic<br />

Study. 5th Edition Elsevier B.V.<br />

Ensure that you start with a fresh growing pure culture; streak for single colony isolation<br />

if necessary.<br />

Chromogenic Agars are used for primary isolation for rapid species identification and<br />

detection of mixed flora, especially from non-sterile sites. Depending on the brand<br />

of chromogenic media presumptive identification of C. albicans, C. tropicalis and P.<br />

kudriavzevii is possible. It is particularly useful for detection of mixed infections.<br />

CHROMagar Candida plate showing chromogenic colour change for C. albicans<br />

(green), C. tropicalis (blue), C. parapsilosis (white) and C. glabrata (mauve).<br />

Germ Tube Test. A rapid screening test for C. albicans and C. dubliniensis. 0.5 mL<br />

of serum, containing 0.5% glucose, is lightly inoculated with the test organism and<br />

incubated at 35 O C for 2-3 hours. On microscopy, the production of germ tubes by the<br />

cells is presumptive for C. albicans and C. dubliniensis.<br />

10 μm<br />

Candida albicans showing production of germ tubes.


36<br />

Descriptions of Medical Fungi<br />

Candida Berkhout<br />

Dalmau Plate Culture: To set up a yeast morphology plate, dip a flamed sterilised<br />

straight wire into a culture and then lightly scratch the wire onto the surface of a<br />

cornmeal/tween 80, rice/tween 80 or yeast morphology agar plate, then place a flamed<br />

coverslip onto the agar surface covering the scratches. Dalmau morphology plates<br />

are examined in-situ directly under the low power of a microscope for the presence<br />

of pseudohyphae which may take up to 4-5 days at 26 O C to develop. C. albicans<br />

also produces characteristic large, round, terminal, thick-walled vesicles (often called<br />

chlamydospores). For best results a light inoculum should be scratched into the agar<br />

surface using a wire.<br />

10 μm<br />

a<br />

5 mm<br />

b<br />

Dalmau plate culture of Candida albicans showing (a) colonies growing out from<br />

scratches on the surface of a cornmeal/tween 80 agar plate, and (b) the production of<br />

large round, thick-walled chlamydospores. Note: A coverslip has been placed onto the<br />

agar surface covering the scratches.<br />

Physiological and Biochemical Tests: including fermentation and assimilation studies<br />

should be performed based on those used at the Centraalbureau voor Schimmelcultures<br />

(CBS), Delft, The Netherlands (Kurtzman et al. 2011). Reliable commercially available<br />

yeast identification systems are the API 20C AUX, API ID 32C, Biolog YT Station and<br />

Vitek 2 YST ID systems. However, they can only be used to identify those species in<br />

their respective databases, and may misidentify yeasts that are not represented.<br />

Other Supplementary Tests include growth at 37 O C, cycloheximide resistance and<br />

hydrolysis of urea.<br />

MALDI-T<strong>OF</strong> MS: The Bruker MALDI-T<strong>OF</strong> database is useful for identification of most<br />

clinical yeasts. The MALDI-T<strong>OF</strong> Vitek MS has been reported to misidentify some<br />

yeasts, notably Candida metapsilosis as Candida parapsilosis (Nobrega et al. 2014).<br />

Molecular Identification: ITS sequencing is useful for the identification of most clinical<br />

yeasts.<br />

References: Barnett et al. (1983), Kurtzman and Fell (1998, 2011), de Hoog et al.<br />

(2000, 2015).


Descriptions of Medical Fungi 37<br />

Candida albicans is a commensal of mucous membranes and the gastrointestinal<br />

tract. Environmental isolations have been made from sources contaminated by human<br />

or animal excreta, such as polluted water, soil, air and plants.<br />

RG-2 organism.<br />

Candida albicans (Robin) Berkhout<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Spherical to subspherical budding blastoconidia, 2-7 x 3-8 µm in size.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Branched pseudohyphae with dense verticils of blastoconidia.<br />

Spherical chlamydospores, mostly terminal, often on a slightly swollen subtending cell,<br />

are formed near the edge of the cover slip.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube + L-Sorbose v L-Arabinose v D-Glucitol v<br />

Fermentation Sucrose v D-Arabinose v α-M-D-glucoside v<br />

Glucose + Maltose + D-Ribose v D-Gluconate v<br />

Galactose v Cellobiose - L-Rhamnose - DL-Lactate +<br />

Sucrose v Trehalose v D-Glucosamine v myo-Inositol -<br />

Maltose + Lactose - N-A-D-glucosamine v 2-K-D-gluconate +<br />

Lactose - Melibiose - Glycerol v D-Glucuronate -<br />

Trehalose v Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose v Ribitol v Urease -<br />

Glucose + Soluble Starch + Galactitol - 0.1% Cycloheximide +<br />

Galactose + D-Xylose + D-Mannitol + Growth at 40 O C +<br />

Key Features: Germ tube positive, production of chlamydospores on Dalmau plate<br />

culture, fermentation of glucose, sugar assimilation profile and a distinctive green<br />

colour on CHROMagar. Note: Germ tube negative variants (previously known as C.<br />

claussenii), and sucrose-negative variants (previously described as C. stellatoidea)<br />

may occur.<br />

Antifungal Susceptibility: C. albicans (Australian National data); MIC µg/mL.<br />

CLSI clinical breakpoints are marked where available (Pfaller and Diekema 2012).<br />

No 64<br />

AmB 1725 2 16 162 644 548 305 48<br />

FLU 1728 2 2 2 83 468 706 314 62 12 11 18 8 40<br />

VORI 1445 739 404 156 75 18 17 18 5 5 1 1 6<br />

POSA 1095 88 416 367 137 50 18 10 6 1 1<br />

ITRA 1728 20 122 443 661 378 43 30 10 3 1 17<br />

ANID 821 5 280 320 146 64 5 0 0 1<br />

MICA 819 470 261 73 11 4<br />

CAS 1171 3 13 207 490 327 112 17 2<br />

5FC 1728 4 147 765 362 164 171 62 16 7 5 3 3 19


38<br />

Descriptions of Medical Fungi<br />

Synonymy: Candida brumptii (Guerra) Langeron & Guerra.<br />

Although most isolates of Candida catenulata originate from human sources, cases of<br />

candidaemia are uncommon.<br />

RG-1 organism.<br />

Candida catenulata Diddens & Lodder<br />

Culture: Colonies (SDA) white to cream-coloured smooth, soft and wrinkled, yeastlike.<br />

Microscopy: Ovoid to cylindrical budding blastoconidia, 1.5-4.5 x 4-12 µm.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: pseudohyphae consisting of chains of ovoid or cylindroid cells,<br />

and sometimes small verticils of ovoid blastoconidia.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose - L-Arabinose - D-Glucitol v<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose v Maltose v D-Ribose v D-Gluconate v<br />

Galactose -,s Cellobiose - L-Rhamnose - DL-Lactate -<br />

Sucrose - Trehalose v D-Glucosamine v myo-Inositol -<br />

Maltose -,s Lactose - N-A-D-glucosamine + 2-K-D-gluconate v<br />

Lactose - Melibiose - Glycerol + D-Glucuronate -<br />

Trehalose - Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol v Urease -<br />

Glucose + Soluble Starch v Galactitol - 0.1% Cycloheximide +,s<br />

Galactose + D-Xylose v D-Mannitol + Growth at 37 O C v<br />

Key Features: Separation from most physiologically similar species can be<br />

accomplished based on positive growth responses on D-mannitol, D-glucitol and<br />

resistance to 0.1% cycloheximide, combined with negative responses for sorbose and<br />

erythritol utilisation or growth in vitamin-free medium.<br />

Antifungal Susceptibility: C. catenulata (Australian National data); MIC µg/mL.<br />

No 64<br />

AmB 3 2 1<br />

FLU 3 1 1 1<br />

VORI 3 1 1 1<br />

POSA 3 2 1<br />

ITRA 3 1 1 1<br />

ANID 3 1 2<br />

MICA 3 1 2<br />

CAS 3 1 2<br />

5FC 3 3


Descriptions of Medical Fungi 39<br />

Candida dubliniensis Sullivan et al.<br />

Candida dubliniensis is an occasional cause of candidaemia and mucosal infection,<br />

especially in HIV patients.<br />

RG-2 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Spherical to subspherical budding blastoconidia, 3-8 x 2-7 µm in size.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Branched pseudohyphae with dense verticils of blastoconidia<br />

and spherical, mostly terminal chlamydospores.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube + L-Sorbose - L-Arabinose - D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose - M-D-glucoside +,s<br />

Glucose + Maltose + D-Ribose - D-Gluconate -<br />

Galactose +,s Cellobiose - L-Rhamnose - DL-Lactate +<br />

Sucrose - Trehalose s,+ D-Glucosamine v myo-Inositol -<br />

Maltose + Lactose - N-A-D-glucosamine + 2-K-D-gluconate +<br />

Lactose - Melibiose - Glycerol w,s,+ D-Glucuronate -<br />

Trehalose v Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose w,+ Ribitol + Urease -<br />

Glucose + Soluble Starch w,+ Galactitol - 0.1% Cycloheximide +<br />

Galactose + D-Xylose s,+ D-Mannitol + Growth at 40 O C +<br />

Key Features: Germ tube positive, similar to C. albicans, except for absence of growth<br />

at 42 O C; glycerol (mostly +), methyl-a-D-glucoside (-), trehalose (-), and D-xylose (-).<br />

Initial colonies dark green on CHROMagar and producing rough colonies on bird seed<br />

agar. ITS sequencing and MALDI-T<strong>OF</strong> can reliably distinguish C. dubliniensis from C.<br />

albicans.<br />

Antifungal Susceptibility: C. dubliniensis (Australian National data); MIC µg/mL.<br />

No 64<br />

AmB 74 1 1 3 28 30 10 1<br />

FLU 74 24 24 21 5<br />

VORI 71 67 4<br />

POSA 63 15 19 18 8 3<br />

ITRA 74 6 22 14 22 6 4<br />

ANID 49 1 7 12 9 12 2 6<br />

MICA 49 4 11 18 8 2 6<br />

CAS 66 8 23 17 9 1 1 7<br />

5FC 74 12 49 7 4 1 1


40<br />

Descriptions of Medical Fungi<br />

Candida glabrata complex<br />

Recently Candida glabrata has been recognised as a species complex consisting of C.<br />

glabrata, C. bracarensis (Correia et al. 2006) and C. nivariensis (Alcoba-Flórez et al.<br />

2005). These three species are phenotypically indistinguishable and are best identified<br />

by molecular methods. C. bracarensis was described based on PCR-fingerprints and<br />

sequence divergence in the D1/D2 domains (Correia et al. 2006). C. nivariensis was<br />

differentiated from other yeasts on the basis of ITS sequences (Borman et al. 2008).<br />

Candida bracarensis Correia, P. Sampaio, James & Pais<br />

RG-2 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ellipsoidal budding blastoconidia, 3.9-6 x 2-4 µm in size. No pseudohyphae<br />

or chlamydospores produced.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: No pseudohyphae produced.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose - L-Arabinose - D-Glucitol -<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose + Maltose - D-Ribose - D-Gluconate +<br />

Galactose - Cellobiose - L-Rhamnose - DL-Lactate -<br />

Sucrose - Trehalose + D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine - 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol v D-Glucuronate -<br />

Trehalose s Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol - Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose - D-Xylose - D-Mannitol - Growth at 40 O C +<br />

Key Features: C. bracarensis has variable API 20C patterns that overlap with C.<br />

nivariensis and some C. glabrata isolates, and has variable results with a rapid trehalose<br />

assay. Note: C. glabrata produces mauve-coloured colonies on CHROMagar, whereas<br />

isolates of C. bracarensis, C. nivariensis, C. norvegensis and C. inconspicua produce<br />

white colonies on CHROMagar (Alcoba-Flórez et al. 2005, Bishop et al. 2008).<br />

Antifungal Susceptibility: C. bracarensis limited data (Australian National data);<br />

MIC µg/mL.<br />

No 64<br />

AmB 3 1 1 1<br />

FLU 3 2 1<br />

VORI 3 2 1<br />

POSA 3 2 1<br />

ITRA 3 2 1<br />

ANID 3 1 2<br />

MICA 3 1 2<br />

CAS 3 2 1<br />

5FC 3 1 1 1


Descriptions of Medical Fungi 41<br />

Candida glabrata complex<br />

RG-2 organism.<br />

Candida glabrata (Anderson) S.A. Meyer & Yarrow<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ovoid to ellipsoidal budding blastoconidia, 3.4 x 2.0 µm in size. No<br />

pseudohyphae or chlamydospores produced.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Ovoid budding yeast cells only. No pseudohyphae produced.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose - L-Arabinose - D-Glucitol -<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose + Maltose - D-Ribose - D-Gluconate +<br />

Galactose - Cellobiose - L-Rhamnose - DL-Lactate v<br />

Sucrose - Trehalose v D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine - 2-K-D-gluconate v<br />

Lactose - Melibiose - Glycerol v D-Glucuronate -<br />

Trehalose v Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol - Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose - D-Xylose - D-Mannitol - Growth at 40 O C +<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern. C. glabrata<br />

is a common yeast species found on the body surface. Approximately 10% of clinical<br />

isolates show azole cross-resistance.<br />

Antifungal Susceptibility: C. glabrata complex (Australian National data); MIC<br />

µg/mL. CLSI clinical breakpoints are marked where available (Pfaller and Diekema<br />

2012).<br />

No 64<br />

AmB 999 1 1 8 25 274 339 283 63 5<br />

FLU 1000 1 1 1 8 21 60 77 316 302 213<br />

VORI 892 1 8 16 45 90 195 278 119 45 68 26 1<br />

POSA 770 3 8 15 40 115 228 231 5 125<br />

ITRA 1000 2 7 30 88 210 316 110 32 6 199<br />

ANID 629 50 215 208 134 5 2 3 10 1 1<br />

MICA 629 167 338 95 7 4 3 2 3 3 2 5<br />

CAS 779 1 34 187 292 185 58 9 3 1 9<br />

5FC 998 3 306 631 27 7 6 5 4 1 1 7


42<br />

Descriptions of Medical Fungi<br />

Candida glabrata complex<br />

Candida nivariensis Alcoba-Flórez et al.<br />

RG-2 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ellipsoidal budding blastoconidia, 3-5 x 1.8-3 µm in size. No pseudohyphae<br />

or chlamydospores produced.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: No pseudohyphae produced.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose - L-Arabinose - D-Glucitol -<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose + Maltose - D-Ribose - D-Gluconate +<br />

Galactose - Cellobiose - L-Rhamnose - DL-Lactate -<br />

Sucrose - Trehalose - D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine - 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate -<br />

Trehalose Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol - Urease -<br />

Glucose + Soluble Starch v Galactitol - 0.1% Cycloheximide -<br />

Galactose - D-Xylose - D-Mannitol - Growth at 40 O C +<br />

C. nivariensis is closely related to C. glabrata and C. bracarensis. These three species<br />

were found to differ by DNA-DNA reassociation experiments, RAPD-typing, AFLPtyping<br />

and D1/D2 and ITS sequence divergence (Alcoba-Flórez et al. 2005, Correia et<br />

al. 2006, Wahyuningsih et al. 2008).<br />

Antifungal Susceptibility: C. nivariensis limited data (Australian National data);<br />

MIC µg/mL.<br />

No 64<br />

AmB 4 1 1 2<br />

FLU 4 1 1 2<br />

VORI 4 2 1 1<br />

POSA 4 1 1 2<br />

ITRA 4 4<br />

ANID 4 1 1 2<br />

MICA 4 2 2<br />

CAS 4 2 2<br />

5FC 4 1 2 1


Descriptions of Medical Fungi 43<br />

Candida haemulonii complex<br />

Candida haemulonii has recently been reclassified as a complex of three phenotypically<br />

identical but genotypically distinct entities: C. haemulonii, C. duobushaemulonii and C.<br />

haemulonii var. vulnera, based on ITS and D1/D2 sequencing. (Cendejas-Bueno et al.<br />

2012, Ramos et al. 2015).<br />

Candida haemulonii (van Uden & Kolipinski) Meyer & Yarrow<br />

RG-1 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ovoid to globose, budding yeast-like cells or blastoconidia, 2-7 x 2-7 µm.<br />

No pseudohyphae produced.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: No pseudohyphae produced.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose - L-Arabinose - D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose - α-M-D-glucoside -<br />

Glucose + Maltose + D-Ribose - D-Gluconate +<br />

Galactose - Cellobiose - L-Rhamnose +,w DL-Lactate -<br />

Sucrose + Trehalose + D-Glucosamine +,s myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine + 2-K-D-gluconate +<br />

Lactose - Melibiose - Glycerol +,s D-Glucuronate -<br />

Trehalose +,s Raffinose +,s Erythritol - Nitrate -<br />

Assimilation Melezitose +,w Ribitol +,s Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose +,w D-Xylose - D-Mannitol + Growth at 37 O C -<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern. Molecular<br />

identification may be required. C. haemulonii has been reported from a few cases of<br />

fungaemia but clinical isolations remain rare.<br />

Antifungal Susceptibility: C. haemulonii (data from Cendejas-Bueno et al. 2012,<br />

Ramos et al. 2015 and Australian National data); MIC µg/mL.<br />

No 64<br />

AmB 30 3 10 7 7 1 2<br />

FLU 30 1 3 1 25<br />

VORI 28 1 1 1 1 20 4<br />

POSA 23 1 2 1 1 3 16<br />

ITRA 30 2 1 2 1 24<br />

ANID 21 13 1 3 1 1 2<br />

MICA 21 2 10 3 2 4<br />

CAS 28 1 2 4 6 1 14<br />

5FC 25 2 4 7 6 2 3 1


44<br />

Descriptions of Medical Fungi<br />

Candida inconspicua (Lodder & Kreger-van Rij) S.A. Meyer & Yarrow<br />

Candida inconspicua is a rare cause of candidaemia.<br />

RG-1 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ovoidal budding blastoconidia, 2.0-5 x 5.0-11.0 µm.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Spherical to ovoid budding yeast cells only. Primitive<br />

pseudohyphae may be produced after 14 days.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose - L-Arabinose - D-Glucitol -<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose - Maltose - D-Ribose - D-Gluconate -<br />

Galactose - Cellobiose - L-Rhamnose - DL-Lactate +<br />

Sucrose - Trehalose - D-Glucosamine + myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine + 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate -<br />

Trehalose - Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol - Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose - D-Xylose - D-Mannitol - Growth at 40 O C +<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern and colonies<br />

are white on Candida CHROMagar.<br />

Antifungal Susceptibility: C. inconspicua limited data available (Guitard et al.<br />

2013, and Australian National data); MIC µg/mL.<br />

No 64<br />

AmB 16 2 2 1 1 3 6 1<br />

FLU 16 2 4 10<br />

VORI 13 2 4 3 2 1 1<br />

POSA 13 2 3 5 2 1<br />

ITRA 15 2 6 6 1<br />

ANID 2 2<br />

MICA 2 2<br />

CAS 13 2 5 5 1<br />

5FC 14 1 1 2 4 3 1 1 1


Descriptions of Medical Fungi 45<br />

RG-1 organism.<br />

Candida parapsilosis complex<br />

Recently Candida parapsilosis has been recognised as a complex of four species:<br />

C. parapsilosis, C. orthopsilosis, C. metapsilosis and Lodderomyces elongisporus<br />

(Tavanti et al. 2005). These four species are phenotypically indistinguishable and are<br />

best identified by ITS sequencing or MALDI-T<strong>OF</strong> MS analysis.<br />

Candida metapsilosis Tavanti et al.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ellipsoid, subglobose to fusiform budding blastoconidia, 4 x 3-6 µm, with<br />

some larger subglobose forms present.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Abundant, much branched pseudohyphae produced.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose + L-Arabinose + D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose - M-D-glucoside +<br />

Glucose + Maltose + D-Ribose + D-Gluconate +<br />

Galactose - Cellobiose - L-Rhamnose - DL-Lactate -<br />

Sucrose - Trehalose + D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine + 2-K-D-gluconate +<br />

Lactose - Melibiose - Glycerol + D-Glucuronate -<br />

Trehalose - Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose + Ribitol + Urease -<br />

Glucose + Soluble Starch v Galactitol - 0.1% Cycloheximide -<br />

Galactose + D-Xylose + D-Mannitol + Growth at 37 O C +<br />

Key Features: Candida metapsilosis cannot be distinguished morphologically from C.<br />

parapsilosis and C. orthopsilosis, but can be identified by ITS sequencing (Asadzadeh<br />

et al. 2009, Borman et al. 2009, Tavanti et al. 2005) and MALDI-T<strong>OF</strong> MS analysis.<br />

Antifungal Susceptibility: C. metapsilosis limited data (Diekema et al. 2009 and<br />

Australian National data); MIC µg/mL.<br />

No 64<br />

AmB 32 1 5 12 11 2 1<br />

FLU 32 1 19 10 2<br />

VORI 32 1 4 22 3 2<br />

POSA 32 1 7 16 6 1<br />

ITRA 2 1 1<br />

ANID 13 2 5 3 2 1<br />

MICA 13 9 3 1<br />

CAS 26 1 5 15 4 1<br />

5FC 2 2<br />

Note: Additional data for ITRA MIC range 0.06-0.5, MIC 90<br />

= 0.25; and 5FC MIC range 0.06-<br />

64, MIC 90<br />

= 0.5 (Gomez-Lopez et al. 2008, Miranda-Zapico et al. 2011).


46<br />

Descriptions of Medical Fungi<br />

Candida parapsilosis complex<br />

Candida orthopsilosis Tavanti et al.<br />

RG-1 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ellipsoid to subglobose budding blastoconidia, 2-5 x 3-7 µm, with some<br />

larger elongated forms present.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Abundant, much-branched pseudohyphae produced.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose + L-Arabinose + D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose - α-M-D-glucoside +<br />

Glucose + Maltose + D-Ribose - D-Gluconate +<br />

Galactose - Cellobiose - L-Rhamnose - DL-Lactate -<br />

Sucrose - Trehalose + D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine + 2-K-D-gluconate +<br />

Lactose - Melibiose - Glycerol + D-Glucuronate -<br />

Trehalose - Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose + Ribitol + Urease -<br />

Glucose + Soluble Starch v Galactitol - 0.1% Cycloheximide -<br />

Galactose + D-Xylose + D-Mannitol + Growth at 37 O C +<br />

Key Features: Candida orthopsilosis cannot be distinguished morphologically from C.<br />

parapsilosis and C. metapsilosis, but can be identified by ITS sequencing (Asadzadeh<br />

et al. 2009, Borman et al. 2009, Tavanti et al. 2005) and MALDI-T<strong>OF</strong> MS analysis.<br />

Antifungal Susceptibility: C. orthopsilosis (Diekema et al. 2009, Canton et al.<br />

2012, 2013 and Australian National data); MIC µg/mL.<br />

No 64<br />

AmB 224 3 3 30 45 60 48 23 9<br />

FLU 196 1 9 45 85 24 16 9 4 3<br />

VORI 176 28 40 51 35 6 13 2 2<br />

POSA 136 5 21 49 41 9 11<br />

ITRA 94 1 2 14 30 32 15<br />

ANID 86 1 2 9 20 44 10<br />

MICA 85 1 1 3 36 37 6 1<br />

CAS 146 2 4 18 45 36 27 10 4<br />

5FC 92 56 22 8 3 1 1 1


Descriptions of Medical Fungi 47<br />

Candida parapsilosis complex<br />

RG-1 organism.<br />

Candida parapsilosis (Ashford) Langeron & Talice<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Predominantly small, globose to ovoid budding blastoconidia, 3-4 x 5-8<br />

µm, with some larger elongated forms present.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Abundant, much-branched pseudohyphae in a delicate treelike<br />

pattern with 2-3 blastoconidia in small clusters at intervals along the pseudohyphae.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose +,s L-Arabinose + D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose - M-D-glucoside +<br />

Glucose + Maltose + D-Ribose v D-Gluconate v<br />

Galactose + Cellobiose - L-Rhamnose - DL-Lactate -<br />

Sucrose + Trehalose + D-Glucosamine v myo-Inositol -<br />

Maltose -,s Lactose - N-A-D-glucosamine + 2-K-D-gluconate +<br />

Lactose - Melibiose - Glycerol + D-Glucuronate -<br />

Trehalose -,s Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose + Ribitol v Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose + D-Xylose + D-Mannitol + Growth at 37 O C +<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern. C. parapsilosis<br />

is commonly found on the skin and is a causative agent of candidaemia.<br />

Antifungal Susceptibility: C. parapsilosis (Australian National data); MIC µg/mL.<br />

CLSI clinical breakpoints are marked where available (Pfaller and Diekema 2012).<br />

No 64<br />

AmB 607 2 8 28 131 142 210 80 2<br />

FLU 608 12 66 122 169 136 67 27 4 4 1<br />

VORI 529 145 98 120 100 48 14 3 1<br />

POSA 437 4 41 126 182 69 13 2<br />

ITRA 608 1 15 80 134 279 84 12 3<br />

ANID 342 3 6 7 45 149 107 18 7<br />

MICA 342 1 3 2 51 164 90 26 5<br />

CAS 459 1 1 5 38 183 177 46 4 1 3<br />

5FC 608 2 1 31 193 176 158 38 5 2 2


48<br />

Descriptions of Medical Fungi<br />

Candida parapsilosis complex<br />

Lodderomyces elongisporus (Recca & Mrak) van der Walt<br />

Lodderomyces elongisporus has been isolated from soft drinks and juice concentrates,<br />

natural fermentations of cocoa, soil, an infected fingernail, human blood infections and<br />

from baby cream. Initially isolates appeared to be atypical forms of C. parapsilosis,<br />

but sequence analysis identified them as L. elongisporus. In view of these findings,<br />

L. elongisporus may be more common among clinical isolates than initially thought<br />

(Lockhart et al. 2008a, Kurtzman).<br />

RG-1 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ellipsoid to elongate budding blastoconidia, 2.6-6.3 x 4-7.4 µm, with<br />

occasional spherical forms present.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Abundant, much-branched pseudohyphae produced.<br />

Ascospore Formation: Asci are unconjugated, persistent, and are transformed from<br />

budding cells. Each ascus forms one, rarely two, long-ellipsoid ascospores. Ascospores<br />

observed on V8 agar after 7-10 days at 25 O C.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose + L-Arabinose - D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose - α-M-D-glucoside +<br />

Glucose + Maltose + D-Ribose - D-Gluconate +,w<br />

Galactose - Cellobiose - L-Rhamnose - DL-Lactate -<br />

Sucrose - Trehalose + D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine + 2-K-D-gluconate +<br />

Lactose - Melibiose - Glycerol + D-Glucuronate -<br />

Trehalose + Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose + Ribitol + Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose + D-Xylose +,w D-Mannitol + Growth at 37 O C +<br />

Key Features: In the absence of ascospores, L. elongisporus cannot be distinguished<br />

physiologically from C. parapsilosis, C. orthopsilosis and C. metapsilosis but can be<br />

identified based on ITS sequencing (Asadzadeh et al. 2009, Borman et al. 2009,<br />

Tavanti et al. 2005) and MALDI-T<strong>OF</strong> MS analysis.


Descriptions of Medical Fungi 49<br />

Candida rugosa complex<br />

Candida rugosa has recently been recognised as a species complex of C. rugosa, C.<br />

pseudorugosa and another as yet undescribed species (Li et al. 2006, Paredes et al.<br />

2012). These species are best identified by ITS sequencing.<br />

RG-1 organism.<br />

Candida rugosa (Anderson) Diddens & Lodder<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ellipsoidal to elongate budding blastoconidia, 5-11 x 1.5-2.5 µm.<br />

Sometimes short pseudohyphae may be produced.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Densely branched pseudohyphae produced.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose v L-Arabinose - D-Glucitol v<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose - Maltose - D-Ribose - D-Gluconate v<br />

Galactose - Cellobiose - L-Rhamnose - DL-Lactate v<br />

Sucrose - Trehalose - D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine + 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate -<br />

Trehalose - Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol - Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose + D-Xylose + D-Mannitol v Growth at 37 O C +<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern (D-Xylose<br />

and Glycerol +ve; Ribitol -ve.). C. rugosa has been associated with catheter related<br />

fungaemia and has been isolated from human and bovine faeces, sea water and soil.<br />

Antifungal Susceptibility: C. rugosa limited data (Diekema et al. 2009, Espinel-<br />

Ingroff et al. 2014 and Australian National data); MIC µg/mL.<br />

No 64<br />

AmB 21 1 1 5 5 7 1 1<br />

FLU 97 1 3 15 24 29 8 8 6 1 2<br />

VORI 80 4 8 38 16 4 7 3<br />

POSA 66 1 4 12 27 9 8 5<br />

ITRA 5 1 1 1 1 1<br />

ANID 20 3 4 4 4 3 2<br />

MICA 20 2 3 5 6 2 2<br />

CAS 25 1 1 2 3 10 1 1 2


50<br />

Descriptions of Medical Fungi<br />

Candida tropicalis is a major cause of septicaemia and disseminated candidiasis. It is<br />

also found as part of the normal human mucocutaneous flora and environmental isolations<br />

have been made from faeces, shrimp, kefir and soil.<br />

RG-2 organism.<br />

Candida tropicalis (Castellani) Berkhout<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Spherical to subspherical budding yeast-like cells or blastoconidia, 3.5-7<br />

x 5.5-10 µm.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Abundant, long, wavy, branched pseudohyphae with numerous<br />

ovoid blastoconidia, budding off. Terminal vesicles (chlamydospores) are not produced.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose v L-Arabinose - D-Glucitol +<br />

Fermentation Sucrose v D-Arabinose - M-D-glucoside v<br />

Glucose + Maltose + D-Ribose v,s D-Gluconate v<br />

Galactose + Cellobiose v L-Rhamnose - DL-Lactate v<br />

Sucrose v Trehalose + D-Glucosamine v myo-Inositol -<br />

Maltose + Lactose - N-A-D-glucosamine + 2-K-D-gluconate +<br />

Lactose - Melibiose - Glycerol v D-Glucuronate -<br />

Trehalose +,s Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose v Ribitol v Urease -<br />

Glucose + Soluble Starch + Galactitol - 0.1% Cycloheximide +<br />

Galactose + D-Xylose + D-Mannitol + Growth at 40 O C +<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern. Colonies are<br />

dark blue on Candida CHROMagar.<br />

Antifungal Susceptibility: C. tropicalis (Australian National data); MIC µg/mL.<br />

CLSI clinical breakpoints are marked where available (Pfaller and Diekema 2012).<br />

No 64<br />

AmB 284 1 8 45 94 97 38 1<br />

FLU 284 1 10 46 98 65 32 14 4 4 10<br />

VORI 251 10 17 37 64 58 34 14 6 5 1 4 1<br />

POSA 190 4 7 28 35 53 33 17 6 2 5<br />

ITRA 284 2 13 22 94 98 43 2 1 1 1 7<br />

ANID 126 9 8 29 67 11 1 1<br />

MICA 126 8 28 69 16 2 1 1 1<br />

CAS 207 1 15 68 67 40 9 3 1<br />

5FC 284 49 139 54 21 5 2 1 1 2 1 2 7


Descriptions of Medical Fungi 51<br />

Chaetomium Kunze ex Fries<br />

The genus Chaetomium contains between 160 and 180 species. All are saprophytic<br />

being isolated from soil, straw, dung and plant debris. Several species are thermophilic<br />

and can grow at temperatures above 37 O C. Chaetomium species are important agents<br />

for the decomposition of cellulose waste and plant materials, and are only rarely<br />

isolated in medical mycology laboratories. RG-1 organisms.<br />

Morphological Description: Chaetomium is a common ascomycete characterised<br />

by the formation of darkly-pigmented, globose, ovoid, barrel to flask-shaped, ostiolate<br />

ascocarps (perithecia) beset with dark-coloured terminal hairs (setae) which are<br />

straight, branched or curved. Asci are clavate to cylindrical, typically eight-spored and<br />

evanescent. Ascospores are one-celled, darkly-pigmented, smooth-walled, of varying<br />

shape, mostly ovoid, ellipsoidal or lemon-shaped. Chlamydospores and solitary conidia<br />

may also be produced.<br />

Molecular Identification: Lee and Hanlin (1999) established the phylogenetic<br />

relationships of Chaetomium based on ribosomal DNA sequences. ITS sequencing<br />

may be useful for identification of some clinical species.<br />

Key Features: Ascomycete producing darkly-pigmented ostiolate perithecia beset<br />

with long dark terminal setae.<br />

References: Ames (1963), Seth (1970), Millner (1975), Domsch et al. (2007), Ellis and<br />

Keane (1981), Ellis (1981), von Arx (1986), de Hoog et al. (2000, 2015).<br />

a<br />

100 μm 10 μm<br />

b<br />

Chaetomium spp. (a) ascocarps (perithecia) and (b) ascus with ascospores.<br />

Antifungal Susceptibility: Chaetomium very limited data (McGinnis and Pasarell<br />

1998a, Serena et al. 2003, Barron et al. 2003, Australian National Data); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 0.125-16 4 VORI 0.125-0.5 0.5<br />

ITRA 0.03-0.25 0.125 POSA


52<br />

Descriptions of Medical Fungi<br />

Chrysosporium Corda<br />

Species of Chrysosporium are occasionally isolated from skin and nail scrapings,<br />

especially from feet, but because they are common soil saprophytes they are usually<br />

considered contaminants. There are about 70 species of Chrysosporium, several are<br />

keratinolytic with some also being thermotolerant, and cultures may closely resemble<br />

some dermatophytes, especially Trichophyton mentagrophytes. Some strains may<br />

also resemble cultures of Histoplasma and Blastomyces.<br />

Morphological Description: Colonies are moderately fast growing, flat, white to tan<br />

to beige in colour, often with a powdery or granular surface texture. Reverse pigment<br />

absent or pale brownish-yellow with age. Hyaline, one-celled conidia are produced<br />

directly on vegetative hyphae by non-specialised conidiogenous cells. Conidia are<br />

typically pyriform to clavate with truncate bases and are formed either intercalary<br />

(arthroconidia), laterally (often on pedicels) or terminally.<br />

Molecular Identification: Chrysosporium is phylogenetically heterogeneous; the<br />

polyphyletic origin of the genus was first demonstrated by Vidal et al. (2000) on the basis<br />

of ITS sequences, and further elaborated by Stchigel et al. (2014). ITS sequencing can<br />

assist in identification of clinical isolates.<br />

Chrysosporium tropicum Carmichael<br />

Morphological Descriptions: Colonies are flat, white to cream-coloured with a<br />

very granular surface. Reverse pigment absent or pale brownish-yellow with age.<br />

Microscopically, conidia are numerous, hyaline, single-celled, clavate to pyriform,<br />

smooth, slightly thick-walled (6-7 x 3.5-4 µm), and have broad truncate bases and<br />

pronounced basal scars. The conidia are formed at the tips of the hyphae, on short or<br />

long lateral branches, or sessile along the hyphae (intercalary). No macroconidia or<br />

hyphal spirals are seen. RG-2 organism.<br />

References: Carmichael (1962), Rebell and Taplin (1970), Sigler and Carmichael<br />

(1976), van Oorschot (1980), Domsch et al. (2007), de Hoog et al. (2000, 2015).<br />

a<br />

10 μm<br />

Chrysosporium tropicum (a) culture and (b) typical pyriform to clavate-shaped conidia<br />

with truncated bases which may be formed either intercalary, laterally or terminally.<br />

b


Descriptions of Medical Fungi 53<br />

Cladophialophora Borelli<br />

The genus Cladophialophora is characterised by: (1) the absence of conidiophores,<br />

“shield cells,” or prominent hila (attachment points); (2) the ability to grow on media<br />

containing cycloheximide; and (3) having dry, non-fragile chains of conidia (Revankar<br />

and Sutton 2010). It has recently been re-evaluated by multilocus sequencing and<br />

currently contains seven species associated with human infection (Badali et al. 2008).<br />

Cladophialophora bantiana is the causative agent of numerous cases of cerebral<br />

phaeohyphomycosis many of which occur in immunocompetent individuals and most<br />

of which are fatal (Chakrabarti et al. 2016). C. carrionii and the recently described C.<br />

samoensis are agents of chromoblastomycosis. Less common species occasionally<br />

implicated in deep and superficial mycoses include, C. arxii, C. boppii, C. devriesii, C.<br />

emmonsii, C. modesta and C. saturnica. C. yegresii is a closely related environmental<br />

sister species to C. carrionii (Revankar and Sutton 2010, de Hoog et al. 1995, 2015).<br />

Cladophialophora bantiana (Saccardo) de Hoog et al.<br />

Synonymy: Xylohypha bantiana (Saccardo) McGinnis, Borelli and Ajello.<br />

Cladosporium bantianum (Sacc.) Borelli.<br />

Cladosporium trichoides Emmons.<br />

Cladophialophora bantiana has been isolated from soil and is a recognised agent of<br />

cerebral phaeohyphomycosis. The fungus is neurotropic and may cause brain abscess<br />

in both normal and immunosuppressed patients and is usually fatal. The fungus is likely<br />

introduced via inhalation and direct transfer to the brain via the paranasal sinuses, or<br />

traumatic head injury.<br />

WARNING: RG-3 organism. Cultures of C. bantiana represent a potential biohazard<br />

to laboratory personnel and must be handled with extreme caution in Class II Biological<br />

Safety Cabinet (BSCII).<br />

Morphological Description: Colonies are moderately fast growing, olivaceousgrey,<br />

suede-like to floccose and grow at temperatures up to 42-43 O C. Conidia are<br />

formed in long, sparsely branched, flexuose, acropetal chains from undifferentiated<br />

conidiophores. Conidia are one-celled (very occasionally two-celled), pale brown,<br />

smooth-walled, ellipsoid to oblong-ellipsoid and are 2-3 x 4-7 µm in size.<br />

C. bantiana may be distinguished from Cladosporium species by the absence of conidia<br />

with distinctly pigmented hila, the absence of shield cells and by growth at >40 O C<br />

(compared with C. carrionii which has a maximum growth temperature of 35-37 O C, and<br />

Cladosporium species which have a maximum of


54<br />

Descriptions of Medical Fungi<br />

Cladophialophora bantiana (Saccardo) de Hoog et al.<br />

a<br />

b<br />

10 μm<br />

Cladophialophora bantiana (a) culture and (b) conidiophore and conidia.<br />

Antifungal Susceptibility: C. bantiana limited data available (Australian National<br />

data); MIC µg/mL.<br />

No. 64<br />

AmB 14 1 3 2 5 1 2<br />

VORI 14 3 2 7 1 1<br />

POSA 12 3 2 4 3<br />

ITRA 14 5 2 5 2<br />

C. bantiana limited data (McGinnis and Pasarell 1998a, 1998b, Espinel-Ingroff et al.<br />

2001, Chakrabarti et al. 2015); MIC µg/mL.<br />

AmB Range 0.03-8; MIC 90<br />

= 2 VORI Range 0.03-1; MIC 90<br />

= 0.125<br />

ITRA Range 0.03-1; MIC 90<br />

= 0.125 POSA Range 0.03-0.125; MIC 90<br />

= 0.125<br />

TERB Range 0.03-1; Geometric Mean = 0.08


Descriptions of Medical Fungi 55<br />

Cladophialophora carrionii (Trejos) de Hoog et al.<br />

Synonymy: Cladosporium carrionii Trejos<br />

Cladophialophora carrionii is a recognised agent of chromoblastomycosis and it<br />

has been isolated from soil and fence posts made from Eucalyptus spp. Cases of<br />

chromoblastomycosis caused by C. carrionii are commonly found in Australia,<br />

Venezuela, Madagascar and South America. Isolates from phaeomycotic cysts and<br />

opportunistic infections have also been reported.<br />

RG-2 organism.<br />

Morphological Description: Colonies are slow growing, reaching 3-4 cm in diameter<br />

after one month, with a compact suede-like to downy surface and are olivaceous-black<br />

in colour. Microscopy shows ascending to erect, olivaceous-green, apically branched,<br />

elongate conidiophores producing branched acropetal chains of conidia. Conidia are<br />

pale olivaceous, smooth-walled or slightly verrucose, limoniform to fusiform, 1.5-3.0 x<br />

2.0-7.0 µm in size. Bulbous phialides with large collarettes and minute, hyaline conidia<br />

are occasionally formed on nutritionally poor media. Maximum growth temperature<br />

35-37 O C.<br />

Molecular Identification: ITS sequencing is recommended (Abliz et al. 2004; de Hoog<br />

et al. 2007).<br />

Key Features: Conidia are smaller and comprise heavily branched chains which fall<br />

apart much more easily than in the other Cladophialophora species.<br />

References: McGinnis (1980), Rippon (1988), de Hoog et al. (1995, 2000, 2015).<br />

Cladophialophora carrionii culture.


56<br />

Descriptions of Medical Fungi<br />

Cladophialophora carrionii (Trejos) de Hoog et al.<br />

10 μm<br />

Cladophialophora carrionii conidiophores and conidia.<br />

Antifungal Susceptibility: C. carrionii limited data available (Australian National<br />

data); MIC µg/mL.<br />

No. 64<br />

AmB 2 1 1<br />

VORI 2 1 1<br />

POSA 2 1 1<br />

ITRA 2 1 1<br />

C. carrionii limited data available (McGinnis and Pasarell 1998a,1998b, Espinel-<br />

Ingroff et al. 2001, Gonzales et al. 2005); MIC µg/mL.<br />

AmB Range 0.06-4; MIC 90<br />

= 1 VORI Range 0.03-0.5; MIC 90<br />

= 0.25<br />

ITRA Range 0.06-0.5; MIC 90<br />

= 0.125 POSA Range 0.06-0.5; MIC 90<br />

= 0.25<br />

TERB Range 0.03-0.125; Geometric Mean = 0.03


Descriptions of Medical Fungi 57<br />

Cladosporium species are ubiquitous worldwide, and commonly isolated from soil and<br />

organic matter. They represent the most frequently isolated airborne fungi. The genus<br />

has undergone a number of revisions. The well-known thermotolerant ‘true humanpathogenic<br />

species, formerly known as C. bantiana, C. carrionii and C. devriesii,<br />

characterised by the absence of conidiophores, and unpigmented conidial scars, were<br />

reclassified in Cladophialophora (de Hoog et al. 1995, Bensch et al. 2012). The remaining<br />

species of medical interest were C. cladosporioides, C. herbarum, C. oxysporum,<br />

and C. sphaerospermum. More recently, extensive revisions based on polyphasic<br />

approaches have recognised 169 species, and demonstrated that C. cladosporioides,<br />

C. herbarum and C. sphaerospermum are species complexes encompassing several<br />

sibling species that can only be distinguished by phylogenetic analyses (Crous et al.<br />

2007, Schubert et al. 2007, Zalar et al. 2007, Bensch et al. 2010, 2012).<br />

Sandoval-Denis et al. (2015) analysed 92 clinical isolates from the United States<br />

using phenotypic and molecular methods, which included sequence analysis of the<br />

ITS and D1/D2 regions, partial EF-1α and actin genes. Surprisingly, the most frequent<br />

species was Cladosporium halotolerans (15%), followed by C. tenuissimum (10%), C.<br />

subuliforme (6%), and C. pseudocladosporioides (5%). However, 40% of the isolates<br />

did not correspond to any known species and were deemed to represent at least 17<br />

new lineages for Cladosporium. The most frequent anatomic site of isolation was the<br />

respiratory tract (55%), followed by superficial (28%) and deep tissues and fluids (15%).<br />

Species of the two recently described Cladosporium-like genera Toxicocladosporium<br />

and Penidiella were also reported for the first time from clinical samples (Sandoval-<br />

Denis et al. 2015).<br />

RG-1 organisms.<br />

Cladosporium Link ex Fries<br />

Morphological Description: Colonies are slow growing, mostly olivaceous-brown<br />

to blackish-brown but also sometimes grey, buff or brown, suede-like to floccose,<br />

often becoming powdery due to the production of abundant conidia. The reverse is<br />

olivaceous-black. Vegetative hyphae, conidiophores and conidia are equally pigmented.<br />

Conidiophores are more or less distinct from the vegetative hyphae, being erect,<br />

straight or flexuose, unbranched or branched only in the apical region, with geniculate<br />

sympodial elongation in some species. Conidia are produced in branched acropetal<br />

chains, being smooth, verrucose or echinulate, one to four-celled, and have a distinct<br />

dark hilum. The term blastocatenate is often used to describe chains of conidia where<br />

the youngest conidium is at the apical or distal end of the chain. Note: The conidia<br />

closest to the conidiophore, and where the chains branch, are usually “shield-shaped”.<br />

The presence of shield-shaped conidia, a distinct hilum, and chains of conidia that<br />

readily disarticulate, are characteristic of the genus Cladosporium.<br />

Key Features: Dematiaceous hyphomycete forming branched acropetal chains of<br />

conidia, each with a distinct hilum.<br />

Molecular Identification: Genus level identification is usually sufficient and<br />

morphological identification can be confirmed by ITS and D1/D2 sequence analysis.<br />

Multilocus gene analysis of the ITS, D1/D2, EF-1α and actin gene loci is necessary for<br />

accurate species identification (Bensche et al. 2012).


58<br />

Descriptions of Medical Fungi<br />

Cladosporium Link ex Fries<br />

a<br />

b<br />

10 μm<br />

Cladosporium cladosporioides complex (a) culture, (b) conidiophores and conidia.<br />

Antifungal Susceptibility: Cladosporium species (Sandoval-Denis et al. 2015 and<br />

Australian National data); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 0.03-2 2 POSA 0.03-4 0.5<br />

ITRA 0.03-16 0.5 VORI 0.25-16 4<br />

References: Ellis (1971, 1976), Domsch et al. (1980), McGinnis (1980), de Hoog et al.<br />

(2000, 2015), Crous et al. (2007), Schubert et al. (2007), Zalar et al. (2007), Bensch et<br />

al. (2010 and 2012), Sandoval-Denis et al. (2015).


Descriptions of Medical Fungi 59<br />

Clavispora lusitaniae Rodrigues de Miranda<br />

Synonymy: Candida lusitaniae van Uden & do Carmo-Sousa.<br />

Clavispora lusitaniae is a known cause of disseminated candidiasis, including<br />

septicaemia and pyelonephritis. C. lusitaniae was first isolated from the digestive<br />

tract of warm-blooded animals and environmental isolations have been made from<br />

cornmeal, citrus peel, fruit juices, and milk from cows with mastitis. RG-2 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ovoid to ellipsoidal budding blastoconidia, 1.5-6.0 x 2.5-10µm.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Abundant pseudohyphae with short chains of blastoconidia.<br />

Molecular Identification: ITS sequencing recommended.<br />

MALDI-T<strong>OF</strong> MS: Able to accurately identify this species.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose + L-Arabinose v D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose - α-M-D-glucoside v<br />

Glucose + Maltose + D-Ribose - D-Gluconate s<br />

Galactose v Cellobiose + L-Rhamnose v DL-Lactate +,w<br />

Sucrose v Trehalose + D-Glucosamine - myo-Inositol -<br />

Maltose v Lactose - N-A-D-glucosamine + 2-K-D-gluconate +<br />

Lactose - Melibiose - Glycerol + D-Glucuronate nd<br />

Trehalose v Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose + Ribitol s Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose + D-Xylose + D-Mannitol + Growth at 40 O C +<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern.<br />

Note: C. lusitaniae may also be difficult to distinguish from Candida tropicalis using<br />

some yeast identification systems.<br />

Antifungal Susceptibility: C. lusitaniae (Diekema et al. 2009, Pfaller et al. 2013,<br />

Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 214 1 9 31 86 74 10 1 1 1<br />

FLU 279 12 33 66 82 33 21 8 4 1 5 4<br />

VORI 262 169 53 15 12 3 5 3 2<br />

POSA 252 5 40 77 83 26 15 2 4<br />

ITRA 99 1 1 3 21 33 20 15 4 1<br />

ANID 175 1 13 32 56 59 13 1<br />

MICA 159 1 2 13 22 57 46 11 6 1<br />

CAS 247 2 4 12 86 91 38 12 1 1<br />

5FC 44 15 20 1 2 2 1 1 2


60<br />

Descriptions of Medical Fungi<br />

Coccidioides immitis/posadasii complex<br />

WARNING: RG-3 organism. Cultures of Coccidioides immitis/posadasii represent a<br />

severe biohazard to laboratory personnel and must be handled with extreme caution<br />

in Class II Biological Safety Cabinet (BSCII).<br />

Coccidioides immitis has been separated into two distinct species: C. immitis and C.<br />

posadasii (Fisher et al. 2002). The two species are morphologically identical and can be<br />

distinguished only by genetic analysis and different rates of growth in the presence of<br />

high salt concentrations (C. posadasii grows more slowly). C. immitis is geographically<br />

limited to California’s San Joaquin Valley region and Mexico, whereas C. posadasii is<br />

found in California, Arizona, Texas, Mexico and South America.<br />

Morphological Description: Colonies of C. immitis and C. posadasii grown at 25 O C<br />

are initially moist and glabrous, but rapidly become suede-like to downy, greyishwhite<br />

with a tan to brown reverse, however considerable variation in growth rate and<br />

culture morphology has been noted. Microscopy shows typical single-celled, hyaline,<br />

rectangular to barrel-shaped, alternate arthroconidia, 2.5-4 x 3-6 µm in size, separated<br />

from each other by a disjunctor cell. This arthroconidial state has been classified in the<br />

genus Malbranchea and is similar to that produced by many non-pathogenic soil fungi,<br />

e.g. Gymnoascus species.<br />

Comment: Coccidioides immitis and C. posadasii dimorphic fungi, existing in living<br />

tissue as spherules and endospores, and in soil or culture in a mycelial form. Culture<br />

identification by either exoantigen test or DNA sequencing is preferred to minimise<br />

exposure to the infectious propagule.<br />

Key Features: Clinical history, tissue pathology, culture identification by ITS sequence<br />

analysis.<br />

20 μm<br />

Coccidioides immitis tissue morphology showing typical endosporulating spherules.<br />

Young spherules have a clear centre with peripheral cytoplasm and a prominent<br />

thick-wall. Endospores (sporangiospores) are later formed within the spherule by<br />

repeated cytoplasmic cleavage. Rupture of the spherule releases endospores into the<br />

surrounding tissue where they re-initiate the cycle of spherule development.<br />

References: Ajello (1957), Steele et al. (1977), McGinnis (1980), Chandler et al.<br />

(1980), Catanzaro (1986), Rippon (1988), de Hoog et al. (2015), Fisher et al. (2002).


Descriptions of Medical Fungi 61<br />

Coccidioides immitis/posadasii complex<br />

Molecular Identification: In endemic areas a DNA probe for recognition of the species<br />

is commercially available (Padhye et al. 1994b). ITS sequencing is recommended for<br />

differentiation of species (Tintelnot et al. 2007, Binnicker et al. 2011).<br />

a<br />

b<br />

5 μm<br />

Coccidioides immitis (a) culture and (b) arthroconidia with disjunctor cells.<br />

Antifungal Susceptibility: C. immitis (Ramani and Chaturvedi 2007); MIC µg/mL.<br />

Antifungal susceptibility testing not recommended. For treatment options see<br />

Clinical Practice Guidelines for the Management of Coccidioidomycosis (Galgiani et<br />

al. 2005).<br />

No. 16<br />

AmB 45 7 25 8 4 1<br />

FLU 45 1 4 10 27 3<br />

ITRA 45 13 9 15 8<br />

POSA 45 25 13 3 4<br />

VORI 45 23 22


62<br />

Descriptions of Medical Fungi<br />

Colletotrichum coccodes (Wallroth) S. Hughes<br />

Over 500 Colletotrichum species have been reported. C. coccodes is a common soil<br />

and plant pathogen widely distributed in Africa, Asia, Australasia, Europe, and the<br />

Americas. It has been reported from a case of human mycotic keratitis.<br />

RG-1 organism.<br />

Morphological Description: Colonies usually darkly pigmented with white aerial<br />

mycelium, consisting of numerous black sclerotia and light brown-coloured conidial<br />

masses, reverse is dark brown. Sclerotia are usually abundant, setose, spherical and<br />

are often confluent. Conidia are straight, fusiform, attenuated at the ends, 16-22 x 3-4<br />

µm. Appressoria are common, clavate, brown, 11-16.5 x 6-9.5 µm, variable in shape.<br />

Molecular Identification: ITS and/or D1/D2 sequencing may be used for species<br />

identification (Cano et al. 2004).<br />

References: Domsch et al. (1980), McGinnis (1980), de Hoog et al. (2000, 2015).<br />

a<br />

b<br />

50 μm<br />

c<br />

20 μm<br />

d<br />

10 μm<br />

Colletotrichum coccodes (a) culture, (b) sclerotia with setae,<br />

(c) conidia and (d) appressoria.


Descriptions of Medical Fungi 63<br />

Synonymy: Entomophthora coronata (Costantin) Kevorkian.<br />

The species of the genus Conidiobolus produce characteristic multinucleate primary<br />

and secondary (replicative) conidia on top of unbranched conidiophores. Each<br />

subspherical conidium is discharged as a result of the pressure developed within the<br />

conidium, and each bears a more or less prominent papilla after discharge (King 1983).<br />

The genus contains 27 species, however only Conidiobolus coronatus, C. incongruus<br />

and C. lamprauges have been reported as causative agents of human and animal<br />

infection. A morphological identification key for clinical isolates was given by Vilela et<br />

al. (2010).<br />

RG-2 organism.<br />

Conidiobolus coronatus (Costantin) Batko<br />

Morphological Description: Colonies of C. coronatus grow rapidly and are flat,<br />

cream-coloured, glabrous becoming radially folded and covered by a fine, powdery,<br />

white surface mycelium and conidiophores. The lid of the petri dish soon becomes<br />

covered with conidia, which are forcibly discharged by the conidiophores. The colour<br />

of the colony may become tan to brown with age. Conidiophores are simple forming<br />

solitary, terminal conidia which are spherical, 10 to 25 µm in diameter, single-celled<br />

and have a prominent papilla. Conidia may also produce hair-like appendages, called<br />

villae. Conidia germinate to produce either: (1) single or multiple hyphal tubes that<br />

may also become conidiophores which bear secondary conidia; or (2) replicate by<br />

producing multiple short conidiophores, each bearing a small secondary conidium.<br />

C. coronatus is commonly present in soil and decaying leaves. It has a worldwide<br />

distribution especially in tropical rain forests of Africa. Human infections are usually<br />

restricted to the rhinofacial area. However, there are occasional reports of dissemination<br />

to other sites. All human infections have been confined to the tropics.<br />

References: Emmons and Bridges (1961), King (1983), McGinnis (1980), Rippon<br />

(1988), Kwon-Chung and Bennett (1992), de Hoog et al. (2000, 2015), Ellis (2005a).<br />

Conidiobolus coronatus culture showing satellite colonies formed<br />

by germinating conidia ejected from the primary colony.


64<br />

Descriptions of Medical Fungi<br />

Conidiobolus coronatus (Costantin) Batko<br />

a<br />

10 μm<br />

b<br />

10 μm<br />

Conidiobolus coronatus (a) spherical conidium with hair-like<br />

appendages (villae) and (b) conidia with prominent papillae.<br />

Coniochaeta hoffmannii (J.F.H. Beyma) Z.U. Khan, Gené & Guarro<br />

Synonymy: Lecythophora hoffmannii (J.F.H. Beyma) W. Gams & McGinnis.<br />

Phialophora hoffmannii (J.F.H. Beyma) Schol-Schwarz.<br />

In response to recent changes in the nomenclature for pleomorphic fungi, Khan<br />

et al. (2013) transferred all Lecythophora species to Coniochaeta. The genus<br />

Coniochaeta is morphologically similar to Phialemonium. It also forms adelophialides,<br />

but in Coniochaeta, these conidiogenous cells show conspicuous collarettes, and the<br />

colonies are usually pink-salmon to dark brown, although discrete phialides like those of<br />

Acremonium may also be present (Perdomo et al. 2011b). Coniochaeta hoffmannii has<br />

been associated with cases of subcutaneous infections, keratitis, sinusitis, peritonitis,<br />

and canine osteomyelitis. Coniochaeta mutabilis has been described to be a causative<br />

agent of human peritonitis, endocarditis, endophthalmitis, and keratitis (Perdomo et al.<br />

2011b).<br />

RG-1 organism.<br />

Morphological Description: Colonies are flat, smooth, moist, pink to orange, with<br />

regular and sharp margin; reverse pink. Hyphae are narrow, hyaline, producing conidia<br />

laterally from small collarettes directly on the hyphae, or from lateral cells which are<br />

sometimes arranged in dense groups; lateral cells flask-shaped or nearly cylindrical.<br />

Collarettes are unpigmented, about 1.5 µm wide. Conidia are hyaline, smooth and thinwalled,<br />

broadly ellipsoidal to cylindrical or allantoid, 3.0-3.5 x 1.5-2.5 µm, produced in<br />

slimy heads.


Descriptions of Medical Fungi 65<br />

Coniochaeta hoffmannii (J.F.H. Beyma) Z.U. Khan, Gené & Guarro<br />

Note: Coniochaeta and Phialemonium species are poorly differentiated morphologically,<br />

and are difficult to identify. They may also be confused with poorly sporulating Fusarium<br />

or Acremonium species.<br />

Molecular Identification: Khan et al. (2013) used a combined sequence data set of<br />

the ITS region, D1/D2, actin and β-tubulin genes to resolve the unique phylogenetic<br />

status of this species.<br />

References: de Hoog (1983), de Hoog et al. (2000, 2015), Perdomo et al. (2011b),<br />

Khan et al. (2013).<br />

a<br />

b<br />

10 µm<br />

b<br />

10 µm<br />

Coniochaeta hoffmannii (a) culture, (b) hyphae with small collarettes and conidia.<br />

Antifungal Susceptibility: Coniochaeta hoffmannii very limited data (McGinnis and<br />

Pasarell 1998a, Perdomo et al. 2011b and Australian National data); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 0.06-2 2 POSA 0.125-1 1<br />

ITRA 0.06-32 1 VORI 0.25-1 1


66<br />

Descriptions of Medical Fungi<br />

Cryptococcus Kützing emend. Phaff & Spencer<br />

The genus Cryptococcus is characterised by globose to elongate yeast-like cells or<br />

blastoconidia that reproduce by narrow-necked budding. Pseudohyphae are absent or<br />

rudimentary. Most species are encapsulated, although the extent of capsule formation<br />

depends on the medium. Under certain conditions of growth, the capsule may contain<br />

starch-like compounds, which are released into the medium by many strains. Within<br />

tissue sections, mucicarmine or Alcian blue stains the capsule of Cryptococcus species<br />

to distinguish it from other yeasts with similar morphologies.<br />

On solid media the cultures are generally mucoid or slimy in appearance; red, orange or<br />

yellow carotenoid pigments may be produced, but young colonies of most species are<br />

usually non-pigmented, and cream in colour. All Cryptococcus species produce urease<br />

and are non-fermentative. Nitrate may be assimilated or not; inositol assimilated. The<br />

genus Cryptococcus differs from the genus Rhodotorula in its inositol assimilation.<br />

Cryptococcosis is a chronic, subacute to acute pulmonary, systemic or meningitic<br />

disease, initiated by the inhalation of infectious propagules (basidiospores and/or<br />

desiccated yeast cells) from the environment. Primary pulmonary infections have no<br />

diagnostic symptoms and are usually subclinical. On dissemination, the fungus usually<br />

shows a predilection for the central nervous system, however skin, bones and other<br />

visceral organs may also become involved. Although C. neoformans and C. gattii are<br />

regarded as the principle pathogenic species, Cryptococcus albidus and C. laurentii<br />

have on occasion also been implicated in human infection.<br />

Molecular Identification: Requires ITS and/or D1/D2 sequencing, particularly for<br />

identification of unusual species.<br />

MALDI-T<strong>OF</strong> MS: Can provide reliable species and subspecies level identification of<br />

Cryptococcus species, but its accuracy is dependent on database quality (Arendrup et<br />

al. 2014).<br />

a<br />

b<br />

5 μm<br />

Cryptococcus neoformans (a) culture appearances on bird seed agar (brown colonies)<br />

and Candida albicans (white colonies) and (b) India Ink preparation of C. neoformans<br />

surrounded by a characteristic wide gelatinous capsule.<br />

References: Rippon (1982), Barnett et al. (1983), Kurtzman et al. (2011), Casadevall<br />

and Perfect (1998), de Hoog et al. (2000, 2015), McTaggart et al. (2013).


Descriptions of Medical Fungi 67<br />

Cryptococcus albidus (Saito) Skinner<br />

Synonymy: Cryptococcus diffluens (Zach) Lodder & Kreger-van Rij.<br />

RG-1 organism.<br />

Culture: Colonies (SDA) are cream-coloured smooth, mucoid, glabrous, yeast-like.<br />

Microscopy: Globose to ovoid budding yeast-like cells, 3.5-8.8 x 5.5-10.2 μm.<br />

Pseudohyphae are absent.<br />

India Ink Preparation: Positive - distinct thin capsules are present.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose v L-Arabinose + D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose v α-M-D-glucoside v<br />

Glucose - Maltose + D-Ribose v D-Gluconate +<br />

Galactose - Cellobiose + L-Rhamnose v DL-Lactate v<br />

Sucrose - Trehalose +,w D-Glucosamine - myo-Inositol +<br />

Maltose - Lactose v N-A-D-glucosamine - 2-K-D-gluconate +<br />

Lactose - Melibiose v Glycerol v D-Glucuronate +<br />

Trehalose - Raffinose + Erythritol v Nitrate +<br />

Assimilation Melezitose + Ribitol v Urease +<br />

Glucose + Soluble Starch v Galactitol v 0.1% Cycloheximide -<br />

Galactose v D-Xylose + D-Mannitol + Growth at 37 O C v<br />

Cryptococcus albidus has variable growth at 37 O C, and infections in humans are<br />

rare. Along with C. laurentii, C. albidus accounts for 80% of the non-neoformans/gattii<br />

infections. Impaired cellular immunity is the most common risk factor with HIV infection<br />

and low CD4 counts a common comorbidity.<br />

Antifungal Susceptibility: C. albidus (Australian National data); MIC µg/mL.<br />

Note: All Cryptococcus species are intrinsically resistant to echinocandins.<br />

No. 64<br />

AmB 1 1<br />

FLU 1 1<br />

VORI 1 1<br />

POSA 1 1<br />

ITRA 1 1<br />

5FC 1 1


68<br />

Descriptions of Medical Fungi<br />

Cryptococcus gattii (Vanbreus. & Takashio) Kwon-Chung & Boekhout<br />

Synonymy: Filobasidiella bacillispora Kwon-Chung.<br />

Cryptococcus neoformans var. gattii Vanbreus. & Takashio.<br />

RG-2 organism.<br />

Cryptococcus gattii has two serotypes (B and C) and was reclassified as a separate<br />

species from C. neoformans in 2002 (Kwon-Chung et al. 2002). C. gattii generally has a<br />

more restricted geographical distribution than C. neoformans, causing human disease<br />

in climates ranging from temperate to tropical Australia, Papua New Guinea, parts of<br />

Africa, India, Southeast Asia, Mexico, Brazil, Paraguay and Southern California, although<br />

recent infections have also been reported from Vancouver Island, Canada and in the<br />

Pacific Northwest, USA (Pfaller & Diekema, 2010, Espinel-Ingroff and Kidd, 2015). C.<br />

gattii has a specific ecological association with numerous species of Eucalyptus trees,<br />

although the Canadian isolates are associated with a range of native non-Eucalyptus<br />

species (Kidd et al. 2007). Historically considered a pathogen in immunocompetent<br />

hosts, a recent review in Australia noted an increase in C. gattii infections in HIVnegative<br />

immunocompromised patients (Chen et al. 2012). Cryptococcosis caused by<br />

C. gattii is often associated with large mass lesions (cryptococcomas) in the lung and/<br />

or brain (Sorrell, 2001).<br />

Canavanine glycine bromothymol blue (CGB) agar (Kwon-Chung et al. 1982) is the<br />

media of choice to differentiate C. gattii from C. neoformans. This simple biotype test is<br />

based on the ability of C. gattii isolates to grow in the presence of L-canavanine and to<br />

assimilate glycine as a sole carbon source. A heavy inoculum is important.<br />

Cryptococcus gattii turns CGB agar blue within 2-5 days;<br />

Cryptococcus neoformans does not grow on this medium


Descriptions of Medical Fungi 69<br />

Cryptococcus gattii (Vanbreus. & Takashio) Kwon-Chung & Boekhout<br />

Culture: Colonies (SDA) cream-coloured smooth, mucoid, yeast-like colonies.<br />

Microscopy: Globose to ovoid budding yeast-like cells 3.0-7.0 x 3.3- 7.9 µm.<br />

India Ink Preparation: Positive - distinct, wide gelatinous capsules are present. Some<br />

strains may not produce apparent capsules from culture.<br />

Dalmau Plate Culture: Budding yeast cells only. No pseudohyphae present.<br />

Bird Seed Agar: Colonies turn dark brown in colour as colonies selectively absorb a<br />

brown pigment from this media. Colonies are often more mucoid when compared with<br />

C. neoformans (Staib, 1987).<br />

Canavanine Glycine Bromothymol Blue (CGB) Agar: Turns blue within 2-5 days.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose - L-Arabinose +,w D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose + M-D-glucoside +<br />

Glucose - Maltose + D-Ribose v D-Gluconate +<br />

Galactose - Cellobiose +,w L-Rhamnose + DL-Lactate -<br />

Sucrose - Trehalose + D-Glucosamine v myo-Inositol +<br />

Maltose - Lactose - N-A-D-glucosamine v 2-K-D-gluconate nd<br />

Lactose - Melibiose - Glycerol - D-Glucuronate nd<br />

Trehalose - Raffinose +,w Erythritol - Nitrate -<br />

Assimilation Melezitose + Ribitol v Urease +<br />

Glucose + Soluble Starch + Galactitol + 0.1% Cycloheximide -<br />

Galactose + D-Xylose + D-Mannitol + Growth at 37 O C +<br />

Key Features: Encapsulated yeast; absence of pseudohyphae; growth at 37 O C;<br />

positive hydrolysis of urea; negative fermentation of sugars and positive assimilation<br />

of glucose, maltose, sucrose, galactose, trehalose, raffinose, inositol, cellobiose,<br />

rhamnose, arabinose, melezitose and xylose, and negative assimilation of nitrate,<br />

lactose, melibiose, erythritol and soluble starch; growth on bird seed (Guizotia<br />

abyssinica seed) or caffeic acid agar - colonies turn a dark brown colour; growth on<br />

CGB agar turning it blue within 2-5 days.<br />

Antifungal Susceptibility: C. gattii (Australian National data); MIC µg/mL.<br />

Note: All Cryptococcus species are intrinsically resistant to echinocandins.<br />

No. 64<br />

AmB 152 1 3 18 52 55 12 10 1<br />

FLU 152 2 13 29 48 38 19 2 1<br />

VORI 130 5 12 32 44 20 13 4<br />

POSA 90 2 9 12 23 26 16 1 1<br />

ITRA 152 7 22 38 61 23 1<br />

5FC 152 2 10 38 52 39 7 4


70<br />

Descriptions of Medical Fungi<br />

RG-1 organism.<br />

Cryptococcus laurentii (Kufferath) Skinner<br />

Culture: Colonies (SDA) are cream-coloured, often becoming a deeper orange-yellow<br />

with age, with a smooth mucoid texture.<br />

Microscopy: Spherical and elongated budding yeast-like cells or blastoconidia, 2.0-<br />

5.5 x 3.0-7.0 μm. No pseudohyphae present.<br />

India Ink Preparation: Positive - narrow but distinct capsules are present.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose v L-Arabinose + D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose + α-M-D-glucoside +<br />

Glucose - Maltose + D-Ribose + D-Gluconate +<br />

Galactose - Cellobiose + L-Rhamnose + DL-Lactate v<br />

Sucrose - Trehalose + D-Glucosamine - myo-Inositol +<br />

Maltose - Lactose + N-A-D-glucosamine - 2-K-D-gluconate +<br />

Lactose - Melibiose + Glycerol v D-Glucuronate +<br />

Trehalose - Raffinose + Erythritol v Nitrate -<br />

Assimilation Melezitose + Ribitol + Urease +<br />

Glucose + Soluble Starch v Galactitol + 0.1% Cycloheximide -<br />

Galactose v D-Xylose + D-Mannitol + Growth at 37 O C -,w<br />

Note: Some strains of C. laurentii may develop brown pigment on bird seed agar and<br />

turn CGB media blue, similar to C. gattii, however C. laurentii assimilates both lactose<br />

and melibiose while C. gattii does not. Along with C. albidus, C. laurentii accounts for<br />

80% of the non-neoformans/gattii infections. Impaired cellular immunity is the most<br />

common risk factor with HIV infection and low CD4 counts a common comorbidity.<br />

Invasive devices are an additional risk factor.<br />

Antifungal Susceptibility: C. laurentii (Australian National data); MIC µg/mL.<br />

Note: All Cryptococcus species are intrinsically resistant to echinocandins.<br />

No. 64<br />

AmB 7 1 2 3 1<br />

FLU 6 1 1 3 1<br />

VORI 6 2 2 1 1<br />

POSA 5 1 1 2 1<br />

ITRA 7 1 2 2 1 1<br />

5FC 7 7


Descriptions of Medical Fungi 71<br />

Cryptococcus neoformans (Sanfelice) Vuillemin<br />

Synonymy: Filobasidiella neoformans Kwon-Chung.<br />

Cryptococcus neoformans var. neoformans (San Felice) Vuill.<br />

RG-2 organism.<br />

This species comprises two varieties: C. neoformans var. grubii (serotype A) and C.<br />

neoformans var. neoformans (serotype D).<br />

C. neoformans var. grubii has a worldwide distribution, causing 95% of all C. neoformans<br />

infections. It has been isolated from various sources in nature and is noted for its<br />

association with accumulations of avian guano, especially with pigeon excreta. The<br />

fungus has also been isolated from the dung of caged birds including canaries, parrots<br />

and budgerigars. Other environmental isolations of C. neoformans var. grubii include<br />

rotting vegetables, fruits and fruit juices, wood, dairy products and soil.<br />

C. neoformans var. neoformans has a more restricted distribution with infections being<br />

more prevalent in Europe, including France, Italy and Denmark, where it accounts<br />

for 30% of isolates. Moreover, C. neoformans var. neoformans infections are more<br />

strongly correlated with older patients, the skin, and the use of corticosteroids (Franzot<br />

et al. 1999).<br />

Culture: Colonies (SDA) cream-coloured smooth, mucoid, yeast-like colonies.<br />

Microscopy: Globose to ovoid budding yeast-like cells 3.0-7.0 x 3.3-7.9 µm.<br />

India Ink Preparation: Positive - distinct, wide gelatinous capsules are present on<br />

direct microscopy. Some strains may not produce apparent capsules from culture.<br />

Dalmau Plate Culture: Budding yeast cells only. No pseudohyphae present.<br />

Bird Seed Agar: Colonies turn dark brown in colour as colonies selectively absorb a<br />

brown pigment from this media (Staib, 1987).<br />

Canavanine Glycine Bromothymol Blue (CGB) Agar: No growth or colour change.<br />

Creatinine Dextrose Bromothymol Blue Thymine (CDBT) Agar may be used<br />

to differentiate C. neoformans var. neoformans and C. neoformans var. grubii. C.<br />

neoformans var. neoformans grows as bright red colonies, turning the medium a bright<br />

orange after 5 days while no colour change is observed for C. neoformans var. grubii<br />

(Irokanulo et al. 1994).


72<br />

Descriptions of Medical Fungi<br />

Cryptococcus neoformans (Sanfelice) Vuillemin<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose - L-Arabinose +,w D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose + α-M-D-glucoside +<br />

Glucose - Maltose + D-Ribose v D-Gluconate +<br />

Galactose - Cellobiose +,w L-Rhamnose + DL-Lactate -<br />

Sucrose - Trehalose + D-Glucosamine v myo-Inositol +<br />

Maltose - Lactose - N-A-D-glucosamine v 2-K-D-gluconate nd<br />

Lactose - Melibiose - Glycerol - D-Glucuronate nd<br />

Trehalose - Raffinose +,w Erythritol - Nitrate -<br />

Assimilation Melezitose + Ribitol v Urease +<br />

Glucose + Soluble Starch + Galactitol + 0.1% Cycloheximide -<br />

Galactose + D-Xylose + D-Mannitol + Growth at 37 O C +<br />

Key Features: Encapsulated yeast; absence of pseudohyphae; growth at 37 O C;<br />

positive hydrolysis of urea; negative fermentation of sugars and positive assimilation<br />

of glucose, maltose, sucrose, galactose, trehalose, raffinose, inositol, cellobiose,<br />

rhamnose, arabinose, melezitose and xylose, and negative assimilation of nitrate,<br />

lactose, melibiose, erythritol and soluble starch; growth on bird seed (Guizotia<br />

abyssinica seed) or caffeic acid agar - colonies turn a dark brown colour; does not<br />

grow on CGB agar (no colour change).<br />

Antifungal Susceptibility: C. neoformans (Australian National data); MIC µg/mL.<br />

Note: All Cryptococcus species are intrinsically resistant to echinocandins.<br />

No. 64<br />

AmB 236 2 5 17 82 57 57 16<br />

FLU 236 1 10 32 40 65 65 16 6 1<br />

VORI 197 27 25 54 63 23 4 1<br />

POSA 152 1 9 29 39 46 24 4<br />

ITRA 236 2 10 37 72 90 20 4 1<br />

5FC 236 1 1 5 20 45 66 61 33 2 2


Descriptions of Medical Fungi 73<br />

Cunninghamella bertholletiae Stadel<br />

Synonymy: Cunninghamella elegans Lendner.<br />

Cunninghamella echinulata var. elegans (Lendner) Lunn & Shipton.<br />

The genus Cunninghamella is characterised by white to grey, rapidly growing colonies,<br />

producing erect, straight, branching sporangiophores. These sporangiophores end in<br />

globose or pyriform-shaped vesicles from which several one-celled, globose to ovoid,<br />

echinulate or smooth-walled sporangiola develop on swollen denticles. Chlamydospores<br />

and zygospores may also be present.<br />

Cunninghamella species are mainly soil fungi of the Mediterranean and subtropical<br />

zones; they are only rarely isolated in temperate regions. The genus now contains<br />

seven species with C. bertholletiae the only known species to cause disease in humans<br />

and animals, often in association with trauma and immunosuppression.<br />

RG-2 organism.<br />

Morphological Description: Colonies are very fast growing, white at first, but<br />

becoming dark grey and powdery with sporangiola development. Sporangiophores<br />

up to 20 μm wide, straight, with verticillate or solitary branches. Vesicles subglobose<br />

to pyriform, the terminal ones up to 40 µm and the lateral ones 10-30 µm in diameter.<br />

Sporangiola are globose (7-11 µm diameter), or ellipsoidal (9-13 x 6-10 μm), verrucose<br />

or short-echinulate, hyaline singly but brownish in mass. Temperature: optimum 25-<br />

30 O C, maximum up to 50 O C.<br />

Key Features: Mucorales, clinical isolates grow at 40 O C, one-celled, globose to ovoid,<br />

echinulate sporangiola borne on swollen terminal or lateral globose to clavate fertile<br />

vesicles.<br />

Molecular Identification: ITS sequencing is recommended (Yu et al. 2015).<br />

References: McGinnis (1980), Weitzman and Crist (1980), Weitzman (1984), Lunn<br />

and Shipton (1983), Domsch et al. (1980), Samson (1969), de Hoog et al. (2000,<br />

2015), Ellis (2005b), Zheng and Chen (2001).<br />

Antifungal Susceptibility: C. bertholletiae (Espinel-Ingroff et al. 2015a, includes<br />

Australian data); MIC µg/mL.<br />

No. 16<br />

AmB 32 1 1 5 16 8 1<br />

POSA 30 4 18 8<br />

ITRA 25 4 4 10 7


74<br />

Descriptions of Medical Fungi<br />

Cunninghamella bertholletiae Stadel<br />

10 μm<br />

10 μm<br />

Cunninghamella bertholletiae showing simple sporangiophores forming a swollen,<br />

terminal vesicle around which single-celled, globose to ovoid sporangiola develop<br />

on swollen denticles.


Descriptions of Medical Fungi 75<br />

The genus Curvularia contains about 80 species, which are mostly soil or plant<br />

pathogens. Recent studies have shown that morphological identification does not<br />

correlate with molecular identification (Manamgoda et al. 2012; Yanagihara et al. 2010,<br />

da Cunha et al. 2013). A phylogenetic analysis of the genera Bipolaris and Curvularia<br />

has resulted in a re-alignment of several species. In particular, clinical isolates previously<br />

identified as Bipolaris species, notably B. australiensis, B. hawaiiensis and B. spicifera<br />

have now been transferred to Curvularia (Manamgoda et al. 2012).<br />

Previously Curvularia lunata was the most frequently reported clinical species, however<br />

other species, such as C. americana, C. brachyspora, C. chlamydospora, C. clavata,<br />

C. hominis, C. inaequalis, C. muehlenbeckiae, C. pallescens, C. pseudolunata, C.<br />

senegalensis and C. verruculosa have now also been reported from clinical cases<br />

(Revankar and Sutton, 2010, da Cunha et al. 2013, Madrid et al. 2014).<br />

RG-1 organisms.<br />

Curvularia Boedijn<br />

Morphological Description: Colonies are fast growing, suede-like to downy, brown to<br />

blackish brown with a black reverse. Conidiophores erect, straight to flexuous, septate,<br />

often geniculate (producing conidia in sympodial succession) sometimes nodulose.<br />

Conidia are ellipsoidal, often curved or lunate, rounded at the ends or sometimes<br />

tapering slightly towards the base, pale brown, medium reddish brown to dark brown,<br />

3–10 (usually 3–5) septa, conidial wall smooth to verrucose. Hilum protuberant in some<br />

species.<br />

Key Features: Dematiaceous hyphomycete producing sympodial, pale brown,<br />

cylindrical or slightly curved phragmoconidia.<br />

Comment: There is no clear morphological boundary between the genera Bipolaris<br />

and Curvularia and some species show intermediate morphology (Manamgoda et al.<br />

2012). ITS and GPDH gene analysis is recommended for definitive identification of<br />

species (Manamgoda et al. 2012).<br />

Molecular Identification: GPDH (Manamgoda et al. 2012) and/or ITS (da Cunha et<br />

al. 2013, Irinyi et al. 2015).<br />

References: Ellis (1971), McGinnis (1980), Rippon (1988), de Hoog et al. (2000, 2015),<br />

Revankar and Sutton (2010), Yanagihara et al. (2010), Manamgoda et al. (2012); da<br />

Cunha et al. (2013), Madrid et al. (2014).<br />

Antifungal Susceptibility: Curvularia australiensis (da Cunha et al. 2012a and<br />

Australian National data); MIC µg/mL.<br />

No 16<br />

AmB 13 1 1 6 3 2<br />

VORI 13 1 2 6 3 1<br />

POSA 11 1 2 6 2<br />

ITRA 13 1 3 2 6 1


76<br />

Descriptions of Medical Fungi<br />

Curvularia australiensis (M.B. Ellis) Manamgoda, L.Cai. & K.D. Hyde<br />

Synonymy: Bipolaris australiensis (M.B. Ellis) Tsuda & Ueyama.<br />

RG-1 organism.<br />

Morphological Description: Colonies grey to blackish-brown, suede-like.<br />

Conidiophores brown, solitary, flexuose or geniculate, smooth-walled, up to 150 µm<br />

long, mostly 3-7 µm wide. Conidia straight, rounded at the ends, pale brown to mid<br />

reddish-brown, mostly 3-, rarely 4-5-distoseptate, ellipsoidal or oblong, 14-40 x 6-11<br />

µm, smooth-walled to finely roughened.<br />

10 μm<br />

Curvularia australiensis showing sympodial development of pale brown, fusiform to<br />

ellipsoidal, pseudoseptate, conidia on a geniculate or zig-zag rachis.


Descriptions of Medical Fungi 77<br />

Curvularia hawaiiensis (Bugnic.) Manamgoda, L. Cai & K.D. Hyde<br />

Synonymy: Bipolaris hawaiiensis Bugnic.<br />

RG-1 organism.<br />

Morphological Description: Colonies powdery to hairy, black. Conidiophores erect,<br />

unbranched, septate, apically flexuose with flat conidial scars on the edges, up to 80<br />

µm long. Conidia smooth- and rather thick-walled, brown, with (3-) 5 (-7) distosepta,<br />

cylindrical to cigar-shaped, 18-35 × 6-9 µm.<br />

Antifungal Susceptibility: C. hawaiiensis (da Cunha et al. 2012a); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 0.125-0.25 0.25 VORI 0.25-2 1<br />

ITRA


78<br />

Descriptions of Medical Fungi<br />

Curvularia lunata Boedijn<br />

Antifungal Susceptibility: Curvularia lunata (Australian National data); MIC µg/<br />

mL.<br />

No 16<br />

AmB 7 1 3 1 1 1<br />

VORI 7 1 2 3 1<br />

POSA 5 1 3 1<br />

ITRA 7 1 1 3 1 1<br />

Curvularia spicifera (Bainier) Boedijn<br />

Synonymy: Bipolaris spicifera Bainier.<br />

RG-1 organism.<br />

Morphological Description: Colonies appearing glassy with sooty powder of conidia,<br />

or hairy if sporulation is poor. Conidiophores erect, unbranched, septate, up to 250 µm<br />

long and 4-8 µm wide, regularly zig-zagged in the apical part, with flat, dark brown scars<br />

on the edges. Conidia brown, cylindrical with rounded ends, medium brown except for<br />

narrow subhyaline spots at the extremites, 20-40 × 9-14 µm, with 3 distosepta.<br />

Antifungal Susceptibility: C. spicifera (da Cunha et al. 2012a); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB


Descriptions of Medical Fungi 79<br />

Cyberlindnera fabianii (Wick.) Minter<br />

Synonymy: Candida fabianii (Hartmann) S.A. Meyer & Yarrow.<br />

Cyberlindnera fabianii is a rare cause of candidaemia.<br />

RG-1 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Spheroidal to ellipsoidal budding blastoconidia, 3.0-6.5 x 2-5.5 µm in<br />

size. No pseudohyphae are produced. Asci when present, are spherical, containing<br />

one to four spherical, finely roughened ascospores.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Spherical to ovoid budding yeast cells and occasional<br />

pseudohyphae produced.<br />

Molecular Identification: ITS sequencing recommended.<br />

MALDI-T<strong>OF</strong> MS: Able to accurately identify this species.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose - L-Arabinose - D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose - α-M-D-glucoside +<br />

Glucose + Maltose + D-Ribose - D-Gluconate +<br />

Galactose - Cellobiose + L-Rhamnose - DL-Lactate +<br />

Sucrose + Trehalose + D-Glucosamine - myo-Inositol -<br />

Maltose +,s Lactose - N-A-D-glucosamine - 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate nd<br />

Trehalose - Raffinose + Erythritol - Nitrate +<br />

Assimilation Melezitose + Ribitol - Urease -<br />

Glucose + Soluble Starch + Galactitol - 0.1% Cycloheximide nd<br />

Galactose - D-Xylose + D-Mannitol + Growth at 37 O C +<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern. Molecular<br />

identification may be required.<br />

Antifungal Susceptibility: C. fabianii very limited data (Pfaller et al. 2015, Australian<br />

National data); MIC µg/mL.<br />

No. 64<br />

AmB 1 1<br />

FLU 1 1<br />

VORI 5 1 3 1<br />

POSA 5 3 2<br />

ITRA 1 1<br />

MICA 4 2 1 1<br />

CAS 1 1<br />

5FC 1 1


80<br />

Descriptions of Medical Fungi<br />

Cylindrocarpon Wollenw.<br />

The genus contains 35 species, mostly from soil and as an occasional human and<br />

animal pathogen. Cylindrocarpon differs from Fusarium by lacking an asymmetrical<br />

foot-cell on the macroconidia.<br />

RG-2 organism (if isolated from humans).<br />

Morphological Description: Colonies are fast growing, hyaline or bright-coloured,<br />

suede-like or woolly. Sporodochia may occasionally be present. Conidiophores<br />

consist of simple or repeatedly verticillate phialides, arranged in brush-like structures.<br />

Phialides are cylindrical to subulate, with small collarettes producing hyaline, smoothwalled<br />

conidia, arranged in slimy masses. Two types of conidia may be produced: (1)<br />

macroconidia which are one to several septate, hyaline, straight or curved, cylindrical<br />

to fusiform, with a rounded apex and flat base; and (2) microconidia which are onecelled,<br />

and usually clearly distinct from the macroconidia. Chlamydospores may be<br />

present or absent, hyaline to brown, spherical, formed singly, in chains or in clumps,<br />

intercalary or terminal.<br />

References: Booth (1966), Domsch et al. (1980), de Hoog et al. (2000, 2015).<br />

a<br />

b<br />

15 µm<br />

c<br />

15 µm<br />

Cylindrocarpon spp. (a) culture, (b) chlamydospores and (c) macroconidia.


Descriptions of Medical Fungi 81<br />

Debaryomyces hansenii (Zopf) Lodder & Kreger-van Rij.<br />

Synonymy: Candida famata (Harrison) S.A. Meyer & Yarrow.<br />

Debaryomyces hansenii is a common environmental isolate. It may be isolated from<br />

human skin, and is only rarely recovered from blood stream infections. RG-1 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ovoid to broadly ellipsoidal budding blastoconidia, 3.5-5 x 2-3.5 µm<br />

in size. No pseudohyphae produced. Asci when present are spherical, persistent,<br />

containing one to two spherical ascospores with rough walls.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Spherical to ovoid budding yeast cells only. Pseudohyphae are<br />

usually lacking.<br />

Molecular Identification: ITS sequencing recommended.<br />

MALDI-T<strong>OF</strong> MS: Able to accurately identify this species.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose v L-Arabinose +,w D-Glucitol +,w<br />

Fermentation Sucrose + D-Arabinose v α-M-D-glucoside +<br />

Glucose -,w Maltose + D-Ribose v D-Gluconate +,w<br />

Galactose -,w Cellobiose + L-Rhamnose v DL-Lactate v<br />

Sucrose -,w Trehalose + D-Glucosamine v myo-Inositol -<br />

Maltose -,w Lactose v N-A-D-glucosamine v 2-K-D-gluconate +<br />

Lactose - Melibiose v Glycerol + D-Glucuronate v<br />

Trehalose -,w Raffinose + Erythritol v Nitrate -<br />

Assimilation Melezitose v Ribitol + Urease -<br />

Glucose + Soluble Starch v Galactitol v 0.1% Cycloheximide v<br />

Galactose + D-Xylose + D-Mannitol + Growth at 40 O C -<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern.<br />

Antifungal Susceptibility: D. hansenii (Diekema et al. 2009, Beyda et al. 2013,<br />

Espinel-Ingroff et al. 2014, Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 27 1 3 7 13 2 1<br />

FLU 76 9 18 13 9 8 11 2 3 3<br />

VORI 79 8 11 27 16 9 4 2 1 1<br />

POSA 72 8 11 20 12 8 8 2 3<br />

ITRA 11 2 1 3 2 1 2<br />

ANID 23 1 2 1 1 0 8 10<br />

MICA 23 1 3 1 2 8 5 3<br />

CAS 26 1 2 3 8 5 4 1 2<br />

5FC 11 1 5 1 1 1 2


82<br />

Descriptions of Medical Fungi<br />

McGinnis et al. (1986b) have reviewed the isolates from human and animal disease<br />

purported to be Drechslera or Helminthosporium and concluded that all pathogenic<br />

isolates examined actually belong to the genera Bipolaris or Exserohilum.<br />

RG-1 organism.<br />

Drechslera Ito<br />

Morphological Description: Colonies are fast growing, suede-like to downy, brown<br />

to blackish brown with a black reverse. Conidia are pale to dark brown, usually<br />

cylindrical or subcylindrical, straight, smooth-walled, and are formed apically through<br />

a pore (poroconidia) in a sympodially elongating, geniculate conidiophore. Conidia<br />

are transversely septate (phragmoconidia), with the septum delimiting the basal cell<br />

formed first during conidium maturation. Germinating is from any or all cells and the<br />

hilum is not protuberant.<br />

Key Features: Dematiaceous hyphomycete producing sympodial, pale brown,<br />

cylindrical or subcylindrical, transversely septate poroconidia.<br />

References: Luttrell (1978), Ellis (1971, 1976), McGinnis (1980), McGinnis et al.<br />

(1986b), Sivanesan (1987), Rippon (1988), de Hoog et al. (2000, 2015). Also see<br />

Descriptions for Bipolaris, Curvularia and Exserohilum.<br />

20 µm<br />

Drechslera spp. conidia.


Descriptions of Medical Fungi 83<br />

Synonymy: Epicoccum purpurascens Ehrenb. ex Schlecht.<br />

Epicoccum nigrum is a cosmopolitan saprophyte of worldwide distribution which is<br />

occasionally isolated as a contaminant from clinical specimens like skin.<br />

RG-1 organism.<br />

Epicoccum nigrum Link<br />

Morphological Description: Colonies are fast growing, suede-like to downy, with a<br />

strong yellow to orange-brown diffusible pigment. When sporulating, numerous black<br />

sporodochia (aggregates of conidiophores) are visible. Conidia are formed singly on<br />

densely compacted, non-specialised, determinant, slightly pigmented conidiophores.<br />

Conidia are globose to pyriform, mostly 15-25 µm diameter with a funnel-shaped base<br />

and broad attachment scar, often seceding with a protuberant basal cell; i.e. aleuric or<br />

rhexolytic dehiscence of conidia. Conidia become multicellular (dictyoconidia), darkly<br />

pigmented and have a verrucose external surface.<br />

Key Features: Dematiaceous hyphomycete producing darkly pigmented, large globose<br />

to pyriform, verrucose dictyoconidia on a sporodochium.<br />

References: Ellis (1971), Domsch et al. (1980), McGinnis (1980), Samson et al. (1995).<br />

a<br />

b<br />

15 µm<br />

Epicoccum nigrum (a) culture and (b) conidia.


84<br />

Descriptions of Medical Fungi<br />

Epidermophyton floccosum (Harz) Langeron et Milochevitch<br />

Epidermophyton floccosum is an anthropophilic dermatophyte with a worldwide<br />

distribution which often causes tinea pedis, tinea cruris, tinea corporis and<br />

onychomycosis. It is not known to invade hair in vivo and no specific growth requirements<br />

have been reported.<br />

RG-2 organism.<br />

Morphological Description: Colonies are usually slow growing, greenishbrown<br />

or khaki-coloured with a suede-like surface, raised and folded in the centre,<br />

with a flat periphery and submerged fringe of growth. Older cultures may develop<br />

white pleomorphic tufts of mycelium. A deep yellowish-brown reverse pigment is<br />

usually present. Microscopic morphology shows characteristic smooth, thin-walled<br />

macroconidia which are often produced in clusters growing directly from the hyphae.<br />

Numerous chlamydospores are formed in older cultures. Microconidia are not formed.<br />

Key Features: Culture characteristics, microscopic morphology and clinical disease.<br />

References: Rebell and Taplin (1970), Mackenzie et al. (1987), Rippon (1988), de<br />

Hoog et al. (2000, 2015).<br />

a<br />

b<br />

20 µm<br />

c<br />

15 µm<br />

Epidermophyton floccosum (a) culture, (b) chlamydospores and (c) macroconidia.


Descriptions of Medical Fungi 85<br />

Exophiala dermatitidis has been isolated from plant debris and soil and is a recognised<br />

causative agent of mycetoma and phaeohyphomycosis in humans (Zeng et al. 2007).<br />

RG-2 organism.<br />

Exophiala dermatitidis (Kano) de Hoog<br />

Morphological Description: Colonies are slow growing, initially yeast-like and black,<br />

becoming suede-like, olivaceous-grey with the development of aerial mycelium with<br />

age. A brown pigment is often produced in the agar. The initial yeast-like phase is<br />

characterised by unicellular, ovoid to elliptical, budding yeast-like cells. The yeastlike<br />

cells are hyaline and thin-walled when young becoming darkly pigmented<br />

(dematiaceous) and thick-walled when mature. With the development of mycelium,<br />

flask-shaped to cylindrical annellides are produced. Conidia are hyaline to pale brown,<br />

one-celled, round to obovoid, 2-4 x 2.5-6 µm, smooth-walled and accumulate in slimy<br />

balls at the apices of the annellides or down their sides. Cultures grow at 42 O C and on<br />

media containing 0.1% cycloheximide.<br />

Molecular Identification: ITS and/or D1/D2 sequencing is recommended for species<br />

identification (Halliday et al. 2015).<br />

References: de Hoog and Hermanides-Nijhof (1977), McGinnis (1980), Hohl et al.<br />

(1983), Nishimura and Miyaji (1983), Matsumoto et al. (1984), Dixon and Polak-Wyss<br />

(1991), de Hoog et al. (2000, 2015).<br />

10 µm<br />

Exophiala dermatitidis annellides and conidia.


86<br />

Descriptions of Medical Fungi<br />

Exophiala dermatitidis (Kano) de Hoog<br />

Antifungal Susceptibility: E. dermatitidis (Duarte et al. 2013 and Australian<br />

National data); MIC µg/mL.<br />

No. 64<br />

AmB 44 1 2 25 16<br />

VORI 44 4 12 20 1 4 2 1<br />

POSA 1 1<br />

ITRA 44 10 13 6 9 4 1 1<br />

E. dermatitidis data from 27 isolates (Fothergill et al. 2009); MIC µg/mL.<br />

AmB Range 0.125-1; MIC 90<br />

= 1 VORI Range 0.06-1; MIC 90<br />

= 0.25<br />

ITRA Range


Descriptions of Medical Fungi 87<br />

RG-2 organism.<br />

Exophiala jeanselmei complex<br />

Exophiala jeanselmei McGinnis & Padhye<br />

Morphological Description: Colonies are initially smooth, greenish-grey to black,<br />

mucoid and yeast-like, becoming raised and developing tufts of aerial mycelium with<br />

age, often becoming dome-shaped and suede-like in texture. Reverse is olivaceousblack.<br />

Numerous ellipsoidal, yeast-like, budding cells are usually present, especially in<br />

young cultures. Scattered amongst these yeast-like cells are larger, inflated, subglobose<br />

to broadly ellipsoidal cells (germinating cells) which give rise to short torulose hyphae<br />

that gradually change into unswollen hyphae. Conidia are formed on lateral pegs either<br />

arising apically or laterally at right or acute angles from essentially undifferentiated<br />

hyphae or from strongly inflated detached conidia. Conidiogenous pegs are 1-3 µm<br />

long, slightly tapering and imperceptibly annellate. Conidia are hyaline, smooth, thinwalled,<br />

broadly ellipsoidal, 3.2-4.4 x 1.2-2.2 µm, and with inconspicuous basal scars.<br />

Cultures grow at 37 O C but not at 40 O C.<br />

a<br />

5 µm<br />

b<br />

15 µm<br />

5 µm<br />

Exophiala jeanselmei complex showing annellides, conidia and conidiogenous<br />

pegs (annellides) on (a) yeast-like cells and (b) torulose hyphae.<br />

References: de Hoog and Hermanides-Nijhof (1977), de Hoog (1977, 1985), McGinnis<br />

and Padhye (1977), McGinnis (1978b, 1980), Domsch et al. (1980), Nishimura and<br />

Miyaji (1983), Matsumoto et al. (1987), Dixon and Polak-Wyss (1991), Badali et al.<br />

2010, de Hoog et al. (2000, 2003, 2006, 2015).


88<br />

Descriptions of Medical Fungi<br />

Exophiala jeanselmei complex<br />

Antifungal Susceptibility: E. jeanselmei limited data (Australian National data);<br />

MIC µg/mL.<br />

No. 64<br />

AmB 9 1 3 2 2 1<br />

VORI 9 2 2 1 1 1 2<br />

POSA 6 1 2 2 1<br />

ITRA 9 2 3 2 2<br />

E. jeanselmei data from eight isolates (Fothergill et al. 2009); MIC µg/mL.<br />

AmB Range 0.125-1; MIC 90<br />

= 1 VORI Range 0.06-0.5; MIC 90<br />

= 0.5<br />

ITRA Range


Descriptions of Medical Fungi 89<br />

Exophiala spinifera complex<br />

Synonymy: Phialophora spinifera Nielsen & Conant<br />

Rhinocladiella spinifera (Nielsen & Conant) de Hoog<br />

E. spinifera has a worldwide distribution and is a recognised causative agent of<br />

mycetoma and phaeohyphomycosis in humans. Zeng et al. (2007) presented an<br />

overview of the medically important Exophiala species.<br />

Recent molecular studies have re-examined Exophiala spinifera and have recognised<br />

two species: E. spinifera and E. attenuata (Vitale and de Hoog, 2002). These two<br />

species are morphologically very similar and can best be distinguished by genetic<br />

analysis.<br />

Molecular Identification: ITS sequencing is recommended for accurate species<br />

identification (Zeng and de Hoog, 2008).<br />

Morphological Description: Conidiogenous cells are predominately annellidic and<br />

erect, multicellular conidiophores that are darker than the supporting hyphae are<br />

present. No growth at 40 O C.<br />

E. spinifera Annellated zones are long with clearly visible, frilled annellations.<br />

E. attenuata Annellated zones are inconspicuous and degenerate.<br />

RG-2 organism.<br />

Exophiala spinifera (Nielsen & Conant) McGinnis<br />

Morphological Description: Colonies are initially mucoid and yeast-like, black,<br />

becoming raised and developing tufts of aerial mycelium with age, finally becoming<br />

suede-like to downy in texture. Reverse is olivaceous-black. Conidiophores are<br />

simple or branched, erect or sub-erect, spine-like with rather thick brown pigmented<br />

walls. Conidia are formed in basipetal succession on lateral pegs either arising<br />

apically or laterally at right or acute angles from the spine-like conidiophores or from<br />

undifferentiated hyphae. Conidiogenous pegs are 1-3 µm long, slightly tapering and<br />

imperceptibly annellate. Conidia are one-celled, subhyaline, smooth, thin-walled, subglobose<br />

to ellipsoidal, 1.0-2.9 x 1.8-2.5 µm, and aggregate in clusters at the tip of each<br />

annellide. Toruloid hyphae and yeast-like cells with secondary conidia are typically<br />

present.<br />

Note: Yeast cells show the presence of capsules in India Ink stained mounts and<br />

cultures will grow on media containing 0.1% cycloheximide. No growth at 40 O C.<br />

References: de Hoog and Hermanides-Nijhof (1977), McGinnis and Padhye (1977),<br />

Domsch et al. (1980), McGinnis (1980), Nishimura and Miyaji (1983), de Hoog (1985),<br />

Matsumoto et al. (1987), Dixon and Polak-Wyss (1991), de Hoog et al. (2000, 2003,<br />

2006, 2015).


90<br />

Descriptions of Medical Fungi<br />

Exophiala spinifera complex<br />

a<br />

b<br />

10 µm<br />

Exophiala spinifera (a) culture showing typical black, mucoid yeast-like growth, and<br />

(b) conidia being formed on erect, multiseptate conidiophores that are darker than the<br />

supporting hyphae, with long annellated zones.<br />

Antifungal Susceptibility: E. spinifera limited data (Australian National data); MIC<br />

µg/mL.<br />

No. 64<br />

AmB 5 1 2 1 1<br />

VORI 5 1 2 2<br />

POSA 4 1 3<br />

ITRA 5 1 4<br />

E. spinifera data from eight isolates (Fothergill et al. 2009); MIC µg/mL.<br />

AmB Range 0.25-1; MIC 90<br />

= 1 VORI Range 0.06-0.5; MIC 90<br />

= 0.25<br />

ITRA Range


Descriptions of Medical Fungi 91<br />

The genus Exserohilum contains about 35 species and may be differentiated from<br />

the closely related genera Bipolaris and Dreschlera by forming conidia with a strongly<br />

protruding truncate hilum (i.e. exserted hilum). In Drechslera species, the hilum does<br />

not protrude; in Bipolaris species the hilum protrudes only slightly. Several species<br />

of Exserohilum have been reported as agents of phaeohyphomycosis, notably E.<br />

rostratum, E. mcginnisii and E. longirostratum. Although recent molecular studies have<br />

demonstrated that the latter two species are probable synonyms of E. rostratum (de<br />

Cunha et al. 2012b, Katragkou et al. 2014). E. rostratum was recently implicated in an<br />

outbreak of fungal meningitis associated with contaminated methylprednisolone in the<br />

United States.<br />

RG-1 organism.<br />

Exserohilum Leonard and Suggs<br />

Exserohilum rostratum (Drechsler) Leonard & Suggs<br />

Morphological Description: Colonies are grey to blackish-brown, suede-like to<br />

floccose in texture and have an olivaceous-black reverse. Conidia are straight,<br />

curved or slightly bent, ellipsoidal to fusiform and are formed apically through a pore<br />

(poroconidia) on a sympodially elongating geniculate conidiophore. Conidia have a<br />

strongly protruding, truncate hilum and the septum above the hilum is usually thickened<br />

and dark, with the end cells often paler than other cells, walls often finely roughed.<br />

Conidial germination is bipolar.<br />

Key Features: Dematiaceous hyphomycete producing sympodial, transverse septate,<br />

ellipsoidal to fusiform conidia with dark bands on both ends and a strongly protruding,<br />

truncate hilum.<br />

Molecular Identification: ITS and D1/D2 sequencing is recommended.<br />

References: Domsch et al. (1980), Alcorn (1983), Adam et al. (1986), McGinnis et al.<br />

(1986b), Rippon (1988), Burges et al. (1987), Dixon and Polak-Wyss (1991), de Hoog<br />

et al. (2000, 2015), de Cunha et al. (2012b), Katragkou et al. (2014).<br />

Antifungal Susceptibility: E. rostratum limited data (Australian National data); MIC<br />

µg/mL.<br />

No. 64<br />

AmB 8 4 2 2<br />

VORI 8 1 1 6<br />

POSA 6 5 1<br />

ITRA 8 4 4<br />

E. rostratum data from 34 isolates (da Cunha et al. 2012); MIC µg/mL.<br />

AmB Range


92<br />

Descriptions of Medical Fungi<br />

Exserohilum rostratum (Drechsler) Leonard & Suggs<br />

20 µm<br />

20 µm<br />

Exserohilum rostratum conidiophores and conidia with a distinctive hilum (arrow).


Descriptions of Medical Fungi 93<br />

Morphologically the genus Fonsecaea is defined by the presence of indistinct<br />

melanised conidiophores with blunt, scattered denticles bearing conidia singly or in<br />

short chains that eventually become branched. de Hoog et al. (2004b) revised the<br />

genus on the basis of ITS sequencing data. Three species from humans are currently<br />

recognised, F. monophora, F. nubica, and F. pedrosoi, although they are morphologically<br />

indistinguishable (Najafzadeh et al. 2009, 2010b; Xi et al. 2009, de Hoog et al. 2015).<br />

All strains grow at 37 O C but not at 40 O C and the three species are recognised as<br />

aetiological agents of chromoblastomycosis.<br />

Morphological Description: Colonies are slow growing, flat to heaped and folded,<br />

suede-like to downy, olivaceous to black with black reverse. Conidiogenous cells pale<br />

olivaceous, arranged in loosely branched systems, with prominent denticles. Conidia<br />

pale olivaceous, clavate to ellipsoidal, in short chains, subhyaline, smooth and thinwalled,<br />

3.5-5 x 1.5-2 µm.<br />

RG-2 organism.<br />

Fonsecaea complex<br />

Molecular Identification: ITS sequencing is recommended for species identification<br />

(Abliz et al. 2003, Najafzadeh et al. 2009, 2010b, Xi et al. 2009).<br />

References: McGinnis (1980), Dixon and Polak-Wyss (1991), de Hoog et al. (2000,<br />

2004b, 2015), Abliz et al. 2003, Najafadeh et al. (2009, 2010a, 2010b).<br />

10 µm<br />

Fonsecaea complex conidiophores and conidia.<br />

Antifungal Susceptibility: F. pedrosoi (Australian National data); MIC µg/mL.<br />

No. 16<br />

AmB 4 1 3<br />

VORI 4 1 3<br />

POSA 4 1 2 1<br />

ITRA 4 1 3


94<br />

Descriptions of Medical Fungi<br />

Fusarium Link ex Fries<br />

Most Fusarium species are soil fungi and have a worldwide distribution. Some are plant<br />

pathogens, causing root and stem rot, vascular wilt or fruit rot. Several species have<br />

emerged as important opportunistic pathogens in humans causing hyalohyphomycosis<br />

(especially in burn victims and bone marrow transplant patients), mycotic keratitis and<br />

onychomycosis (Guarro 2013). Other species cause storage rot and are important<br />

mycotoxin producers.<br />

Multi-locus sequence analysis of EF-1α, β-tubulin, calmodulin, and RPB2 have revealed<br />

the presence of multiple cryptic species within each “morphospecies” of medically<br />

important fusaria (Balajee et al. 2009). For instance, Fusarium solani represents a<br />

complex (i.e. F. solani complex) of over 45 phylogenetically distinct species of which<br />

at least 20 are associated with human infections. Similarly, members of the Fusarium<br />

oxysporum complex are phylogenetically diverse, as are members of the Fusarium<br />

incarnatum-equiseti complex and Fusarium chlamydosporum complex (Balajee et al.<br />

2009, Tortorano et al. 2014, Salah et al. 2015).<br />

Currently the genus Fusarium comprises at least 300 phylogenetically distinct<br />

species, 20 species complexes and nine monotypic lineages (Balajee et al. 2009,<br />

O’Donnell et al. 2015). Most of the identified opportunistic Fusarium pathogens<br />

belong to the F. solani complex, F. oxysporum complex and F. fujikuroi complex. Less<br />

frequently encountered are members of the F. incarnatum-equiseti, F. dimerum and F.<br />

chlamydosporum complexes, or species such as F. sporotrichioides (O’Donnell et al.<br />

2015, van Diepeningen et al. 2015).<br />

Morphological Description: Colonies are usually fast growing, pale or brightcoloured<br />

(depending on the species) with or without a cottony aerial mycelium. The<br />

colour of the thallus varies from whitish to yellow, pink, red or purple shades. Species<br />

of Fusarium typically produce both macro- and microconidia from slender phialides.<br />

Macroconidia are hyaline, two to several-celled, fusiform to sickle-shaped, mostly with<br />

an elongated apical cell and pedicellate basal cell. Microconidia are one or two-celled,<br />

hyaline, smaller than macroconidia, pyriform, fusiform to ovoid, straight or curved.<br />

Chlamydospores may be present or absent.<br />

Identification of Fusarium species is often difficult due to the variability between isolates<br />

(e.g. in shape and size of conidia and colony colour) and because not all features<br />

required are always well developed (e.g. the absence of macroconidia in some isolates<br />

after subculture). Note: Sporulation may need to be induced in some isolates and a<br />

good slide culture is essential. The important characters used in the identification of<br />

Fusarium species are as follows.<br />

1. Colony growth diameters on potato dextrose agar and/or potato sucrose agar after<br />

incubation in the dark for four days at 25 O C.<br />

2. Culture pigmentation on potato dextrose agar and/or potato sucrose agar after<br />

incubation for 10-14 days with daily exposure to light.<br />

3. Microscopic morphology including shape of the macroconidia; presence or<br />

absence of microconidia; shape and mode of formation of microconidia; nature<br />

of the conidiogenous cell bearing microconidia; and presence or absence of<br />

chlamydospores.


Descriptions of Medical Fungi 95<br />

Fusarium Link ex Fries<br />

a<br />

b<br />

Cultures of (a) Fusarium oxysporum complex showing purple pigmentation<br />

and (b) Fusarium fujikuroi complex showing pink pigmentation.<br />

Molecular Identification: Current species identification is on the basis of multilocus<br />

sequence data (Guarro 2013, O’Donnell et al. 2015, van Diepeningen et al. 2015).<br />

Internet-accessible validated databases dedicated to the identification of fusaria via<br />

nucleotide BLAST queries are available at FUSARIUM-ID at Pennsylvania State<br />

University (http://www.fusariumdb.org) and Fusarium MLST at the CBS-KNAW Fungal<br />

Biodiversity Centre (http://www.cbs.knaw.nl/Fusarium/).<br />

For sequence-based identification of Fusarium species (O’Donnell et al. 2015).<br />

1. Use EF-1α, RPB1 and/or RPB2. Use of at least two independent loci will increase<br />

the accuracy of identification.<br />

2. Fusarium MLST or FUSARIUM-ID are the recommended sequence databases,<br />

rather than GenBank.<br />

3. Ensure sequences are carefully edited and free of ambiguities.<br />

4. Ensure the species names associated with the top BLASTn matches are the same.<br />

If multiple species names have similar scores it may be necessary to sequence<br />

additional loci.<br />

Note: ITS and D1/D2 sequences are too conserved to resolve species limits of most<br />

fusaria. O’Donnell et al. (2015) recommend avoiding ITS or D1/D2 sequences from an<br />

unknown isolate to query GenBank, because >50% of the sequences from Fusarium<br />

species are misidentified in this database.<br />

Identifications based on morphology and/or ITS and D1/D2 sequences should<br />

be reported as species complexes. Sequencing of EF-1α, RPB1 and/or RPB2 is<br />

required for accurate species identification.<br />

MALDI-T<strong>OF</strong> MS: A comprehensive ‘in-house’ database of reference spectra allows<br />

accurate identification of Fusarium species complexes (Lau et al. 2013).<br />

References: Booth (1971, 1977), Domsch et al. (1980), McGinnis (1980), Burgess<br />

and Liddell (1983), Rippon (1988), Samson et al. (1995), de Hoog et al. (2000, 2015),<br />

O’Donnell et al. (2008, 2009a, 2009b, 2015), Balajee et al. (2009), Guarro (2013),<br />

Geiser et al. (2013), van Diepeningen et al. (2015), Salah et al. (2015), Tortorano et al.<br />

(2014).


96<br />

Descriptions of Medical Fungi<br />

Fusarium chlamydosporum complex<br />

Fusarium chlamydosporum complex contains five phylogenetically distinct species<br />

and is common in soils and the rhizosphere of numerous vascular plants worldwide.<br />

It is occasionally isolated from human and animal infections (O’Donnell et al. 2009b,<br />

Guarro 2013).<br />

RG-1 organisms.<br />

Morphological Description: Colonies growing rapidly, with abundant aerial mycelium,<br />

deep pink, red or ochraceous to brownish; reverse carmine red or tan to brown.<br />

Sporodochia orange, flesh-coloured or ochraceous. Conidiophores scattered over the<br />

aerial mycelium, branched; numerous polyblastic conidiogenous cells are present.<br />

Macroconidia rarely produced and appearing only on sporodochial phialides, usually<br />

three-(some up to five)-septate, slightly curved, 30-38 x 3.0-4.5 μm, with no distinct<br />

foot-shaped cell. Microconidia and blastoconidia fusiform, rounded apically and<br />

tapered towards the base, single-celled to one-(some up to three)-septate, 6-26 x 2-4<br />

μm. Chlamydospores abundant, intercalary, often roughened.<br />

Fusarium chlamydosporum complex, culture showing pink to<br />

ochraceous to brownish surface and a carmine red reverse.<br />

Antifungal Susceptibility: F. chlamydosporum complex (Australian National<br />

data); MIC µg/mL.<br />

No. 64<br />

AmB 5 1 2 2<br />

VORI 5 1 2 2<br />

POSA 4 2 2<br />

ITRA 4 1 1 1 1


Descriptions of Medical Fungi 97<br />

Fusarium dimerum complex<br />

The Fusarium dimerum complex contains 12 phylogenetically distinct species including<br />

F. delphinoides, F. penzigii and F. dimerum. These are regarded as cosmopolitan<br />

saprotrophs in soil and on plant materials (Domsch et al. 2007). They have also been<br />

isolated from human corneal ulcers after trauma and from disseminated or localised<br />

infections in immunocompromised patients (Schroers et al. 2009, Guarro 2013).<br />

RG-1 organisms.<br />

Morphological Description: Colonies growing slowly; surface usually orange to<br />

deep apricot due to confluent conidial slime; aerial mycelium sometimes floccose and<br />

whitish. Conidiophores loosely branched, with short, often swollen phialides, 10-18 x<br />

4-5 μm. Macroconidia strongly curved and pointed at the apex, mostly one-(some up to<br />

three)-septate, 5-25 (-32) x 1.5-4.2 μm. Microconidia absent. Chlamydospores mostly<br />

intercalary, exceptionally terminal, spherical to ovoidal, 6-12 μm diam, smooth-walled,<br />

single or in chains.<br />

a<br />

b<br />

20 μm<br />

Fusarium dimerum complex (a) culture showing orange to deep apricot<br />

colour due to confluent conidial slime, and (b) macroconidia.<br />

Antifungal Susceptibility: F. dimerum complex (Australian National data); MIC<br />

µg/mL.<br />

No. 64<br />

AmB 21 1 6 6 5 6 2 1<br />

VORI 20 5 8 6 3 3 1<br />

POSA 18 1 4 2 11<br />

ITRA 21 3 24


98<br />

Descriptions of Medical Fungi<br />

Synonymy: Gibberella fujikuroi complex.<br />

Fusarium fujikuroi complex<br />

Fusarium fujikuroi complex consists of 50 phylogenetically distinct species including<br />

13 of which have been reported to cause human infection; F. acutatum, F. ananatum,<br />

F. andiyazi, F. fujikuroi, F. guttiforme, F. napiforme, F. nygamai, F. verticillioides, F.<br />

proliferatum, F. sacchari, F. subglutinans, F. temperatum and F. thapsinum (Guarro,<br />

2013, Al-Hatmi et al. 2015).<br />

RG-1 organisms.<br />

Morphological Descriptrion: Colonies growing rapidly, pink or vinaceous to violet;<br />

aerial mycelium abundant. Sporodochia present or absent, when present they are tan<br />

to orange. Conidiophores usually erect and branched. Macroconidia abundant, falcate<br />

to rather straight, three to five-septate, with a distinct foot-cell, 27-73 x 3.4-5.2 μm.<br />

Blastoconidia straight or slightly curved, two to three-septate, fusiform to lanceolate,<br />

with a somewhat pointed, often slightly asymmetrical apical cell and a truncate basal<br />

cell, 16-43 x 3.0-4.5 μm. Microconidia produced on polyphialides and aggregating<br />

in heads, usually unicellular, ovoidal, ellipsoidal or allantoid, 4-20 x 1.5-4.5 μm.<br />

Chlamydospores absent.<br />

Antifungal Susceptibility: F. fujikuroi complex (Castanheir et al. 2012, Australian<br />

National data); MIC µg/mL.<br />

No. 64<br />

AmB 39 3 19 15 2<br />

VORI 39 2 12 14 8 3<br />

POSA 39 4 9 7 3 16<br />

ITRA 30 2 4 2 22<br />

Fusarium incarnatum-equiseti complex<br />

Fusarium incarnatum-equiseti complex consists of 40 phylogenetically distinct species.<br />

They occasionally cause infections in humans and animals (O’Donnell et al. 2009b,<br />

Guarro 2013).<br />

RG-1 organisms.<br />

Morphological Description: Colonies growing rapidly; aerial mycelium floccose,<br />

at first whitish, later becoming avellaneous to buff-brown; reverse pale, becoming<br />

peach-coloured. Conidiophores scattered in the aerial mycelium, loosely branched;<br />

polyblastic conidiogenous cells abundant. Sporodochial macroconidia slightly<br />

curved, with foot-cell, three to seven-septate, 20-46 x 3.0-5.5 µm. Conidia on aerial<br />

conidiophores (blastoconidia) usually borne singly on scattered denticles, fusiform<br />

to falcate, mostly three to five-septate, 7.5-35 x 2.5-4.0 µm. Microconidia sparse or<br />

absent. Chlamydospores sparse, spherical, 10-12 µm diameter, becoming brown,<br />

intercalary, single or in chains.


Descriptions of Medical Fungi 99<br />

Fusarium incarnatum-equiseti complex<br />

Antifungal Susceptibility: F. incarnatum-equiseti complex (Australian National<br />

data); MIC µg/mL.<br />

No. 64<br />

AmB 6 1 1 3 1<br />

VORI 6 1 1 3 1<br />

POSA 5 1 3 1<br />

ITRA 4 1 3<br />

Fusarium oxysporum complex<br />

This complex contains at least five phylogenetically distinct species and accounts for<br />

about 20% of human infections caused by fusaria (Guarro 2013, Tortorano et al. 2014,<br />

Salah et al. 2015). All are ubiquitous soil borne pathogens responsible for vascular<br />

wilts, rots, and damping-off diseases of a broad range of plants. A number of these<br />

fusaria are also clinically important, causing localised or deeply invasive life threatening<br />

infections in humans and other animals (O’Donnell et al. 2009a). Mortality in patients<br />

who are persistently and severely neutropenic is typically 100% (Nucci and Anaissie,<br />

2007).<br />

RG-2 organisms.<br />

Morphological Description: Colonies growing rapidly, 4.5 cm in four days, aerial<br />

mycelium white, becoming purple, with discrete orange sporodochia present in some<br />

strains; reverse hyaline to dark blue or dark purple. Conidiophores are short, single,<br />

lateral monophialides in the aerial mycelium, later arranged in densely branched<br />

clusters. Macroconidia are fusiform, slightly curved, pointed at the tip, mostly three<br />

septate, basal cells pedicellate, 23-54 x 3-4.5 µm. Microconidia are abundant, never<br />

in chains, mostly non-septate, ellipsoidal to cylindrical, straight or often curved, 5-12<br />

x 2.3-3.5 µm. Chlamydospores are terminal or intercalary, hyaline, smooth or rough<br />

walled, 5-13 µm. In contrast to F. solani complex, the phialides are short and mostly<br />

non-septate.<br />

Antifungal Susceptibility: F. oxysporum complex (Australian National data); MIC<br />

µg/mL.<br />

No. 64<br />

AmB 52 1 4 6 12 14 9 4 2<br />

VORI 49 5 6 10 19 7 3<br />

POSA 37 1 1 7 6 1 15 6<br />

ITRA 39 3 1 1 34


100<br />

Descriptions of Medical Fungi<br />

Fusarium oxysporum complex<br />

a<br />

15 µm<br />

b<br />

15 µm<br />

Fusarium oxysporum complex (a)microconidia on short phialides<br />

and (b) macroconidia.<br />

Fusarium solani complex<br />

Antifungal Susceptibility: F. solani complex (Australian National data); MIC µg/<br />

mL.<br />

No. 64<br />

AmB 130 1 2 18 29 40 29 11<br />

VORI 126 7 24 36 34 25<br />

POSA 114 2 1 2 1 71 37<br />

ITRA 93 2 1 90


Descriptions of Medical Fungi 101<br />

The Fusarium solani complex contains at least 60 species and accounts for about 50%<br />

of human infections caused by fusaria (Guarro 2013, Tortorano et al. 2014, Salah et al.<br />

2015). All are ubiquitous soil borne pathogens responsible for vascular wilts, rots, and<br />

damping-off diseases of a broad range of plants. A number of these fusaria, notably<br />

F. keratoplasticum, F. petroliphilum, F. lichenicola and F. solani are clinically important,<br />

causing localised or deeply invasive life threatening infections in humans and other<br />

animals (Guarro 2013, O’Donnell et al. 2008).<br />

RG-2 organisms.<br />

Fusarium solani complex<br />

Morphological Description: Colonies growing rapidly, 4.5 cm in four days, aerial<br />

mycelium white to cream, becoming bluish-brown when sporodochia are present.<br />

Macroconidia are formed after 4-7 days from short multiple branched conidiophores<br />

which may form sporodochia. They are three to five-septate (usually three-septate),<br />

fusiform, cylindrical, often moderately curved, with an indistinct pedicellate foot cell and<br />

a short blunt apical cell, 28-42 x 4-6 µm. Microconidia are usually abundant, cylindrical<br />

to oval, one to two-celled and formed from long lateral phialides, 8-16 x 2-4.5 µm.<br />

Chlamydospores are hyaline, globose, smooth to rough-walled, borne singly or in pairs<br />

on short lateral hyphal branches or intercalary, 6-10 µm.<br />

a<br />

15 µm<br />

b<br />

15 µm<br />

c<br />

15 µm<br />

Fusarium solani complex (a) microconidia on long phialides,<br />

(b) macroconidia and (c) chlamydospores.


102<br />

Descriptions of Medical Fungi<br />

Geotrichum candidum Link.<br />

Synonymy: Galactomyces candidus de Hoog & M.Th. Smith.<br />

The genus Geotrichum and related species have undergone extensive taxonomic<br />

revision (de Hoog and Smith 2004, 2011a, 2011b, 2011c). The three species of prime<br />

interest to medical mycology are Geotrichum candidum (Galactomyces candidus),<br />

Magnusiomyces capitatus (previously known as Geotrichum capitatum), and<br />

Saprochaete clavata (previously known as Geotrichum clavatum).<br />

Geotrichum candidum is an extremely common fungus with a worldwide distribution. It<br />

is commonly isolated from soil, air, water, milk, silage, plant tissues, and the digestive<br />

tract in humans and other mammals (Pottier et al. 2008). Pulmonary involvement is the<br />

most frequently reported form of the disease in humans and animals, but bronchial,<br />

oral, vaginal, cutaneous and alimentary infections have also been noted (Arendrup et<br />

al. 2014).<br />

RG-1 organism.<br />

Morphological Description: Colonies are fast growing, flat, white to cream, dry and<br />

finely suede-like with no reverse pigment. Hyphae are hyaline, septate, branched<br />

and break up into chains of hyaline, smooth, one-celled, subglobose to cylindrical<br />

arthroconidia. Arthroconidia are 6-12 x 3-6 µm in size and are released by the separation<br />

of a double septum.<br />

Note: True blastoconidia production is not found in this genus. This characteristic<br />

distinguishes Geotrichum from Trichosporon, which usually does produce blastoconidia.<br />

Molecular Identification: ITS sequencing recommended for accurate species<br />

identification (de Hoog and Smith 2004).<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose + L-Arabinose - D-Glucitol +<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose v Maltose - D-Ribose - D-Gluconate -<br />

Galactose v Cellobiose - L-Rhamnose - DL-Lactate +<br />

Sucrose - Trehalose - D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine nd 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate nd<br />

Trehalose - Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol v Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide nd<br />

Galactose + D-Xylose + D-Mannitol + Growth at 37 O C v<br />

References: Gueho (1979), Domsch et al. (1980), McGinnis (1980), Barnett et al.<br />

(1983), Buchta and Otcenasek (1988), Samson et al. (1995), de Hoog et al. (1986,<br />

2015), de Hoog and Smith (2004, 2011a, 2011b, 2011c).


Descriptions of Medical Fungi 103<br />

Geotrichum candidum Link.<br />

15 µm<br />

15 µm<br />

Geotrichum candidum arthroconidium formation. Hyphal elements are progressively<br />

compartmentalised by fragmentation of septa. Conidial secession is by the centripetal<br />

separation (schizolysis) of a so called double septum and concomitant rupture of the<br />

original outer hyphal wall layer.<br />

Antifungal Susceptibility: Geotrichum candidum limited data (Sfakianakis et al.<br />

2007, Henrich et al. 2009, Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 6 2 4<br />

FLU 5 4 1<br />

VORI 6 2 2 1 1<br />

POSA 3 3<br />

ITRA 5 1 2 2<br />

ANID 2 2<br />

MICA 2 2<br />

CAS 3 1 2<br />

5FC 5 1 3 1


104<br />

Descriptions of Medical Fungi<br />

Gliocladium species have a worldwide distribution and are commonly isolated from a<br />

wide range of plant debris and soil.<br />

RG-1 organism.<br />

Gliocladium Corda<br />

Morphological Description: The genus Gliocladium is often described as a<br />

counterpart of Penicillium with slimy conidia. Colonies are fast growing, suede-like to<br />

downy in texture, white at first, sometimes pink to salmon, becoming pale to dark green<br />

with sporulation. The most characteristic feature of the genus is the distinctive erect,<br />

often densely penicillate conidiophores with phialides which bear slimy, one-celled<br />

hyaline to green, smooth-walled conidia in heads or columns. Although, penicillate<br />

conidiophores are always present, Gliocladium species may also produce verticillate<br />

branching conidiophores which can be confused with Verticillium or Trichoderma.<br />

Key Features: Hyphomycete producing distinctive erect penicillate conidiophores with<br />

phialides bearing clusters of single-celled conidia.<br />

References: Domsch et al. (2007), McGinnis (1980), Onions et al. (1981), Rippon<br />

(1988), de Hoog et al. (2000).<br />

10 µm<br />

Gliocladium spp. conidiophore and conidia.


Descriptions of Medical Fungi 105<br />

The genus Graphium is characterised by the formation of synnemata which consist<br />

of a more or less compact group of erect conidiophores that are cemented together,<br />

usually splaying out and bearing conidia at the apex. Graphium species are commonly<br />

found on woody plant material. Graphium basitruncatum has been reported as causing<br />

fungaemia in an immunosuppressed child post stem-cell transplantation (El Feghaly<br />

et al. 2012).<br />

Note: Many other fungi such as Scedosporium species may also produce synnemata.<br />

RG-1 organism.<br />

Graphium Corda<br />

Morphological Description: Synnemata are darkly pigmented, erect and occur<br />

solitarily or in clusters. Conidia are hyaline, one-celled, smooth, subglobose to ovoid<br />

and are usually aggregated in slimy heads at the apex of the synnemata. Colonies are<br />

effuse, grey, olivaceous brown or black.<br />

Molecular diagnostics: The genus is phylogenetically close to Scedosporium but ITS<br />

sequencing can be used to resolve all species (Okada et al. 2000, Lackner and de<br />

Hoog 2011).<br />

Key Features: Dematiaceous hyphomycete producing erect synnemata with apical<br />

aggregates of single-celled conidia in slimy heads.<br />

References: Barron (1968), Ellis (1971), McGinnis (1980), de Hoog et al. (2000, 2015).<br />

20 µm<br />

Graphium spp. synnemata and conidia.


106<br />

Descriptions of Medical Fungi<br />

Histoplasma capsulatum Darling<br />

WARNING: RG-3 organism. Cultures of Histoplasma capsulatum represent a severe<br />

biohazard to laboratory personnel and must be handled with extreme caution in a<br />

Class II Biological Safety Cabinet (BSCII).<br />

Histoplasma capsulatum has a worldwide distribution, however the Mississippi-Ohio<br />

River Valley in the USA is recognised as a major endemic region. Environmental<br />

isolations of the fungus have been made from soil enriched with excreta from<br />

chicken, starlings and bats. Histoplasmosis is an intracellular mycotic infection of the<br />

reticuloendothelial system caused by the inhalation of the fungus. Approximately 95%<br />

of cases of histoplasmosis are inapparent, subclinical or benign. The remaining 5% of<br />

cases may develop chronic progressive lung disease, chronic cutaneous or systemic<br />

disease or an acute fulminating fatal systemic disease. All stages of this disease may<br />

mimic tuberculosis. Sporadic cases have been reported in Australia.<br />

Morphological Description: Histoplasma capsulatum exhibits thermal dimorphism<br />

growing in living tissue or in culture at 37 O C as a budding yeast-like fungus and in soil<br />

or culture at temperatures below 30 O C as a mould.<br />

Colonies at 25 O C are slow growing, white or buff-brown, suede-like to cottony with<br />

a pale yellow-brown reverse. Other colony types are glabrous or verrucose, and a<br />

red pigmented strain has been noted (Rippon, 1988). Microscopic morphology shows<br />

the presence of characteristic large, rounded, single-celled, 8-14 µm in diameter,<br />

tuberculate macroconidia formed on short, hyaline, undifferentiated conidiophores.<br />

Small, round to pyriform microconidia, 2-4 µm in diameter, borne on short branches or<br />

directly on the sides of the hyphae may also be present.<br />

Colonies at 37 O C grown on brain heart infusion (BHI) agar containing blood are<br />

smooth, moist, white and yeast-like. Microscopically, numerous small round to oval<br />

budding yeast-like cells, 3-4 x 2-3 µm in size are observed.<br />

Three varieties of Histoplasma capsulatum are recognised, depending on the clinical<br />

disease: var. capsulatum is the common cause of histoplasmosis; var. duboisii is the<br />

African type and var. farciminosum causes lymphangitis in horses. Histoplasma isolates<br />

may also resemble species of Sepedonium and Chrysosporium. Traditionally, positive<br />

identification required conversion of the mould form to the yeast phase by growth at<br />

37 O C on enriched media, however for laboratory safety, culture identification by either<br />

exoantigen test or DNA sequencing is now preferred.<br />

Key Features: Clinical history, tissue morphology, culture morphology and positive<br />

exoantigen test or DNA sequencing.<br />

Molecular Identification: A probe for species recognition is commercially available<br />

(Padhye et al. 1992, Chemaly et al. 2001) and Elias et al. (2012) developed a multiplex-<br />

PCR for identification from cultures. Scheel et al. (2014) developed a loop-mediated<br />

isothermal amplification (LAMP) assay for detection directly in clinical samples which is<br />

affordable and useful in resource poor facilities. ITS sequencing may also be used for<br />

accurate identification (Estrada-Bárcenas et al. 2014, Irinyi et al. 2015).<br />

References: McGinnis (1980), Chandler et al. (1980), George and Penn (1986),<br />

Rippon (1988), de Hoog et al. (2000, 2015).


Descriptions of Medical Fungi 107<br />

Histoplasma capsulatum Darling<br />

a<br />

b<br />

20 µm<br />

Histoplasma capsulatum (a) culture and (b) microscopic morphology of the saprophytic<br />

or mycelial form showing characteristic large, rounded, single-celled, tuberculate<br />

macroconidia and smaller microconidia.<br />

Antifungal Susceptibility: H. capsulatum (Espinel-Ingroff 2003, Gonzalez et al.<br />

2005, Sabatelli et al. 2006, Brilhante et al. 2012, Kathuria et al. 2014); MIC µg/mL.<br />

Antifungal<br />

Filamentous form<br />

Yeast form<br />

Range MIC 90<br />

Range MIC 90<br />

AmB


108<br />

Descriptions of Medical Fungi<br />

Hortaea werneckii (Horta) Nishimura & Miyaji<br />

Synonymy: Phaeoannellomyces werneckii (Horta) McGinnis & Schell; Exophiala<br />

werneckii (Horta) v. Arx; Cladosporium werneckii Horta.<br />

Hortaea werneckii is a common saprophytic fungus believed to occur in soil, compost,<br />

humus and on wood in humid tropical and subtropical regions and is the causative<br />

agent of tinea nigra in humans. RG-1 organism.<br />

Morphological Description: Colonies are slow growing, initially mucoid, yeast-like<br />

and shiny-black. However with age they develop abundant aerial mycelia and become<br />

dark olivaceous in colour. Microscopically, colonies consist of brown to dark olivaceous,<br />

septate hyphal elements and numerous two-celled, pale brown, cylindrical to spindleshaped<br />

yeast-like cells that taper towards the ends to form an annellide. Most yeastlike<br />

cells also have prominent darkly-pigmented septa. Annellides may also arise from<br />

the hyphae. Conidia are one to two-celled, cylindrical to spindle-shaped, hyaline to<br />

pale brown and usually occur in aggregated masses. Chlamydospores also present.<br />

Key Features: Hyphomycete, two-celled yeast-like cells producing annelloconidia.<br />

Molecular Identification: An ITS-primer specific for H. werneckii was developed by<br />

Abliz et al. (2003). ITS sequencing can also assist identification.<br />

References: Mok (1982), McGinnis (1980), McGinnis et al. (1985), Rippon (1988), de<br />

Hoog et al. (2000), Abliz et al. (2003), Ng et al. (2005).<br />

Antifungal Susceptibility: H. wernickii (Ng et al. 2005, Formoso et al. 2015); MIC<br />

µg/mL.<br />

No. 64<br />

AmB 6 1 1 1 1 2<br />

FLU 6 2 1 3<br />

VORI 6 2 1 3<br />

POSA 4 2 2<br />

ITRA 8 2 2 2 2<br />

a<br />

20 µm<br />

b<br />

Hortaea werneckii (a) culture and (b) conidia.


Descriptions of Medical Fungi 109<br />

Kluyveromyces marxianus (Hansen) van der Walt<br />

Synonymy: Candida kefyr (Beijerinck) van Uden & Buckley.<br />

Candida pseudotropicalis (Castellani) Basgal.<br />

Kluyveromyces marxianus is a rare cause of candidiasis and is usually associated<br />

with superficial cutaneous manifestations rather than systemic disease. Environmental<br />

isolations have been made from cheese and dairy products. RG-1 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Short-ovoid to long-ovoid, budding blastoconidia, 3.0-6.5 x 5.5-11.0 µm,<br />

sometimes becoming elongate (up to 16.0 µm).<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Abundant, long, wavy, branched pseudohyphae usually formed,<br />

with ovoid blastoconidia, budding off singly, in pairs or chains, often in a verticillated<br />

position. Note: In some strains pseudohyphae may be scarce or almost absent.<br />

Molecular Identification: ITS sequencing recommended.<br />

MALDI-T<strong>OF</strong> MS: Able to accurately identify this species.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose v L-Arabinose v D-Glucitol v<br />

Fermentation Sucrose + D-Arabinose - α-M-D-glucoside -<br />

Glucose + Maltose - D-Ribose v D-Gluconate -<br />

Galactose +,s Cellobiose v L-Rhamnose - DL-Lactate +<br />

Sucrose + Trehalose -,w D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose v N-A-D-glucosamine - 2-K-D-gluconate -<br />

Lactose v Melibiose - Glycerol s D-Glucuronate nd<br />

Trehalose - Raffinose + Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol s Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide +<br />

Galactose s D-Xylose s D-Mannitol v Growth at 40 O C +<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern.<br />

Antifungal Susceptibility: K. marxianus (Diekema et al. 2009, Pfaller et al. 2013,<br />

Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 95 6 28 50 10 1<br />

FLU 111 1 22 62 19 6 1<br />

VORI 111 82 23 4 2<br />

POSA 106 1 6 23 33 31 10 2<br />

ITRA 37 5 6 12 12 2<br />

ANID 86 1 14 37 29 4<br />

MICA 71 4 29 42 4 1 1<br />

CAS 106 12 70 13 5 3 2 1<br />

5FC 21 4 8 2 1 1 3 2


110<br />

Descriptions of Medical Fungi<br />

Lasiodiplodia theobromae (Pat.) Griffon & Maublanc<br />

Synonymy: Botryosphaeria rhodina (Berk & Curt) v. Arx.<br />

Botryodiplodia theobromae Patouillard.<br />

Lasiodiplodia theobromae is a well known plant pathogen reported from about 500<br />

host plants, mainly confined to an area 40 O north to 40 O south of the equator. It has<br />

also been associated with ulcerated human cornea, lesions on nail and subcutaneous<br />

tissue.<br />

RG-1 organism.<br />

Morphological Description: Colonies are greyish sepia to mouse grey to black,<br />

fluffy with abundant aerial mycelium; reverse fuscous to black. Pycnidia are simple or<br />

compound, often aggregated, stromatic, ostiolate, frequently setose, up to 5 mm wide.<br />

Conidiophores are hyaline, simple, sometimes septate, rarely branched cylindrical,<br />

arising from the inner layers of cells lining the pycnidial cavity. Conidiogenous cells<br />

are hyaline, simple, cylindrical to sub-obpyriform, holoblastic, annellidic. Conidia are<br />

initially unicellular, hyaline, granulose, sub-ovoid to ellipsoid-oblong, thick-walled, base<br />

truncate; mature conidia one-septate, cinnamon to fawn, often longitudinally striate, 20-<br />

30 x 10-15 µm. Paraphyses when present are hyaline, cylindrical, sometimes septate,<br />

up to 50 µm long.<br />

Key Features: Coelomycete, with pycnidia producing characteristic two-celled, dark<br />

brown, striated conidia.<br />

Molecular Identification: Recommended genetic marker: EF-1α (de Hoog et al. 2015).<br />

ITS sequencing is useful for identifying clinically important species (Bagyalakshmi<br />

2008).<br />

References: de Hoog et al. (2000, 2015), Liu et al. (2012), Phillips et al. (2013).<br />

a<br />

b<br />

10 µm<br />

Lasiodiplodia theobromae (a) pycnidia growing on carnation leaf agar, and<br />

(b) mature two-celled dark brown conidia with typical longitudinal striations.


Descriptions of Medical Fungi 111<br />

Synonymy: Mycocladus corymbifera (Cohn) J.H. Mirza.<br />

Absidia corymbifera (Cohn) Saccardo & Trotter.<br />

Mucor corymbifera Cohn.<br />

Recent taxonomic revisions of the genus Absidia have placed the thermotolerant<br />

species into the genus Lichtheimia (Hoffmann et al. 2007, 2009). The genus Lichtheimia<br />

currently contains five mostly saprophytic plant decaying and soil-borne species.<br />

Lichtheimia corymbifera is the principle pathogen causing human and animal infections,<br />

however L. ramosa and L. ornata have also been reported as human pathogens (often<br />

misidentified morphologically as L. corymbifera) (Alastruey-Izquierdo et al. 2010).<br />

RG-2 organism.<br />

Lichtheimia corymbifera (Cohn) Vuill.<br />

Morphological Description: Colonies are fast growing, floccose, white at first<br />

becoming pale grey with age, and up to 1.5 cm high. Sporangiophores are hyaline to<br />

faintly pigmented, simple or sometimes branched, arising solitarily or in groups. Subsporangial<br />

septa are absent or rare. Rhizoids are very sparingly produced and may be<br />

difficult to find without the aid of a dissecting microscope to examine the colony on the<br />

agar surface. Sporangia are small (10-40 µm in diameter) and are typically pyriform in<br />

shape with a characteristic conical-shaped columella and pronounced apophysis, often<br />

with a short projection at the top. Sporangiospores vary from subglobose to oblongellipsoidal<br />

(3-7 x 2.5-4.5 µm), hyaline to light grey and smooth-walled. Intercalary giant<br />

cells may also be present. Temperature: optimum 35-37 O C; maximum 46 O C.<br />

Key Features: Mucorales, small pyriform-shaped sporangia with a characteristic<br />

conical-shaped columellae and pronounced apophyses, rapid growth at 40 O C.<br />

Comment: Morphological characteristics alone are not sufficient to reliably differentiate<br />

between L. corymbifera, L. ramosa and L. ornata. While L. ornata develops large, densely<br />

branched giant cells and L. ramosa has more ellipsoidal to cylindrical sporangiospores<br />

and a faster growth rate, these characters are often difficult to interpret. Molecular<br />

methods are needed to accurately separate these species.<br />

Molecular Identification: Species recognition in Lichtheimia is based on ITS and/or<br />

D1/D2 sequencing (Garcia-Hermoso et al. 2009, Alastruey-Izquierdo et al. 2010).<br />

MALDI-T<strong>OF</strong> MS: Direct identification of Lichtheimia species was described by Schrödl<br />

et al. (2012).<br />

References: Ellis and Hesseltine (1965, 1966), Hesseltine and Ellis (1964a, 1966),<br />

Nottebrock et al. (1974), O’Donnell (1979), Samson et al. (1995), Domsch et al. (1980),<br />

McGinnis (1980), Ellis (2005b), de Hoog et al. (2000, 2015).


112<br />

Descriptions of Medical Fungi<br />

Lichtheimia corymbifera (Cohn) Vuill.<br />

a<br />

b<br />

15 μm<br />

c<br />

Lichtheimia corymbifera (a) culture and (b) typical pyriform-shaped sporangium with<br />

a conical-shaped columella and pronounced apophysis (arrow), and (c) Grocott’s<br />

methenamine silver (GMS) stained tissue section from a lung biopsy showing a typical<br />

sporangium of L. corymbifera.<br />

Antifungal Susceptibility: L. corymbifera (Espinel-Ingroff et al. 2015a, Australian<br />

National data); MIC µg/mL.<br />

No. 16<br />

AmB 136 7 17 36 53 19 3 1<br />

POSA 112 3 9 26 51 21 1 1<br />

ITRA 93 5 10 24 21 23 6 3 1


Descriptions of Medical Fungi 113<br />

Synonymy: Scedosporium prolificans (Hennebert & Desai) Gueho & de Hoog.<br />

Scedosporium inflatum Malloch & Salkin.<br />

Lomentospora prolificans (formerly Scedosporium prolificans) is phylogenetically and<br />

morphologically distinguishable from Scedosporium species (Lennon et al. 1994,<br />

Lackner et al. 2014a).<br />

L. prolificans appears to occupy a restricted geographic range, with infections occurring<br />

mainly in Australia, Spain, and the United States (Heath et al. 2009, Revankar and<br />

Sutton, 2010). L. prolificans infections are refractory to antifungal therapy and are<br />

associated with high mortality. Major risk factors include malignancy, cystic fibrosis,<br />

and solid organ transplantation. The main clinical presentations are disseminated<br />

infection and pulmonary mycoses, followed by bone and joint infections (Cortez et al.<br />

2008, Heath et al. 2009, Rodriguez-Tudela et al. 2009, Revankar and Sutton, 2010).<br />

RG-2 organism.<br />

Lomentospora prolificans Hennebert & Desai<br />

Morphological Description: Colonies are rapid growing, flat, spreading, olive-grey<br />

to black and have a suede-like to downy surface texture. Conidia are borne in small<br />

groups on distinctive basally swollen, flask-shaped conidiophores, which occur singly<br />

or in clusters along the vegetative hyphae. Conidia are aggregated in slimy heads,<br />

single-celled, hyaline to pale-brown, ovoid to pyriform, 3-7 x 2-5 µm, and have smooth<br />

thick walls. Growth at 45 O C.<br />

Key Features: Dematiaceous hyphomycete with initial black pasty colony, conidiophores<br />

with distinctly swollen bases, and the conidial mass forms apical aggregates of conidia.<br />

A Graphium synanamorph is absent and there is no growth on media containing<br />

cycloheximide (actidione).<br />

Molecular Identification: Recommended genetic markers: ITS and β-tubulin.<br />

References: Malloch and Salkin (1984), Salkin et al. (1988), Rippon (1988), Wilson<br />

et al. (1990), Gueho and de Hoog (1991), Lennon et al. (1994), Gilgado et al. (2005),<br />

Rainer and de Hoog (2006), Revankar and Sutton (2010), Lackner et al. (2014a), de<br />

Hoog et al. (2015).<br />

Antifungal Susceptibility: L. prolificans (Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 190 1 2 17 47 93 30<br />

VORI 183 1 6 31 62 83<br />

POSA 105 1 1 103<br />

ITRA 191 2 189


114<br />

Descriptions of Medical Fungi<br />

Lomentospora prolificans Hennebert & Desai<br />

a<br />

b<br />

20 µm<br />

Lomentospora prolificans (a) culture and (b) conidiophores and conidia.<br />

Synergy testing results for L. prolificans (Australian National data).<br />

Antifungal Combination No. ∑FIC < 0.5 (S) ∑FIC >0.5-4 (I) ∑FIC >4 (A)<br />

VORI/TERB 109 94 (86%) 14 (14%) 0<br />

ITRA/TERB 93 56 (60%) 37 (40%) 0<br />

S = synergy, I = indifferent (synergy or antagonism not demonstrated), A = antagonism.


Descriptions of Medical Fungi 115<br />

Lophophyton gallinae (Megnin) Matruchot & Dassonville<br />

Synonymy: Microsporum gallinae (Megnin) Grigorakis.<br />

Lophophyton gallinae is a zoophilic fungus causing fowl favus in chickens and other<br />

fowl, affecting the comb and wattles producing “white comb” lesions. A rare cause of<br />

tinea in humans. Invaded hairs show a sparse ectothrix infection but do not fluoresce<br />

under Wood’s ultra-violet light. RG-2 organism.<br />

Morphological Description: Colonies are flat with a suede-like texture and are<br />

white with a pinkish tinge in colour. Some cultures show radial folding. An orangepink<br />

“strawberry” reverse pigment is usually present. Macroconidia, when present, are<br />

usually five to six celled, thin to thick-walled, slightly echinulate, cylindrical to clavate<br />

with a narrow base and blunt tip, 15-60 x 6-10 µm. Microconidia are ovoidal to pyriform<br />

in shape.<br />

Key Features: Macroconidial morphology, culture characteristics and clinical lesions<br />

in chickens.<br />

Molecular Identification: ITS sequencing is recommended.<br />

References: Rebell and Taplin (1970), Rippon (1988), Gräser et al. (2008), Cafarchia<br />

et al. (2013), de Hoog et al. (2015, 2016).<br />

a<br />

b<br />

Lophophyton gallinae (a) culture and (b) macroconidia.<br />

20 µm


116<br />

Descriptions of Medical Fungi<br />

Madurella complex<br />

The genus Madurella was originally based on tissue morphology (mycetoma with black<br />

grains) and the formation of sterile cultures on mycological media. Initially two species<br />

were described, M. mycetomatis and M. grisea. However recent molecular studies have<br />

recognised five species: Madurella mycetomatis, Trematosphaeria grisea (formerly<br />

M. grisea), M. fahalii, M. pseudomycetomatis and M. tropicana (Desnos-Ollivier et al.<br />

2006, de Hoog et al. 2004a, 2012). All species have been isolated from soil and are<br />

major causative agents of mycetoma. RG-2 organism.<br />

Madurella mycetomatis (Laveran) Brumpt<br />

Morphological Description: Colonies are slow growing, flat and leathery at first,<br />

white to yellow to yellowish-brown, becoming brownish, folded and heaped with age,<br />

and with the formation of aerial mycelia. A brown diffusible pigment is characteristically<br />

produced in primary cultures. Although most cultures are sterile, two types of conidiation<br />

have been observed, the first being flask-shaped phialides that bear rounded conidia,<br />

the second being simple or branched conidiophores bearing pyriform conidia (3-5 µm)<br />

with truncated bases. The optimum temperature for growth of this mould is 37 O C.<br />

Grains of Madurella mycetomatis (tissue microcolonies) are brown or black, 0.5-1.0<br />

mm in size, round or lobed, hard and brittle, composed of hyphae which are 2-5 µm in<br />

diameter, with terminal cells expanded to 12-15 (30) µm in diameter.<br />

M. mycetomatis can be distinguished from Trematosphaeria grisea by growth at 37 O C<br />

and its inability to assimilate sucrose.<br />

Key Features: Black grain mycetoma, growth at 37 O C, diffusible brown pigment<br />

produced on culture and the occasional presence of phialides.<br />

References: McGinnis (1980), Chandler et al. (1980), Rippon (1988), de Hoog et al.<br />

(2000, 2004a, 2012, 2015), Desnos-Ollivier et al. (2006).<br />

a<br />

b<br />

20 µm<br />

Madurella mycetomatis (a) culture showing brown diffusible<br />

pigment, and (b) phialides and conidia.


Descriptions of Medical Fungi 117<br />

Molecular Identification: ITS sequencing is recommended for species separation<br />

(Ahmed et al. 2014b, Desnos-Olliver et al. 2006, Irinyi et al. 2015). A five locus<br />

phylogenetic analysis was performed by Ahmed et al. (2014a) using the ITS, D1/D2,<br />

RPB2 and EF-1α genes.<br />

Trematosphaeria grisea (MacKinnon et al.) S.A. Ahmed et al.<br />

Synonymy: Madurella grisea Mackinnon, Ferrada and Montemayer.<br />

RG-2 organism.<br />

Madurella complex<br />

Morphological Description: Colonies are slow growing, dark, leathery, folded<br />

with radial grooves and with a light brown to greyish surface mycelium. With age,<br />

colonies become dark brown to reddish-brown and have a brownish-black reverse.<br />

Microscopically, cultures are sterile, although hyphae of two widths have been<br />

described, thin at 1-3 µm in width or broad at 3-5 µm in width. The optimum temperature<br />

for growth of T. grisea is 30 O C; this fungus does not grow at 37 O C.<br />

Trematosphaeria grisea can be distinguished from Madurella mycetomatis by the<br />

inability to grow at 37 O C and to assimilate lactose.<br />

Key Features: Black grain mycetoma, no growth at 37 O C, no diffusible brown pigment<br />

produced on culture and absence of conidia.<br />

References: McGinnis (1980), Chandler et al. (1980), Rippon (1988), de Hoog et al.<br />

(2000, 2015), Ahmed et al. (2014b), Desnos-Olliver et al. (2006), Irinyi et al. (2015).<br />

100 µm<br />

Trematosphaeria grisea grains (tissue microcolonies) are black, round to<br />

lobed, soft to firm, up to 1.0 mm, with two distinctive zones, a hyaline to<br />

weakly pigmented central zone and a deeply pigmented periphery.


118<br />

Descriptions of Medical Fungi<br />

Magnusiomyces capitatus (de Hoog et al.) de Hoog & M.Th. Smith<br />

Synonymy: Saprochaete capitata (Diddens & Lodder) de Hoog & M.Th. Smith;<br />

Geotrichum capitatum (Diddens & Lodder) v. Arx; Trichosporon capitatum Diddens &<br />

Lodder; Blastoschizomyces capitis (Diddens & Lodder) Salkin et al.<br />

Based on a phylogenetic analysis of rDNA gene sequences, de Hoog and Smith (2004)<br />

transferred Geotrichum capitatum to the genus Magnusiomyces. Magnusiomyces<br />

capitatus occurs quite commonly in humans, usually as a transient component of<br />

normal skin flora and sputum. Systemic infections including pulmonary, fungaemia and<br />

endocarditis have been reported in immunosuppressed patients. RG-1 organism.<br />

Morphological Description: Colonies are moderately fast growing, flat, whitish, and<br />

finely suede-like with no reverse pigment. Hyphae are profusely branched at acute<br />

angles, with terminal and intercalary conidiogenous cells which form long, cicatrised<br />

rachids on which conidia are borne. Conidia are hyaline, smooth, one-celled, cylindrical<br />

to clavate, with a rounded apex and flat base, 7-10 x 2.5-3.5 µm. Rectangular<br />

arthroconidia are also often present.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose v L-Arabinose - D-Glucitol +<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose - Maltose - D-Ribose - D-Gluconate -<br />

Galactose - Cellobiose - L-Rhamnose - DL-Lactate +<br />

Sucrose - Trehalose - D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine nd 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate nd<br />

Trehalose - Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol - Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide +<br />

Galactose + D-Xylose - D-Mannitol + Growth at 40 O C +<br />

Molecular Identification: ITS sequencing recommended (de Hoog and Smith 2004).<br />

Note: Magnusiomyces capitatus and Saprochaete clavata are human pathogens<br />

that are closely related and are frequently mistaken for each other. Desnos-Ollivier<br />

et al. (2014) proposed species-specific carbon assimilation patterns and MALDI-T<strong>OF</strong><br />

MS profiles to enable the identification of S. clavata, M. capitatus and Geotrichum<br />

candidum to the species level.<br />

References: de Hoog and Smith (2004, 2011c), de Hoog et al. (2015), Garcia-Ruiz et<br />

al. (2013), Desnos-Ollivier et al. (2014), Arendrup et al. (2014).<br />

Antifungal Susceptibility: M. capitatus limited data (Garcia-Ruiz et al. 2013,<br />

Australian National data); MIC µg/mL. Note: Isolates of M. capitatus are intrinsically<br />

resistant to echinocandins (Arendrup et al. 2014).<br />

No. 64<br />

AmB 9 1 1 4 3<br />

FLU 9 1 1 2 3 2<br />

VORI 8 2 2 2 1 1<br />

POSA 8 2 1 2 2 1<br />

ITRA 9 1 4 4<br />

5FC 4 2 1 1


Descriptions of Medical Fungi 119<br />

Malassezia Baillon<br />

Malassezia species are basidiomycetous yeasts and form part of the normal skin flora<br />

of humans and animals. The genus now includes 14 species of which 13 are lipid<br />

dependent. These include M. caprae (goat, horse), M. cuniculi (rabbit), M. dermatis<br />

(human), M. equina (horse, cow), M. furfur (human, cow, elephant, pig, monkey, ostrich,<br />

pelican), M. globosa (human, cheetah, cow), M. japonica (human), M. nana (cat, cow,<br />

dog), M. obtusa (human), M. pachydermatis (dog, cat, carnivores, birds), M. restricta<br />

(human), M. slooffiae (human, pig, goat, sheep), M. sympodialis (human, horse, pig<br />

sheep) and M. yamatoensis (human) (Cabanes et al. 2011).<br />

M. sympodialis, M. globosa, M. slooffiae and M. restricta are the most frequently found<br />

species responsible for colonisation of humans (Arendrup et al. 2014).<br />

Malassezia species may cause various skin manifestations including pityriasis versicolor,<br />

seborrhoeic dermatitis, dandruff, atopic eczema and folliculitis. M. pachydermatis is<br />

known to cause external otitis in dogs. Fungaemia due to lipid-dependent Malassezia<br />

species usually occurs in patients with central line catheters receiving lipid replacement<br />

therapy, especially in infants (Tragiannides et al. 2010, Gaitanis et al. 2012, Arendrup<br />

et al. 2014).<br />

Note: With the exception of M. pachydermatis, the primary isolation and culture of<br />

Malassezia species is challenging because in vitro growth must be stimulated by<br />

natural oils or other fatty substances. The most common method used is to overlay<br />

Sabouraud’s dextrose agar (SDA) containing cycloheximide (actidione) with olive oil or<br />

alternatively to use a more specialised media like modified Leeming and Notham agar<br />

(Kaneko et al. 2007), or modified Dixon’s agar (see specialised culture media).<br />

However, CHROMagar Malassezia medium is now commercially available for the<br />

primary isolation and differentiation of the most common Malassezia species.<br />

Comment: For clinical management at the level of the individual patient, species<br />

identification is less important, although it is obviously needed for epidemiological<br />

surveillance and outbreak investigation (Arendrup et al. 2014).<br />

a<br />

b<br />

10 μm<br />

Malassezia furfur (a) culture on modified Dixon’s agar and (b) direct microscopy of skin<br />

scrapings showing characteristic clusters of thick-walled round, budding yeast-like cells<br />

and short angular hyphal forms (the so called “spaghetti and meatballs” appearance)<br />

typically seen in pityriasis versicolor.


120<br />

Descriptions of Medical Fungi<br />

RG-1 organisms.<br />

Malassezia Baillon<br />

Morphological Description: On media like modified Dixon’s agar, colonies are cream<br />

to yellowish, smooth or lightly wrinkled, glistening or dull, and with the margin being<br />

either entire or lobate. Malassezia is characterised by globose, oblong-ellipsoidal to<br />

cylindrical yeast cells. Reproduction is by budding on a broad base and from the same<br />

site at one pole (unipolar).<br />

Molecular Identification: ITS and D1/D2 sequencing may be used for accurate<br />

species identification (de Hoog et al. 2015).<br />

MALDI-T<strong>OF</strong> MS: Capable of identifying all 14 Malassezia species in concordance with<br />

those of ITS sequence analyses (Kolecka et al. 2014).<br />

Identification criteria for the differentiation of Malassezia species (de Hoog et. al. 2015).<br />

+ Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Species Buds SDA 40 O C<br />

Cremophor<br />

EL<br />

Tween<br />

80<br />

Tween<br />

60<br />

Tween<br />

40<br />

Tween<br />

20<br />

Esculine<br />

M. caprae narrow - - - + + + + + +<br />

M. couiculi narrow - + - - - - - nd +<br />

M. dermatis wide - + nd + + + + nd +<br />

M. equina narrow - - - + + + + + +<br />

M. furfur wide - + + + + + w +<br />

M. globosa narrow - - - - - - - +<br />

M. japonica wide - - nd - + w - nd +<br />

M. nana narrow - v nd w + + v nd +<br />

M. obtusa wide - - - - - - + +<br />

M. pachydermatis wide + + -,w + + -,w v v<br />

M. restricta narrow - - - - - - - -<br />

M. slooffiae wide - + - +,w + + - +<br />

M. sympodialis narrow - + -,w + + + + + +<br />

M. yamatoensis wide - - nd + + + + nd +<br />

Antifungal Susceptibility: Very limited data available. Special growth conditions<br />

are needed for antifungal susceptibility testing; data from Nakamura et al. (2000),<br />

Velegraki et al. (2004) and Miranda et al. (2007); MIC µg/mL.<br />

MIC µg/mL<br />

MIC µg/mL<br />

Antifungal Range MIC Antifungal<br />

90<br />

Range MIC 90<br />

FLU 0.125->64 4 (8) AmB 0.03-16 1 (8)<br />

Catalase<br />

References: Guillot and Gueho (1995), Gueho et al. (1996), Guillot et al. (1996, 2000),<br />

Boekhout et al. (2010), Cafarchia et al. (2011), de Hoog et al. (2015).<br />

ITRA 0.03-16 0.125 VORI 0.03-16 0.125 (1)<br />

KETO 0.03-4 0.25 POSA 0.03-32 0.125 (2)<br />

Note: There is no standardised method and results are often variable, therefore<br />

susceptibility testing for guiding treatment is not recommended.


Descriptions of Medical Fungi 121<br />

Malbranchea pulchella Sacc. & Penz.<br />

Malbranchea species are soil fungi of worldwide distribution which microscopically may<br />

resemble Coccidioides immitis/posadasii. Note: Culture identification by exoantigen<br />

test or ITS sequencing is the method of choice for identification of C. immitis/posadasii.<br />

RG-1 organism.<br />

Morphological Description: Colonies are white to sulphur-yellow to ochre-brown<br />

in colour, suede-like in texture, with a reddish-brown reverse, and often a reddish<br />

diffusible pigment. Microscopic morphology shows typical hyaline, one-celled,<br />

cylindrical, truncate, alternate arthroconidia produced in tightly coiled, terminal fertile<br />

branches of the hyphae. Arthroconidia are released by lysis of the disjunctor cells.<br />

These arthroconidia may be perceived as a yellow dust when released at maturity.<br />

Key Features: Hyphomycete producing alternate arthroconidia with disjunctor cells.<br />

References: Cooney and Emerson (1964), Sigler and Carmichael (1976), McGinnis<br />

(1980), Rippon (1988), de Hoog et al. (2000, 2015).<br />

20 µm<br />

Malbranchea pulchella arthroconidia produced in tightly<br />

coiled, terminal fertile branches of the hyphae.


122<br />

Descriptions of Medical Fungi<br />

Meyerozyma guilliermondii (Wick.) Kurtzman & M. Suzuki<br />

Synonymy: Candida guilliermondii (Castellani) Langeron & Guerra.<br />

Meyerozyma guilliermondii has been isolated from numerous human infections, mostly<br />

of cutaneous origin. It is also found on normal skin and in sea water, faeces of animals,<br />

fig wasps, buttermilk, leather, fish and beer. RG-1 organism.<br />

Culture: White to cream-coloured smooth, glabrous, yeast-like colonies.<br />

Microscopy: Spherical to subspherical budding yeast-like cells or blastoconidia, 2.0-<br />

4.0 x 3.0-6.5 µm.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Branched pseudohyphae with dense verticils of blastoconidia.<br />

Molecular Identification: ITS sequencing recommended.<br />

MALDI-T<strong>OF</strong> MS: Able to accurately identify this species.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose v L-Arabinose v D-Glucitol v<br />

Fermentation Sucrose + D-Arabinose v α-M-D-glucoside v<br />

Glucose + Maltose + D-Ribose + D-Gluconate v<br />

Galactose v Cellobiose v L-Rhamnose v DL-Lactate v<br />

Sucrose + Trehalose + D-Glucosamine + myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine + 2-K-D-gluconate +<br />

Lactose - Melibiose v Glycerol + D-Glucuronate nd<br />

Trehalose + Raffinose + Erythritol - Nitrate -<br />

Assimilation Melezitose v Ribitol + Urease -<br />

Glucose + Soluble Starch - Galactitol v 0.1% Cycloheximide v<br />

Galactose + D-Xylose + D-Mannitol v Growth at 37 O C v<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern.<br />

Antifungal Susceptibility: M. guilliermondii (Diekema et al. 2009, Australian<br />

National data); MIC µg/mL. CLSI clinical breakpoints are marked where available<br />

(Pfaller and Diekema 2012).<br />

No. 64<br />

AmB 198 1 1 2 9 70 70 27 9 2 1 1 5<br />

FLU 199 1 4 5 75 75 7 11 4 7<br />

VORI 197 2 11 23 88 52 9 3 1 2 6<br />

POSA 194 1 9 14 47 79 30 4 4 6<br />

ITRA 24 6 3 12 2 1<br />

ANID 120 1 2 6 5 9 41 48 8<br />

MICA 13 3 1 4 10 14 39 29 8 1<br />

CAS 175 2 9 24 35 63 25 8 2 3 4<br />

5FC 24 8 12 1 1 1 1


Descriptions of Medical Fungi 123<br />

Microsphaeropsis arundinis (S. Ahmad) B. Sutton<br />

Microsphaeropsis arundinis is a coelomycete that is ubiquitous in soil and fresh water.<br />

It typically inhabits terrestrial plant hosts and has a well-known association with Aruno<br />

donax, a garden escape weed known as ‘giant reed’ or ‘elephant grass’. M. arundinis<br />

is an emerging cause of phaeohyphomycosis in cats and immunosuppressed humans.<br />

RG-1 organism.<br />

Morphological Description: Colonies growing slowly, with dense aerial mycelium,<br />

initially greenish-grey, later becoming dark brown to grey-brown. Hyphae are septate,<br />

pigmented, and irregularly shaped, with swollen segments up to 4 μm in diameter.<br />

Pycnidia are subspherical, 250-350 μm in diameter; with a pseudoparenchymatous<br />

wall composed of very densely packed cells that appear angular in cross section<br />

(textura angularis). Conidiogenous cells ampulliform, up to 5 μm long. Conidia brown,<br />

thick- and smooth-walled, cylindrical, 3.5-4.5 × 1.0-1.5 μm.<br />

Key Features: Coelomycete, with ostiolate pycnidia, ampulliform conidiogenous cells,<br />

and small, smooth-walled, brown, cylindrical conidia.<br />

Molecular Identification: ITS and D1/D2 sequencing is recommended, especially as<br />

it may take many weeks before pycnidia are produced in culture (Reppas et al. 2015).<br />

References: Kluger et al. (2004), Pendle et al. (2004), Krockenberger et al. (2010),<br />

Hall et al. (2013), Reppas et al. (2015), de Hoog et al. (2015).<br />

a<br />

b<br />

Microsphaeropsis arundinis (a) culture and (b) pigmented<br />

septate hyphae, with swollen segments.


124<br />

Descriptions of Medical Fungi<br />

Microsporum Gruby<br />

A recent multilocus phylogenetic study the has reviewed the taxonomy of the<br />

dermatophytes. Arthroderma now contains 21 species, Epidermophyton one species,<br />

Lophophyton one species, Microsporum three species, Nannizzia nine species<br />

and Trichophyton 16 species. In addition, two new genera have been introduced:<br />

Guarromyces containing one species and Paraphyton three species (de Hoog et al.<br />

2016).<br />

The genus Microsporum is now restricted to just three species: M. audouinii, M.<br />

canis and M. ferrugineum. The remaining geophilic and zoophilic species, previously<br />

considered Microsporum species, have been transferred to the genera Lophophyton<br />

and Nannizzia.<br />

Microsporum species may form both macro- and microconidia, although they are not<br />

always present. Cultures are mostly granular to cottony, yellowish to brownish, with a<br />

cream-coloured or brown colony reverse. Macroconidia are hyaline, multiseptate, with<br />

thick rough cell walls, and are clavate, fusiform or spindle-shaped. Microconidia are<br />

single-celled, hyaline, smooth-walled, and are predominantly clavate in shape.<br />

Note: Strains of M. canis often do not produce macroconidia and/or microconidia on<br />

primary isolation media and subcultures onto polished rice grains or lactritmel agar<br />

are recommended to stimulate sporulation. These non-sporulating strains of M. canis<br />

are often erroneously identified as M. audouinii and it is surprising just how many<br />

laboratories have difficulty in differentiating M. canis and M. audouinii.<br />

Molecular Identification: ITS sequencing is recommended (Gräser et al. 1998, 2000,<br />

Brillowska-Dabrowska et al. 2013).<br />

References: Rebell and Taplin (1970), Rippon (1988), McGinnis (1980), Domsch et<br />

al. (1980), Ajello (1977), Weitzman et al. (1986), Mackenzie et al. (1986), Kane et<br />

al. (1997), de Hoog et al. (2000, 2015), Gräser et al. (1999a, 2008). Cafarchia et al.<br />

(2013).<br />

a<br />

b<br />

(a) Microsporum audouinii showing poor growth on rice grains, usually being<br />

visible only as a brown discolouration. (b) Microsporum canis on rice grains<br />

showing good growth, yellow pigmentation and sporulation.


Descriptions of Medical Fungi 125<br />

Microsporum audouinii is an anthropophilic fungus causing non-inflammatory infections<br />

of the scalp and skin, especially in children. Once the cause of epidemics of tinea<br />

capitis in Europe and North America, it is now less common. Invaded hairs show an<br />

ectothrix infection and usually fluoresce a bright greenish-yellow under Wood’s ultraviolet<br />

light. Only rarely found in Australasia, most reports are in fact misidentified nonsporulating<br />

strains of M. canis.<br />

RG-2 organism.<br />

Microsporum audouinii Gruby<br />

Morphological Description: Colonies are flat, spreading, greyish-white to light tanwhite<br />

in colour, and have a dense suede-like to downy surface, suggestive of mouse<br />

fur in texture. Reverse can be yellow-brown to reddish-brown in colour. Some strains<br />

may show no reverse pigment. Macroconidia and microconidia are rarely produced,<br />

most cultures are sterile or produce only occasional thick-walled terminal or intercalary<br />

chlamydospores. When present, macroconidia may resemble those of M. canis but<br />

are usually longer, smoother and more irregularly fusiform in shape; microconidia,<br />

when present, are pyriform to clavate in shape and are similar to those seen in other<br />

species of Microsporum, Lophophyton and Nannizzia. Pectinate (comb-like) hyphae<br />

and racquet hyphae (a series of hyphal segments swollen at one end) may also be<br />

present.<br />

a<br />

b<br />

10 µm<br />

Microsporum audouinii (a) Culture and (b) a thick-walled intercalary<br />

chlamydospore. Note: Macroconidia and microconidia are only rarely produced.<br />

Confirmatory Tests:<br />

Growth on Rice Grains: Very poor or absent, usually being visible only as a brown<br />

discolouration. This is one of the features which distinguish M. audouinii from M. canis.<br />

Reverse Pigment on Potato Dextrose Agar: Salmon to pinkish-brown (M. canis is<br />

bright yellow).<br />

BCP Milk Solids Glucose Agar: Both M. canis and M. audounii demonstrate profuse<br />

growth, but only M. audouinii shows a rapid pH change to alkaline (purple colour).<br />

Vitamin Free Agar (Trichophyton Agar No.1): Good growth indicating no special<br />

nutritional requirements. Cultures are flat, white, suede-like to downy, with a yellowbrown<br />

reverse. Note: Growth of some strains of M. audouinii is enhanced by the<br />

presence of thiamine (Trichophyton agar No.4).<br />

Hair Perforation Test: Negative after 28 days.<br />

Key Features: Absence of conidia, poor or no growth on polished rice grains, inability<br />

to perforate hair in vitro, and culture characteristics.


126<br />

Descriptions of Medical Fungi<br />

Microsporum canis is a zoophilic dermatophyte of worldwide distribution and is a<br />

frequent cause of ringworm in humans, especially children. Invades hair, skin and<br />

rarely nails. Cats and dogs are the main sources of infection. Invaded hairs show<br />

an ectothrix infection and fluoresce a bright greenish-yellow under Wood’s ultra-violet<br />

light.<br />

RG-2 organism.<br />

Morphological Description: Colonies are flat, spreading, white to cream-coloured, with<br />

a dense cottony surface which may show some radial grooves. Colonies usually have<br />

a bright golden yellow to brownish yellow reverse pigment, but non-pigmented strains<br />

may also occur. Macroconidia are typically spindle-shaped with 5-15 cells, verrucose,<br />

thick-walled and often have a terminal knob, 35-110 x 12-25 µm. A few pyriform to<br />

clavate microconidia are also present. Macroconidia and/or microconidia are often not<br />

produced on primary isolation media and it is recommended that subcultures be made<br />

onto lactritmel agar and/or boiled polished rice grains to stimulate sporulation.<br />

Confirmatory Tests:<br />

Microsporum canis Bodin<br />

Growth on Rice Grains: good growth of white aerial mycelium with production of<br />

yellow pigment. Microscopy reveals numerous macroconidia and microconidia similar<br />

to those described above.<br />

Reverse Pigment on Potato Dextrose Agar: Bright yellow (both M. audouinii and M.<br />

canis var. equinum are salmon to pinkish-brown).<br />

Vitamin Free Agar (Trichophyton Agar No.1): Good growth indicating no special<br />

nutritional requirements. Cultures are flat, white, suede-like to downy, with a yellow to<br />

pale yellow-brown reverse.<br />

Hair Perforation Test: Positive at 14 days.<br />

Key Features: Distinctive macroconidia and culture characteristics. Abundant growth<br />

and sporulation on polished rice grains and in vitro perforation of hair.<br />

Microsporum canis culture showing a bright golden yellow reverse pigment


Descriptions of Medical Fungi 127<br />

Microsporum canis Bodin<br />

20 µm<br />

Microsporum canis typical spindle-shaped macroconidia.<br />

Microsporum canis dysgonic strains are rare but may also occur. These dysgonic strains<br />

typically have a heaped and folded, yellow-brown thallus and macroconidia are usually<br />

absent. However, typical colonies and macroconidia of M. canis are usually produced<br />

by this variant when subcultured onto polished rice grains. Note: The dysgonic type<br />

colony of M. canis is similar to that of Microsporum ferrugineum.


128<br />

Descriptions of Medical Fungi<br />

Microsporum canis Bodin<br />

Supplementary description for Microsporum canis var. distortum, a dysgonic variant of<br />

M. canis with distinctive distorted macroconidia. Abundant growth and sporulation on<br />

rice grains.<br />

Microsporum canis var. distortum is a zoophilic fungus known to cause infections<br />

in cats, dogs and other animals. It is a rare cause of tinea capitis in New Zealand,<br />

Australia and North America. Clinical disease is similar to M. canis. Invaded hairs show<br />

an ectothrix infection and fluoresce a bright greenish-yellow under Wood’s ultra-violet<br />

light.<br />

a<br />

b<br />

10 μm<br />

Microsporum canis var. distortum (a) culture and (b) distorted macroconidia.<br />

Microsporum canis var. equinum is now considered to be a genotypic synonym of<br />

Microsporum canis (de Hoog et al. 2000). This variant of M. canis is a rare cause of<br />

ringworm of horses. Invaded hairs show an ectothrix infection and fluoresce a bright<br />

greenish-yellow under Wood’s ultra-violet light. Rarely infects man or other animal<br />

species. Reported from Australia, Europe and North America.<br />

a<br />

b<br />

10 μm<br />

Microsporum canis var. equinum (a) colonies are pale buff to pale salmon with, a buff<br />

to pinkish-buff to yellow-brown reverse. (b) Macroconidia are small, broad, irregular,<br />

spindle-shaped, 18-60 x 5-15 µm with rough thick walls and few septa. Microconidia<br />

are pyriform to clavate in shape, 3-9 x 1.5-3.5 µm, but are rarely produced.


Descriptions of Medical Fungi 129<br />

Microsporum ferrugineum Ota<br />

Microsporum ferrugineum is an anthropophilic fungus causing epidemic juvenile tinea<br />

capitis in humans. The clinical features are similar to those of infections caused by M.<br />

audouinii. Invaded hairs show an ectothrix infection and fluoresce a greenish-yellow<br />

under Wood’s ultra-violet light. Reported from Asia (including China and Japan),<br />

Russia, Eastern Europe and Africa.<br />

RG-2 organism.<br />

Morphological Description: Colonies are slow growing, forming a waxy, glabrous,<br />

convoluted thallus with a cream to buff-coloured surface and no reverse pigment.<br />

Note: Surface pigmentation may vary from cream to yellow to deep red and a flatter<br />

white form sometimes occurs. Cultures rapidly become downy and pleomorphic.<br />

Microscopic morphology is negative, microconidia or macroconidia are not produced.<br />

However, irregular branching hyphae with prominent cross walls (“bamboo hyphae”)<br />

and chlamydospores are seen. “Bamboo hyphae” are a characteristic of this species.<br />

Key Features: Clinical history, culture characteristics and distinctive “bamboo hyphae”.<br />

a<br />

b<br />

20 µm<br />

Microsporum ferrugineum (a) culture and (b) “bamboo hyphae”.


130<br />

Descriptions of Medical Fungi<br />

Mortierella wolfii Mehrotra & Baijal<br />

The genus Mortierella has now been placed in a separate order, the Mortierellales<br />

(Cavalier-Smith 1998). The genus contains about 90 recognised species, however<br />

Mortierella wolfii is probably the only pathogenic species being an important causal<br />

agent of bovine mycotic abortion, pneumonia and systemic mycosis in New Zealand,<br />

Australia, Europe and USA. RG-2 organism.<br />

Morphological Description: Cultures are fast growing, white to greyish-white, downy,<br />

often with a broadly zonate or lobed (rosette-like) surface appearance and no reverse<br />

pigment. Sporangiophores are typically erect, delicate, 80-250 µm in height, 6-20 µm<br />

wide at the base, arising from rhizoids or bulbous swellings on the substrate hyphae<br />

and terminating with a compact cluster of short acrotonous (terminal) branches.<br />

Sporangia are usually 15-48 µm in diameter, with transparent walls and a conspicuous<br />

collarette is usually present following dehiscence of the sporangiospores. Columellae<br />

are generally lacking and sporangiospores are single-celled, short-cylindrical, 6-10 x<br />

3-5 µm, with a double membrane. Chlamydospores with or without blunt appendages<br />

(amoeba-like) may be present, zygospores have not been observed. Temperature:<br />

grows well at 40-42 O C; maximum 48 O C.<br />

Key Features: Mucorales, rapid growth at 40 O C (thermotolerant), and characteristic<br />

delicate acrotonous branching sporangia without columellae.<br />

References: Domsch et al. (1980), McGinnis (1980), Rippon (1988), de Hoog et al.<br />

(2000, 2015).<br />

a<br />

b<br />

20 µm<br />

b<br />

20 µm<br />

Mortierella wolfii (a) culture showing a broadly zonate or lobed rosette-like surface<br />

appearance, and (b) sporangium, showing a sporangiophore, wide at the base, arising<br />

from rhizoids, and acrotonous (terminal) branches, collarettes and sporangiospores.


Descriptions of Medical Fungi 131<br />

Mucor Micheli ex Staint-Amans<br />

The genus Mucor contains about 50 recognised taxa, many of which have widespread<br />

occurrence and are of considerable economic importance (Zycha et al. 1969, Schipper<br />

1978, Domsch et al. 1980). However, only a few thermotolerant species are of medical<br />

importance and human infections are only rarely reported. Most infections reported<br />

list M. circinelloides and similar species such as M. indicus, M. ramosissimus, M.<br />

irregularis and M. amphibiorum as the causative agents. However, M. hiemalis and<br />

M. racemosus have also been reported as infectious agents, although their inability to<br />

grow at temperatures above 32 O C raises doubt as to their validity as human pathogens<br />

and their pathogenic role may be limited to cutaneous infections (Scholer et al. 1983,<br />

Goodman and Rinaldi 1991, Kwon-Chung and Bennett 1992, de Hoog et al. 2000,<br />

2015).<br />

Maximum temperature for growth of the reported pathogenic species of Mucor.<br />

Species Max temp. ( O C) Pathogenicity<br />

M. amphibiorum 36 Animals, principally amphibians<br />

M. circinelloides 37 Animals, occasionally humans<br />

M. hiemalis 30 Questionable cutaneous infections only<br />

M. indicus 42 Humans and animals<br />

M. irregularis 38 Humans<br />

M. racemosus 32 Questionable<br />

M. ramosissimus 36 Humans and animals<br />

Morphological Description: Colonies are very fast growing, cottony to fluffy, white to<br />

yellow, becoming dark-grey, with the development of sporangia. Sporangiophores are<br />

erect, simple or branched, forming large (60-300 µm in diameter), terminal, globose<br />

to spherical, multispored sporangia, without apophyses and with well-developed<br />

subtending columellae. A conspicuous collarette (remnants of the sporangial<br />

wall) is usually visible at the base of the columella after sporangiospore dispersal.<br />

Sporangiospores are hyaline, grey or brownish, globose to ellipsoidal, and smoothwalled<br />

or finely ornamented. Chlamydospores and zygospores may also be present.<br />

Key Features: Mucorales, large, spherical, non-apophysate sporangia with<br />

pronounced columellae and conspicuous collarette at the base of the columella<br />

following sporangiospore dispersal.<br />

Molecular Identification: ITS sequencing recommended (Walther et al. 2012).<br />

References: Schipper (1978), Domsch et al. (1980), McGinnis (1980), Onions et al.<br />

(1981), Scholer et al. (1983), Rippon (1988), Goodman and Rinaldi (1991), Samson<br />

et al. (1995), de Hoog et al. (2000, 2015), Schipper and Stalpers (2003), Ellis (2005b).


132<br />

Descriptions of Medical Fungi<br />

Mucor Micheli ex Staint-Amans<br />

20 µm 20 µm<br />

Mucor spp. showing sporangia, columella with inconspicuous<br />

collarette (arrow) and sporangiospores.<br />

RG-2 organism.<br />

Mucor amphibiorum Schipper<br />

Morphological Description: Colonies are greyish-brown, slightly aromatic and<br />

do not grow at 37 O C (maximum temperature for growth is 36 O C). Sporangiophores<br />

are hyaline, erect and mostly unbranched, rarely sympodially branched. Sporangia<br />

are dark-brown, up to 75 µm in diameter, and are slightly flattened with a diffluent<br />

membrane. Columellae are subglobose to ellipsoidal or pyriform, up to 60 x 50 µm,<br />

with small collarettes. Sporangiospores are smooth-walled, spherical, and 3.5-5.5 µm<br />

in diameter. Zygospores, when formed by compatible mating types, are spherical to<br />

slightly compressed, up to 70 x 60 µm in diameter, with stellate projections.<br />

Comment: Mucor amphibiorum is distinguished by poor branching of the<br />

sporangiophores and by globose sporangiospores. Ethanol and nitrates are not<br />

assimilated (Schipper 1978, Scholer et al. 1983, Hoog et al. 2000, 2015).<br />

Antifungal Susceptibility: M. amphibiorum very limited data (Australian National<br />

data); MIC µg/mL.<br />

No. 16<br />

AmB 1 1<br />

POSA 1 1<br />

ITRA 1 1


Descriptions of Medical Fungi 133<br />

M. circinelloides is a common and variable species that includes four formae:<br />

circinelloides, lusitanicus, griseocyanus and janssenii (Schipper 1978, Scholer et al.<br />

1983).<br />

RG-1 organism.<br />

Mucor circinelloides v. Tiegh<br />

Morphological Description: Colonies are floccose, pale greyish-brown and grow<br />

poorly at 37 O C (maximum growth temperature 37 O C). Sporangiophores are hyaline and<br />

mostly sympodially branched with long branches erect and shorter branches becoming<br />

circinate (coiled). Sporangia are spherical, varying from 20-80 µm in diameter, with<br />

small sporangia often having a persistent sporangial wall. Columellae are spherical<br />

to ellipsoidal and are up to 50 µm in diameter. Sporangiospores are hyaline, smoothwalled,<br />

ellipsoidal, and 4.5-7 x 3.5-5 µm in size. Chlamydospores are generally absent.<br />

Zygospores are only produced in crosses of compatible mating types and are reddishbrown<br />

to dark-brown, spherical with stellate spines, up to 100 µm in diameter and have<br />

equal to slightly unequal suspensor cells.<br />

Comment: M. circinelloides differs from other species of Mucor in its formation of<br />

short circinated, branched sporangiophores bearing brown sporangia and its ability<br />

to assimilate ethanol and nitrates (Schipper 1976, Scholer et al. 1983, Samson et al.<br />

1995, de Hoog et al. 2000, Schipper and Stalpers 2003).<br />

Antifungal Susceptibility: M. circinelloides (Espinel-Ingroff et al. 2015a, Australian<br />

National data); MIC µg/mL.<br />

No. 16<br />

AmB 123 1 4 14 42 44 18<br />

POSA 120 2 2 9 21 49 26 5 2 4<br />

ITRA 48 4 3 7 12 15 5 3<br />

Mucor circinelloides culture.


134<br />

Descriptions of Medical Fungi<br />

Mucor circinelloides v. Tiegh<br />

20 µm 20 µm<br />

Mucor circinelloides sporangia showing circinate sporangiophores, also note<br />

the columella with inconspicuous collarette (arrow) and sporangiospores.<br />

Mucor indicus Lendner<br />

Morphological Description: Colonies are characteristically deep-yellow, aromatic<br />

and have a maximum growth temperature of 42 O C. Sporangiophores are hyaline to<br />

yellowish, erect or rarely circinate and repeatedly sympodially branched, with long<br />

branches. Sporangia are yellow to brown, up to 75 µm in diameter, with diffluent<br />

membranes. Columellae are subglobose to pyriform, often with truncate bases, up to<br />

40 µm high. Sporangiospores are smooth-walled, subglobose to ellipsoidal, and 4-5<br />

µm in diameter. Chlamydospores are produced in abundance, especially in the light.<br />

Zygospores are black, spherical up to 100 µm in diameter, with stellate spines and<br />

unequal suspensor cells. RG-1 organism.<br />

Comment: Mucor indicus differs from other species of Mucor by its characteristic<br />

deep-yellow colony colour, growth at over 40 O C, assimilating ethanol, but not nitrate,<br />

and thiamine dependence (Schipper 1978, de Hoog et al. 2000, Schipper and Stalpers<br />

2003).<br />

Antifungal Susceptibility: M. indicus (Espinel-Ingroff et al. 2015a, Australian<br />

National data); MIC µg/mL.<br />

No. 16<br />

AmB 10 1 3 4 1 1<br />

POSA 10 2 3 3 1 1


Descriptions of Medical Fungi 135<br />

Synonymy: Rhizomucor variabilis Zheng & Chen.<br />

RG-2 organism.<br />

Morphological Description: Colonies are whitish to ochraceous, with buff-ochre<br />

reverse. Sporangiophores arising from hyphae or from stolons; rhizoids abundant.<br />

Sporangiophores are hyaline, up to 2 mm long, 9-23 μm wide, simple or once branched,<br />

with branches terminating at a higher level than the main stems; branches all ending in<br />

a sporangium. Sporangia subspherical to spherical, up to 100 μm diameter. Columella<br />

spherical, ellipsoidal to cylindrical, about 40 μm wide, sometimes lobed, with or without<br />

an apophysis. Sporangiospores are hyaline, smooth-walled, very variable, mostly<br />

subspherical to ellipsoidal, 3-11 × 2-7 μm. Chlamydospores are abundant. Maximum<br />

growth temperature 38 O C.<br />

Comment: Mucor irregularis differs from other species of Mucor by having abundant<br />

rhizoids of different sizes and sporangiospores of highly variable shape, mostly<br />

subspherical to ellipsoidal (Lu et al. 2009, 2013). This species is closely related to<br />

Mucor hiemalis (Voigt et al. 1999).<br />

RG-1 organism.<br />

Mucor irregularis Stchigel et al.<br />

Mucor ramosissimus Samutsevich<br />

Morphological Description: Colonial growth is restricted, greyish and does not<br />

grow at 37 O C (maximum temperature for growth is 36 O C). Sporangiophores are<br />

hyaline, slightly roughened, tapering towards the apex and are erect with repeated<br />

sympodial branching. Sporangia are grey to black, globose or somewhat flattened,<br />

up to 80 µm in diameter and have very persistent sporangial walls. Columellae are<br />

applanate (flattened), up to 40-50 µm in size and are often absent in smaller sporangia.<br />

Sporangiospores are faintly brown, smooth-walled, subglobose to broadly ellipsoidal,<br />

5-8 x 4.5-6 µm in size. Oidia may be present in the substrate hyphae, chlamydospores<br />

and zygospores are absent. Assimilation of ethanol is negative and that of nitrate is<br />

positive.<br />

Comment: Mucor ramosissimus differs from other species of Mucor by its low, restricted<br />

growth on any medium, extremely persistent sporangial walls, columellae that are<br />

applanate or absent in smaller sporangia (often resembling Mortierella species), short<br />

sporangiophores that repeatedly branch sympodially as many as 12 times, and the<br />

occurrence of racket-shaped enlargements in the sporangiophores (Hesseltine and<br />

Ellis 1964b, Schipper 1976, Scholer et al. 1983, de Hoog et al. 2000, Schipper and<br />

Stalpers 2003).<br />

Antifungal Susceptibility: M. ramosissimus (Espinel-Ingroff et al. 2015a, Australian<br />

National data); MIC µg/mL.<br />

No. 16<br />

AmB 19 2 4 3 6 1 1<br />

POSA 13 4 4 2 2 1


136<br />

Descriptions of Medical Fungi<br />

Myrmecridium schulzeri (Sacc.) Arzanlou et al.<br />

Synonymy: Ramichloridium schulzeri Stahel ex de Hoog.<br />

Ramichloridium schulzeri was placed in a new genus, Myrmecridium by Arzanlou et al.<br />

(2007). M. schulzeri is an uncommon soil saprophyte of worldwide distribution. It has<br />

also been isolated from plant detritus and as a contaminant of bronchoscopy fluid. It<br />

is the causative agent of “Golden Tongue” syndrome reported by Rippon et al. (1985).<br />

RG-1 organism.<br />

Morphological Description: Colonies growing moderately rapidly, consisting of a<br />

rather compact, flat, submerged mycelium, pale orange, locally with some powdery,<br />

brownish aerial mycelium; reverse pink to orange. Conidiophores are erect, straight,<br />

unbranched, thick-walled, reddish-brown, up to 250 µm high, gradually becoming paler<br />

towards the apex, of variable length, elongating sympodially during conidiogenesis, with<br />

scattered, pimple-shaped conidium bearing denticles which have unpigmented scars.<br />

Conidia are subhyaline, smooth-walled or slightly rough-walled, ellipsoidal, obovoidal<br />

or fusiform, 6.5-10 x 3-4 µm, usually with an acuminate base and unpigmented scars.<br />

Molecular Identification: ITS and D1/D2 sequencing may be used for accurate<br />

species identification (Halliday et al. 2015).<br />

Note: Myrmecridium species can be distinguished from other Ramichloridium-like fungi<br />

by having entirely hyaline vegetative hyphae, and widely scattered, pimple-shaped<br />

denticles on the long hyaline rachis. The conidial sheath is also visible in lactic acid<br />

mounts with bright field microscopy Arzanlou et al. (2007).<br />

References: de Hoog (1977), Rippon et al. (1985), de Hoog et al. (2000, 2015),<br />

Arzanlou et al. (2007).<br />

10 µm<br />

20 µm<br />

Myrmecridium schulzeri conidiophores showing sympodial development of conidia.


Descriptions of Medical Fungi 137<br />

Nannizzia fulva (Uriburu) Stockdale<br />

Synonymy: Microsporum fulvum Uriburu.<br />

Nannizzia fulva is a geophilic fungus of worldwide distribution which may cause<br />

occasional infections in humans and animals. Clinical disease is similar to N. gypsea<br />

but less common. Invaded hairs show a sparse ectothrix infection but do not fluoresce<br />

under Wood’s ultra-violet light. RG-1 organism.<br />

Morphological Description: Colonies are fast growing, flat, suede-like, tawny-buff<br />

to pinkish-buff in colour and frequently have a fluffy white advancing edge. A dark red<br />

under surface is occasionally seen, otherwise it is colourless to yellow brown. Abundant<br />

thin-walled, elongate, ellipsoidal macroconidia are formed which closely resemble<br />

those of N. gypsea, except they are longer and more bullet-shaped (clavate) with three<br />

to six septa. Numerous spiral hyphae, which are often branched are seen. Numerous<br />

pyriform to clavate microconidia are also produced but these are not diagnostic.<br />

Key Features: Macroconidial morphology and culture characteristics.<br />

Molecular Identification: ITS sequencing is recommended.<br />

References: Rebell and Taplin (1970), Rippon (1988), Gräser et al. (2008), Cafarchia<br />

et al. (2013), de Hoog et al. (2015, 2016).<br />

a<br />

b<br />

20 µm<br />

Nannizzia fulva (a) culture and (b) macroconidia.


138<br />

Descriptions of Medical Fungi<br />

Nannizzia gypsea (Nannizzi) Stockdale<br />

Synonymy: Microsporum gypseum (Bodin) Guiart & Grigorakis.<br />

Nannizzia gypsea is a geophilic fungus with a worldwide distribution which may cause<br />

infections in animals and humans, particularly children and rural workers during warm<br />

humid weather. Usually produces a single inflammatory skin or scalp lesion. Invaded<br />

hairs show an ectothrix infection but do not fluoresce under Wood’s ultra-violet light.<br />

RG-1 organism.<br />

Morphological Description: Colonies are usually flat, spreading, suede-like to<br />

granular, with a deep cream to tawny-buff to pale cinnamon-coloured surface. Many<br />

cultures develop a central white downy umbo (dome) or a fluffy white tuft of mycelium<br />

and some also have a narrow white peripheral border. A yellow-brown pigment, often<br />

with a central darker brown spot, is usually produced on the reverse, however a reddishbrown<br />

reverse pigment may be present in some strains. Cultures produce abundant,<br />

symmetrical, ellipsoidal, thin-walled, verrucose, four to six-celled macroconidia. The<br />

terminal or distal ends of most macroconidia are slightly rounded, while the proximal<br />

ends (point of attachment to hyphae) are truncate. Numerous clavate-shaped<br />

microconidia are also present, but these are not diagnostic.<br />

Key Features: Distinctive macroconidia and culture characteristics.<br />

Molecular Identification: ITS sequencing is recommended, especially for the<br />

separation of N. gypsea and N. incurvata which are morphologically similar.<br />

References: Rebell and Taplin (1970), Rippon (1988), Gräser et al. (2008), Cafarchia<br />

et al. (2013), de Hoog et al. (2015, 2016).<br />

a<br />

b<br />

20 µm<br />

Nannizzia gypsea (a) culture and (b) macroconidia.


Descriptions of Medical Fungi 139<br />

Nannizzia nana (Fuentes) Gräser & de Hoog<br />

Synonymy: Microsporum nanum Fuentes<br />

Nannizzia nana is a geophilic and zoophilic fungus frequently causing chronic noninflammatory<br />

lesions in pigs and a rare cause of tinea in humans. Also present in soil of<br />

pig-yards. Infections in man are usually contracted directly from pigs or fomites. Invaded<br />

hairs may show a sparse ectothrix or endothrix infection but do not fluoresce under<br />

Wood’s ultra-violet light. The geographical distribution is worldwide. RG-2 organism.<br />

Morphological Description: Colonies are flat, cream to buff in colour with a suedelike<br />

to powdery surface texture. Young colonies have a brownish-orange pigment which<br />

deepens into a dark reddish-brown with age. Cultures produce numerous small ovoid<br />

to pyriform macroconidia with one to three (mostly two) cells, with relatively thin, finely<br />

echinulate (rough) walls, and broad truncate bases. Many macroconidia are borne on<br />

conidiophores (stalks) which do not stain readily. Occasional clavate microconidia are<br />

present, which distinguishes N. nana from some species of Chrysosporium.<br />

Key Features: Distinctive macroconidia and culture characteristics.<br />

Molecular Identification: ITS sequencing is recommended.<br />

References: Rebell and Taplin (1970), Rippon (1988), Gräser et al. (2008), Cafarchia<br />

et al. (2013), de Hoog et al. (2015, 2016).<br />

a<br />

b<br />

20 µm<br />

Nannizzia nana (a) culture and (b) macroconidia.


140<br />

Descriptions of Medical Fungi<br />

Nannizzia persicolor (Sabouraud) Stockdale<br />

Synonymy: Microsporum persicolor (Sabouraud) Guiart & Grigorakis<br />

Nannizzia persicolor is a zoophilic fungus often occurring as a saprophyte on voles<br />

and bats. A rare cause of tinea corporis in humans. Not known to invade hair in vivo,<br />

but produces hair perforations in vitro. Distribution: Africa, Australia, Europe and North<br />

America. RG-2 organism.<br />

Morphological Description: Colonies are generally flat, white to pinkish in colour, with<br />

a suede-like to granular texture and peripheral fringe. Reverse pigmentation is orange<br />

to red. Macroconidia are thin-walled, cigar-shaped, four to seven-celled, 40-60 x 6-8<br />

µm but are only rarely produced. Microconidia are abundant, spherical to pyriform.<br />

Key Features: Microscopic morphology and culture characteristics.<br />

Molecular Identification: ITS sequencing is recommended.<br />

References: Rebell and Taplin (1970), Rippon (1988), Gräser et al. (2008), Cafarchia<br />

et al. (2013), de Hoog et al. (2015, 2016).<br />

a<br />

b<br />

Nannizzia persicolor (a) culture and (b) microconidia.<br />

20 µm


Descriptions of Medical Fungi 141<br />

Synonymy: Hendersonula toruloidea Nattras.<br />

Scytalidium dimidiatum (Penzig) Sutton & Dyko.<br />

Scytalidium hyalinum Campbell & Mulder.<br />

Neoscytalidium dimidiatum is a coelomycete and recognised agent of onychomycosis<br />

and superficial skin infections, especially in tropical regions. The primary isolation<br />

of this fungus from clinical specimens may be difficult as isolates are sensitive to<br />

cycloheximide (actidione), which is commonly added to primary isolation media used<br />

for culturing skin scrapings.<br />

The taxonomy of this species has been very confusing; the conidial state of S.<br />

dimidiatum was originally described under the name Hendersonula toruloidea.<br />

However it is phylogenetically remote from Scytalidium and has now been placed<br />

into Neoscytalidium (Crous et al. 2006, Machouart et al. 2012). Colourless mutants<br />

(previously known as Scytalidium hyalinum) often occur and have now been listed as<br />

variety N. dimidiatum var. hyalinum (Madrid et al. 2009).<br />

Note: Nattrassia mangiferae, previously considered the teleomorph form of S.<br />

dimidiatum, is now considered a distinct species, and has been placed in the genus<br />

Neofusicoccum, based on the lack of an arthoconidial anamorph (Crous et al. 2006,<br />

Machouart et al. 2012).<br />

RG-2 organism.<br />

Neoscytalidium dimidiatum (Penzig) Crous & Slippers<br />

Morphological Description: Cultures are effuse, hairy, dark grey to blackish-brown, or<br />

white to greyish, with a cream-coloured to deep ochraceous-yellow reverse. Colourless<br />

mutants often occur. Arthroconidia are typically present in chains of one to two-cells,<br />

darkly pigmented, 3.5-5 x 6.5-12 µm, produced by the holothallic fragmentation of<br />

undifferentiated hyphae. Pycnidia, only occasionally formed in older cultures are black,<br />

ostiolate and contain numerous hyaline, flask-shaped phialides. Phialoconidia are at<br />

first one-celled and hyaline, later becoming three-celled, brown, with the centre cell<br />

darker than the end cells and are ovoid to ellipsoidal in shape.<br />

Molecular Identification: ITS and D1/D2 sequencing may be used for accurate<br />

species identification (Halliday et al. 2015).<br />

MALDI-T<strong>OF</strong> MS: Alshawa et al. (2012) developed a spectral database for 12 different<br />

species of dermatophytes which also included Neoscytalidium dimidiatum and N.<br />

dimidiatum var. hyalinum. Correct identification of the species was obtained for 18/21<br />

Neoscytalidium isolates (85.7%).<br />

References: McGinnis (1980), Moore (1986), Rippon (1988), Frankel and Rippon<br />

(1989), Sutton and Dyko (1989) Madrid et al. (2009), de Hoog et al. (2000, 2015),<br />

Crous et al. (2006), Machouart et al. (2012), Alshawa et al. (2012).


142<br />

Descriptions of Medical Fungi<br />

Neoscytalidium dimidiatum (Penzig) Crous & Slippers<br />

20 µm<br />

Neoscytalidium dimidiatum showing chains of one to two<br />

celled, darkly pigmented arthroconidia.<br />

Antifungal Susceptibility: Neoscytalidium dimidiatum limited data (Australian<br />

National data); MIC µg/mL.<br />

No. 64<br />

AmB 3 1 1 1<br />

VORI 3 1 1 1<br />

POSA 3 1 1 1<br />

ITRA 3 1 1 1<br />

N. dimidiatum data from 24 isolates (Madrid et al. 2009); MIC µg/mL.<br />

AmB Range 0.125-0.5; Geometric mean = 0.32<br />

VORI Range 0.06-4; Geometric mean = 1.37<br />

POSA Range 0.06-32; Geometric mean = 6.15<br />

N. dimidiatum data from 10 isolates (Espinel-Ingroff et al. 2001); MIC µg/mL.<br />

ITRA Range 0.03-32; Geometric mean = 0.65


Descriptions of Medical Fungi 143<br />

Ochroconis de Hoog and Arx<br />

Recently, the genus Ochroconis has undergone taxonomic revision, and the most<br />

relevant species, Ochroconis gallopava, was transferred to the new genus Verruconis<br />

(Samerpitak et al. 2014). Ochroconis species are mesophilic saprobes, with an<br />

optimum growth temperature between 15 and 30 O C and an inability to grow at 37 O C.<br />

Nevertheless, some species have been isolated from clinical specimens; such as<br />

Ochroconis tshawytschae, O. mirabilis, O. cordanae, O. olivacea and O. ramosa<br />

(Samerpitak et al. 2014, Seyedmousavi et al. 2014, Giraldo et al. 2014). Note: Species<br />

described in the literature from clinical cases as O. constricta or O. humicola are<br />

probably O. musae (Samerpitak et al. 2014).<br />

Giraldo et al. (2014) reported on the occurrence of Ochroconis and Verruconis species in<br />

clinical specimens from the United States. V. gallopava was the most common species<br />

(69%), followed by O. mirabilis (22%). Other species isolated were O. cordanae, O.<br />

olivacea and O. ramosa. The most common anatomical site of isolation was the lower<br />

respiratory tract (59%), followed by superficial (22%) and deep tissues (20%) (Giraldo<br />

et al. 2014).<br />

RG-1 organisms.<br />

Morphological Description: Colonies restricted, velvety to funiculose, brown to<br />

olivaceous, often with rust-brown reverse. Hyphae smooth- or somewhat rough-walled,<br />

pale olivaceous. Conidiophores slightly or conspicuously differentiated, cylindrical,<br />

often flexuose, producing conidia on scattered, cylindrical to conical denticles. After<br />

detachment an inconspicuous frill often remains both on the denticle and on the<br />

conidium base. Conidia one to four-celled, pale olivaceous brown, smooth- or roughwalled,<br />

ellipsoidal, cylindrical, clavate or cuneiform. Maximum growth temperature<br />

around 33 O C.<br />

Molecular Identification: Giraldo et al. (2014) used sequence analyses of the 18S,<br />

ITS, D1/2, actin, and β-tubulin genes, while Seyedmousavi et al. (2014), used ITS<br />

sequence analyses to identify species.<br />

References: de Hoog et al. (2015), Samerpitak et al. (2014), Seyedmousavi et al.<br />

(2014), Giraldo et al. (2014).<br />

Antifungal Susceptibility: Ochroconis mirabilis variable data for posaconazole,<br />

caspofungin and anidulafungin from Seyedmousavi et al. (2014) 1 and Giraldo et al<br />

(2014) 2 ; MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

0.06-0.25<br />

AmB 1-32 32 POSA<br />

1 -<br />

0.5-32 2 32<br />

ITRA 0.25-32 32 VORI 2-32 32<br />

CAS<br />

0.5-1 1<br />

1-32 2 -<br />

4<br />

ANID<br />

0.03-0.125 1<br />

0.015-32 2 -<br />

4<br />

MICA 0.06-0.50 0.25 TERB 0.015-0.125 0.02


144<br />

Descriptions of Medical Fungi<br />

Onychocola canadensis Sigler<br />

Onychocola canadensis is an uncommon cause of distal and lateral subungual or white<br />

superficial onychomycosis. However, it may sometimes be present in an abnormalappearing<br />

nail as an insignificant finding, not acting as a pathogen.<br />

RG-2 organism.<br />

Morphologial Description: Colonies grow slowly and are velvety to lanose, white to<br />

yellowish, with a brownish reverse. Arthroconidia are cylindrical to broadly ellipsoidal,<br />

one to two-celled, hyaline to subhyaline, 4-16 x 2-5 µm in size, forming long chains.<br />

Older cultures may show broad, brown, rough-walled hyphae.<br />

Key Features: Slow growing, white, arthroconidial mould isolated from nails.<br />

References: Sigler and Congly (1990), Sigler et al. (1994), Gupta et al. (1998), de<br />

Hoog et al. (2000, 2015).<br />

a<br />

b<br />

10 µm<br />

Onychocola canadensis (a) culture and (b) arthroconidia.


Descriptions of Medical Fungi 145<br />

Paecilomyces Bain<br />

The genus Paecilomyces may be distinguished from the closely related genus Penicillium<br />

by having long slender divergent phialides and colonies that are never typically<br />

green. Paecilomyces species are common environmental moulds and are seldom<br />

associated with human infection. However, the species, P. variotii and P. marquandii<br />

are emerging as causative agents of mycotic keratitis and of hyalohyphomycosis in the<br />

immunocompromised patient. Note: Paecilomyces lilacinus has been transferred to<br />

Purpureocillium lilacinum (Luangsa-ard et al. 2011).<br />

Morphological Description: Colonies are fast growing, powdery or suede-like, gold,<br />

green-gold, yellow-brown, lilac or tan, but never green or blue-green as in Penicillium.<br />

Phialides are swollen at their bases, gradually tapering into a rather long and slender<br />

neck, and occur solitarily, in pairs, as verticils, and in penicillate heads. Long, dry chains<br />

of single-celled, hyaline to dark, smooth or rough, ovoid to fusoid conidia are produced<br />

in basipetal succession from the phialides.<br />

Molecular Identification: Molecular phylogeny based on 18S rDNA sequences was<br />

done by Luangsa-ard et al. (2004); the genus is polyphyletic.<br />

Key Features: Long slender divergent phialides and culture pigmentation.<br />

References: Samson (1974), Domsch et al. (1980), McGinnis (1980), Onions et al.<br />

(1981), Rippon (1988), de Hoog et al. (2000, 2015).<br />

a<br />

b<br />

Cultures of (a) Paecilomyces marquandii and (b) Paecilomyces variotii<br />

showing colony pigmentation.


146<br />

Descriptions of Medical Fungi<br />

Paecilomyces marquandii (Massee) Hughes<br />

Paecilomyces marquandii is a soil fungus of worldwide distribution from temperate to<br />

tropical regions.<br />

RG-1 organism.<br />

Morphological Description: Colonies are fast growing, suede-like to floccose, pale<br />

vinaceous to violet-coloured, with a yellow to orange yellow reverse. Conidiophores are<br />

erect, arising from submerged hyphae, 50-300 µm in length, bearing loose whorls of<br />

branches and phialides. Conidiophore stipes are 2.5-3.0 µm wide, hyaline and smoothwalled.<br />

Phialides are swollen at their bases, tapering into a thin, distinct neck. Conidia<br />

are ellipsoidal to fusiform, smooth-walled to slightly roughened, hyaline to purple in<br />

mass, 2.5-3.0 x 2-2.2 µm. Spherical to ellipsoidal chlamydospores, 3-5 µm diameter<br />

are present. No growth at 37 O C.<br />

Key Features: Colony pigmentation with yellow reverse pigment, phialides with swollen<br />

bases, smooth conidiophore stipes, presence of chlamydospores, and no growth at<br />

37 O C. Note: Purpureocillium lilacinum has no yellow reverse pigment, rough-walled<br />

conidiophore stipes, absence of chlamydospores and growth at 37 O C.<br />

References: Samson (1974), Domsch et al. (1980, 2007), de Hoog et al. (2000, 2015).<br />

20 µm<br />

Paecilomyces marquandii conidiophores, phialides and conidia.<br />

Antifungal Susceptibility: P. marquandii (Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 7 1 1 1 1 3<br />

VORI 4 1 2 1<br />

POSA 2 1 1<br />

ITRA 7 1 2 2 1 1


Descriptions of Medical Fungi 147<br />

Paecilomyces variotii is a common environmental mould that is widespread in composts,<br />

soils and food products. It is known from substrates including food, indoor air, wood,<br />

soil and carpet dust.<br />

RG-2 organism.<br />

Paecilomyces variotii Bain<br />

Morphological Description: Colonies are fast growing, powdery to suede-like,<br />

funiculose or tufted, and yellow-brown or sand-coloured. Conidiophores bearing<br />

dense, verticillately arranged branches bearing phialides. Phialides are cylindrical or<br />

ellipsoidal, tapering abruptly into a long and cylindrical neck. Conidia are subspherical,<br />

ellipsoidal to fusiform, hyaline to yellow, smooth-walled, 3-5 x 2-4 µm and are produced<br />

in long divergent chains. Chlamydospores are usually present, singly or in short chains,<br />

brown, subspherical to pyriform, 4-8 µm in diameter, thick-walled to slightly verrucose.<br />

Key Features: Yellow-brown colony pigmentation, cylindrical phialides, and presence<br />

of chlamydospores.<br />

a<br />

10 µm b 5 µm c 10 µm<br />

Paecilomyces variotii (a) conidiophores, phialides, (b) conidia<br />

and (c) terminal chlamydospores.<br />

Antifungal Susceptibility: P. variotii (Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 17 7 3 4 2 1<br />

VORI 17 1 2 1 12 1<br />

POSA 17 5 3 2 5 2<br />

ITRA 17 4 5 5 3


148<br />

Descriptions of Medical Fungi<br />

Paracoccidioides brasiliensis/lutzii Complex<br />

WARNING: RG-3 organism. Cultures of Paracoccidioides brasiliensis/lutzii represent<br />

a biohazard to laboratory personnel and should be handled with extreme caution in a<br />

Class II Biological Safety Cabinet (BSCII).<br />

Recently P. brasiliensis has been recognised as two species: P. brasiliensis and P.<br />

lutzii (Teixeira et al. 2014, Theodoro et al. 2012). P. brasiliensis/lutzii is geographically<br />

restricted to areas of South and Central America. The two species are morphologically<br />

very similar; conidia of P. lutzii are elongated whereas those from P. brasiliensis are<br />

pyriform. Molecular confirmation is recommended.<br />

Molecular Identification: ITS sequencing is recommended (Imai et al. 2000)<br />

Morphological Description: Colonies grown at 25 O C are slow growing and variable<br />

in morphology. Colonies may be flat, wrinkled and folded, glabrous, suede-like or<br />

downy in texture, white to brownish with a tan or brown reverse. Microscopically, a<br />

variety of conidia may be seen, including pyriform microconidia, chlamydospores and<br />

arthroconidia. However, none of these are characteristic of the species, and most<br />

strains may grow for long periods of time without the production of conidia.<br />

On blood agar at 37 O C, the mycelium converts to the yeast phase and colonies are white<br />

to tan, moist and glabrous and become wrinkled, folded and heaped. Microscopically,<br />

numerous large, 20-60 μm, round, narrow base budding yeast cells are present. Single<br />

and multiple budding occurs, the latter are thick-walled cells that form the classical<br />

“steering wheel” or “mickey mouse” structures that are diagnostic for this fungus,<br />

especially in methenamine silver stained tissue sections.<br />

Key Features: Clinical history, tissue pathology, culture identification with conversion<br />

to yeast phase at 37 O C, however molecular identification is now recommended.<br />

References: McGinnis (1980), Chandler et al. (1980), Rippon (1988), de Hoog et al.<br />

(2000, 2015).<br />

20 µm<br />

20 µm<br />

Paracoccidioides brasiliensis/lutzii showing multiple,<br />

narrow base budding yeast cells “steering wheels”.<br />

Antifungal Susceptibility: P. brasiliensis very limited data (McGinnis et al. 1997).<br />

Antifungal MIC Range µg/mL Antifungal MIC Range µg/mL<br />

AmB 0.03-4 ITRA


Descriptions of Medical Fungi 149<br />

Synonymy: Microsporum cookei Ajello; Microsporum racemosum Borelli.<br />

Lophophyton cookei is a geophilic fungus which has been isolated from the hair of<br />

small mammals showing no clinical lesions. Infection has been reported in rodents,<br />

dogs and rarely in humans. It is not known to invade hair in vivo, but produces hair<br />

perforations in vitro. RG-1 organism.<br />

Morphological Description: Colonies are flat, spreading, buff to pale brown, powdery<br />

to suede-like, with a slightly raised and folded centre and some radial grooves. Reverse<br />

pigment dark reddish-brown. Numerous large, very thick-walled, echinulate (rough)<br />

elliptical macroconidia with predominantly five to six septa but may be from two to eight<br />

septa. Occasional spiral hyphae may be seen. Moderate numbers of mainly slender<br />

clavate with some pyriform microconidia are present.<br />

Key Features: The macroconidia of L. cookei are quite characteristic and diagnostic;<br />

the thick walls and usually larger size of the macroconidia distinguish it from N. gypsea.<br />

Confirmatory Tests:<br />

Paraphyton cookei (Ajello) Gräser & de Hoog<br />

Vitamin Free Agar (Trichophyton Agar No.1): Good growth indicating no special<br />

nutritional requirements, pinkish-buff-coloured, suede-like colony with a deep magenta<br />

red reverse.<br />

Hair Perforation Test: Positive.<br />

Key Features: Distinctive macroconidial morphology and culture characteristics.<br />

Molecular Identification: ITS sequencing is recommended.<br />

References: Rebell and Taplin (1970), Rippon (1988), Gräser et al. (2008), Cafarchia<br />

et al. (2013), de Hoog et al. (2015, 2016).<br />

a<br />

b<br />

20 µm<br />

Lophophyton cookei (a) culture and (b) macroconidia.


150<br />

Descriptions of Medical Fungi<br />

Penicillium is a very large and ubiquitous genus which currently contains 354 accepted<br />

species (Visagie et al. 2014). Many species are common contaminants on various<br />

substrates and are known as potential mycotoxin producers. Correct identification is<br />

therefore important when studying possible Penicillium contamination of food. Human<br />

pathogenic species are rare, however opportunistic infections leading to mycotic<br />

keratitis, otomycosis and endocarditis (following insertion of valve prosthesis) have<br />

been reported (Lyratzopoulos et al. 2002). Note: Penicillium marneffei and other<br />

subgenus Biverticillium species have been transferred to the genus Talaromyces<br />

(Samson et al. 2011b).<br />

RG-1 organisms.<br />

Penicillium Link: Fries<br />

Morphological Description: Colonies are usually fast growing, in shades of green,<br />

sometimes white, mostly consisting of a dense felt of conidiophores. Microscopically,<br />

chains of single-celled conidia are produced in basipetal succession from a specialised<br />

conidiogenous cell called a phialide. The term basocatenate is often used to describe<br />

such chains of conidia where the youngest conidium is at the basal or proximal end of<br />

the chain. In Penicillium, phialides may be produced singly, in groups or from branched<br />

metulae, giving a brush-like appearance (a penicillus). The penicillus may contain both<br />

branches and metulae (penultimate branches which bear a whorl of phialides). All cells<br />

between the metulae and the stipes of the conidiophores are referred to as branches.<br />

The branching pattern may be either simple (non-branched or monoverticillate),<br />

one-stage branched (biverticillate-symmetrical), two-stage branched (biverticillateasymmetrical)<br />

or three- to more-staged branched. Conidiophores are hyaline, smooth<br />

or rough-walled. Phialides are usually flask-shaped, consisting of a cylindrical basal<br />

part and a distinct neck, or lanceolate (with a narrow basal part tapering to a somewhat<br />

pointed apex). Conidia are in long dry chains, divergent or in columns, are globose,<br />

ellipsoidal, cylindrical or fusiform, hyaline or greenish, smooth or rough-walled. Sclerotia<br />

are produced by some species.<br />

phialides<br />

metulae<br />

branches<br />

a b c d<br />

Morphological structures and types of conidiophore branching in Penicillium.<br />

(a) Monoverticillate; (b) Biverticillate; (c) Terverticillate; (d) Quaterverticillate<br />

(see Visagie et al. 2014).


Descriptions of Medical Fungi 151<br />

Penicillium Link:Fries<br />

For identification, isolates are usually inoculated at three points on Czapek Dox agar<br />

and 2% Malt extract agar and incubated at 25 O C. Most species sporulate within 7 days.<br />

Microscopic mounts are best made using a cellotape flag or a slide culture preparation<br />

mounted in lactophenol cotton blue. A drop of alcohol is usually needed to remove<br />

bubbles and excess conidia (Samson et al. 1995).<br />

Molecular Identification: ITS and/or β-tubulin loci are recommended for identification<br />

of Penicillium species (Visagie et al. 2014, Yilmaz et al. 2014).<br />

Key Features: Hyphomycete, flask-shaped phialides arranged in groups from branched<br />

metulae forming a penicillus.<br />

References: Raper and Thom (1949), Pitt (1979), Domsch et al. (1980), McGinnis<br />

(1980), Onions et al. (1981), Ramirez (1982), Samson et al. (1995, 2011b), de Hoog et<br />

al. (2000, 2015), Visagie et al. (2014).<br />

a<br />

20 µm b<br />

20 µm<br />

(a) P. verrucosum var. cyclopium conidiophores showing two-stage branching.<br />

(b) P. cheresanum simple conidiophore showing chains of single-celled conidia.<br />

Antifungal Susceptibility: Penicillium sp. (Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 28 2 3 5 12 5 1<br />

VORI 26 2 4 6 3 2 8 1<br />

POSA 26 1 3 4 5 3 6 4<br />

ITRA 27 3 4 1 6 8 1 4


152<br />

Descriptions of Medical Fungi<br />

Phaeoacremonium parasiticum (Ajello et al.) W. Gams et al.<br />

Synonymy: Phialophora parasiticum Ajello, Gerog & Wang.<br />

The genus Phaeoacremonium initially accommodated isolates with features similar to<br />

those seen in both Acremonium and Phialophora. It differs from the former by having<br />

pigmented hyphae and conidiophores and from the latter by having indistinct collarettes<br />

and warty conidiogenous cells (Revankar and Sutton, 2010).<br />

Phaeoacremonium currently consists of 46 species with P. parasiticum and P. krajdenii<br />

recognised as the predominant species associated with human infections (Mostert<br />

et al. 2005). Other species have also been isolated from clinical cases i.e. P. alvesii,<br />

P. amstelodamense, P. griseorubrum, P. minimum, P. rubrigenum, P. tardicrescens,<br />

and P. venezuelense. Infections caused by P. parasiticum include subcutaneous<br />

abscesses, thorn-induced arthritis, and disseminated infection (Revankar and Sutton,<br />

2010, Gramaje et al. 2015).<br />

RG-2 organism.<br />

Morphological Description: Cultures usually slow growing, suede-like with radial<br />

furrows, initially whitish-grey becoming olivaceous-grey with age. Hyphae hyaline, later<br />

becoming brown and some becoming rough-walled. Phialides are brown, thick-walled,<br />

slender, acular to cylindrical slightly tapering towards the tip, 15-50 μm long, often<br />

proliferating, with small, funnel-shaped collarettes. Conidia, often in balls, are hyaline,<br />

thin-walled, cylindrical to sausage-shaped, 3-6 x 1-2 μm, later inflating. Maximum<br />

growth temperature 40°C.<br />

Molecular Identification: ITS and β-tubulin sequencing (Mostert et al. 2006, Gramaje<br />

et al. 2015).<br />

Key Features: The identification of the different Phaeoacremonium species can be<br />

done by combining cultural, morphological and sequence data (Mostert et al. 2006,<br />

Gramaje et al. 2015).<br />

References: de Hoog et al. (2000, 2015), Revankar and Sutton (2010), Mostert et al.<br />

(2005, 2006), Gramaje et al. (2015), Badali et al. (2015).<br />

10 µm<br />

a<br />

b<br />

Phaeoacremonium parasiticum (a) colony and (b) a phialide<br />

with a small, funnel-shaped collarette.


Descriptions of Medical Fungi 153<br />

Phaeoacremonium parasiticum (Ajello et al.) W. Gams et al.<br />

20 µm<br />

Phaeoacremonium parasiticum phialides and conidia.<br />

Antifungal Susceptibility: Phaeoacremonium parasiticum (Badali et al. 2015,<br />

Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 56 1 7 13 17 13 5<br />

VORI 56 2 8 21 23 1 2<br />

POSA 53 1 5 17 23 6 1<br />

ITRA 56 1 3 3 1 1 47<br />

Antifungal Susceptibility: Phialophora verrucosa limited data (Australian National<br />

data); MIC µg/mL.<br />

No. 64<br />

AmB 4 1 1 2<br />

VORI 4 1 3<br />

POSA 4 2 2<br />

ITRA 4 1 3 1<br />

P. verrucosa data from 25 isolates (McGinnis and Pasarell 1998a); MIC µg/mL.<br />

AmB Range 0.03-4; Geometric mean = 0.36<br />

VORI Range 0.03-0.5; Geometric mean = 0.12<br />

ITRA Range 0.03-0.5; Geometric mean = 0.07


154<br />

Descriptions of Medical Fungi<br />

The genus Phialophora contains more than 40 species, most are saprophytes<br />

commonly found in soil or on decaying wood. Some human pathogens with phialidic<br />

conidiogenesis previously assigned to Phialophora have been moved to other genera,<br />

namely, Phaeoacremonium and Pleurostomophora. P. verrucosa, P. americana, P.<br />

bubakii, P. europaea and P. reptans remain of medical interest (Revankar and Sutton<br />

2010). Both P. verrucosa and P. americana produce their conidia from phialides with<br />

conspicuous darkened collarettes, however sequencing has demonstrated a close<br />

relatedness, suggesting that these species may be synonymous (de Hoog et al. 1999).<br />

P. verrucosa is primarily an agent of chromoblastomycosis although other reported<br />

infections include endocarditis, keratitis, and osteomyelitis.<br />

RG-2 organism.<br />

Phialophora verrucosa Medlar<br />

Morphological Description: Colonies (SDA) are slow growing, initially dome-shaped,<br />

later becoming flat, suede-like and olivaceous to black in colour. Phialides are flaskshaped<br />

or elliptical with distinctive funnel-shaped, darkly pigmented collarettes. Conidia<br />

are ellipsoidal, smooth-walled, hyaline, mostly 3.0-5.0 x 1.5-3.0 μm, and aggregate in<br />

slimy heads at the apices of the phialide.<br />

Key Features: Characteristic flask-shaped phialides with distinctive funnel-shaped,<br />

darkly pigmented collarettes.<br />

Molecular Identification: ITS sequencing recommended (de Hoog et al. 1999).<br />

References: Ellis (1971), McGinnis (1978a, 1980), Domsch et al. (1980), de Hoog et<br />

al. (1999, 2000, 2015), Revankar and Sutton (2010).<br />

20 µm<br />

20 µm<br />

Phialophora verrucosa phialides and conidia.


Descriptions of Medical Fungi 155<br />

Members of the genus Phoma have a worldwide distribution and are ubiquitous in<br />

nature, with over 200 species having been described from soil, as saprophytes on<br />

various plants, and as pathogens to plants and humans.<br />

RG-1 organism.<br />

Phoma Saccardo<br />

Morphological Description: Colonies are spreading, greyish-brown, powdery or<br />

suede-like and produce large, globose, membranous to leathery, darkly pigmented,<br />

ostiolate pycnidia. Conidia are produced in abundance within the pycnidia on narrow<br />

thread-like phialides, which are hardly differentiated from the inner pycnidial wall cells.<br />

Conidia are globose to cylindrical, one-celled, hyaline, and are usually extruded in<br />

slimy masses from the apical ostiole.<br />

Molecular Identification: ITS, D1/D2, β-tubulin and 18S sequencing has been used<br />

to identify Phoma species (de Gruyter et al. 2009, Aveskamp et al. 2010). Note: Public<br />

sequence databases, particularly GenBank, contain many sequences from incorrectly<br />

identified species, making identifications of coelomycetous fungi very difficult, without<br />

confirmatory morphological studies.<br />

Key Features: Coelomycete, ostiolate pycnidia producing masses of slimy, hyaline,<br />

single-celled conidia.<br />

References: Punithalingam (1979), McGinnis (1980), Sutton (1980), Rippon (1988),<br />

Montel et al. (1991), Samson et al. (1995), de Hoog et al. (2000, 2015).<br />

20 µm<br />

Phoma spp. pycnidia with apical ostiole.


156<br />

Descriptions of Medical Fungi<br />

Pichia kudriavzevii Boidin et al.<br />

Synonymy: Candida krusei (Castellani) Berkhout.<br />

Issatchenkia orientalis Kudryavtesev.<br />

Pichia kudriavzevii is regularly associated with some forms of infant diarrhoea<br />

and occasionally with systemic disease. It has also been reported to colonise the<br />

gastrointestinal, respiratory and urinary tracts of patients with granulocytopenia.<br />

Environmental isolations have been made from beer, milk products, skin, faeces of<br />

animals and birds. RG-2 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Predominantly small, elongated to ovoid blastoconidia, 2-5 x 4-5 µm.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Abundant long, wavy, branched pseudohyphae with elongated<br />

to ovoid blastoconidia, budding off in verticillate branches.<br />

Molecular Identification: ITS sequencing recommended.<br />

MALDI-T<strong>OF</strong> MS: Able to accurately identify this species.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose - L-Arabinose - D-Glucitol -<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose + Maltose - D-Ribose - D-Gluconate -<br />

Galactose - Cellobiose - L-Rhamnose - DL-Lactate +<br />

Sucrose - Trehalose - D-Glucosamine + myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine + 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate -<br />

Trehalose - Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol - Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide v<br />

Galactose - D-Xylose - D-Mannitol - Growth at 40 O C +<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern. Colonies<br />

pale pink on Candida CHROMagar.<br />

Antifungal Susceptibility: P. kudriavzevii (Australian National data); MIC µg/mL.<br />

CLSI clinical breakpoints are marked where available (Pfaller and Diekema 2012).<br />

Note: P. kudriavzevii is intrinsically resistant to fluconazole (No CLSI breakpoints).<br />

No. 64<br />

AmB 152 1 2 25 41 61 18 4<br />

FLU 152 1 9 45 77<br />

VORI 135 5 16 62 40 7 3 2<br />

POSA 115 1 8 18 49 30 6 1 2<br />

ITRA 152 1 3 16 73 47 10 2<br />

ANID 86 5 35 26 19 1<br />

MICA 86 1 0 32 34 19<br />

CAS 115 1 15 55 29 13 1 1<br />

5FC 152 3 2 5 15 92 33 2


Descriptions of Medical Fungi 157<br />

Synonymy: Candida norvegensis Dietrichson ex van Uden & Buckley.<br />

Pichia norvegensis is a very rare clinical isolate that has been reported as a causative<br />

agent of peritonitis and disseminated candidiasis in a patient on CAPD.<br />

RG-1 organism.<br />

Pichia norvegensis Leask & Yarrow<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Ovoid to ellipsoid, budding blastoconidia, 2.0-4 x 3-10 µm.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Spherical to ovoid budding yeast cells only. Abundant<br />

pseudohyphae produced.<br />

Molecular Identification: ITS sequencing recommended.<br />

MALDI-T<strong>OF</strong> MS: Able to accurately identify this species.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose - L-Arabinose - D-Glucitol -<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose s Maltose - D-Ribose - D-Gluconate -<br />

Galactose - Cellobiose + L-Rhamnose - DL-Lactate w<br />

Sucrose - Trehalose - D-Glucosamine + myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine - 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate -<br />

Trehalose - Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol - Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose - D-Xylose - D-Mannitol - Growth at 37 O C +<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern.<br />

Antifungal Susceptibility: P. norvegensis very limited data (Guitard et al. 2013,<br />

Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 8 1 6 1<br />

FLU 8 1 2 1 4<br />

VORI 5 2 1 2<br />

POSA 5 2 3<br />

ITRA 7 1 4 1 1<br />

ANID 1 1<br />

MICA 1 1<br />

CAS 4 2 2<br />

5FC 5 1 1 1 1 1


158<br />

Descriptions of Medical Fungi<br />

The genus Pithomyces contains about 50 species commonly isolated from a wide range<br />

of plant material, also from air, soil, hay, sawn timber and ceiling plaster. Pithomyces<br />

chartarum has long been reported as causing facial eczema in sheep. However, recent<br />

molecular studies have identified at least two additional species P. sacchari and P.<br />

maydicus (de Cunha et al. 2014). Most human isolates are recovered from skin, nail,<br />

respiratory and sinus specimens.<br />

RG-1 organism.<br />

Pithomyces chartarum (Berk. & M.A. Curtis) M.B. Ellis<br />

Morphological Description: Colonies are fast growing, suede-like to downy<br />

and black. Conidiophores are pale olive, smooth or verrucose, 2.5-10 x 2-3.5 µm.<br />

Conidiogenous cells integrated, intercalary or terminal, indeterminate, with one to two<br />

loci of similar width in the conidiogenous cells. Conidia are muriform, medium to dark<br />

brown, echinulate to verrucose, three-(some up to five)-euseptate, slightly constricted<br />

at the septa, with one or both median cells divided by longitudinal septa, thick-walled,<br />

broadly ellipsoidal, apex obtuse, base truncate and characteristically with part of the<br />

conidiogenous cell remaining attached as a small pedicel, 18-29 x 10-17 µm.<br />

Key Features: Dematiaceous hyphomycete with multicelled conidia produced on<br />

small peg-like branches of the vegetative hyphae.<br />

Molecular Identification: ITS and D1/D2 sequencing recommended (de Cunha et al.<br />

2014).<br />

References: Ellis (1971, 1976), Domsch et al. (1980), Rippon (1988), de Hoog et al.<br />

(2000, 2015), de Cunha et al. (2014).<br />

15 µm<br />

Pithomyces chartarum conidiophores and conidia.


Descriptions of Medical Fungi 159<br />

Pleurostomophora richardsiae (Nannf.) Mostert et al.<br />

Synonymy: Phialophora richardsiae (Nannf.) Conant.<br />

Vijaykrishna et al. (2004) separated Pleurostomophora richardsiae from Phialophora<br />

based on molecular data. P. richardsiae is a soft rot fungus of wood and is an<br />

uncommon cause of human infection, usually through traumatic implantation causing<br />

subcutaneous phaeohyphomycosis. RG-2 organism.<br />

Morphological Description: Colonies grow rapidly, and are powdery to woolly or<br />

tufted, greyish-brown with a grey-brown to olivaceous-black reverse. Two conidial<br />

types are produced: (1) hyaline conidia which are allantoid or cylindrical, 3-6 x 1.5-2.5<br />

μm in size, formed on inconspicuous, peg-like phialides on thin-walled hyphae; and<br />

(2) brown, thick-walled conidia which are spherical to subspherical, 2.5-3.5 x 2-3 μm,<br />

formed on dark brown, slender, tapering phialides with flaring collarettes.<br />

Key Features: P. richardsiae is characterised microscopically by phialides with<br />

prominent flaring collarettes bearing globose, brown conidia while phialides with<br />

indistinct collarettes bear pale allantoid to cylindrical conidia.<br />

Molecular Identification: ITS sequencing is recommended (Vijaykrishna et al. 2004).<br />

References: Ellis (1971), McGinnis (1978a, 1980), Domsch et al. (1980), de Hoog et<br />

al. (2000, 2015), Vijaykrishna et al. (2004), Revankar and Sutton (2010).<br />

2<br />

10 µm<br />

1<br />

10 µm<br />

Pleurostomophora richardsiae phialides producing two types of conidia.<br />

(1) hyaline conidia, and (2) brown, thick-walled conidia (arrows).<br />

Antifungal Susceptibility: Pleurostomophora richardsiae limited data (Australian<br />

National data); MIC µg/mL.<br />

No. 64<br />

AmB 7 1 1 4 1<br />

VORI 7 3 4<br />

POSA 5 1 2 2<br />

ITRA 7 2 3 2 1<br />

P. richardsiae data from 11 isolates (McGinnis and Pasarell 1998a); MIC µg/mL.<br />

AmB Range 0.125-1; Geometric mean = 0.73<br />

VORI Range 0.25-2; Geometric mean = 0.64<br />

ITRA Range 0.03-2; Geometric mean = 0.44


160<br />

Descriptions of Medical Fungi<br />

Prototheca species are achlorophyllous algae with phylogenetic affinities to the genus<br />

Chlorella. To date only P. wickerhamii and P. zopfii have been involved in human or<br />

animal infections (Lass-Florl and Mayr 2007).<br />

RG-1 organisms.<br />

Prototheca Kruger<br />

Morphological Description: Colonies are smooth, moist, white to cream and yeastlike.<br />

Cultures are sensitive to cycloheximide (actidione) and optimal growth occurs<br />

at 25-30 O C. Mycelium and conidia are absent. Vegetative cells are globose to ovoid,<br />

hyaline, varying in size from approximately 3-30 µm, and have a relatively thick and<br />

highly refractile wall. No budding cells are present; reproduction is by the development<br />

of large sporangia (theca) which contain from 2-20 or more small sporangiospores<br />

(endospores or autospores) which are asexually produced by nuclear division and<br />

cleavage of the cytoplasm.<br />

Molecular Identification: ITS and D1/D2 sequencing is recommended (Wang et al.<br />

2014).<br />

Key Features: Achlorophyllous algae reproducing by sporangia (theca) and<br />

sporangiospores (autospores). Prototheca species which can be differentiated by<br />

assimilation tests and morphological criteria as outlined below. The API 20C yeast<br />

identification strip may be used for species identification.<br />

References: Kaplan (1977), McGinnis (1980), Rippon (1988), Pore (1985), Ueno et al.<br />

(2005), Lass-Florl and Mayr (2007), Wang et al. (2014).<br />

Colony morphology<br />

P. wickerhamii P. zopfii P. stagnora<br />

Hemispheric, with<br />

smooth margin<br />

Flat, rough,<br />

corrugated margin<br />

Flat, with smooth<br />

margin<br />

Cell diameter μm 3-10 7-30 7-14<br />

Growth at 37 O C + + -<br />

Glucose + + +<br />

Trehalose + - -<br />

L-propanol - + +/-<br />

Acetate (pH 5) - + +/-<br />

Galactose + - +<br />

Capsule - - +<br />

5 µm<br />

Prototheca wickerhamii thecae and autospores.


Descriptions of Medical Fungi 161<br />

Prototheca Kruger<br />

Antifungal Susceptibility: P. wickerhamii (Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 9 3 3 2 1<br />

VORI 7 2 4 1<br />

POSA 7 1 2 4<br />

ITRA 9 2 5 2<br />

Purpureocillium lilacinum (Thom) Luangsa-ard et al.<br />

Synonymy: Paecilomyces lilacinus (Thom) Samson.<br />

Purpureocillium lilacinum is commonly isolated from soil, decaying vegetation, insects,<br />

nematodes and as a laboratory contaminant. It is also a causative agent of infection in<br />

human and other vertebrates (Luangsa-ard et al. 2011).<br />

RG-1 organism.<br />

Morphological Description: Colonies are fast growing, suede-like to floccose,<br />

vinaceous to violet-coloured. Conidiophores are erect 400-600 µm in length, bearing<br />

branches with densely clustered phialides. Conidiophore stipes are 3-4 µm wide, yellow<br />

to purple and rough-walled. Phialides are swollen at their bases, gradually tapering into<br />

a slender neck. Conidia are ellipsoidal to fusiform, smooth-walled to slightly roughened,<br />

hyaline to purple in mass, 2.5-3.0 x 2-2.2 µm, and are produced in divergent chains.<br />

Chlamydospores are absent. Growth at 38 O C.<br />

Molecular Identification: ITS sequencing is recommended (Atkins et al. 2005,<br />

Luangsa-ard et al. 2011).<br />

Key Features: Colony pigmentation, phialides with swollen bases, pigmented and<br />

rough-walled conidiophore stipes, absence of chlamydospores and growth at 37 O C.<br />

Note: Paecilomyces marquandii differs by having a yellow reverse pigment, smooth<br />

conidiophore stipes, presence of chlamydospores, and no growth at 37 O C.<br />

References: Samson (1974), Domsch et al. (1980), McGinnis (1980), Onions et al.<br />

(1981), Rippon (1988), de Hoog et al. (2000, 2015), Perdomo et al. (2013).<br />

Antifungal Susceptibility: P. lilacinum (Australian National data); MIC µg/mL.<br />

No. 16<br />

AmB 52 1 1 6 6 24 14<br />

VORI 50 8 30 9 1 2<br />

POSA 37 2 7 4 23 1<br />

ITRA 53 5 13 18 2 7 2 6


162<br />

Descriptions of Medical Fungi<br />

Purpureocillium lilacinum (Thom) Luangsa-ard et al.<br />

a<br />

b<br />

10 µm b<br />

10 µm<br />

Purpureocillium lilacinum (a) culture and (b) conidiophores,<br />

phialides and conidia. Note: Rough-walled conidiophore (arrow).


Descriptions of Medical Fungi 163<br />

Quambalaria cyanescens<br />

Synonymy: Sporothrix cyanescens de Hoog & de Vries.<br />

Cerinosterus cyanescens (de Hoog & de Vries) R.T. Moore<br />

The genus Quambalaria contains five species, including Q. cyanescens, Q. pitereka,<br />

Q. eucalypti, Q. coyrecup and Q. simpsonii. Q. cyanescens is a hyaline basidiomycete<br />

isolated from a broad range of ecological niches, including air, soil, and insect larvae as<br />

well as in association with diverse plant sources, including Corymbia and Eucalyptus<br />

species from Australia. Q. cyanescens appears to be an emerging opportunistic<br />

pathogen in immunocompromised or debilitated individuals. It has been isolated from<br />

human skin and subcutaneous infections, blood, nosocomial infections in patients with<br />

pneumonia, peritonitis and invasive pulmonary infection (Jackson et al. 1990, Tambini<br />

et al. 1996, Schmidt et al. 2000, Kuan et al. 2015).<br />

RG-1 organism.<br />

Morphological Description: Colonies are restricted, farinose or velvety, often compact<br />

and somewhat cerebriform, snow-white, and later often exuding a pH-dependent, deep<br />

blue/violet pigment into the agar. Conidiogenous cells are undifferentiated, cylindrical,<br />

of variable size (1.5-3.0 µm wide), apically with a cluster of small denticles, the cluster<br />

often repeatedly proliferating and forming similar clusters. Conidia are hyaline, smoothwalled<br />

or finely verrucose, obovoidal, 3-4 µm long, somewhat larger (3.5-6.5 µm long)<br />

when bearing secondary conidia.<br />

Molecular Identification: ITS and D1/D2 sequencing recommended.<br />

Key Features: Q. cyanescens is similar to Sporothrix, but the conidial scars are<br />

very small and cultures are thin and fragile. In fresh cultures the diffusible pigment<br />

is characteristic. Sporothrix schenckii forms tough colonies, which finally become<br />

blackish-brown.<br />

References: de Hoog and de Vries (1973), de Beer et al. (2006), Simpson (2000), de<br />

Hoog et al. (2015).<br />

b<br />

b<br />

a<br />

b<br />

10 µm<br />

b<br />

c<br />

Quambalaria cyanescens (a) culture, (b) conidiogenous cells with small<br />

denticles and conidia, and (c) mature conidium bearing secondary conidia.


164<br />

Descriptions of Medical Fungi<br />

Rhinocladiella Nannfeldt<br />

Rhinocladiella contains six to eight species, with five species of medical interest;<br />

R. aquaspersa, R. atrovirens, R. basitona, R. mackenziei (formerly Ramichloridium<br />

mackenziei) and R. similis. R. mackenziei is a frequently fatal neurotropic organism<br />

and appears to be restricted to individuals residing in, or immigrating from, Middle<br />

Eastern countries (Revankar and Sutton 2010).<br />

RG-1 organism.<br />

Rhinocladiella atrovirens Nannfeldt<br />

Morphological Description: Colonies are restricted, velvety or lanose, olivaceous,<br />

often slightly mucoid at the centre; reverse dark olivaceous green to blackish.<br />

Conidiophores are short, brown, thick-walled. Conidiogenous cells are cylindrical,<br />

intercalary or free, 9-19 x 1.6-2.2 µm; denticulate rachis up to 15 µm long, with<br />

crowded, flat or butt-shaped, unpigmented conidial denticles. Conidia are hyaline, thinand<br />

smooth-walled, short-cylindrical, with truncate basal scars, 3.7-5.5 x 1.2-1.8 µm.<br />

Budding cells, if present, are hyaline, thin-walled, broadly ellipsoidal, 3.0-4.3 x 1.7-2.5<br />

µm. Germinating cells are inflated, spherical to subspherical, 4.5-6.0 µm. An annellidic<br />

Exophiala synanamorph may be present.<br />

Molecular Identification: ITS sequencing is recommended for accurate species<br />

identification (Taj-Aldeen et al. 2010).<br />

References: de Hoog (1977, 1983), Schell et al. (1983), de Hoog et al. (2000, 2015).<br />

a<br />

c<br />

10 µm 10 µm b b 10 µm<br />

Rhinocladiella atrovirens (a) culture, (b) conidiophores showing a terminal<br />

denticulate rachis with conidia, and (c) budding yeast cells.


Descriptions of Medical Fungi 165<br />

Rhinocladiella Nannfeldt<br />

Antifungal Susceptibility: Rhinocladiella atrovirens very limited data (McGinnis<br />

and Pasarell 1998a); MIC µg/mL.<br />

Antifungal Range Antifungal Range Antifungal Range<br />

AmB 0.03-0.25 ITRA 0.03-0.06 VORI 0.03-0.5<br />

Rhinocladiella mackenziei (Campbell & Al-Hedaithy) Arzanlou & Crous<br />

WARNING: RG-3 organism. Cultures of R. mackenziei represent a potential biohazard<br />

to laboratory personnel and must be handled with extreme caution in a Class II Biological<br />

Safety Cabinet (BSCII).<br />

R. mackenziei is an extremely rare, neurotropic organism that causes fatal brain<br />

lesions, mostly in patients who are immunocompromised or suffered from underlying<br />

metabolic diseases. The species is typically restricted to the Middle East, in an arid<br />

zone between Israel and Pakistan (Kanj et al. 2001, Khan et al. 2002, Taj-Aldeen et al.<br />

2010), with a single autochtonous case in India (Badali et al. 2010). Cases in the USA<br />

and Europe invariably were found in patients originating from the Middle East (Sutton<br />

et al. 1998, Revankar and Sutton 2010, de Hoog et al. 2015).<br />

Morphological Description: Colonies growing moderately rapidly, velvety, olivaceousbrown.<br />

Conidiophores arising at right angles from creeping hyphae, stout, thick-walled,<br />

brown, 3.0-4.5 µm wide, 10-25 µm long, apically with short-cylindrical denticles. Conidia<br />

brown, ellipsoidal, 8.5-12.0 × 4-5 µm, with a prominent, wide basal scar.<br />

Molecular Identification: ITS and D1/D2 sequencing is recommended for accurate<br />

species identification (Taj-Aldeen et al. 2010).<br />

References: de Hoog (1977, 1983), Schell et al. (1983), de Hoog et al. (2000, 2015),<br />

Taj-Aldeen et al. (2010), Revankar and Sutton (2010).<br />

Antifungal Susceptibility: R. mackenziei very limited data compiled from six case<br />

reports (Taj-Aldeen et al. 2010); MIC µg/mL.<br />

No. 32<br />

AmB 7 1 3 2 1<br />

VORI 3 1 1 1<br />

POSA 5 2 2 1<br />

ITRA 7 5 1 1


166<br />

Descriptions of Medical Fungi<br />

The genus Rhizomucor is distinguished from Mucor by the presence of stolons and<br />

poorly developed rhizoids at the base of the sporangiophores and by the thermophilic<br />

nature of its two species: R. miehei and R. pusillus. Both of these species are potential<br />

human and animal pathogens and were originally classified in the genus Mucor.<br />

Rhizomucor pusillus is cosmopolitan and both R. miehei and R. pusillus have been<br />

reported as pathogens to humans and animals, the latter to a greater extent.<br />

References: Cooney and Emerson (1964), Schipper (1978), Domsch et al. (1980),<br />

McGinnis (1980), Ellis and Keane (1981), Scholer et al. (1983), de Hoog et al. (2000,<br />

2015), Schipper and Stalpers (2003) and Ellis (2005b).<br />

Identification of most Mucorales is based primarily on the morphology of the sporangia;<br />

i.e. arrangement and number of sporangiospores, shape, colour, presence or absence<br />

of columellae and apophyses, as well as the arrangement of the sporangiophores and<br />

the presence or absence of rhizoids. Growth temperature tests can also be especially<br />

helpful in identifying and differentiating members of the genera Rhizomucor, Rhizopus<br />

and Lichtheimia.<br />

Rhizomucor miehei (Cooney and Emerson) Schipper<br />

Synonymy: Mucor miehei Lindt.<br />

This species has been reported as a rare cause of bovine mastitis (Scholer et al. 1983)<br />

and is similar in many respects to R. pusillus.<br />

RG-1 organism.<br />

Rhizomucor Lucet & Costantin<br />

Morphological Description: All strains are homothallic forming numerous zygospores,<br />

which are reddish-brown to blackish-brown, globose to slightly compressed, up to 50<br />

µm in diameter, with stellate warts and equal suspensor cells. Colony colour is a dirty<br />

grey rather than brown, and sporangia have spiny walls, are up to 50-60 µm in diameter,<br />

with columellae rarely larger than 30 µm in diameter. Growth is stimulated by thiamine,<br />

with no assimilation of sucrose and maximum growth temperature is 54-58 O C.<br />

Key Features: Growth at 45 O C, the formation of numerous zygospores, a dirty grey<br />

culture colour and a partial growth requirement for thiamine.<br />

Synonymy: Mucor pusillus Lindt.<br />

Rhizomucor pusillus (Lindt) Schipper<br />

This species is a rare human pathogen. It has been reported from cases of pulmonary,<br />

disseminated and cutaneous types of infection. It is more often associated with animal<br />

disease, especially bovine abortion. Rhizomucor pusillus has a worldwide distribution<br />

and is commonly associated with compost heaps.


Descriptions of Medical Fungi 167<br />

RG-2 organism.<br />

Rhizomucor pusillus (Lindt) Schipper<br />

Morphological Description: Cultures are characterised by compact, low growing<br />

(2-3 mm high), grey to greyish brown-coloured mycelium and by the development of<br />

typical sympodially branched, hyaline to yellow-brown sporangiophores (8-15 μm in<br />

diameter), always with a septum below the sporangium. Sporangia are globose (40-60<br />

µm in diameter), each possessing an oval or pear-shaped columella (20-30 µm), often<br />

with a collarette. Sporangiospores are hyaline, smooth-walled, globose to subglobose,<br />

occasionally oval (3-5 µm), and are often mixed with crystalline remnants of the<br />

sporangial wall. Chlamydospores are absent. Zygospores are rough-walled, reddish<br />

brown to black, 45-65 µm in diameter and may be produced throughout the aerial<br />

hyphae in matings between compatible isolates. Temperature growth range: minimum<br />

20-27 O C; optimum 35-55 O C; maximum 55 O C. There is positive assimilation of sucrose<br />

and no thiamine dependence.<br />

Key Features: Mucorales, growth at 45 O C (thermophilic), poorly developed stolons<br />

and rhizoids, branching sporangiophores with a septum below the sporangium, darkcoloured<br />

sporangia without apophyses and smooth-walled globose to subglobose<br />

sporangiospores.<br />

20 µm<br />

b<br />

a<br />

15 µm<br />

Rhizomucor pusillus (a) sporangiophores, collumellae and (b) primitive rhizoids.<br />

Antifungal Susceptibility: R. pusillus (Espinel-Ingroff et al. 2015a, Australian<br />

National data); MIC µg/mL.<br />

No. 32<br />

AmB 36 3 8 9 12 1 2 1<br />

POSA 36 1 5 11 11 7 1<br />

ITRA 18 3 6 4 3 1 1


168<br />

Descriptions of Medical Fungi<br />

Rhizopus Ehrenberg ex Corda.<br />

A molecular phylogenetic study of the genus Rhizopus by Abe et al. (2010) recognised<br />

eight species; R. caespitosus, R. delemar, R. homothallicus, R. microsporus, R. arrhizus<br />

(R. oryzae), R. reflexus, R. schipperae and R. stolonifer. Based on these results and<br />

confirmed by Dolatabadi et al. (2014) the previous varieties of Rhizopus microsporus<br />

(R. microsporus var. oligosporus and R. microsporus var. rhizopodiformis) have been<br />

reduced to synonyms. In addition, R. azygosporus has been reduced to a synonym of<br />

R. microsporus. Finally, the controversy surrounding which species name to use for<br />

R. oryzae - R. arrhizus has been resolved in favour of the latter (Ellis 1985, de Hoog<br />

et al. 2015). Thus the important medical pathogens have now been reduced to just<br />

R. arrhizus and R. microsporus. These two species are the most common causative<br />

agents of mucormycosis, accounting for some 60% of the reported cases.<br />

Morphological Description: The genus Rhizopus is characterised by the presence of<br />

stolons and pigmented rhizoids, the formation of sporangiophores, singly or in groups<br />

from nodes directly above the rhizoids, and apophysate, columellate, multispored,<br />

generally globose sporangia. After spore release the apophyses and columella often<br />

collapse to form an umbrella-like structure. Sporangiospores are globose to ovoid,<br />

one-celled, hyaline to brown and striate in many species. Colonies are fast growing<br />

and cover an agar surface with a dense cottony growth that is at first white becoming<br />

grey or yellowish brown with sporulation.<br />

Molecular Identification: ITS sequencing is recommended but sequences must be<br />

compared to those of quality controlled reference strains with updated species names.<br />

(Alvarez et al. 2009 and Abe et al. 2010).<br />

References: Domsch et al. (1980), McGinnis (1980), Onions et al. (1981), Scholer<br />

et al. (1983), Schipper (1984), Schipper and Stalpers (1984, 2003), Yuan and Jong<br />

(1984), Ellis (1985, 1986), Rippon (1988), Kwon-Chung and Bennett (1992), Samson<br />

et al. (1995), Schipper et al. (1996), de Hoog et al. (2000, 2015), Ellis (2005b), Alvarez<br />

et al. (2009), Abe et al. (2010), Dolatabadi et al. (2014).<br />

Antifungal Susceptibility: R. arrhizus (Espinel-Ingroff et al. 2015a, Australian<br />

National data); MIC µg/mL.<br />

No. 32<br />

AmB 280 1 7 21 30 67 112 39 3<br />

POSA 370 1 5 14 84 161 65 29 4 4 2<br />

ITRA 238 5 9 42 93 41 34 2 4 9<br />

Antifungal Susceptibility: R. microsporus (Espinel-Ingroff et al. 2015a, Australian<br />

National data); MIC µg/mL.<br />

No. 32<br />

AmB 189 2 23 22 95 72 22 5<br />

POSA 180 3 12 34 60 21 4 1 2<br />

ITRA 117 1 1 8 27 39 24 3 1 13


Descriptions of Medical Fungi 169<br />

Synonymy: Rhizopus oryzae Went & Prinsen Geerligs.<br />

RG-2 organism.<br />

Rhizopus arrhizus Fischer<br />

Morphological Description: Colonies are very fast growing, about 5-8 mm high,<br />

with some tendency to collapse, white cottony at first becoming brownish grey to<br />

blackish-grey depending on the amount of sporulation. Sporangiophores up to 1500<br />

µm in length and 18 µm in width, smooth-walled, non-septate, simple or branched,<br />

arising from stolons opposite rhizoids usually in groups of three or more. Sporangia are<br />

globose, often with a flattened base, greyish black, powdery in appearance, up to 175<br />

µm in diameter and many spored. Columellae and apophysis together are globose,<br />

subglobose or oval, up to 130 µm in height collapsing to an umbrella-like form after<br />

spore release. Sporangiospores are angular, subglobose to ellipsoidal, with striations<br />

on the surface, and up to 8 µm in length. No growth at 45 O C; good growth at 40 O C.<br />

a<br />

b<br />

20 µm<br />

c<br />

100 µm<br />

Rhizopus arrhizus (a) culture, (b) columellae and (c) sporangia<br />

showing sporangiospores, sporangiophores and rhizoids.


170<br />

Descriptions of Medical Fungi<br />

Rhizopus microsporus v. Tiegh<br />

Synonymy: Rhizopus azygosporus Yuan & Jong.<br />

Rhizopus microsporus var. microsporus Tiegh.<br />

Rhizopus microsporus var. oligosporus (Saito) Schipper & Stalpers.<br />

Rhizopus microsporus var. rhizopodiformis (Cohn) Schipper & Stalpes.<br />

Rhizopus microsporus var. chinensis (Saito) Schipper & Stalpers.<br />

RG-2 organism.<br />

Morphological Description: Colonies are dark greyish-brown, up to 10 mm high<br />

producing simple rhizoids. Sporangiophores are brownish, up to 400 µm high and 10<br />

µm wide, and may be produced in groups of one to four, usually in pairs. Sporangia<br />

are greyish-black, spherical, up to 100 µm in diameter. Columellae are subglobose to<br />

globose to conical comprising 80% of the sporangium. Sporangiospores are angular<br />

to broadly ellipsoidal or subglobose, up to 5-9 µm in length and are distinctly striate.<br />

Chlamydospores may be present. Zygospores are dark red–brown, spherical, up<br />

to 100 µm in diameter, with stellate projections and unequal suspensor cells. Some<br />

strains may be homothallic and produce azygospores. There is good growth at 45 O C,<br />

with a maximum of 50-52 O C.<br />

30 µm<br />

Rhizopus microsporus sporangia showing sporangiospores,<br />

columellae, sporangiophores and rhizoids.


Descriptions of Medical Fungi 171<br />

Rhodotorula Harrison<br />

Rhodotorula species are common environmental basidiomycetous yeasts, which can<br />

be found in soil, ocean and lake water, fruit juice and milk, and on shower curtains<br />

and toothbrushes. Today, the genus contains 46 species of which three have been<br />

described as rare human pathogens: R. mucilaginosa (formerly known as R. rubra), R.<br />

glutinis and R. minuta (Arendrup et al. 2014).<br />

Rhodotorula mucilaginosa is a common airborne contaminant of skin, lungs, urine and<br />

faeces. R. mucilaginosa is a known cause of fungal peritonitis in patients on continuous<br />

ambulatory peritoneal dialysis (CAPD). This is usually due to saprophytic colonisation<br />

of catheters or dialysis machinery and removal of the source of contamination usually<br />

leads to clearing of the symptoms. This species accounts for the majority of the<br />

infections (74–79%) followed by R. glutinis (7.7%) (Tuon and Costa 2008, Arendrup et<br />

al. 2014).<br />

Molecular Identification: In many clinical cases species identification requires ITS<br />

and/or D1/D2 sequencing (Duboc de Almeida et al. 2008, Tuon and Costa 2008,<br />

Arendrup et al. 2014).<br />

MALDI-T<strong>OF</strong> MS: reliably identifies clinically relevant Rhodotorula spp.<br />

References: McGinnis (1980), Barnett et al. (1983), Kreger-Van Rij (1984), Rippon<br />

(1988), Kurtzman and Fell (1988), de Hoog et al. (2000, 2015), Spiliopoulou et al.<br />

(2012), Duboc de Almeida et al. (2008), Tuon and Costa (2008), Arendrup et al. (2014).<br />

Rhodotorula mucilaginosa culture.


172<br />

Descriptions of Medical Fungi<br />

RG-1 organism.<br />

Rhodotorula glutinis (Fresenius) Harrison<br />

Morphological Description: Colonies are coral red to salmon-coloured or slightly<br />

orange, smooth to wrinkled, highly glossy to semi-glossy. Mucoid to pasty to slightly<br />

tough, yeast-like colonies. Ovoidal to globose or more elongate budding yeast-like<br />

cells or blastoconidia, 2.3-5.0 x 4.0-10 µm.<br />

India Ink Preparation: Small capsules present.<br />

Molecular Identification: Requires sequencing of the ITS and/or D1/D2 regions.<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern. Common<br />

saprophyte however cases of fungaemia have been reported.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No data<br />

Germ Tube - L-Sorbose v L-Arabinose v D-Glucitol v<br />

Fermentation Sucrose + D-Arabinose v α-M-D-glucoside<br />

Glucose - Maltose + D-Ribose v D-Gluconate +<br />

Galactose - Cellobiose v L-Rhamnose v DL-Lactate v<br />

Sucrose - Trehalose + D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine - 2-K-D-gluconate nd<br />

Lactose - Melibiose - Glycerol v D-Glucuronate nd<br />

Trehalose - Raffinose v Erythritol - Nitrate -<br />

Assimilation Melezitose + Ribitol v Urease +<br />

Glucose + Soluble Starch - Galactitol v 0.1% Cycloheximide v<br />

Galactose v D-Xylose v D-Mannitol v Growth at 37 O C v<br />

Antifungal Susceptibility: R. glutinis (Diekema et al. 2005, Australian National<br />

data); MIC µg/mL. Note: Rhodotorula species are intrinsically resistant to azoles<br />

and echinocandins (Arendrup et al. 2014).<br />

No. 64<br />

AmB 38 1 2 4 3 28<br />

FLU 38 38<br />

VORI 36 1 1 12 12 9 1<br />

POSA 34 2 9 18 4 1<br />

ITRA 38 1 1 3 9 11 2 3 8<br />

5FC 38 17 12 8 1


Descriptions of Medical Fungi 173<br />

Rhodotorula mucilaginosa (Jorgensen) Harrison<br />

Synonymy: Rhodotorula rubra (Demme) Lodder.<br />

RG-1 organism.<br />

Morphological Description: Colonies are coral pink, usually smooth, sometimes<br />

reticulate, rugose or corrugated, moist to mucoid, yeast-like colonies. Spherical to<br />

elongate budding yeast-like cells or blastoconidia, 2.5-6.5 x 6.5-14.0 µm.<br />

India Ink Preparation: Small capsules present.<br />

Molecular Identification: Requires sequencing of the ITS and/or D1/D2 regions.<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern. Common<br />

saprophyte however cases of peritonitis and fungaemia have been reported.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose v L-Arabinose v D-Glucitol v<br />

Fermentation Sucrose + D-Arabinose v α-M-D-glucoside<br />

Glucose - Maltose v D-Ribose v D-Gluconate +<br />

Galactose - Cellobiose v L-Rhamnose v DL-Lactate v<br />

Sucrose - Trehalose + D-Glucosamine v myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine - 2-K-D-gluconate nd<br />

Lactose - Melibiose - Glycerol v D-Glucuronate nd<br />

Trehalose - Raffinose + Erythritol v Nitrate -<br />

Assimilation Melezitose v Ribitol v Urease +<br />

Glucose + Soluble Starch - Galactitol v 0.1% Cycloheximide -<br />

Galactose v D-Xylose + D-Mannitol v Growth at 40 O C +<br />

Antifungal Susceptibility: R. mucilaginosa (Diekema et al. 2005, Australian<br />

National data); MIC µg/mL. Note: Rhodotorula species are intrinsically resistant to<br />

azoles and echinocandins (Arendrup et al. 2014).<br />

No. 64<br />

AmB 39 4 4 5 25 1<br />

FLU 39 39<br />

VORI 37 2 3 6 11 10 4 1<br />

POSA 37 1 3 8 17 2 2 4<br />

ITRA 39 3 8 15 4 1 2 6<br />

5FC 39 1 14 12 11 1


174<br />

Descriptions of Medical Fungi<br />

Saccharomyces cerevisiae Meyen ex Hansen<br />

Synonomy: Candida robusta Diddens & Lodder<br />

Saccharomyces cerevisiae, commonly known as Baker’s yeast, may be found as a<br />

harmless and transient digestive commensal and coloniser of mucosal surfaces of<br />

normal individuals. The anamorphic state of S. cerevisiae is sometimes referred to as<br />

Candida robusta. This species is phylogenetically closely related to Candida glabrata<br />

and shares many clinical and microbiological characteristics to this species (Arendrup<br />

et al. 2014). S. cerevisiae may be involved in mucosal infections like vaginitis,<br />

and in bloodstream infections, particularly in fluconazole-exposed patients. Note:<br />

Saccharomyces boulardii, a genetically similar subtype that is used as a probiotic<br />

for prevention and treatment of various sorts of diarrhoea and recurrent Clostridium<br />

difficile-associated diarrhoea should be avoided in immunocompromised hosts<br />

(Enache-Angoulvant and Hennequin 2005, Arendrup et al. 2014). RG-1 organism.<br />

Morphological Description: Colonies are white to cream, smooth, glabrous and<br />

yeast-like. Large globose to ellipsoidal budding yeast-like cells or blastoconidia, 3.0-<br />

10.0 x 4.5-21.0 µm. No capsules present on India Ink preparation.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No data<br />

Germ Tube - L-Sorbose - L-Arabinose - D-Glucitol -<br />

Fermentation Sucrose + D-Arabinose - M-D-glucoside<br />

Glucose + Maltose + D-Ribose - D-Gluconate -<br />

Galactose v Cellobiose - L-Rhamnose - DL-Lactate v<br />

Sucrose + Trehalose + D-Glucosamine - myo-Inositol -<br />

Maltose v Lactose - N-A-D-glucosamine - 2-K-D-gluconate nd<br />

Lactose - Melibiose v Glycerol - D-Glucuronate nd<br />

Trehalose - Raffinose + Erythritol - Nitrate -<br />

Assimilation Melezitose v Ribitol - Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose v D-Xylose - D-Mannitol - Growth at 37 O C v<br />

Molecular Identification: ITS and/or D1/D2 sequencing is recommended (McCullough<br />

et al. 1988).<br />

References: McGinnis (1980), Barnett et al. (1983), Kreger-Van Rij (1984), Rippon<br />

(1988), Kurtzman and Fell (1988), de Hoog et al. (2000, 2015).<br />

Antifungal Susceptibility: S. cerevisiae (Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 41 1 1 8 14 11 6<br />

FLU 42 3 2 4 6 5 12 10<br />

VORI 38 5 8 7 8 8 2<br />

POSA 35 1 2 4 10 11 5 2<br />

ITRA 42 2 6 9 10 10 1 1 3<br />

ANID 32 3 3 17 6 3<br />

MICA 32 1 21 9 1<br />

CAS 36 2 3 11 14 5 1<br />

5FC 42 7 32 1 1 1


Descriptions of Medical Fungi 175<br />

Saksenaea vasiformis Complex<br />

The genus Saksenaea is characterised by the formation of flask-shaped sporangia with<br />

columellae and simple, darkly pigmented rhizoids. It is an emerging human pathogen<br />

(Holland, 1997) that is most often associated with cutaneous or subcutaneous lesions<br />

after trauma. Until recently, Saksenaea vasiformis was the only known species with a<br />

worldwide distribution in association with soil. However S. vasiformis has recently been<br />

split into three species; S. vasiformis, S erythrospora and S. oblongispora (Alvarez et<br />

al. 2010b). All three species have been isolated from clinical samples, but as yet no<br />

proven case reports have been published on the new species (de Hoog et al. 2015).<br />

Saksenaea vasiformis Saksena<br />

Morphological Description: Colonies are fast growing, downy, white with no reverse<br />

pigment, and made up of broad, non-septate hyphae typical of a mucormycetous<br />

fungus. Sporangia are typically flask-shaped with a distinct spherical venter and longneck,<br />

arising singly or in pairs from dichotomously branched, darkly pigmented rhizoids.<br />

Collumellae are prominent and dome-shaped. Sporangiospores are small, oblong, 1-2<br />

x 3-4 µm, and are discharged through the neck following the dissolution of an apical<br />

mucilaginous plug. RG-2 organism.<br />

Key Features: Mucorales, unique flask-shaped sporangia, failure to sporulate on<br />

primary isolation media.<br />

Molecular Identification: ITS sequencing is required for differentiation of species<br />

within the complex, and may be necessary to achieve identification in a timely manner<br />

(Alvarez et al. 2010b, Walther et al. 2012, Halliday et al. 2015).<br />

Comment: Laboratory identification of this fungus may be difficult or delayed because of<br />

the mould’s failure to sporulate on primary isolation media or on subsequent subculture<br />

onto potato dextrose agar. Sporulation may be stimulated by using the agar block<br />

method described by Ellis and Ajello (1982), Ellis and Kaminski (1985) and Padhye<br />

and Ajello (1988), although this may still take a period of days to weeks. Failure to<br />

sporulate prohibits antifungal susceptibility testing.<br />

The agar block method to induce sporulation<br />

of Saksenaea spp. and Apophysomyces<br />

spp.<br />

A small block of agar is cut from a well<br />

established culture grown on PDA and is<br />

placed in the centre of petri dish containing<br />

1% agar in distilled water. After 21 days<br />

at 26 O C sporangia might be formed at the<br />

periphery of the petri dish.<br />

Antifungal Susceptibility: S. vasiformis very limited data, due to poor sporulation<br />

(Sun et al. 2002 and Australian national data); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 0.125-2 2 POSA 0.016-0.25 0.25<br />

ITRA 0.016-0.03 0.03 VORI 0.5-4 4


176<br />

Descriptions of Medical Fungi<br />

Saksenaea vasiformis Saksena<br />

15 µm<br />

Saksenaea vasiformis showing a typically flask-shaped sporangium.<br />

References: Saksena (1953), Ellis and Hesseltine (1966), Ajello et al. (1976), Ellis and<br />

Ajello (1982), Ellis and Kaminski (1985), Pritchard et al. (1986), Padhye et al. (1988),<br />

Padhye and Ajello (1988), Goldschmied-Reouven et al. (1989), de Hoog et al. (2000)<br />

and Ellis (2005b).


Descriptions of Medical Fungi 177<br />

Saprochaete clavata (de Hoog et al.) de Hoog & M.Th. Smith<br />

Synonmy: Geotrichum clavatum de Hoog, M.Th. Smith & Guého.<br />

Saprochaete clavata (formerly known as Geotrichum clavatum), which is also closely<br />

related to Magnusiomyces capitatus (formerly known as Geotrichum capitatum or<br />

Saprochaeta capitata), has only very infrequently been described as involved in invasive<br />

human infection. However, an outbreak of invasive infections caused by Saprochaete<br />

clavata in haematology patients has been reported (Vaux et al. 2013).<br />

RG-1 organism.<br />

Morphological Description: Colonies are moderately fast growing, flat, whitish and<br />

butyrous. True hyphae are abundant, soon breaking up into rectangular arthroconidia<br />

of variable size, 2.8-4 3 x 6-20 μm. Sympodial conidiogenesis is occasionally present.<br />

Terminal parts of hyphae may swell and become thick-walled.<br />

Note: Saprochaete clavata and Magnusiomyces capitatus are human pathogens that<br />

are closely related and are frequently mistaken for each other.<br />

Molecular Identification: ITS sequencing is recommended for accurate species<br />

identification.<br />

MALDI-T<strong>OF</strong> MS: reliably identifies S. clavata, M. capitatus and Geotrichum candidum<br />

to the species level (Desnos-Ollivier et al. 2014).<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose + L-Arabinose - D-Glucitol -<br />

Fermentation Sucrose - D-Arabinose - M-D-glucoside -<br />

Glucose - Maltose - D-Ribose - D-Gluconate -<br />

Galactose - Cellobiose + L-Rhamnose - DL-Lactate +,w<br />

Sucrose - Trehalose - D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine nd 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate nd<br />

Trehalose - Raffinose - Erythritol - Nitrate -<br />

Assimilation Melezitose - Ribitol - Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose + D-Xylose - D-Mannitol - Growth at 37 O C +<br />

Antifungal Susceptibility: Saprochaete clavata (Vaux et al. 2013); MIC µg/mL.<br />

Note: S. clavata is intrinsically resistant to echinocandins.<br />

Antifungal Range Median Antifungal Range Median<br />

AmB 0.125-1 0.5 POSA 0.125-1 0.5<br />

5FC


178<br />

Descriptions of Medical Fungi<br />

Sarocladium W. Gams & D. Hawksw.<br />

Based on a recent molecular phylogenetic study the taxonomy of Acremonium was<br />

reviewed and some medically important species have been transferred to Sarocladium;<br />

i.e. S. kiliense (formerly A. kiliense) and S. strictum (formerly A. strictum). Although these<br />

genera are morphologically similar they are phylogenetically distant. Sarocladium can<br />

be morphologically differentiated from Acremonium by its elongated phialides rising<br />

solitary on vegetative hyphae or on conidiophores that are sparsely or repeatedly<br />

branched, the production of abundant adelophialides and elongated conidia. (Glenn et<br />

al. 1996, Summerbell et al. 2011, Giraldo et al. 2015).<br />

Sarocladium strictum (W. Gams) Summerbell<br />

Sarocladium strictum is commonly found in soil and plant debris. Cutaneous, CAPDrelated<br />

peritonitis and invasive infections in immunosuppressed patients have been<br />

reported.<br />

RG-1 organism.<br />

Morphological Description: Colonies growing rapidly, moist to slimy, pink or orange;<br />

reverse remaining colourless or turning pink to orange. Conidiophores simple,<br />

occasionally branched. Phialides slender, arising from submerged or slightly fasciculate<br />

aerial hyphae, 20-65 × 1.4-2.5 μm. Submerged sporulation frequent from reduced<br />

phialides. Conidia grouped in slimy heads, cylindrical or ellipsoidal, 3.3-5.5 × 0.9-1.8<br />

μm, hyaline.<br />

Molecular Identification: Summerbell et al. (2011) revised the genus and recommends<br />

using D1/D2 sequences for phylogenetic analysis and sequence-based identification.<br />

References: Glenn et al. (1996), Summerbell et al. (2011), Giraldo et al. (2015), de<br />

Hoog et al. (2015).<br />

a<br />

b<br />

10 µm<br />

Sarocladium strictum (a) colony and (b) slender phialides with conidia in slimy heads.<br />

Antifungal Susceptibility: S. strictum very limited data (Australian National data);<br />

MIC µg/mL.<br />

No. 32<br />

AmB 5 1 2 2<br />

VORI 5 1 1 1 2<br />

POSA 5 1 2 1 1<br />

ITRA 5 1 1 3


Descriptions of Medical Fungi 179<br />

Scedosporium Sacc. ex Castell. & Chalm.<br />

The taxonomy of this genus has been subject to change on the basis of sequence data;<br />

Scedosporium apiospermum and Scedosporium boydii (formerly Pseudallescheria<br />

boydii) are now recognised as separate species and along with S. aurantiacum are<br />

the principal human pathogens (Lackner et al. 2014a). The majority of infections<br />

are mycetomas, the remainder include infections of the eye, ear, central nervous<br />

system, internal organs and more commonly the lungs. Scedosporium dehoogii and<br />

S. minutispora are mainly isolated from environmental samples and have been rarely<br />

reported from clinical cases (Gilgado et al. 2005, Rainer and de Hoog 2006, Cortez et<br />

al. 2008, Kaltseis et al. 2009).<br />

Scedosporium prolificans has been transferred to the genus Lomentospora. L. prolificans<br />

is phylogenetically and morphologically distinct from the remaining Scedosporium<br />

species (Lennon et al. 1994, Lackner et al. 2014a).<br />

Morphological identification of Scedosporium species has become increasingly<br />

unreliable and molecular identification methods are now recommended. The conidial<br />

states of S. apiospermum and S. boydii are morphologically indistinguishable; although<br />

the latter is homothallic and produces ascocarps. S. aurantiacum also exhibits similar<br />

conidial morphology but most strains produce a pale to bright yellow diffusible pigment<br />

on potato dextrose agar.<br />

Molecular Identification: Recommended genetic markers are ITS and β-tubulin<br />

(Lackner et al. 2012a).<br />

MALDI-T<strong>OF</strong> MS: A comprehensive ‘in-house’ database of reference spectra allows<br />

accurate identification of Scedosporium and Lomentospora species (Lau et al. 2013,<br />

Sitterlé et al. 2014).<br />

References: McGinnis (1980), Domsch et al. (1980), McGinnis et al. (1982), Campbell<br />

and Smith (1982), Rippon (1988), de Hoog et al. (2000, 2015), Gilgado et al. (2005),<br />

Rainer and de Hoog (2006), Guarro et al. (2006), Lackner et al. (2014a).<br />

Scedosporium apiospermum (Saccardo) Castellani and Chalmers<br />

Synonymy: Pseudallescheria apiosperma Gilgado, Gené, Cano & Guarro<br />

RG-2 organism.<br />

Morphological Description: Colonies are fast growing, greyish-white, suede-like<br />

to downy with a greyish-black reverse. Numerous single-celled, pale-brown, broadly<br />

clavate to ovoid conidia, 4-9 x 6-10 µm, rounded above with truncate bases are<br />

observed. Conidia are borne singly or in small groups on elongate, simple or branched<br />

conidiophores or laterally on hyphae. Conidial development can be described as<br />

annellidic, although the annellations (ring-like scars left at the apex of an annellide after<br />

conidial secession) are extremely difficult to see. Erect synnemata may be present in<br />

some isolates. Optimum temperature for growth is 30-37 O C.


180<br />

Descriptions of Medical Fungi<br />

Scedosporium apiospermum (Saccardo) Castellani and Chalmers<br />

a<br />

20 µm<br />

10 µm<br />

b<br />

c<br />

Scedosporium apiospermum (a) conidiophores and conidia,<br />

(b) culture and (c) synnemata.<br />

Antifungal Susceptibility: S. apiospermum (Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 167 2 2 10 45 51 40 17<br />

VORI 163 5 27 57 47 23 1 3<br />

POSA 112 2 6 22 49 27 3 1 2<br />

ITRA 167 2 5 21 86 38 7 8


Descriptions of Medical Fungi 181<br />

Scedosporium aurantiacum Gilgado et al.<br />

Morphological Description: Most isolates produce a light yellow diffusible pigment<br />

on potato dextrose agar after a few days incubation. Conidiogenous cells and<br />

conidia are similar in shape and size to S. apiospermum, and the two can best be<br />

distinguished by genetic analysis. Conidiogenous cells arising from undifferentiated<br />

hyphae are cylindrical to slightly flask-shaped, producing slimy heads of one-celled<br />

, smooth-walled, subhyaline, obovoid or sub-cylindrical conidia. 5-14 x 2-5 um. Erect<br />

synnemata may be present in some isolates, but the teleomorph is unknown. Optimum<br />

temperature for growth 37-40 O C, max 45 O C. RG-2 organism.<br />

a<br />

b<br />

Culture reverse (PDA) of (a) S. apiospermum and (b) S. aurantiacum showing<br />

production of a light yellow diffusible pigment that is typical of S. aurantiacum.<br />

20 µm<br />

Scedosporium aurantiacum conidiophores (annellides) and conidia.<br />

Antifungal Susceptibility: S. aurantiacum (Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 11 1 7 3<br />

VORI 11 5 5 1<br />

POSA 11 4 4 2 1<br />

ITRA 11 4 4 3


182<br />

Descriptions of Medical Fungi<br />

Synonymy: Pseudallescheria boydii (Shear) McGinnis, A.A. Padhye & Ajello.<br />

RG-2 organism.<br />

Scedosporium boydii (Shear) Gilgado et al.<br />

Morphological Description: Colonies are fast growing, greyish-white, suede-like<br />

to downy with a greyish-black reverse. Numerous single-celled, pale-brown, broadly<br />

clavate to ovoid conidia, 4-9 x 6-10 µm, rounded above with truncate bases are<br />

observed. Conidia are borne singly or in small groups on elongate, simple or branched<br />

conidiophores or laterally on hyphae. Cleistothecia (non-ostiolate ascocarps) are<br />

yellow-brown to black, spherical, 50-200 μm in diameter, and are mostly submerged<br />

in the agar and are composed of irregularly interwoven brown hyphae. When crushed<br />

cleistothecia release numerous, faintly brown, ellipsoidal ascospores, 4-5 x 7- 9 µm<br />

in size. Erect synnemata may be present in some isolates. Optimum temperature for<br />

growth is 30-37 O C.<br />

Note: S. boydii is homothallic and is recognised by smaller cleistothecia (50-200 μm)<br />

whereas S. apiospermum is heterothallic (requires mating of two strains) and has<br />

larger cleistothecia, 140-480 μm (Gilgado et al. 2010).<br />

a<br />

Scedosporium boydii (a) culture and a (b) cleistothecium.<br />

b<br />

20 µm<br />

Antifungal Susceptibility: S. boydii vs S. apiospermum (Lackner et al. 2014b).<br />

S. boydii S. apiospermum<br />

Antifungal (MIC µg/mL)<br />

Range MIC 90<br />

Range MIC 90<br />

AmB 0.5->16 >16 0.5->16 >16<br />

VORI 0.125-2 2 0.25->8 2<br />

POSA 0.125->16 >16 0.25->16 >16<br />

ITRA 0.125->16 >16 0.25->16 >16


Descriptions of Medical Fungi 183<br />

Schizophyllum commune Fries<br />

Schizophyllum commune is a common basidiomycete bracket fungus found on rotten<br />

wood, and is an occasional human pathogen, principally associated with sinusitis,<br />

allergic bronchopulmonary mycosis and as a contaminant from respiratory specimens.<br />

However the introduction of DNA sequencing and/or MALDI-T<strong>OF</strong> MS identification in<br />

the clinical laboratory has seen many more cases of S. commune fungal rhinosinusitis<br />

identified (Michel et al. 2012, Chowdhary et al. 2013a, 2014a,b).<br />

RG-1 organism.<br />

Morphological Description: Colonies on 2% malt extract agar are spreading, woolly,<br />

whitish to pale greyish-brown, soon forming macroscopically visible fruiting bodies.<br />

Some isolates may take up to 12 weeks to form fruiting bodies. Fruit bodies are sessile,<br />

kidney-shaped, lobed with split gills on the lower side. Hyphae are hyaline, wide<br />

and have clamp connections. Basidia bear four basidiospores on erect sterigmata.<br />

Basidiospores hyaline, smooth-walled, elongate with lateral scar at lower end, 6-7 x<br />

2-3 µm.<br />

Note: Many clinical isolates of S. commune are monokaryotic and do not show clamp<br />

connections, therefore any white, rapidly growing, sterile isolate showing good growth<br />

at 37 O C with tolerance to benomyl, susceptibility to cycloheximide, and a pronounced<br />

odour should be suspected of being S. commune (Sigler et al. 1995).<br />

Molecular Identification: Sequencing of the ITS and D1/D2 regions is recommended<br />

(Buzina et al. 2001, Won et al. 2012, Chowdhary et al. 2013b, Michel et al. 2012),<br />

however the number of well identified nucleotide sequences of these fungi in the<br />

GenBank database remains limited.<br />

MALDI-T<strong>OF</strong> MS: Michel et al. (2012), Chowdhary et al. (2014b), Huguenin et al. (2015)<br />

provide identification procedures, however the number of mass spectral profiles to be<br />

found in MALDI-T<strong>OF</strong> libraries remains limited.<br />

References: McGinnis (1980), Rippon (1988), Sigler et al. (1995), de Hoog et al.<br />

(2015).<br />

Schizophyllum commune basidiocarps growing on malt extract agar.<br />

Antifungal Susceptibility: S. commune (Chowdhary et al. 2013b); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 0.03-2 1 FLU 2-64 64<br />

ITRA 0.03-8 1 VORI 0.06-2 0.5


184<br />

Descriptions of Medical Fungi<br />

Scopulariopsis Bain<br />

Most members of the genus Scopulariopsis are soil fungi, which are frequently isolated<br />

from food, paper and other materials. They also occur as laboratory contaminants.<br />

Several species have been reported as causative agents of onychomycosis and<br />

hyalohyphomycosis (Sandoval-Denis et al. 2013). The most common species seen in<br />

the clinical laboratory is S. brevicaulis, followed by S. gracilis S. brumptii, Microascus<br />

cinereus, S. candida complex, and M. cirrosus (Sandoval-Denis et al. 2013).<br />

Morphological Description: Colonies are fast growing, varying in colour from white,<br />

cream, grey, buff to brown and black, but are predominantly light brown. Microscopic<br />

morphology shows chains of single-celled conidia produced in basipetal succession<br />

from a specialised conidiogenous cell called an annellide. Once again, the term<br />

basocatenate can be used to describe such chains of conidia where the youngest<br />

conidium is at the basal end of the chain. In Scopulariopsis, annellides may be solitary,<br />

in groups, or organised into a distinct penicillus. Conidia are globose to pyriform, usually<br />

truncate, with a rounded distal portion, smooth to rough, and hyaline to brown in colour.<br />

RG-2 for species isolated from humans.<br />

Key Features: Hyphomycete, conidia often shaped like light globes, basocatenate<br />

arising from annellides.<br />

Molecular Identification: D1/D2 and EF-1α sequence analysis can be useful for the<br />

identification of the most common clinically relevant species (Sandoval-Denis et al.<br />

2013).<br />

References: Morton and Smith (1963), McGinnis (1980), Rippon (1988), Samson et<br />

al. (1995), Domsch et al. (2007), de Hoog et al. (2000, 2015).<br />

10 µm<br />

Scopulariopsis brevicaulis conidiophores (annellides) and conidia.<br />

Antifungal Susceptibility: S. brevicaulis (Skora et al. 2014); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 4-16 >16 VORI 8->16 >16<br />

ITRA >16 >16 TERB 0.5-16 4


Descriptions of Medical Fungi 185<br />

Sepedonium species are common soil fungi and parasites of fleshy fungi, however<br />

Yogo et al. (2014) reported an intra-abdominal infection in an immunosuppressed<br />

patient. Sepedonium species closely resemble Histoplasma capsulatum. Note: For<br />

laboratory safety, culture identification to exclude Histoplasma capsulatum should be<br />

performed by either exoantigen test or DNA sequencing.<br />

RG-1 organism.<br />

Sepedonium Link ex Greville<br />

Morphological Description: Colonies are moderately fast growing, usually white<br />

to golden yellow, suede-like to downy, becoming fluffy with age. Conidiophores are<br />

hyaline and non-specialised, resembling short branches of the vegetative hyphae.<br />

Conidia are terminal, solitary, or in clusters, one-celled, globose to ovoid, 7-17 µm,<br />

hyaline to amber, smooth to verrucose and usually with a thick wall.<br />

Key Features: Hyphomycete, producing large, thick-walled, one-celled, verrucose,<br />

globose, terminal conidia from non-specialised conidiophores, resembling the<br />

macroconidia seen in Histoplasma capsulatum.<br />

Molecular Identification: ITS and D1/D2 sequencing may be used for accurate<br />

species identification (Halliday et al. 2015).<br />

References: McGinnis (1980), Rippon (1988), Yogo et al. (2014).<br />

15 µm<br />

Sepedonium spp. showing large, globose, thick-walled,<br />

one-celled, verrucose, terminal conidia.


186<br />

Descriptions of Medical Fungi<br />

Sporothrix schenckii complex<br />

It is now recognised that Sporothrix schenckii is a species complex of five distinct<br />

species: S. schenckii sensu strictu, S. brasiliensis, S. globosa, S. mexicana and S.<br />

luriei (Marimon et al. 2007, Romeo et al. 2011, Barros et al. 2011, Oliveira et al. 2014,<br />

Zhang et al. 2015b). RG-2 organism.<br />

Sporothrix schenckii complex is a dimorphic fungus and has a worldwide distribution,<br />

particularly in tropical and temperate regions. It is commonly found in soil and on decaying<br />

vegetation and is a well-known pathogen of humans and animals. Sporotrichosis is<br />

primarily a chronic mycotic infection of the cutaneous or subcutaneous tissues and<br />

adjacent lymphatics characterised by nodular lesions which may suppurate and<br />

ulcerate. Infections are caused by the traumatic implantation of the fungus into the<br />

skin, or very rarely, by inhalation into the lungs. Secondary spread to articular surfaces,<br />

bone and muscle is not infrequent, and the infection may also occasionally involve the<br />

central nervous system, lungs or genitourinary tract.<br />

Sporothrix schenckii Hektoen & Perkins<br />

Morphological Description: Colonies at 25 O C, are slow growing, moist and glabrous,<br />

with a wrinkled and folded surface. Some strains may produce short aerial hyphae<br />

and pigmentation may vary from white to cream to black. Conidiophores arise at right<br />

angles from thin septate hyphae and are usually solitary, erect and tapered toward the<br />

apex. Conidia are formed in clusters on tiny denticles by sympodial proliferation at the<br />

apex of the conidiophore, their arrangement often suggestive of a flower. As the culture<br />

ages, conidia are subsequently formed singly along the sides of both conidiophores<br />

and undifferentiated hyphae. Conidia are ovoid or elongated, 3-6 x 2-3 µm, hyaline,<br />

one-celled and smooth-walled. In some isolates, solitary, darkly-pigmented, thickwalled,<br />

one-celled, obovate to angular conidia may be observed along the hyphae. On<br />

brain heart infusion (BHI) agar containing blood at 37 O C, colonies are glabrous, white<br />

to greyish-yellow and yeast-like consisting of spherical or oval budding yeast cells.<br />

Molecular Identification: DNA sequencing using ITS, D1/D2, β-tubulin, calmodulin<br />

and chalcone synthase genes is recommended for species identification (Marimon<br />

et al. 2007, Romeo et al. 2011, Barros et al. 2011, Oliveira et al. 2014, Zhang et al.<br />

2015b).<br />

MALDI-T<strong>OF</strong> MS: Oliverira et al. (2015) established a MALDI-T<strong>OF</strong> protocol and<br />

reference database for the identification of Sporothrix species.<br />

Key Features: Hyphomycete characterised by thermal dimorphism and clusters of<br />

ovoid, denticulate conidia produced sympodially on short conidiophores.<br />

References: McGinnis (1980), Domsch et al. (1980), Rippon (1988), de Hoog et al.<br />

(1985, 2000, 2015).<br />

Antifungal Susceptibility: S. schenckii variable data for amphotericin B and azoles;<br />

testing of individual strains recommended (Alvarado-Ramirez and Torres-Rodriguez<br />

2007, Marimon et al. 2008, Silveira et al. 2009, Oliveira et al. 2011, Ottonelli Stopiglia<br />

et al. 2014, Rodrigues et al. 2014, Australian National data); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 0.03->16 >16 POSA 0.03->16 8<br />

ITRA 0.03->16 >16 VORI 0.125->16 >16<br />

TERB 0.03-1 0.5


Descriptions of Medical Fungi 187<br />

Sporothrix schenckii complex<br />

10 µm<br />

Sporothrix schenckii PAS stained tissue section showing budding yeast-like cells.<br />

a<br />

b<br />

10 µm<br />

Sporothrix schenckii (a) culture at 25 O C and (b) budding yeast cells in BHI at 37 O C.<br />

10 µm<br />

Sporothrix schenckii conidiophores and conidia at 25 O C.


188<br />

Descriptions of Medical Fungi<br />

Most species of Stemphylium are plant pathogens with occasional isolates from soil,<br />

they are rarely seen in the clinical laboratory.<br />

RG-1 organism.<br />

Stemphylium Wallroth<br />

Morphological Description: Colonies are rapid growing, brown to olivaceous-black or<br />

greyish and suede-like to floccose. Microscopically, solitary, darkly pigmented, terminal,<br />

multicellular conidia (dictyoconidia) are formed on a distinctive conidiophore with a<br />

darker terminal swelling. Note: The conidiophore proliferates percurrently through the<br />

scar where the terminal conidium (poroconidium) was formed. Conidia are pale to<br />

mid-brown, oblong, rounded at the ends, ellipsoidal, obclavate or subspherical and<br />

are smooth or in part verrucose. Stemphylium should not be confused with Ulocladium<br />

which produces similar dictyoconidia from a sympodial conidiophore, not from a<br />

percurrent conidiogenous cell as in Stemphylium.<br />

Molecular Identification: ITS sequencing (Woudenberg et al. 2013).<br />

Key Features: Dematiaceous hyphomycete producing darkly pigmented, dictyoconidia<br />

from the swollen end of a percurrent conidiophore.<br />

References: Ellis (1971, 1976), Rippon (1988), de Hoog et al. (2000).<br />

20 µm<br />

Stemphylium spp. conidiophores and conidia.


Descriptions of Medical Fungi 189<br />

Syncephalastrum racemosum Cohn<br />

The genus Syncephalastrum is characterised by the formation of cylindrical<br />

merosporangia on a terminal swelling of the sporangiophore. Sporangiospores are<br />

arranged in a single row within the merosporangia. Syncephalastrum racemosum is the<br />

type species of the genus and a potential human pathogen; however, well-documented<br />

cases are lacking. It is found mainly from soil and dung in tropical and subtropical<br />

regions. It can also be a laboratory aerial contaminant. The sporangiophore and<br />

merosporangia of Syncephalastrum species may also be mistaken for an Aspergillus<br />

species, if the isolate is not examined carefully. RG-2 organism.<br />

Morphological Description: Colonies are very fast growing, cottony to fluffy, white to<br />

light grey, becoming dark grey with the development of sporangia. Sporangiophores are<br />

erect, stolon-like, often producing adventitious rhizoids, and show sympodial branching<br />

(racemose branching) producing curved lateral branches. The main stalk and branches<br />

form terminal, globose to ovoid vesicles which bear finger-like merosporangia directly<br />

over their entire surface. At maturity, merosporangia are thin-walled, evanescent<br />

and contain five to ten (up to 18) globose to ovoid, smooth-walled sporangiospores<br />

(merospores). Maximum growth temperature 40 O C.<br />

Key Features: Mucorales, producing sympodially branching sporangiophores with<br />

terminal vesicles bearing merosporangia.<br />

References: Domsch et al. (1980), McGinnis (1980), Onions et al. (1981), Rippon<br />

(1988), Samson et al. (1995), de Hoog et al. (2000, 2015), Ellis (2005b).<br />

a<br />

30 µm<br />

b<br />

10 µm<br />

Syncephalastrum racemosum (a) merosporangia and (b) merospores.<br />

Antifungal Susceptibility: S. racemosum (Espinel-Ingroff et al. 2015a); MIC µg/mL.<br />

No. 16<br />

AmB 35 8 16 3 6 2<br />

POSA 36 1 2 4 10 11 5 1 2<br />

ITRA 26 4 3 5 7 4 1 2


190<br />

Descriptions of Medical Fungi<br />

Talaromyces marneffei (Segretain et al.) Samson et al.<br />

Synonymy: Penicillium marneffei Segretain et al.<br />

WARNING: RG-3 organism. Cultures of Talaromyces marneffei may represent a<br />

biohazard to laboratory personnel and should be handled with caution in a class II<br />

Biological Safety Cabinet (BSCII). T. marneffei exhibits thermal dimorphism and is<br />

endemic in Southeast Asia and the southern region of China.<br />

Samson et al. (2011b) redefined Talaromyces by combining Penicillium subgenus<br />

Biverticillium into Talaromyces based upon phylogenetic analysis of the ITS and RPB1<br />

loci. The genus contains 88 species that were placed into seven sections based on a<br />

multigene phylogeny of the ITS, β-tubulin and RPB2 regions (Yilmaz et al. 2014). T.<br />

marneffei is the only known dimorphic species in the genus, producing filamentous<br />

growth at 25 O C and a yeast phase at 37 O C (Andrianopoulos 2002).<br />

Molecular Identification: ITS sequencing is recommended, as well as β-tubulin as a<br />

secondary molecular marker for identification (Yilmaz et al. 2014).<br />

Morphological Description: Colonies at 25 O C are fast growing, suede-like to downy,<br />

white with yellowish-green conidial heads. Colonies become greyish-pink to brown<br />

with age and produce a diffusible brownish-red to wine-red pigment. Conidiophores<br />

generally biverticillate and sometimes monoverticillate; hyaline, smooth-walled and<br />

bear terminal verticils of three to five metulae, each bearing three to seven phialides.<br />

Phialides are acerose to flask-shaped. Conidia are globose to subglobose, 2-3 µm in<br />

diameter, smooth-walled and are produced in basipetal succession from the phialides.<br />

On brain heart infusion (BHI) agar containing blood incubated at 37 O C, colonies are<br />

rough, glabrous, tan-coloured and yeast-like. Microscopically, yeast cells are spherical<br />

to ellipsoidal, 2-6 µm in diameter, and divide by fission rather than budding. Numerous<br />

short hyphal elements are also present.<br />

Histopathology: Tissue sections show small, oval to ellipsoidal yeast-like cells, 3 µm in<br />

diameter, either packed within histiocytes or scattered through the tissue. Occasional,<br />

large, elongated sausage-shaped cells, up to 8 µm long, with distinctive septa may be<br />

present.<br />

Key Features: Talaromyces marneffei is the only dimorphic species of Talaromyces,<br />

which grows as a yeast at 37 O C. It produces a red soluble pigment on general media<br />

and conidiophores have flask-shaped to acerose phialides.<br />

References: Pitt (1979), Ramirez (1982), de Hoog et al. (2000, 2015), Andrianopoulos<br />

(2002), Lyratzopoulos et al. (2002), Samson et al. (2011b), Visagie et al. (2014), Yilmaz<br />

et al. (2014).


Descriptions of Medical Fungi 191<br />

Talaromyces marneffei (Segretain et al.) Samson et al.<br />

a<br />

b<br />

5 µm<br />

c<br />

15 µm 5 µm<br />

c<br />

Talaromyces marneffei (a) colony, (b) a giemsa stained touch smear showing<br />

typical septate yeast-like cells (arrow), (c) phialides and conidia.<br />

Antifungal Susceptibility: T. marneffei very limited data (Australian National data);<br />

MIC µg/mL.<br />

No. 8<br />

AmB 5 2 2 1<br />

VORI 5 3 2<br />

POSA 4 3 1<br />

ITRA 5 1 4


192<br />

Descriptions of Medical Fungi<br />

Torulaspora delbrueckii (Lindner) Lindner<br />

Synonymy: Candida colliculosa (Hartmann) S.A. Meyer & Yarrow.<br />

Torulaspora delbrueckii is a rare cause of candidaemia.<br />

RG-1 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Spherical to ellipsoidal budding blastoconidia, 2-6 x 3-7 µm in size.<br />

Ascospores may be produced on 5% malt extract or cornmeal agar after 5-30 days at<br />

25 O C.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Budding yeast cells only. No pseudohyphae or true hyphae<br />

produced.<br />

Molecular Identification: ITS sequencing recommended.<br />

MALDI-T<strong>OF</strong> MS: Able to accurately identify this species.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Germ Tube - L-Sorbose v L-Arabinose - D-Glucitol v<br />

Fermentation Sucrose v D-Arabinose - α-M-D-glucoside v<br />

Glucose + Maltose v D-Ribose - D-Gluconate v<br />

Galactose v Cellobiose - L-Rhamnose - DL-Lactate v<br />

Sucrose v Trehalose -,s D-Glucosamine - myo-Inositol -<br />

Maltose v Lactose - N-A-D-glucosamine - 2-K-D-gluconate +<br />

Lactose - Melibiose - Glycerol v D-Glucuronate v<br />

Trehalose v Raffinose v Erythritol - Nitrate -<br />

Assimilation Melezitose v Ribitol v Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose v D-Xylose v D-Mannitol + Growth at 37 O C v<br />

Key Features: asci containing one to four spheroidal ascospores, strong growth at<br />

37 O C and a variable sugar assimilation profile.<br />

Antifungal Susceptibility: T. delbrueckii very limited data (Australian National<br />

data); MIC µg/mL.<br />

No. 64<br />

AmB 2 2<br />

FLU 2 2<br />

VORI 2 1 1<br />

POSA 2 1 1<br />

ITRA 2 2<br />

CAS 1 1<br />

5FC 2 2


Descriptions of Medical Fungi 193<br />

Trichoderma Persoon ex Grey<br />

Trichoderma is a very common genus especially in soil and decaying wood. Gliocladium<br />

(with strongly convergent phialides) and Verticillium (with straight and moderately<br />

divergent phialides) are closely related genera. Trichoderma infections in humans<br />

have been associated mostly with peritoneal dialysis, organ transplantation, and<br />

haematologic disorders (Sandoval-Denis et al. 2014b).<br />

Morphological Description: Colonies are fast growing, at first white and downy, later<br />

developing yellowish-green to deep green compact tufts, often only in small areas<br />

or in concentric ring-like zones on the agar surface. Conidiophores are repeatedly<br />

branched, irregularly verticillate, bearing clusters of divergent, often irregularly bent,<br />

flask-shaped phialides. Conidia are mostly green, sometimes hyaline, with smooth or<br />

rough walls and are formed in slimy conidial heads (gloiospora) clustered at the tips of<br />

the phialides. RG-1 organism.<br />

Key Features: Hyphomycete with repeatedly branched conidiophores bearing clusters<br />

of divergent, flask-shaped phialides.<br />

Molecular Identification: Species identification is based on multilocus sequence data<br />

using ITS, EF-1α, Chi18-5, and actin genes (Sandoval-Denis et al. 2014b).<br />

References: Domsch et al. (1980), McGinnis (1980), Rippon (1988), Samson et al.<br />

(1995), de Hoog et al. (2000, 2015), Sandoval-Denis et al. (2014b).<br />

20 µm<br />

Trichoderma harzianum species complex phialides and conidia.<br />

Antifungal Susceptibility: Trichoderma spp.(Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 3 2 1<br />

VORI 3 1 1 1<br />

POSA 2 2<br />

ITRA 3 2 1<br />

Trichoderma spp. data from 73 isolates (Sandoval-Denis et al. 2014b); MIC µg/mL.<br />

AmB Range 0.03-8; MIC 90<br />

= 2 VORI Range 0.125-32; MIC 90<br />

= 4<br />

ITRA Range 1-32; MIC 90<br />

= 32 POSA Range 1-32; MIC 90<br />

= 32


194<br />

Descriptions of Medical Fungi<br />

Trichophyton Malmsten<br />

Rippon (1988) accepted 22 species and four varieties in the genus Trichophyton based<br />

on morphology. DNA sequences now play a prominent role in delineating phylogenetic<br />

relationships, and as such species concepts in Trichophyton have changed. Sixteen<br />

species are now recognised in the genus. The descriptions and species concepts<br />

provided in this publication are based upon a combination of traditional morphological<br />

criteria and the current (2016) recognised phylogenetic species (de Hoog et al. 2016).<br />

The genus Trichophyton is characterised morphologically by the development of both<br />

smooth-walled macro- and microconidia. Macroconidia are mostly borne laterally<br />

directly on the hyphae or on short pedicels, and are thin- or thick-walled, clavate to<br />

fusiform, and range from 4-8 x 8-50 µm in size. Macroconidia are few or absent in many<br />

species. Microconidia are spherical, pyriform to clavate or of irregular shape and range<br />

from 2-3 x 2-4 µm in size. The presence of microconidia differentiates this genus from<br />

Epidermophyton, and the smooth-walled, mostly sessile macroconidia differentiates it<br />

from Lophophyton, Microsporum and Nannizzia.<br />

In practice, two groups may be recognised on direct microscopy:<br />

1. Those species that usually produce microconidia; macroconidia may or may<br />

not be present i.e. T. rubrum, T. interdigitale, T. mentagrophytes, T. equinum, T.<br />

eriotrephon, T. tonsurans, and to a lesser extent T. verrucosum, which may produce<br />

conidia on some media. In these species the shape, size and arrangement of the<br />

microconidia is the most important character. Culture characteristics are also useful.<br />

2. Those species that usually do not produce conidia. Chlamydospores or other<br />

hyphal structures may be present, but microscopy is generally non-diagnostic; i.e.<br />

T. verrucosum, T. violaceum, T. concentricum, T. schoenleinii and T. soudanense.<br />

Culture characteristics and clinical information such as the site, appearance of the<br />

lesion, geographic location, travel history, animal contacts and even occupation are<br />

most important.<br />

Many laboratories have used growth on additional media and/or confirmatory tests to<br />

help differentiate between species of Trichophyton, especially isolates of T. rubrum, T.<br />

interdigitale, T. mentagrophytes and T. tonsurans. These include growth characteristics<br />

on media such as Littman oxgall agar, lactritmel agar, potato dextrose agar, Sabouraud’s<br />

agar with 5% Salt, 1% peptone agar, bromocresol purple-milk solids glucose agar<br />

(BCP), Trichophyton agars No. 1-5, hydrolysis of urea and hair perforation tests.<br />

Molecular Identification: ITS and EF-1α sequencing is recommended for accurate<br />

species identification (Gräser et al. 1998, 1999b, 2000a, 2008; Irinyi et al. 2015;<br />

Mirhendi et al. 2015).<br />

MALDI-T<strong>OF</strong> MS: Methods reported by Erhard et al. (2008), Nenoff et al. (2011),<br />

Cassange et al. (2011), l’Ollivier et al. (2013), Calderaro et al. (2014), Packeu et al.<br />

(2013, 2014).<br />

References: Rebell and Taplin (1970), Ajello (1972), Vanbreusegham et al. (1978),<br />

Rippon (1988), McGinnis (1980), Domsch et al. (1980), Kane et al. (1997), Chen et al.<br />

(2011), de Hoog et al. (2000, 2015, 2016).


Descriptions of Medical Fungi 195<br />

Trichophyton concentricum is an anthropophilic fungus which causes chronic<br />

widespread non-inflammatory tinea corporis known as tinea imbricata because of the<br />

concentric rings of scaling it produces. It is not known to invade hair. Infections among<br />

Europeans are rare. Distribution is restricted to the Pacific Islands of Oceania, South<br />

East Asia and Central and South America.<br />

RG-2 organism.<br />

Morphological Description: Colonies are slow growing, raised and folded, glabrous<br />

becoming suede-like, mostly white to cream-coloured, but sometimes orange-brown<br />

coloured, often deeply folded into the agar which may produce splitting of the medium<br />

in some cultures. Reverse is buff to yellow-brown to brown in colour. Cultures consist<br />

of broad, much-branched, irregular, often segmented, septate hyphae which may<br />

have “antler” tips resembling T. schoenleinii. Chlamydospores are often present in<br />

older cultures. Microconidia and macroconidia are not usually produced, although<br />

some isolates will produce occasional clavate to pyriform microconidia. Note: Hyphal<br />

segments may artificially resemble macroconidia.<br />

Confirmatory Tests:<br />

Trichophyton concentricum Blanchard<br />

Hydrolysis of Urea: Negative after 7 days.<br />

Vitamin Free Agar (Trichophyton Agar No.1): Growth occurs on vitamin free agar<br />

(T1) but is usually slightly better on media containing thiamine i.e. T3 = T1 + thiamine<br />

and inositol, and T4 = T1 + thiamine. The slight enhancement of growth in the presence<br />

of thiamine helps to distinguish T. concentricum from T. schoenleinii, although this<br />

does not occur in all strains.<br />

Hair Perforation Test: Negative at 28 days.<br />

Key Features: Clinical disease, geographical distribution and culture characteristics.<br />

a<br />

b<br />

20 µm<br />

Trichophyton concentricum (a) culture showing a typical slow growing, heaped and<br />

folded, glabrous to suede like colony, and (b) the formation of typical “balloon-shaped”<br />

chlamydospores. Note: Microconidia and macroconidia are usually not produced.


196<br />

Descriptions of Medical Fungi<br />

Trichophyton equinum (Matruchot & Dassonville) Gedoelst<br />

Synonomy: Trichophyton equinum var. autotrophicum J.M.B. Smith et al.<br />

Trichophyton equinum is a zoophilic fungus causing ringworm in horses and rare<br />

infections in humans. It has a worldwide distribution except for the autotrophicum strain<br />

which is restricted to Australia and New Zealand. Invaded hairs show an ectothrix<br />

infection but do not fluoresce under Wood’s ultra-violet light.<br />

RG-2 organism.<br />

Morphological Description: Colonies are usually flat, but some may develop gentle<br />

folds or radial grooves, white to buff in colour, suede-like to downy in texture, and are<br />

similar to T. mentagrophytes. Cultures usually have a deep-yellow submerged fringe<br />

and reverse which later becomes dark red in the centre. Microscopically: abundant<br />

microconidia which may be clavate to pyriform and sessile or spherical and stalked<br />

are formed laterally along the hyphae. Macroconidia are only rarely produced, but<br />

when present are clavate, smooth, thin-walled and of variable size. Occasional nodular<br />

organs may be present and the microconidia often undergo a transformation to produce<br />

abundant chlamydospores in old cultures.<br />

Confirmatory Tests:<br />

Lactritmel Agar: Flat spreading, white to cream-coloured, powdery to granular surface<br />

with a central downy papilla, and deep brownish-red reverse. Microscopic morphology<br />

as described above.<br />

Hydrolysis of Urea: Positive in 4-5 days.<br />

a<br />

b<br />

Nutritional Tests on Trichophyton Agars: (a) Most strains require nicotinic acid for<br />

growth except those from Australia and New Zealand, which are autotrophic (b). T1 =<br />

vitamin free agar, T5 = vitamin free + nicotinic acid agar.<br />

Hair Perforation Test: Negative; but positive for the autotrophicum strains.<br />

Molecular Diagnostics: Species identification supported by ITS sequencing (Gräser<br />

et al. 1999a; Chen et al. 2011).<br />

Key Features: Microscopic morphology, culture characteristics, nicotinic acid<br />

requirement and clinical lesions in horses. Most strains require nicotinic acid for growth<br />

except autotrophic strains.


Descriptions of Medical Fungi 197<br />

Trichophyton equinum (Matruchot & Dassonville) Gedoelst<br />

a<br />

c<br />

b<br />

20 µm<br />

d<br />

Trichophyton equinum (a) culture, (b) microconidia, (c) macroconidia<br />

and (d) nodular organs.


198<br />

Descriptions of Medical Fungi<br />

Trichophyton eriotrephon Papegaaij, Nederl. Tijdschr. Geneesk<br />

Synonymy: Trichophyton mentagrophytes var. erinacei J.M.B. Smith & Marples.<br />

Trichophyton erinacei (Smith & Marples) Quaife.<br />

Trichophyton eriotrephon is a zoophilic fungus associated with hedgehogs and the<br />

epidermal mites, harboured by hedgehogs. Human infections occur most frequently on<br />

the exposed parts of the body, but tinea of the scalp and nails can also occur. Invaded<br />

hairs show an ectothrix infection but do not fluoresce under Wood’s ultra-violet light.<br />

The distribution of this fungus is New Zealand and Europe.<br />

RG-2 organism.<br />

Morphological Description: Colonies are white, flat, powdery, sometimes downy<br />

to fluffy with a brilliant lemon-yellow reverse. Numerous large clavate microconidia<br />

are borne on the sides of hyphae. Macroconidia are smooth-walled, two to six-celled,<br />

clavate, variable in size, and may have terminal appendages. Macroconidia are much<br />

shorter than those seen in T. mentagrophytes.<br />

Confirmatory Tests:<br />

Lactritmel Agar: White suede-like to powdery colony with brilliant yellow reverse.<br />

Numerous large slender clavate microconidia.<br />

Hydrolysis of Urea: Negative at 7 days.<br />

Vitamin Free Agar (Trichophyton Agar No.1): Good growth indicating no special<br />

nutritional requirements. Colonies are white suede-like to powdery with no reverse<br />

pigment.<br />

Hair Perforation Test: Positive.<br />

Key Features: Include microscopic morphology and culture characteristics; slender<br />

clavate microconidia and brilliant lemon-yellow reverse pigment on SDA; and negative<br />

hydrolysis of urea.<br />

Positive “in vitro” hair perforation test.


Descriptions of Medical Fungi 199<br />

Trichophyton eriotrephon Papegaaij, Nederl. Tijdschr. Geneesk<br />

a<br />

c<br />

b<br />

20 µm<br />

Trichophyton eriotrephon (a) culture, (b) microconidia and (c) macroconidia.<br />

Trichophyton eriotrephon is generally distinguished from T. mentagrophytes by: (a)<br />

its microscopic morphology showing numerous large slender clavate microconidia<br />

borne on the sides of hyphae and its smooth, thin-walled clavate macroconidia; (b)<br />

its brilliant lemon yellow reverse pigment on SDA and lactritmel agar; (c) its lack of<br />

reverse pigment on Sabouraud’s salt agar; and (d) its negative hydrolysis of urea.


200<br />

Descriptions of Medical Fungi<br />

Synonymy: T. mentagrophytes var. interdigitale (Priestley) Moraes, Anais Bras.<br />

Trichophyton interdigitale is an anthropophilic fungus which is a common cause of<br />

tinea pedis, particularly the vesicular type, tinea corporis, and sometimes superficial<br />

nail plate invasion in humans. It is not known to invade hair in vivo but produces hair<br />

perforations in vitro. Distribution is worldwide. This species may be regarded as a<br />

clonal offshoot of the zoophilic T. mentagrophytes (de Hoog et al. 2016).<br />

RG-2 organism.<br />

Morphological Description: Colonies are usually flat, white to cream in colour with<br />

a powdery to suede-like surface and yellowish to pinkish brown reverse pigment,<br />

often becoming a darker red-brown with age. Numerous subspherical to pyriform<br />

microconidia, occasional spiral hyphae and spherical chlamydospores are present,<br />

the latter being more abundant in older cultures. Occasional slender, clavate, smoothwalled,<br />

multiseptate macroconidia are also present in some cultures.<br />

Confirmatory Tests:<br />

Littman Oxgall Agar: Raised white downy colony with no reverse pigment.<br />

Lactritmel Agar: Macroscopic and microscopic features as described above.<br />

Sabouraud’s Dextrose Agar with 5% Salt: Heaped and folded, buff-coloured, suedelike<br />

surface with a dark reddish-brown submerged fringe and brown reverse.<br />

1% Peptone Agar: Flat, white to cream, suede-like surface with raised white downy<br />

centre. No reverse pigment.<br />

Vitamin Free Agar (Trichophyton Agar No.1): Good growth indicating no special<br />

nutritional requirements, flat cream powdery surface with central downy tuft. Reverse<br />

pale pinkish-brown.<br />

Hydrolysis of Urea: Positive within 7 days (usually 3 to 5 days).<br />

Hair Perforation Test: Positive.<br />

Trichophyton interdigitale Priestley<br />

Key Features: Culture characteristics, microscopic morphology and in vitro perforation<br />

of human hair.<br />

Trichophyton interdigitale can be distinguished from T. rubrum and T. mentagrophytes<br />

by: (a) its culture characteristics and microscopic morphology on SDA and/or lactritmel<br />

agar; (b) its growth and colony morphology on Sabouraud’s salt agar (colonies of T.<br />

interdigitale and T. mentagrophytes unlike T. rubrum, grow very well on this medium<br />

and usually produce a distinctive dark reddish-brown reverse pigment); (c) a positive<br />

urease test (within 7 days), a positive hair perforation test and the production of a<br />

yellow-brown to pinkish-brown reverse pigment on pigment stimulation media like<br />

lactritmel and Trichophyton No.1 agars; (d) T. interdigitale demonstrates profuse growth<br />

and alkalinity on BCP milk solids agar; (e) on 1% peptone agar T. interdigitale has a<br />

suede-like to downy surface whereas T. mentagrophytes has a characteristic granular<br />

appearance.


Descriptions of Medical Fungi 201<br />

Trichophyton interdigitale Priestley<br />

Trichophyton interdigitale culture.<br />

a<br />

b<br />

20 µm<br />

A dysgonic variant of Trichophyton interdigitale (formerly var. nodulare); (a) culture<br />

showing distinctive bright yellow to apricot-coloured colonies with a suede-like to<br />

powdery surface and a bright yellow-brown to orange reverse; and (b) microscopically<br />

characteristic “nodular organs” are observed in the vegetative hyphae. Usually, no<br />

conidia are seen but some isolates, especially with subculture may produce subspherical<br />

to pyriform microconidia similar to those of T. interdigitale.


202<br />

Descriptions of Medical Fungi<br />

Trichophyton interdigitale Priestley<br />

20 µm<br />

Trichophyton interdigitale microconidia, macroconidia and spiral hyphae.<br />

20 µm<br />

Trichophyton interdigitale microconidia, chlamydospores and spiral hyphae.


Descriptions of Medical Fungi 203<br />

Trichophyton mentagrophytes (Robin) Blanchard<br />

Synonymy: T. mentagrophytes var. mentagrophytes (Robin) Sabour.<br />

Trichophyton mentagrophytes is a zoophilic fungus with a worldwide distribution and<br />

a wide range of animal hosts including mice, guinea-pigs, kangaroos, cats, horses,<br />

sheep and rabbits. Produces inflammatory skin or scalp lesions in humans, particularly<br />

in rural workers. Kerion of the scalp and beard may occur. Invaded hairs show an<br />

ectothrix infection but do not fluoresce under Wood’s ultra-violet light. Distribution is<br />

worldwide.<br />

RG-2 organism.<br />

Morphological Description: Colonies are generally flat, white to cream in colour,<br />

with a powdery to granular surface. Some cultures show central folding or develop<br />

raised central tufts or pleomorphic suede-like to downy areas. Reverse pigmentation is<br />

usually a yellow-brown to reddish-brown colour. Numerous single-celled microconidia<br />

are formed, often in dense clusters. Microconidia are hyaline, smooth-walled, and<br />

are predominantly spherical to subspherical in shape, however occasional clavate<br />

to pyriform forms may occur. Varying numbers of spherical chlamydospores, spiral<br />

hyphae and smooth, thin-walled, clavate-shaped, multicelled macroconidia may also<br />

be present.<br />

Confirmatory Tests:<br />

Littman Oxgall Agar: Raised greyish-white, suede-like to downy colony. Some<br />

cultures may show a diffusible yellow to brown pigment.<br />

Lactritmel Agar: Cultures are flat, white to cream in colour, with a powdery to granular<br />

surface. Some cultures develop raised central tufts or pleomorphic downy areas.<br />

Reverse pigmentation is yellow-brown to pinkish-brown to red-brown. Microscopic<br />

morphology similar to that described above, with predominantly spherical microconidia,<br />

often formed in dense clusters, and varying numbers of spherical chlamydospores,<br />

spiral hyphae and smooth, thin-walled, clavate, multiseptate macroconidia.<br />

Sabouraud’s Dextrose Agar with 5% Salt: Cultures are heaped and folded, buff to<br />

brown in colour, with a suede-like surface texture and characteristically have a very<br />

dark reddish-brown submerged peripheral fringe and reverse pigmentation.<br />

1% Peptone Agar: Flat, cream-coloured, powdery to granular colony with no reverse<br />

pigment.<br />

Hydrolysis of Urea: Positive within 7 days (usually 3 to 5 days).<br />

Vitamin Free Agar (Trichophyton Agar No.1): Good growth indicating no special<br />

nutritional requirements. Cultures are flat, cream-coloured, with a powdery to suedelike<br />

surface, and have a reddish-brown reverse pigmentation.<br />

Hair Perforation Test: Positive within 14 days.<br />

Key Features: Culture characteristics, microscopic morphology and clinical disease<br />

with known animal contacts. T. mentagrophytes can be distinguished from T. interdigitale<br />

by: (a) its granular appearance on the 1% Peptone agar; (b) its microscopic morphology<br />

of more spherical microconidia and generally greater numbers of macroconidia; and<br />

(c) a yellow to brown diffusible pigment is often seen on the Littman oxgall agar. Both<br />

T. interdigitale and T. mentagrophytes demonstrate profuse growth and alkalinity on<br />

BCP milk solids agar.


204<br />

Descriptions of Medical Fungi<br />

Trichophyton mentagrophytes (Robin) Blanchard<br />

a<br />

a<br />

b<br />

20 µm<br />

c<br />

20 µm d<br />

20 µm<br />

Trichophyton mentagrophytes (a) cultures, (b) macroconidia,<br />

(c) microconidia and (d) hyphal spirals.


Descriptions of Medical Fungi 205<br />

Trichophyton quinckeanum (Zopf) MacLeod & Münde<br />

Synonymy: T. mentagrophytes var. quinckeanum (Zopf) J.M.B. Smith & Austwick.<br />

Trichophyton quinckeanum causes “mouse favus” on mice, seen as thick, yellow,<br />

saucer-shaped crusted lesions up to 1 cm in diameter called scutula. Invaded hairs<br />

are rarely seen but they may show either ectothrix or endothrix infection. Infected<br />

human hairs do not fluoresce under Wood’s ultra-violet light, but very occasional hairs<br />

from experimental lesions in guinea pigs have shown a pale yellow fluorescence. The<br />

geographical distribution of this dermatophyte is difficult to establish, but is probably<br />

worldwide. It is often associated with mice plagues in the Australian Wheat Belt.<br />

RG-2 organism.<br />

Morphological Description: Colonies are generally flat, white to cream in colour, with<br />

a powdery to granular surface. Some cultures show central folding or develop raised<br />

central tufts or pleomorphic suede-like to downy areas. Reverse pigmentation is usually<br />

a yellow-brown to reddish-brown colour. Numerous microconidia are borne laterally<br />

along the sides of hyphae, and are predominantly slender clavate when young. With<br />

age the microconidia become broader and pyriform to spherical in shape. Occasional<br />

to moderate numbers of smooth, thin-walled, multiseptate, clavate to cigar-shaped<br />

macroconidia may be present. Varying numbers of spherical chlamydospores and<br />

spiral hyphae may also be present.<br />

Confirmatory Tests:<br />

Littman Oxgall Agar: Raised, dome-like bluish-grey, suede-like colony with a narrow<br />

flat, greyish-white, suede-like border. No diffusible or reverse pigment is present.<br />

Lactritmel Agar: Flat, white to cream, suede-like to powdery colony with either no<br />

reverse pigment or a pale yellow-brown to pinkish-brown reverse. Numerous slender<br />

clavate to pyriform (depending on age of subculture) microconidia and occasional to<br />

moderate numbers of smooth, thin-walled, clavate macroconidia are present.<br />

Sabouraud’s Dextrose Agar with 5% salt: Heaped and much folded white suede-like<br />

colony with very pale yellow-brown reverse. No submerged fringe.<br />

1% Peptone Agar: Raised white suede-like to downy colony with no reverse pigment.<br />

Hydrolysis of Urea: Positive within 7 days (usually very rapid 2-3 days).<br />

Vitamin Free Agar (Trichophyton Agar No.1): Flat, white to cream, suede-like colony<br />

with either no reverse pigment or a pale yellow-brown reverse, i.e. no special nutritional<br />

requirements.<br />

Hair Perforation Test: Positive in 7 to 10 days.<br />

Key Features: Culture characteristics, microscopic morphology, and rapid urease test.


206<br />

Descriptions of Medical Fungi<br />

Trichophyton quinckeanum (Zopf) MacLeod & Münde<br />

Trichophyton quinckeanum may be distinguished from T. mentagrophytes by: (a) its<br />

characteristic culture appearance on Littman oxgall agar (i.e. raised, dome-like, bluishgrey<br />

suede-like colony with a narrow flat, greyish-white, suede-like border and no<br />

diffusible or reverse pigment) and on Sabouraud’s 5% salt agar (typically heaped and<br />

folded white suede-like colony, but with no distinctive dark reddish-brown submerged<br />

fringe and reverse pigment as seen on T. mentagrophytes); (b) microscopic morphology<br />

showing numerous slender clavate with some pyriform microconidia and moderate<br />

numbers of smooth thin-walled, clavate macroconidia; and (c) a rapid urease test,<br />

usually positive within 2 to 3 days.<br />

a<br />

b<br />

20 µm<br />

Trichophyton quinckeanum (a) culture and (b) microconidia and macroconidia.


Descriptions of Medical Fungi 207<br />

Trichophyton rubrum (Castellani) Semon<br />

Synonymy: Trichophyton fischeri Kane.<br />

Trichophyton raubitschekii Kane, Salkin, Weitzman & Smitka.<br />

Trichophyton kanei Summerbell.<br />

all varieties of T. rubrum.<br />

Trichophyton rubrum is an anthropophilic fungus that has become the most widely<br />

distributed dermatophyte of humans. It frequently causes chronic infections of skin,<br />

nails and rarely scalp. Granulomatous lesions may sometimes occur. Infected hairs do<br />

not fluoresce under Wood’s ultraviolet light, and microscopically may show endothrix<br />

or ectothrix type of invasion.<br />

Morphologically T. rubrum exhibits a spectrum of overlapping characters; for example<br />

culture surface texture may vary from downy to suede-like; culture surface pigmentation<br />

may vary from white to cream to deep red; culture reverse pigmentation may vary<br />

from colourless to yellowish to yellow-brown to wine red; numbers of microconidia<br />

range from none to scanty to many; shape of microconidia vary from slender clavate to<br />

pyriform; numbers of macroconidia range from none to scanty to many and may or may<br />

not have terminal projections. This is why so many varieties or synonyms have been<br />

described in the past. However, molecular evidence by Gräser et al. (1999b) reveals<br />

little variation between strains of T. rubrum and determined that the species is largely<br />

clonal.<br />

The majority of isolates, especially those causing tinea pedis and onychomycosis, are<br />

characterised by the production of scanty to moderate numbers of slender clavate<br />

microconidia and no macroconidia (formerly the “downy strain”). Some isolates,<br />

usually from cases of tinea corporis, are characterised by the production of moderate<br />

to abundant numbers of clavate to pyriform microconidia and moderate to abundant<br />

numbers of thin-walled, cigar-shaped macroconidia (formerly the “granular strain”).<br />

It should be stressed that intermediate strains do occur as many culture and<br />

morphological characteristics overlap.<br />

a b c d<br />

Trichophyton rubrum culture morphology; (a) “downy strain” with typical wine-red<br />

reverse; (b) “Y variety” with both yellow and red pigmentation; (c) “var. flava” with<br />

yellow pigmentation; (d) “granular strain” with red surface and reverse pigmentation.


208<br />

Descriptions of Medical Fungi<br />

RG-2 organism.<br />

Morphological Description: Colonies are mostly flat to slightly raised, white to cream,<br />

suede-like to downy, with either no reverse pigment or a yellow-brown to wine-red<br />

reverse. Most cultures show scanty to moderate numbers of slender clavate to pyriform<br />

microconidia. Macroconidia are usually absent, but when present are smooth, thinwalled<br />

multiseptate, slender and cylindrical to cigar-shaped. Older cultures may show<br />

numerous chlamydospores with a few clavate to pyriform microconidia.<br />

Note: On primary isolation some cultures may lack reverse pigmentation and fail to<br />

produce microconidia. These need to be subcultured onto media like lactritmel agar or<br />

potato dextrose agar, which stimulate pigmentation and sporulation. If sporulation still<br />

fails subculture the fungus onto Trichophyton agar No.1.<br />

Confirmatory Tests:<br />

Trichophyton rubrum (Castellani) Semon<br />

Littman Oxgall Agar: Raised, greyish-white, suede-like to downy colony with no<br />

reverse pigment. Some cultures may show a greenish-yellow diffusible pigment.<br />

Lactritmel Agar: Flat, white to rose pink, downy to granular colonies with a deep winered<br />

reverse pigment.<br />

Sabouraud’s Dextrose Agar with 5% salt: Very stunted, white to cream, downy to<br />

glabrous colony with a pale yellow-brown reverse pigment.<br />

1% Peptone Agar: Flat, white to cream, downy to glabrous colony with no reverse<br />

pigment.<br />

BCP Agar: restricted colony growth and neutral (unchanged) pH. Colonies typically<br />

demonstrate red pigment on reverse<br />

Hydrolysis of Urea: Typically, negative at 7 days (some may be positive).<br />

Vitamin Free Agar (Trichophyton Agar No.1): Good growth indicating no special<br />

vitamin requirements. Colonies are flat, white to cream, suede-like to downy with a<br />

deep wine-red reverse pigment.<br />

Hair Perforation Test: Negative at 28 days.<br />

Key Features: Include clinical history, culture characteristics, microscopic morphology<br />

and failure to perforate hair in vitro.


Descriptions of Medical Fungi 209<br />

Trichophyton rubrum (Castellani) Semon<br />

a<br />

b<br />

Note: Trichophyton rubrum produces both red and yellow pigments. Culture colours<br />

may range from none to dark red to dark yellow, with all combinations in between. The<br />

images above show the same strain (a) grown on lactritmel agar that promotes red<br />

pigmentation and (b) on mycobiotic agar that shows an underlying yellow pigmentation.<br />

20 µm<br />

20 µm<br />

Trichophyton rubrum showing typical slender clavate microconidia.


210<br />

Descriptions of Medical Fungi<br />

Trichophyton rubrum (Castellani) Semon<br />

20 µm<br />

20 µm<br />

Trichophyton rubrum showing slender clavate microconidia and<br />

cigar-shaped macroconidia, some with terminal appendages.


Descriptions of Medical Fungi 211<br />

Trichophyton schoenleinii (Lebert) Langeron & Milochevitch<br />

Trichophyton schoenleinii is an anthropophilic fungus causing favus in humans. Favus<br />

is a chronic, scarring form of tinea capitis characterised by saucer-shaped crusted<br />

lesions or scutula and permanent hair loss. Invaded hairs remain intact and fluoresce<br />

a pale greenish-yellow under Wood’s ultra-violet light. Favus was once common in<br />

Eurasia and North Africa, however its incidence is now in decline.<br />

RG-2 organism.<br />

Morphological Description: Colonies are slow growing, waxy or suede-like with a<br />

deeply folded honey-comb-like thallus and some subsurface growth. The thallus is<br />

cream-coloured to yellow to orange brown. Cultures are difficult to maintain in their<br />

typical convoluted form, and rapidly become flat and downy. No reverse pigmentation is<br />

present. No macroconidia and microconidia are seen in routine cultures, and numerous<br />

chlamydospores may be present in older cultures. Characteristic antler “nail head”<br />

hyphae, also known as “favic chandeliers”, may be observed. A few distorted clavate<br />

microconidia may be formed by some isolates when grown on polished rice grains.<br />

Key Features: Clinical history, culture characteristics and microscopic morphology<br />

showing “favic chandeliers”.<br />

a<br />

b<br />

20 µm<br />

Trichophyton schoenleinii (a) culture and (b) “favic chandeliers”.


212<br />

Descriptions of Medical Fungi<br />

Trichophyton soudanense Joyeux.<br />

Trichophyton soudanense is an anthropophilic fungus that is a frequent cause of<br />

tinea capitis in Africa. Invaded hairs show an endothrix infection but do not fluoresce<br />

under Wood’s ultra-violet light. Distribution is mainly in Africa with imported cases now<br />

reported from Europe, Brazil, Australia and USA.<br />

RG-2 organism.<br />

Morphological Description: Colonies are slow-growing with a flat to folded, suedelike<br />

surface. Often there is a broad fringe of submerged growth. Surface mycelium and<br />

reverse pigment are characteristically a deep apricot-orange in colour. Microscopically,<br />

the hyphae often show reflexive or right-angle branching. Pyriform microconidia may<br />

occasionally be present and numerous chlamydospores are often found in older<br />

cultures.<br />

Note: T. soudanense demonstrates profuse growth and alkalinity on BCP milk solids<br />

agar. A thin halo of clearing usually appears in the milk solids around the colony edge<br />

at 7-10 days.<br />

Key Features: clinical history, culture characteristics and microscopic morphology<br />

showing reflexive hyphal branching and endothrix invasion of hair.<br />

a<br />

b<br />

20 µm<br />

Trichophyton soudanense (a) culture and (b) “reflex” hyphal branching.


Descriptions of Medical Fungi 213<br />

Trichophyton tonsurans is an anthropophilic fungus with a worldwide distribution which<br />

causes inflammatory or chronic non-inflammatory finely scaling lesions of skin, nails<br />

and scalp. It is a common cause of tinea capitis in Australian Aborigines and African<br />

Americans. Invaded hairs show an endothrix infection and do not fluoresce under<br />

Wood’s ultra-violet light.<br />

RG-2 organism.<br />

Morphological Description: Colonies show considerable variation in texture and<br />

colour. They may be suede-like to powdery, flat with a raised centre or folded, often<br />

with radial grooves. The colour may vary from pale-buff to yellow, (the sulfureum form<br />

which resembles Epidermophyton floccosum), to dark-brown. The reverse colour<br />

varies from yellow-brown to reddish-brown to deep mahogany. Hyphae are relatively<br />

broad, irregular, much branched with numerous septa. Numerous characteristic<br />

microconidia varying in size and shape from long clavate to broad pyriform, are borne<br />

at right angles to the hyphae, which often remain unstained by lactophenol cotton blue.<br />

Very occasional smooth, thin-walled, irregular, clavate macroconidia may be present<br />

on some cultures. Numerous swollen giant forms of microconidia and chlamydospores<br />

are produced in older cultures.<br />

Confirmatory Tests:<br />

Mycosel Agar: Chlamydospore-like structures produced by 5 days is characteristic of<br />

T. tonsurans (Mochizuki et al. 2013).<br />

Hydrolysis of Urea: positive at 5 days.<br />

Trichophyton tonsurans Malmsten<br />

T1<br />

T4<br />

a<br />

b<br />

Nutritional Tests on Trichophyton Agars: results demonstrate a partial requirement<br />

for (a) thiamine. T1 = vitamin free agar, (b) T4 = vitamin free + thiamine agar.<br />

Hair Perforation Test: Positive within 14 days.<br />

Key Features: microscopic morphology, culture characteristics, endothrix invasion of<br />

hairs and partial thiamine requirement.<br />

References: Rebell and Taplin (1970), Ajello (1972), Vanbreusegham et al. (1978),<br />

Rippon (1988), McGinnis (1980), Domsch et al. (1980), Kane et al. (1997) and de<br />

Hoog et al. (2000, 2015).


214<br />

Descriptions of Medical Fungi<br />

Trichophyton tonsurans Malmsten<br />

Trichophyton tonsurans cultures show considerable variation in texture and colour. The<br />

colour may vary from pale-buff to yellow to dark-brown. The reverse colour varies from<br />

yellow-brown to reddish-brown to deep mahogany.<br />

b<br />

a<br />

20 µm<br />

Trichophyton tonsurans (a) hyphae, microconidia and (b) macroconidia.


Descriptions of Medical Fungi 215<br />

Trichophyton verrucosum is a zoophilic fungus causing ringworm in cattle. Infections in<br />

humans result from direct contact with cattle or infected fomites and are usually highly<br />

inflammatory involving the scalp, beard or exposed areas of the body. Invaded hairs<br />

show an ectothrix infection and fluorescence under Wood’s ultra-violet light has been<br />

noted in cattle but not in humans. Distribution is worldwide. RG-2 organism.<br />

Morphological Description: Colonies are slow growing, small, button or disk-shaped,<br />

white to cream-coloured, with a suede-like to velvety surface, a raised centre, and<br />

flat periphery with some submerged growth. Reverse pigment may vary from nonpigmented<br />

to yellow. Broad, irregular hyphae with many terminal and intercalary<br />

chlamydospores. Chlamydospores are often in chains. The tips of some hyphae are<br />

broad and club-shaped, and occasionally divided, giving the so-called “antler” effect.<br />

When grown on thiamine-enriched media, occasional strains produce clavate to<br />

pyriform microconidia borne singly along the hyphae. Macroconidia are only rarely<br />

produced, but when present have a characteristic tail or string bean shape.<br />

Confirmatory Tests:<br />

Trichophyton verrucosum Bodin<br />

Growth at 37 O C: Unlike other dermatophytes, growth is enhanced at 37 O C.<br />

Nutritional Requirements: all strains require thiamine and approximately 80% require<br />

thiamine and inositol. There is no growth on casein vitamin free agar (T1), minimal<br />

submerged growth on T1 + inositol (T2), good growth on T1 + inositol and thiamine<br />

(T3) and good growth on T1 + thiamine only (T4).<br />

All strains produce typical chains of chlamydospores, often referred to as “chains of<br />

pearls”, especially when grown on BCP milk solids glucose agar at 37 O C. When grown<br />

at 25 O C on milk solids glucose agar a “halo”-like zone of peripheral clearing of milk<br />

solids occurs within 7 days.<br />

Microscopic examination of young colonies (4 to 5 day old), grown from a very small<br />

inoculum, on Sabouraud’s dextrose agar containing 0.5% yeast extract and incubated<br />

at 30 O C, show characteristic terminal vesicles (not chlamydospores) at the tips of<br />

hyphae. The number of vesicles produced is greater from primary inoculations of skin<br />

scrapings or hairs than from subcultures.<br />

Key Features: Culture characteristics and requirements for thiamine and inositol, large<br />

ectothrix invasion of hair, clinical lesions and history.<br />

a<br />

b<br />

Trichophyton verrucosum (a) young button-shaped colonies and (b) mature culture.


216<br />

Descriptions of Medical Fungi<br />

Trichophyton verrucosum Bodin<br />

a<br />

20 µm<br />

b<br />

20 µm<br />

20 µm<br />

c<br />

30 µm<br />

d<br />

Trichophyton verrucosum showing (a) clavate to pyriform microconidia, (b)<br />

characteristic rat tail or string bean-shaped macroconidia, (c) terminal vesicles<br />

at the tips of hyphae in young colonies and (d) chains of chlamydospores.


Descriptions of Medical Fungi 217<br />

Synonymy: Trichophyton yaoundei Cochet & Doby-Dubois.<br />

All varieties of T. violaceum.<br />

Trichophyton violaceum is an anthropophilic fungus causing inflammatory or chronic<br />

non-inflammatory finely scaling lesions of skin, nails, beard and scalp, producing the<br />

so-called “black dot” tinea capitis. Distribution is worldwide, particularly in the Near<br />

East, Eastern Europe, Russia and North Africa. Invaded hairs show an endothrix<br />

infection and do not fluoresce under Wood’s ultra-violet light.<br />

RG-2 organism.<br />

Trichophyton violaceum Sabouraud apud Bodin<br />

Morphological Description: Colonies are very slow growing, glabrous or waxy, heaped<br />

and folded and deep violet in colour. Cultures often become pleomorphic, forming<br />

white sectors. Occasional non-pigmented strains may occur. Hyphae are relatively<br />

broad, tortuous, much branched and distorted. Young hyphae usually stain well in<br />

lactophenol cotton blue, whereas older hyphae stain poorly and show small central fat<br />

globules and granules. Typically, no conidia are present, although occasional pyriform<br />

microconidia have been observed on enriched media. Numerous chlamydospores are<br />

usually present, especially in older cultures.<br />

Nutritional Requirements: T. violaceum has a partial nutrient requirement for thiamine.<br />

There is minimal growth on casein vitamin-free agar (T1), and slightly better growth on<br />

vitamin-free agar plus thiamine (T4). The partial requirement for thiamine separates<br />

this organism from T. rubrum, and other species that may produce purple pigmented<br />

colonies.<br />

Key Features: Culture characteristics, partial thiamine requirement and endothrix hair<br />

invasion.<br />

a<br />

b<br />

20 µm<br />

Trichophyton violaceum (a) culture and (b) chlamydospores.


218<br />

Descriptions of Medical Fungi<br />

Trichosporon Behrend<br />

Trichosporon species are urease-positive, non-encapsulated basidiomycetous yeasts<br />

characterised by the development of hyaline, septate hyphae that fragment into oval or<br />

rectangular arthroconidia. Some blastoconidia are also seen. The colonies are usually<br />

raised and have a waxy appearance, which develop radial furrows and irregular folds.<br />

They are widely distributed in the environment and many have different habitats, usually<br />

occupying narrow ecological niches. Some are soil borne and others are associated<br />

with humans and animals (Colombo et al. 2011, Sugita 2011, Arendrup et al. 2014).<br />

The genus has undergone major taxonomic revision (Gueho et al. 1992, de Hoog et<br />

al. 2000, Rodriguez-Tudela et al. 2005). In particular, the name Trichosporon beigelii is<br />

now obsolete, and previously described infections reported in the literature under this<br />

name could in fact be due to any one of the species listed below.<br />

Six species are of clinical significance: T. asahii, T. asteroides, T. cutaneum, T.<br />

inkin, T. mucoides and T. ovoides. Other species reported from human and animal<br />

infections include T. dermatis, T. domesticum, T. faecale, T. jirovecii, T. loubieri and T.<br />

mycotoxinovorans (Rodriguez-Tudela et al. 2005, Chagas-Neto et al. 2008, Colombo<br />

et al. 2011).<br />

Trichosporon species are a minor component of normal skin flora, and are widely<br />

distributed in nature. They are regularly associated with the soft nodules of white piedra,<br />

and have been involved in a variety of opportunistic infections in the immunosuppressed<br />

patient. Disseminated infections are most frequently (75%) caused by T. asahii<br />

(Arendrup et al. 2014) and have been associated with leukaemia, organ transplantation,<br />

multiple myeloma, aplastic anaemia, lymphoma, solid tumours and AIDS. Disseminated<br />

infections are often fulminate and widespread, with lesions occurring in the liver, spleen,<br />

lungs and gastrointestinal tract. Infections in non-immunosuppressed patients include<br />

endophthalmitis after surgical extraction of cataracts, endocarditis usually following<br />

insertion of prosthetic cardiac valves, peritonitis in patients on continuous ambulatory<br />

peritoneal dialysis (CAPD), and intravenous drug abuse.<br />

Note: Genus identification is mandatory for clinical management and should be<br />

performed and provided in a timely manner. Species identification remains difficult and<br />

requires molecular analysis or MALDI-T<strong>OF</strong> MS (with an extensive database) (Arendrup<br />

et al. 2014).<br />

Molecular Identification: ITS and D1/D2 sequencing is required for accurate species<br />

identification (Arendrup et al. 2014).<br />

MALDI-T<strong>OF</strong> MS: A promising identification tool to accurately identify species (with an<br />

extensive database) (Kolecka et al. 2013).<br />

Comment: The API 20C yeast identification system is recommended for sugar<br />

assimilation tests.<br />

References: Kurtzman and Fell (1988), Gueho et al. (1992), de Hoog et al. (2000,<br />

2015), Rodriguez-Tudela et al. (2005), Chagas-Neto et al. (2008), Guo et al. (2011),<br />

Xiao et al. (2013).


Descriptions of Medical Fungi 219<br />

Trichosporon Behrend<br />

Key to medically important species (de Hoog et al. 2000).<br />

1. Growth with melibiose 2<br />

No growth with melibiose 3<br />

2. Tolerant to cycloheximide T. mucoides<br />

Not tolerant to cycloheximide<br />

T. cutaneum<br />

3. Growth with myo-inositol, no growth with L-arabinose T. inkin<br />

No growth with myo-inositol, growth with L-arabinose 4<br />

4. Colony with very slow growth; thallus consisting of clumps<br />

of meristematic cells<br />

T. asteroides<br />

Colonies and microscopy otherwise 5<br />

5. Appressoria present in slide cultures T. ovoides<br />

Appressoria absent in slide cultures 6<br />

6. Arthroconidia barrel-shaped; thallus not meristematic T. asahii<br />

Arthroconidia elongate, or thallus meristematic<br />

T. asteroides<br />

Trichosporon asahii Akagi ex Sugita et al.<br />

RG-2 organism.<br />

Morphological Description: Colonies are white to cream-coloured, powdery, suedelike<br />

to farinose with radial furrows and irregular folds. Budding cells and lateral conidia<br />

are absent. Arthroconidia are barrel-shaped. Appressoria absent. This species<br />

assimilates L-arabinose but not melibiose. Growth at 37 O C. Most common species,<br />

especially from invasive infections.<br />

Assimilation Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Glucose + Melibiose - L-Rhamnose + D-Glucitol v<br />

Galactose + Raffinose - D-Glucosamine + α-M-D-glucoside +<br />

L-Sorbose v Melezitose v N-A-D-glucosamine + D-Gluconate +<br />

Sucrose v Soluble Starch v Glycerol v DL-Lactate v<br />

Maltose + D-Xylose v Erythritol + myo-Inositol v<br />

Cellobiose + L-Arabinose + Ribitol v Nitrate -<br />

Trehalose + D-Arabinose + Galactitol - 2-K-D-gluconate +<br />

Lactose + D-Ribose + D-Mannitol v D-Glucuronate +


220<br />

Descriptions of Medical Fungi<br />

Trichosporon asahii Akagi ex Sugita et al.<br />

a<br />

b<br />

20 µm<br />

Trichosporon asahii (a) culture and (b) hyphae and arthroconidia.<br />

Antifungal Susceptibility: T. asahii (Australian National data); MIC µg/mL.<br />

No. 64<br />

AmB 22 4 4 6 4 3 1<br />

FLU 22 1 3 11 7<br />

VORI 20 1 1 7 10 1<br />

POSA 19 1 1 11 4 2<br />

ITRA 22 3 15 3 1<br />

Note: Non-T. asahii isolates appear to be more susceptible than T. asahii isolates to<br />

AmB, FLU, and ITRA, while the new triazoles are active against both T. asahii and<br />

non-T. asahii isolates (Rodriguez-Tudela et al. 2005).<br />

Trichosporon inkin. Colonies are<br />

restricted, white, finely cerebriform with<br />

a granular covering, without marginal<br />

zone, often cracking the media.


Descriptions of Medical Fungi 221<br />

Trichosporon asteroides (Rischin) Ota<br />

Morphological Description: Colonies are restricted, dry, cream-coloured, cerebriform,<br />

with radial furrows and irregular folds. The meristematic form is punctiform, brownish<br />

and consists of hyphae which swell and become multiseptate which may fall apart<br />

into smaller packets. Budding cells and lateral conidia are absent. Arthroconidia are<br />

elongate and hyphae are often present. Appressoria absent. This species assimilates<br />

L-arabinose but not myo-inositol. Growth at 37 O C is variable. Uncommon species<br />

usually associated with superficial infections. RG-2 organism.<br />

Assimilation Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Glucose + Melibiose - L-Rhamnose + D-Glucitol v<br />

Galactose + Raffinose - D-Glucosamine v α-M-D-glucoside +<br />

L-Sorbose v Melezitose + N-A-D-glucosamine + D-Gluconate +<br />

Sucrose + Soluble Starch + Glycerol + DL-Lactate +<br />

Maltose + D-Xylose + Erythritol + myo-Inositol -<br />

Cellobiose + L-Arabinose + Ribitol v Nitrate -<br />

Trehalose + D-Arabinose + Galactitol - 2-K-D-gluconate +<br />

Lactose + D-Ribose + D-Mannitol v D-Glucuronate +<br />

Trichosporon cutaneum (de Beurmann et al.) Ota<br />

Morphological Description: Colonies are cream-coloured, cerebriform, glabrous, with<br />

radial furrows and irregular folds. Budding cells abundant in primary cultures; hyphae<br />

developing in older cultures. Arthroconidia are cylindrical to ellipsoidal. Appressoria<br />

absent. This species assimilates melibiose; not tolerant to 0.1% (variable tolerance<br />

to 0.01%) cycloheximide. No growth at 37 O C. Uncommon species usually associated<br />

with superficial infections. RG-2 organism.<br />

Assimilation Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Glucose + Melibiose + L-Rhamnose + D-Glucitol +<br />

Galactose + Raffinose + D-Glucosamine v α-M-D-glucoside +<br />

L-Sorbose v Melezitose + N-A-D-glucosamine + D-Gluconate +<br />

Sucrose + Soluble Starch + Glycerol + DL-Lactate +<br />

Maltose + D-Xylose + Erythritol + myo-Inositol +<br />

Cellobiose + L-Arabinose + Ribitol + Nitrate -<br />

Trehalose + D-Arabinose v Galactitol - 2-K-D-gluconate +<br />

Lactose + D-Ribose + D-Mannitol + D-Glucuronate +<br />

Trichosporon inkin (Oho ex Ota) do Carmo-Sousa & van Uden<br />

Morphological Description: Colonies are restricted, white, finely cerebriform with a<br />

granular covering, without marginal zone, often cracking the media. Budding cells and<br />

lateral conidia absent. Arthroconidia are long cylindrical. Appressoria present in slide<br />

cultures. Sarcinae present on media with high sugar-content. This species assimilates<br />

myo-inositol but not melibiose and is tolerant to 0.01% (variable tolerance to 0.1%)<br />

cycloheximide. Growth at 37 O C. Usually associated with white piedra on pubic hairs.<br />

RG-2 organism.


222<br />

Descriptions of Medical Fungi<br />

Trichosporon inkin (Oho ex Ota) do Carmo-Sousa & van Uden<br />

Assimilation Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Glucose + Melibiose - L-Rhamnose - D-Glucitol -<br />

Galactose v Raffinose - D-Glucosamine v α-M-D-glucoside +<br />

L-Sorbose v Melezitose + N-A-D-glucosamine + D-Gluconate +<br />

Sucrose + Soluble Starch + Glycerol v DL-Lactate +<br />

Maltose + D-Xylose + Erythritol + myo-Inositol +<br />

Cellobiose + L-Arabinose v Ribitol - Nitrate -<br />

Trehalose + D-Arabinose v Galactitol - 2-K-D-gluconate +<br />

Lactose + D-Ribose + D-Mannitol v D-Glucuronate +<br />

Trichosporon mucoides Gueho & M.Th. Smith<br />

Morphological Description: Colonies are moist and glabrous, white, cerebriform,<br />

heaped and folded. Budding cells present in primary cultures. Broadly clavate, terminal<br />

or lateral blastoconidia often present, becoming thick-walled with age. Arthroconidia<br />

are barrel-shaped. Appressoria absent. This species assimilates melibiose and is<br />

tolerant to 0.01% (variable tolerance to 0.1%) cycloheximide. Growth at 37 O C. Common<br />

species associated with superficial infections, white piedra and onychomycosis.<br />

RG-2 organism.<br />

Assimilation Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Glucose + Melibiose + L-Rhamnose + D-Glucitol +<br />

Galactose + Raffinose + D-Glucosamine + α-M-D-glucoside +<br />

L-Sorbose + Melezitose + N-A-D-glucosamine + D-Gluconate +<br />

Sucrose + Soluble Starch + Glycerol + DL-Lactate +<br />

Maltose + D-Xylose + Erythritol + myo-Inositol +<br />

Cellobiose + L-Arabinose + Ribitol + Nitrate -<br />

Trehalose + D-Arabinose + Galactitol + 2-K-D-gluconate +<br />

Lactose + D-Ribose + D-Mannitol + D-Glucuronate +<br />

Trichosporon ovoides Behrend<br />

Morphological Description: Colonies are restricted, white, granular, folded at the<br />

centre, with a flat marginal zone. Budding cells and lateral conidia absent. Arthroconidia<br />

are cylindrical. Appressoria present in slide cultures. This species does not assimilate<br />

melibiose, but tolerates 0.01% cycloheximide. Growth at 37 O C is variable. Uncommon<br />

species usually associated with superficial infections, like white piedra. RG-2 organism.<br />

Assimilation Tests: + Positive, - Negative, v Variable, w Weak, s Slow<br />

Glucose + Melibiose - L-Rhamnose + D-Glucitol v<br />

Galactose + Raffinose v D-Glucosamine v α-M-D-glucoside +<br />

L-Sorbose v Melezitose v N-A-D-glucosamine + D-Gluconate +<br />

Sucrose + Soluble Starch + Glycerol v DL-Lactate +<br />

Maltose + D-Xylose + Erythritol + myo-Inositol +<br />

Cellobiose + L-Arabinose v Ribitol - Nitrate -<br />

Trehalose v D-Arabinose v Galactitol - 2-K-D-gluconate +<br />

Lactose + D-Ribose + D-Mannitol + D-Glucuronate +


Descriptions of Medical Fungi 223<br />

Trichothecium roseum has a worldwide distribution and is often isolated from decaying<br />

plant substrates, soil, seeds of corn, and food-stuffs (especially flour products). It is<br />

occasionally isolated as a saprophyte in the clinical laboratory.<br />

RG-1 organism.<br />

Trichothecium roseum (Persoon) Link<br />

Morphological Description: Colonies are moderately fast growing, flat, suede-like to<br />

powdery, initially white but becoming rosy, pink or orange with age. The conidiophores<br />

are indistinguishable from the vegetative hyphae until the first conidium is produced.<br />

They are erect, unbranched, often septate near the base, more or less roughwalled,<br />

bearing basipetal zig-zag (alternating) chains of conidia at the apex. Note:<br />

The conidiophore is progressively shortened with the formation of each conidium i.e.<br />

retrogressive conidial development. Conidia are two-celled ellipsoidal to pyriform, with<br />

an obliquely truncate basal scar, hyaline, smooth to delicately roughened and thickwalled.<br />

Comment: T. roseum should not be confused with Nannizzia nanum. Colonies of the<br />

latter may be pinkish-buff in colour and also produce ovoid to pear-shaped, mostly twocelled<br />

macroconidia with thin, verrucose walls. However, N. nanum usually produces<br />

a red-brown reverse pigment and the two-celled macroconidia are sessile and formed<br />

singly, they are not produced in basipetal chains as in T. roseum.<br />

Molecular Identification: Summerbell et al. (2011) revised the genus using D1/D2<br />

sequences for phylogenetic analysis and sequence based identification.<br />

Key Features: Hyphomycete, basipetal zig-zag chains of two-celled conidia showing<br />

retrogressive development where the conidiophore becomes progressively shorter.<br />

References: McGinnis (1980), Domsch et al. (2007), Rippon (1988), Samson et al.<br />

(1995), Summerbell et al. (2011).<br />

20 µm<br />

Trichothecium roseum conidiophores showing retrogressive conidial development.


224<br />

Descriptions of Medical Fungi<br />

Species of Ulocladium should not be confused with other poroconidial genera such as<br />

Stemphylium, Alternaria, Bipolaris, Exserohilum, Dreschlera and Curvularia. A human<br />

case of keratitis has been reported (Badenoch et al. 2006).<br />

RG-1 organism.<br />

Ulocladium Preuss<br />

Morphological Description: Colonies are rapid growing, brown to olivaceousblack<br />

or greyish and suede-like to floccose. Microscopically, numerous, usually<br />

solitary, multicelled conidia (dictyoconidia) are formed through a pore (poroconidia)<br />

by a sympodially elongating geniculate conidiophore. Conidia are typically obovoid<br />

(narrowest at the base), dark brown and often rough-walled. Seven species have been<br />

described, all being saprophytes.<br />

Molecular Identification: ITS sequencing is sufficient for genus identification<br />

(Badenoch et al. 2006) and Woudenberg et al. 2013).<br />

References: Ellis (1970, 1976), Domsch et al. (1980), Rippon (1988), Samson et al.<br />

(1995), de Hoog et al. (2000).<br />

10 µm<br />

Ulocladium spp. conidiophores and conidia.<br />

Antifungal Susceptibility: Ulocladium sp. very limited data (Pujol et al. 2000,<br />

Australian National data); MIC µg/mL.<br />

Antifungal Range Antifungal Range Antifungal Range<br />

AmB 1->16 VORI 0.25 ITRA 0.06->16


Descriptions of Medical Fungi 225<br />

Veronaea botryosa Ciferri & Montemartini<br />

This genus is very similar to Rhinocladiella, however the conidia are typically twocelled.<br />

Occasional skin infections have been reported from humans (Revankar and<br />

Sutton 2010).<br />

RG-2 organism.<br />

Morphological Description: Colonies grow rapidly and are suede-like to downy,<br />

greyish-brown to blackish-brown. Conidiophores are erect, straight or flexuose,<br />

occasionally branched and are usually geniculate, due to the sympodial development of<br />

the conidia. They are smooth-walled, pale to medium olivaceous-brown, up to 250 µm<br />

long and 2-4 µm wide. Conidia are pale brown, two-celled, cylindrical with a truncated<br />

base, smooth-walled or slightly verrucose, 5-12 x 3-4 µm.<br />

Molecular Identification: Arzanlou et al. (2007) used D1/D2 and ITS sequence data<br />

in a phylogenetic revision.<br />

References: Ellis (1971), de Hoog et al. (2000, 2015), Revankar and Sutton (2010).<br />

10 µm<br />

Veronaea botryosa conidiophores and conidia.<br />

Antifungal Susceptibility: V. botryosa limited data (Badali et al. 2013); MIC µg/mL.<br />

Antifungal Range MIC 90<br />

Antifungal Range MIC 90<br />

AmB 8-16 16 POSA 0.03-0.25 0.25<br />

ITRA 0.25-1 1 VORI 1-8 4


226<br />

Descriptions of Medical Fungi<br />

Verruconis gallopava (W.B. Cooke) Samerpitak & de Hoog<br />

Synonymy: Ochroconis gallopava (W.B. Cooke) de Hoog.<br />

Verruconis species are thermophilic, with Verruconis gallopava occurring in hot<br />

environments, such as thermal soils, broiler house litter, hot springs, and self-heated<br />

waste (Samerpitak et al. 2014). V. gallopava is neurotropic and is a recognised agent<br />

of human brain infections and is responsible for encephalitis in poultry and wild birds<br />

dogs and cats (Seyedmousavi et al. 2014). Occasional human pulmonary infections<br />

in immunocompetent hosts have also been reported (Samerpitak et al. 2014,<br />

Seyedmousavi et al. 2014, Giraldo et al. 2014).<br />

RG-2 organism.<br />

Morphological Description: Colonies are smooth to suede-like, dry, flat, tobaccobrown<br />

to brownish-black with a dark brown diffusible pigment. Hyphae are brown with<br />

relatively thick walls. Conidiophores are mostly cylindrical to acicular, sometimes poorly<br />

differentiated, bearing a few conidia at the tip. Conidia are two-celled, subhyaline to pale<br />

brown, smooth-walled to verrucose, cylindrical to clavate, constricted at the septum,<br />

11-18 x 2.5-4.5 µm in size, with the apical cell wider than the basal cell. A remnant of<br />

a denticle may also be seen at the conidial base. Optimum growth at 35 O C, tolerant to<br />

40 O C.<br />

Molecular Identification: ITS sequencing can identify species. Additional genes<br />

include β-tubulin, actin, and the D1/D2 region (Giraldo et al. 2014, Seyedmousavi et<br />

al. 2014).<br />

References: Domsch et al. (1980), McGinnis (1980), de Hoog et al. (2000, 2015),<br />

Samerpitak et al. (2014), Seyedmousavi et al. (2014) and Giraldo et al. (2014).<br />

Antifungal Susceptibility: V. gallopava limited data (Australian National data); MIC<br />

µg/mL.<br />

No. 64<br />

AmB 8 1 2 2 1 2<br />

VORI 8 1 3 3 1<br />

POSA 7 1 2 1 1 1 1<br />

ITRA 8 3 1 3 1<br />

V. gallopava data for 11 isolates (Seyedmousavi et al. 2014); MIC µg/mL.<br />

AmB Range 0.125-4; MIC 90<br />

= 0.5 VORI Range 0.5-2; MIC 90<br />

= 2<br />

ITRA Range 0.016-4; MIC 90<br />

= 0.5 POSA Range


Descriptions of Medical Fungi 227<br />

Verruconis gallopava (W.B. Cooke) Samerpitak & de Hoog<br />

a<br />

10 µm<br />

b<br />

10 µm<br />

Verruconis gallopava (a) culture and (b) hyphae, conidiophores and conidia.


228<br />

Descriptions of Medical Fungi<br />

Members of this genus are often isolated from the environment. It has been reported<br />

as a rare agent of mycotic keratitis.<br />

RG-1 organism.<br />

Verticillium Nees ex Link<br />

Morphological Description: Colonies are fast growing, suede-like to downy, white to<br />

pale yellow in colour, becoming pinkish-brown, red, green or yellow with a colourless,<br />

yellow or reddish-brown reverse. Conidiophores are usually well differentiated and<br />

erect, verticillately branched over most of their length, bearing whorls of slender awlshaped<br />

divergent phialides. Conidia are hyaline or brightly-coloured, mostly one-celled,<br />

and are usually borne in slimy heads (glioconidia).<br />

Molecular Identification: ITS, actin, EF-1α, GPDH and tryptophan synthase genes<br />

have been used to identify all recognised Verticillium species (Inderbitzin et al. 2013).<br />

Key Features: Hyphomycete, verticillate branched conidiophores bearing whorls of<br />

awl-shaped, divergent phialides.<br />

References: Domsch et al. (1980), McGinnis (1980), Rippon (1988), Samson et al.<br />

(1995), de Hoog et al. (2000, 2015).<br />

20 µm<br />

Verticillium spp conidiophores, phialides and conidia.


Descriptions of Medical Fungi 229<br />

Wickerhamomyces anomalus (E.C. Hansen) Kurtzman et al.<br />

Synonymy: Candida pelliculosa Redaelli.<br />

Wickerhamomyces anomalus has been reported from cases of candidaemia and<br />

catheter related infections in humans and has been isolated from soil, grains, fruit and<br />

warm-blooded animals.<br />

RG-1 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Spherical to ellipsoidal budding blastoconidia, 2-4 x 2-6 µm. Pseudohyphae<br />

may be present. Asci when present, containing one to four hat-shaped ascospores.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Spherical to ellipsoidal budding yeast cells and abundant<br />

pseudohyphae in most strains.<br />

Molecular Identification: ITS sequencing recommended.<br />

MALDI-T<strong>OF</strong> MS: Able to accurately identify this species.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose - L-Arabinose v D-Glucitol +<br />

Fermentation Sucrose + D-Arabinose - α-M-D-glucoside +<br />

Glucose + Maltose + D-Ribose v D-Gluconate v<br />

Galactose v Cellobiose + L-Rhamnose - DL-Lactate +<br />

Sucrose + Trehalose + D-Glucosamine - myo-Inositol -<br />

Maltose v Lactose - N-A-D-glucosamine - 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate nd<br />

Trehalose - Raffinose + Erythritol + Nitrate +<br />

Assimilation Melezitose + Ribitol v Urease -<br />

Glucose + Soluble Starch + Galactitol - 0.1% Cycloheximide -<br />

Galactose v D-Xylose v D-Mannitol + Growth at 37 O C v<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern<br />

Antifungal Susceptibility: W. anomalus (Diekema et al. 2009, Australian National<br />

data); MIC µg/mL.<br />

No. 64<br />

AmB 42 1 5 14 21 2<br />

FLU 43 2 8 25 8<br />

VORI 42 1 2 1 22 13 3<br />

POSA 42 1 1 7 4 15 12 2<br />

ITRA 3 3<br />

ANID 16 2 10 3 1<br />

MICA 16 5 9 2<br />

CAS 39 1 16 17 4 1<br />

5FC 3 1 2


230<br />

Descriptions of Medical Fungi<br />

Yarrowia lipolytica (Wickerham et al.) van der Walt & von Arx.<br />

Synonymy: Candida lipolytica (F.C. Harrison) Diddens & Lodder.<br />

Yarrowia lipolytica is a rare cause of candidaemia.<br />

RG-1 organism.<br />

Culture: Colonies (SDA) white to cream-coloured smooth, glabrous, yeast-like.<br />

Microscopy: Spherical, ellipsoidal to elongate budding blastoconidia, 3-5 x 3-15 µm.<br />

India Ink Preparation: Negative - no capsules present.<br />

Dalmau Plate Culture: Pseudohyphae and true hyphae are produced.<br />

Molecular Identification: ITS sequencing recommended.<br />

MALDI-T<strong>OF</strong> MS: Able to accurately identify this species.<br />

Physiological Tests: + Positive, - Negative, v Variable, w Weak, s Slow, nd No Data<br />

Germ Tube - L-Sorbose v L-Arabinose - D-Glucitol +<br />

Fermentation Sucrose - D-Arabinose - α-M-D-glucoside -<br />

Glucose - Maltose - D-Ribose v D-Gluconate v<br />

Galactose - Cellobiose w,- L-Rhamnose - DL-Lactate +<br />

Sucrose - Trehalose - D-Glucosamine - myo-Inositol -<br />

Maltose - Lactose - N-A-D-glucosamine + 2-K-D-gluconate -<br />

Lactose - Melibiose - Glycerol + D-Glucuronate nd<br />

Trehalose - Raffinose - Erythritol + Nitrate -<br />

Assimilation Melezitose - Ribitol v Urease -<br />

Glucose + Soluble Starch - Galactitol - 0.1% Cycloheximide -<br />

Galactose v D-Xylose - D-Mannitol + Growth at 37 O C v<br />

Key Features: Germ tube negative yeast and sugar assimilation pattern.<br />

Antifungal Susceptibility: Y. lipolytica limited data (Diekema et al. 2009, Australian<br />

National data); MIC µg/mL.<br />

No. 64<br />

AmB 19 1 1 1 6 5 4 1<br />

FLU 19 1 1 8 6 2 1<br />

VORI 19 1 6 9 2 1<br />

POSA 19 2 2 9 5 1<br />

ITRA 3 1 1 1<br />

ANID 12 1 3 5 2 1<br />

MICA 12 7 3 2<br />

CAS 17 6 10 1<br />

5FC 3 1 1 1


Descriptions of Medical Fungi 231<br />

KOH with Calcofluor White.<br />

MICROSCOPY STAINS & TECHNIQUES<br />

For the direct microscopic examination of skin scrapings, hairs, nails and other<br />

clinical specimens for fungal elements. This is a very sensitive method, however, a<br />

fluorescence microscope with ultraviolet filters is required (Hageage and Harrington,<br />

1984; Hollander et al. 1984; Monheit et al. 1984; Harrington and Hageage 2003).<br />

Solution A: Potassium hydroxide reagent<br />

Potassium hydroxide<br />

Glycerine<br />

Distilled water<br />

10 g<br />

10 mL<br />

80 mL<br />

Solution B: Calcofluor white reagent<br />

Calcofluor white<br />

0.5 g<br />

Evans blue<br />

0.02 g<br />

Distilled water<br />

50 mL<br />

Mix one drop of each solution on the centre of a clean microscope slide.<br />

Place the specimen in the solution and cover with a coverslip.<br />

KOH with Chlorazol Black.<br />

For the direct microscopic examination of skin scrapings, hairs, nails and other clinical<br />

specimens for fungal elements. Note: Parker Quink ink is no longer available.<br />

Polysciences list this product as “Chlorazol black E” product number 02730-25.<br />

Potassium hydroxide<br />

10 g<br />

Chlorazol Black E (0.1%)<br />

10 mL<br />

Glycerol<br />

10 mL<br />

Distilled water<br />

80 mL<br />

Using sterile technique, remove a small portion of the specimen with an<br />

inoculation needle and mount in a drop of KOH on a clean microscope<br />

slide. Cover with a coverslip, squash the preparation with the butt of the<br />

inoculation needle and then blot off the excess fluid.<br />

India Ink Mounts.<br />

For the direct microscopic examination of CSF for Cryptococcus species. Place a drop<br />

of India Ink on the specimen, mix well with a sterile loop, and cover with a coverslip.<br />

Best brands to use are “Pelikan” or “Talons” India Ink.


232<br />

Descriptions of Medical Fungi<br />

MICROSCOPY STAINS & TECHNIQUES<br />

Lactophenol Cotton Blue (LPCB)<br />

For the staining and microscopic identification of fungi.<br />

Cotton Blue (Aniline Blue)<br />

0.05 g<br />

Phenol Crystals<br />

20 g<br />

Glycerol<br />

40 mL<br />

Lactic acid<br />

20 mL<br />

Distilled water<br />

20 mL<br />

This stain is prepared over two days.<br />

1. On the first day, dissolve the cotton blue in the distilled water. Leave<br />

overnight to eliminate insoluble dye.<br />

2. On the second day, wearing gloves, add the phenol crystals to the lactic<br />

acid in a glass beaker. Place on magnetic stirrer until the phenol is<br />

dissolved.<br />

3. Add the glycerol.<br />

4. Filter the cotton blue and distilled water solution into the phenol/glycerol/<br />

lactic acid solution. Mix and store at room temperature.<br />

Direct Microscopic Mounts or Squash Preparations<br />

Using sterile technique, remove a small portion of the colony with an inoculation needle<br />

and mount in a drop of lactophenol cotton blue on a clean microscope slide. Cover with<br />

a coverslip, squash the preparation with the butt of the inoculation needle and then blot<br />

off the excess fluid.<br />

Cellotape Flag Preparations<br />

An excellent technique for the rapid mounting of sporulating fungi because it keeps<br />

more of the reproductive structures intact.<br />

1. Using clear 2 cm wide cellotape and a wooden applicator stick (orange<br />

stick) make a small cellotape flag (2 x 2 cm).<br />

2. Using sterile technique, gently press the sticky side of the flag onto the<br />

surface of the culture.<br />

3. Remove and apply a drop of 95% alcohol to the flag, this acts as a<br />

wetting agent and also dissolves the adhesive glue holding the flag to<br />

the applicator stick.<br />

4. Place the flag onto a small drop of lactophenol cotton blue on a clean<br />

glass slide, remove the applicator stick and discard, add another drop of<br />

stain, cover with a coverslip, gently press and mop up any excess stain.


Descriptions of Medical Fungi 233<br />

Slide Culture Preparations<br />

MICROSCOPY STAINS & TECHNIQUES<br />

In order to accurately identify many fungi it is essential to observe the precise<br />

arrangement of the conidiophores and the way in which spores are produced (conidial<br />

ontogeny). Riddel’s simple method of slide culturing (Mycologia 1950; 42:265-270)<br />

permits fungi to be studied virtually in situ with as little disturbance as possible. A<br />

simple modification of this method using a single agar plate is described below.<br />

One plate of nutrient agar; potato dextrose is recommended, however,<br />

some fastidious fungi may require harsher media to induce sporulation like<br />

cornmeal agar or Czapek Dox agar.<br />

1. Using a sterile blade cut out an agar block (7 x 7 mm) small enough to<br />

fit under a coverslip.<br />

2. Flip the block up onto the surface of the agar plate.<br />

3. Inoculate the four sides of the agar block with spores or mycelial<br />

fragments of the fungus to be grown.<br />

4. Place a flamed coverslip centrally upon the agar block.<br />

5. Incubate the plate at 26 O C until growth and sporulation have occurred.<br />

6. Remove the cover slip from the agar block.<br />

7. Apply a drop of 95% alcohol as a wetting agent.<br />

8. Gently lower the coverslip onto a small drop of lactophenol cotton blue<br />

on a clean glass slide.<br />

9. The slide can be left overnight to dry and later sealed with fingernail<br />

polish.<br />

10. When sealing with nail polish use a coat of clear polish followed by one<br />

coat of red-coloured polish.<br />

Simple agar block method, inoculated on four sides with cover slip<br />

on top. Make at least two slides per culture.


234<br />

Descriptions of Medical Fungi<br />

Bird Seed Agar<br />

SPECIALISED CULTURE MEDIA<br />

For selective isolation of Cryptococcus neoformans and C. gattii.<br />

Guizotia abyssinica (niger seed) 50 g Glucose 1 g<br />

KH 2<br />

PO 4<br />

(potassium dihydrogen 1 g Creatinine 1 g<br />

orthophosphate)<br />

Bacto agar (BD 214010) 15 g Distilled water 1000 mL<br />

Penicillin G (20 units/mL) 1 mL Gentamicin (40 mg/mL) 1 mL<br />

1. Grind seeds of Guizotia abyssinica as finely as possible with an electric grinder<br />

and add to 1000 mL distilled water in a stainless steel jug.<br />

2. Boil for 30 minutes, then pass through filter paper and adjust volume to 1000<br />

mL.<br />

3. Add remaining ingredients (except Bacto agar) to filtrate and dissolve.<br />

If required: Cool to room temperature and adjust pH to 5.5.<br />

Dispense into 500 mL bottles.<br />

4. Add 7.5 g Bacto agar to each 500 mL bottle.<br />

5. Autoclave 121 O C for 20 minutes.<br />

6. Cool to 48 O C and add 0.5 mL Penicillin G and 0.5 mL Gentamicin to each 500<br />

mL bottle of bird seed agar.<br />

7. Mix gently and pour into 90 mm plastic petri dishes.<br />

Bromocresol Purple Milk Solids Glucose Agar (BCP)<br />

For the differentiation of Trichophyton species (Fischer and Kane 1971; Summerbell<br />

et al. 1988).<br />

Solution A:<br />

Distilled water<br />

1000 mL<br />

Skim milk powder (Carnation Brand)<br />

80 g<br />

Bromcresol (or bromocresol) purple<br />

2 mL<br />

(1.6% solution in alcohol)<br />

Dissolve in 2 litre flask and autoclave 121 O C for 10 minutes.<br />

Solution B:<br />

Glucose 40 g Distilled water 200 mL<br />

Dissolve and autoclave at 121 O C for 10 minutes.<br />

Solution C:<br />

Bacto agar (BD 214010) 30 g Distilled water 800 mL<br />

Soak for 15 minutes in 3 litre flask; autoclave at 121 O C for 10 minutes.<br />

To make media add solution A and B to solution C. Adjust final pH to 6.6.<br />

Aseptically dispense for slopes (7 mL amounts into bottles).


Descriptions of Medical Fungi 235<br />

Creatinine Dextrose Bromothymol Blue Thymine (CDBT) Agar<br />

For differentiation of C. neoformans var. neoformans and C. neoformans var. grubii<br />

(Irokanulo et al. 1994).<br />

Solution A:<br />

Creatinine 1 g Dextrose 0.5 g<br />

KH 2<br />

PO 4<br />

1 g MgSO 4<br />

.7H 2<br />

O 0.5 g<br />

Thymine 0.1 g Distilled water 980 mL<br />

1. Dissolve ingredients in small beaker and adjust pH to 5.6<br />

2. Store in refrigerator.<br />

Solution B (Aqueous Bromothymol Blue):<br />

Bromothymol Blue 0.4 g 0.01N NaOH 64 mL<br />

Distilled water<br />

36 mL<br />

1. Dissolve the Bromothymol Blue in the NaOH<br />

2. Add to the water.<br />

To prepare medium (1 litre for plates):<br />

SPECIALISED CULTURE MEDIA<br />

Solution A 980 mL Solution B 20 mL<br />

Bacto agar (BD 214010) 20 g<br />

Autoclave to 121 O C for 15 minutes, cool to 48 O C and dispense as plates.<br />

L-Canavanine Glycine Bromothymol Blue (CGB) Agar<br />

For differentiation of C. neoformans and C. gattii (Kwon-Chung et al. 1982).<br />

Solution A:<br />

Glycine Univar 10 g KH 2<br />

PO 4<br />

1 g<br />

MgSO 4<br />

1 g Thiamine HCl 1 mg<br />

L-canavanine sulphate 30 mg Distilled water 100 mL<br />

1. Dissolve ingredients in small beaker and adjust pH to 5.6<br />

2. Filter sterilise solution using 0.22 µm filter.<br />

3. Store in refrigerator.<br />

Solution B (Aqueous Bromothymol Blue):<br />

Bromothymol Blue 0.4 g 0.01N NaOH 64 mL<br />

Distilled water<br />

36 mL<br />

1. Dissolve the Bromothymol Blue in the NaOH<br />

2. Add to the water.<br />

To prepare medium (1 litre for plates):<br />

Distilled water 880 mL Solution B 20 mL<br />

Bacto agar (BD 214010) 20 g<br />

1. Autoclave to 121 O C for 15 minutes, cool to 48 O C.<br />

2. For plates add 100 mL of the filtered solution A and mix. Dispense as<br />

plates.


236<br />

Descriptions of Medical Fungi<br />

SPECIALISED CULTURE MEDIA<br />

Cornmeal Agar<br />

For routine cultivation and identification of fungi.<br />

Cornmeal agar (Oxoid CM0103)<br />

8.5 g<br />

Distilled water<br />

500 mL<br />

1. Mix dry ingredients into 100 mL water, boil remaining water.<br />

2. Add boiling water to mixture and bring to boil.<br />

3. Autoclave for 10 minutes at 121 O C, then slope on racks.<br />

Cornmeal Glucose Sucrose Yeast Extract Agar<br />

For mucormycete sporulation.<br />

Cornmeal agar (Oxoid CM0103)<br />

17 g<br />

Dextrose (Glucose)<br />

2 g<br />

Sucrose<br />

3 g<br />

Yeast extract<br />

1 g<br />

Distilled water<br />

1000 mL<br />

1. Soak dry ingredients in 100 mL water, then boil remaining water.<br />

2. Add boiling water to mixture and bring to boil.<br />

3. Dispense for slopes.<br />

4. Autoclave for 10 minutes at 121 O C, remove and then angle to form<br />

slopes on racks.<br />

Czapek Dox Agar<br />

For routine cultivation of fungi, especially Aspergillus, Penicillium, and non-sporulating<br />

moulds.<br />

Czapek Dox agar (Oxoid CM97)<br />

45.4 g<br />

Distilled water<br />

1000 mL<br />

1. Soak the ingredients in small amount of water.<br />

2. Bring remaining water to boil, add to soaking ingredients and bring to<br />

the boil again, stirring continuously.<br />

3. Dispense for slopes as required.<br />

4. Autoclave at 121 O C for 10 minutes, remove and slope or pour for<br />

plates as required.<br />

Modified Dixon’s Agar<br />

For primary isolation and cultivation of Malassezia species.<br />

Malt extract (Oxoid L39) 9 g Bacto Tryptone 1.5 g<br />

Ox-bile Desiccated (Oxoid L50) 5 g Tween 40 2.5 mL<br />

Oleic acid 0.5 g Glycerol 0.5 mL<br />

Bacto agar 3 g Distilled water 250 mL<br />

1. Soak ingredients in a little of the water.<br />

2. Bring remaining water to boil, add to the soaking ingredients and bring<br />

to the boil again constantly stirring.<br />

3. Dispense for slopes (7 mL) into 30 mL bottles.<br />

4. Autoclave at 121 O C for 10 minutes and then angle to form slopes on<br />

racks.


Descriptions of Medical Fungi 237<br />

SPECIALISED CULTURE MEDIA<br />

Hair Perforation Test<br />

For the differentiation of Trichophyton species.<br />

Blonde pre-pubital hair cut into short pieces (1 cm) 10-20 hairs<br />

Distilled water<br />

5 mL<br />

1. Autoclave hair at 121 O C for 10 minutes and store in sterile container.<br />

2. Place 10-20 short pieces of hair in 5 mL water in vial.<br />

3. Inoculate with small fragments of the test fungus.<br />

4. Incubate at room temperature.<br />

5. Individual hairs are removed at intervals up to 4 weeks and examined<br />

microscopically in lactophenol cotton blue. Isolates of T. mentagrophytes<br />

produce marked localised areas of pitting and marked erosion whereas<br />

those of T. rubrum do not.<br />

Lactritmel Agar<br />

For the production of pigment by Trichophyton species.<br />

Skimmed milk powder<br />

7 g<br />

(use only Dutch Jug skimmed milk powder)<br />

Honey<br />

10 g<br />

Cornmeal agar (Oxoid CM0103)<br />

17 g<br />

Chloramphenicol<br />

0.05 g<br />

Distilled water<br />

1000 mL<br />

1. Weigh skimmed milk into stainless steel jug. Slowly add some water,<br />

mixing milk into smooth paste. Continue adding small quantities of<br />

water until powder is dissolved (about 150 mL).<br />

2. Weigh other ingredients into skimmed milk and allow to soak.<br />

3. Boil remaining water, and with it wash out honey from beaker.<br />

4. Add to other ingredients and boil.<br />

5. Dispense for slopes (7 mL).<br />

6. Autoclave for 10 minutes at 121 O C.<br />

7. On removal from autoclave allow to stand 5 minutes then shake and<br />

then angle to form slopes on racks.<br />

Note: Do not filter or adjust pH.<br />

Littman Oxgall Agar<br />

For the differentiation of Trichophyton species.<br />

Littman oxgall agar (US Biological L3025)<br />

27.5 g<br />

Distilled water<br />

500 mL<br />

1. Soak agar in 100 mL of water. Boil remaining 400mL in a separate<br />

container.<br />

2. When water has boiled add to soaking agar and bring back to the boil,<br />

stirring constantly.<br />

3. Dispense for slopes.<br />

4. Autoclave for 10 minutes at 121 O C, remove and then angle to form<br />

slopes on racks.


238<br />

Descriptions of Medical Fungi<br />

SPECIALISED CULTURE MEDIA<br />

Malt Extract Agar<br />

For routine cultivation and identification of fungi.<br />

Oxoid Malt Extract (L39)<br />

20 g<br />

Bacto agar (BD 214010)<br />

20 g<br />

Distilled water<br />

1000 mL<br />

1. Dissolve malt extract in a beaker and adjust the solution to pH 6.5 with<br />

NaOH.<br />

2. Soak agar in small quantity of solution. Bring malt extract solution to<br />

the boil, stirring constantly.<br />

3. Add to soaking agar. Bring to boil, stirring constantly.<br />

4. Dispense for slopes as required.<br />

5. Autoclave at 121 O C for 10 minutes, remove and then angle to form<br />

slopes on racks or pour for plates as required.<br />

1% Peptone Agar<br />

For the differentiation of Trichophyton species.<br />

Tryptone Peptone (BD 211705)<br />

5 g<br />

Bacto agar (BD 214010)<br />

10 g<br />

Distilled water<br />

500 mL<br />

1. Soak agar and peptone in about 50 mL of water.<br />

2. Boil remaining water, add to the soaking ingredients and bring back to<br />

the boil again.<br />

3. Dispense for slopes (7 mL).<br />

4. Autoclave for 10 minutes at 121 O C, then angle to form slopes.<br />

Potato Dextrose Agar.<br />

For routine cultivation and identification of fungi.<br />

Potato dextrose agar (Oxoid CM139)<br />

39 g<br />

Distilled water<br />

1000 mL<br />

1. Soak potato dextrose agar in 100 mL of the water.<br />

2. Boil remaining water, add to soaking potato dextrose agar, bring back<br />

to the boil, stirring constantly.<br />

3. Dispense for slopes as required.<br />

4. Autoclave at 121 O C for 15 minutes. Remove and slope or pour for<br />

plates as required.<br />

Rice Grain Slopes.<br />

To induce sporulation and for differentiation of M. audouinii and M. canis.<br />

Polished rice grains<br />

Distilled water<br />

1. Place ~ 0.5 teaspoon rice grains into wide neck 20 mL glass vials.<br />

2. Add 8 mL distilled water to each vial.<br />

3. Loosely cap the vials, then angle to form slopes on racks ensuring rice<br />

grains are evenly distributed.<br />

4. Autoclave racks at 121 O C for 15 minutes.


Descriptions of Medical Fungi 239<br />

SPECIALISED CULTURE MEDIA<br />

Sabouraud’s Dextrose Agar (SDA) with Cycloheximide (0.05%),<br />

Chloramphenicol and Gentamicin<br />

For the primary isolation and cultivation of dermatophytes.<br />

Sabouraud’s dextrose agar (Oxoid CM41)<br />

65 g<br />

Cycloheximide (Actidione)<br />

0.5 g<br />

Chloramphenicol<br />

0.05 g<br />

Gentamicin (40mg/mL)<br />

0.56 mL<br />

Yeast extract<br />

5 g<br />

Distilled water<br />

1000 mL<br />

1. Soak all ingredients, except gentamicin, in 100 mL water.<br />

2. Boil remaining water, add to soaking ingredients, and bring to boil to<br />

dissolve, stirring well to prevent from burning.<br />

3. Add the gentamicin. Mix well.<br />

4. Dispense for slopes as required.<br />

5. Autoclave at 121 O C for 10 minutes. Remove and slope, or pour for plates<br />

as required.<br />

Sabouraud’s Dextrose Agar with Chloramphenicol and Gentamicin.<br />

For primary isolation and routine culture of yeasts and moulds.<br />

Sabouraud’s dextrose agar (Oxoid CM41)<br />

65 g<br />

Chloramphenicol<br />

0.05 g<br />

Gentamicin (40mg/mL)<br />

0.56 mL<br />

Distilled water<br />

1000 mL<br />

See above method for Sabouraud dextrose agar with cycloheximide,<br />

chloramphenicol and gentamicin.<br />

Sabouraud’s Dextrose Agar with 5% Salt.<br />

For the differentiation of Trichophyton species.<br />

Sabouraud dextrose agar (Oxoid CM41)<br />

Sodium Chloride NaCl (Univar 465)<br />

Distilled water<br />

1. Soak dry ingredients in approximately 100 mL water.<br />

2. Bring remaining water to boil, add to soaking ingredients.<br />

3. Dispense into McCartney bottles for slopes (7 mL).<br />

4. Autoclave at 121 O C for 10 minutes, then slope on racks.<br />

32.5 g<br />

25 g<br />

500 mL


240<br />

Descriptions of Medical Fungi<br />

Tap Water Agar<br />

For the stimulation of sporulation in Apophysomyces and Saksenaea isolates.<br />

Bacto agar (BD 214010)<br />

15 g<br />

Distilled water<br />

1000 mL<br />

1. Add agar to water in stainless steel jug, allow to soak.<br />

2. Dispense for slopes.<br />

3. Autoclave at 121 O C for 10 minutes, remove and slope.<br />

Urea Agar Slopes with 0.5% Glucose.<br />

For the differentiation of urease producing organisms.<br />

Urea glucose broth base:<br />

Urea, broth base (Oxoid CM71)<br />

0.9 g<br />

Glucose<br />

5 g<br />

Distilled water<br />

450 mL<br />

1. Add the Urea broth base and glucose to the distilled water in a 500mL<br />

beaker.<br />

2. Dispense in 5 x 90 mL amounts.<br />

3. Autoclave at 121 O C for 20 mins.<br />

4. When cool, label and store in the fridge.<br />

Method to make slopes:<br />

SPECIALISED CULTURE MEDIA<br />

40% Urea Solution (Oxoid SR 20) 10 mL<br />

Bacto agar (BD 214010)<br />

3 g<br />

Distilled water<br />

100 mL<br />

1. Add agar to 100 mL of distilled water in a 250 mL pyrex bottle.<br />

2. Autoclave at 121 O C for 15 minutes and place in 50 O C water bath.<br />

3. When cool add 90 mL of the Urea broth with glucose and the 10 mL<br />

of 40% urea solution to agar and dispense in 3 mL aliquots and angle<br />

bottles to form slopes on racks.<br />

Vitamin Free Agar (Trichophyton Agar No.1)<br />

For the differentiation of Trichophyton species.<br />

Trichophyton agar No. 1 (BD 287710)<br />

11.8 g<br />

Distilled water<br />

200 mL<br />

1. Add agar to water in stainless steel jug, allow to soak.<br />

2. Bring to boil to dissolve, stirring constantly.<br />

3. Once boiled remove immediately to avoid discolouration.<br />

4. Dispense for slopes.<br />

5. Autoclave at 121 O C for 10 minutes, remove and slope.


Descriptions of Medical Fungi 241<br />

REFERENCES<br />

Abe, A., K. Asano and T. Sone. 2010. A molecular phylogeny-based taxonomy of the genus<br />

Rhizopus. Biosci. Biotechnol. Biochem. 74: 1325-1331.<br />

Abliz, P., K. Fukushima, K. Takizawa et al. 2003. Specific oligonucleotide primers for<br />

identification of Hortaea werneckii, a causative agent of tinea nigra. Diagn. Microbiol.<br />

Infect. Dis. 46: 89-93.<br />

Abliz, P., K. Fukushima, K., Takizawa et al. 2003. Rapid identification of the genus Fonsecaea<br />

by PCR with specific oligonucleotide primers. J. Clin. Microbiol. 41: 873-876.<br />

Abliz, P., K. Fukushima, K. Takizawa et al. 2004. Specific oligonucleotide primers for<br />

identification of Cladophialophora carrionii, a causative agent of chromoblastomycosis. J.<br />

Clin. Microbiol. 42: 404-407.<br />

Adam, R.D., M.L. Paquin, E.A. Petersen et al. 1986. Phaeohyphomycosis caused by the<br />

fungal genera Bipolaris and Exserohilum. A report of 9 cases and review of the literature.<br />

Medicine. 65: 203-217.<br />

Ahmed, S.A., W.W.J. van de Sande, D.A. Stevens et al. 2014a. Revision of agents of blackgrain<br />

eumycetoma in the order Pleosporales. Persoonia 33: 141-154.<br />

Ahmed, S.A., B.H.G. van den Ende, A.H. Fahal et al. 2014b. Rapid Identification of Black<br />

Grain Eumycetoma Causative Agents Using Rolling Circle Amplification. PLoS Negl. Trop.<br />

Dis. 8(12): e3368.<br />

Ajello, L. 1957. Coccidioides immitis: Isolation procedures and diagnostic criteria. Proceedings<br />

of symposium on Coccidioidomycosis. Public Health Publication No. 575, CDC Atlanta,<br />

USA.<br />

Ajello, L. 1977. Taxonomy of the dermatophytes: a review of their imperfect and perfect states.<br />

In “Recent Advances in Medical and Veterinary Mycology” (K. Iwata, ed.), pp. 289-297.<br />

University Park Press, Baltimore, Maryland, USA.<br />

Ajello, L., D.F. Dean and R.S. Irwin. 1976. The zygomycete Saksenaea vasiformis as a<br />

pathogen of humans with a critical review of the etiology of zygomycosis. Mycologia. 68:<br />

52-62.<br />

Alastruey-Izquierdo A, K. Hoffman, G.S. de Hoog et al. 2010. Species recognition and clinical<br />

relevance of the zygomycetous genus Lichtheimia (syn, Absidia pro parte, Mycocladus). J.<br />

Clin. Microbiol. 48: 2154-2170.<br />

Alcoba-Flórez, J., S. Méndez-Álvarez, J. Cano et al. 2005. Phenotypic and molecular<br />

characterization of Candida nivariensis sp. nov., a possible new opportunistic fungus. J.<br />

Clin. Microbiol. 43: 4107-4111.<br />

Alcorn, J.L. 1983. Genetic concepts in Drechslera, Bipolaris and Exserohilum. Mycotaxon.<br />

17: 1-86.<br />

Al-Hatmi, A.M.S., A.D. van Diepeningen, I. Curfs-Breuker et al. 2015. Specific antifungal<br />

susceptibility profiles of opportunists in the Fusarium fujikuroi complex. J. Antimicrob.<br />

Chemother. 70: 1068-1071.<br />

Al-Mohsen, I.Z., D.A. Sutton, L. Sigler et. al. 2000. Acrophialophora fusispora brain abscess<br />

in a child with acute lymphoblastic leukaemia: review of cases and taxonomy. J. Clin.<br />

Microbiol. 38: 4569-4576.<br />

Alshawa, K., J.L. Beretti, C. Lacroix et al. 2012. Successful identification of clinical<br />

dermatophyte and Neoscytalidium species by matrix-assisted laser desorption ionizationtime<br />

of flight mass spectrometry. J. Clin. Microbiol. 50: 2277-2281.<br />

Alvarado-Ramirez, E. and J.M. Torres-Rodriguez. 2007. In vitro susceptibility of Sporothrix<br />

schenckii to six antifungal agents using three different methods. Antimicrob. Agents<br />

Chemother. 60: 658-661.<br />

Alvarez, E., D.A. Sutton, J. Cano et al. 2009. Spectrum of zygomycete species identified in<br />

clinically significant specimens in the United States. J. Clin. Microbiol. 47: 1650-1656.


242<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Alvarez, E., A.M. Stchigel, J. Cano, et al. 2010. Molecular phylogenetic diversity of the<br />

emerging mucoralean fungus Apophysomyces: proposal of three new species. Revta<br />

Iberoam. Micol. 27: 80-89.<br />

Alvarez, E., D. Garcia-Hermoso, D.A. Sutton, et al. 2010. Molecular phylogeny and proposal<br />

of two new species of the emerging pathogenic fungus Saksenaea. J. Clin. Microbiol. 48:<br />

4410-4416.<br />

Alshawa, K., J.L. Beretti, C. Lacroix et al. 2012. Successful identification of clinical<br />

dermatophyte and Neoscytalidium species by matrix-assisted laser desorption ionizationtime<br />

of flight mass spectrometry. J. Clin. Microbiol. 50: 2277-2281.<br />

Ames, L.M. 1963. A monograph of the Chaetomiaceae. U.S. Army Research and Development<br />

Serial. 2: 1-125.<br />

Andrianopoulos, A. 2002. Control of morphogenesis in the human fungal pathogen Penicillium<br />

marneffei. Int. J. Med. Microbiol. 292: 331-347.<br />

Arendrup, M.C., T. Boekhout, M. Akova et al. 2014. ESCMID and ECMM joint clinical<br />

guidelines for the diagnosis and management of rare invasive yeast infections. Clin.<br />

Microbiol. Infect. 20 (Suppl. 3): 76-98.<br />

Arzanlou, M., J.Z. Groenewald, W. Gams. et al. 2007. Phylogenetic and morphotaxonomic<br />

revision of Ramichloridium and allied Genera. Stud. Mycol. 58: 57-93.<br />

Asadzadeh, M., S. Ahmad, N. Al-Sweih et al. 2009. Rapid molecular differentiation and<br />

genotypic heterogeneity among Candida parapsilosis and Candida orthopsilosis strains<br />

isolated from clinical specimens in Kuwait. J. Med. Microbiol. 58: 745-52.<br />

Atkins, S.D., I.M. Clark, S. Pande et al. 2005. The use of real-time PCR and species-specific<br />

primers for the identification and monitoring of Paecilomyces lilacinus. FEMS Microbiol.<br />

Ecol. 51: 257-264.<br />

Aveskamp, M.M., J. de Gruyter, J.H.C. Woudenberg et al. 2010. Highlights of the<br />

Didymellaceae: A polyphasic approach to characterise Phoma and related pleosporalean<br />

genera. Stud. Mycol. 65: 1-60.<br />

Badali, H., C. Gueidan, M.J. Najafzadeh et al. 2008. Biodiversity of the genus Cladophialophora.<br />

Stud. Mycol. 61: 175-191.<br />

Badali, H., G.S. de Hoog, I. Curfs-Breuker et al. 2010. Use of amplified fragment length<br />

polymorphism to identify 42 Cladophialophora strains related to cerebral phaeohyphomycosis<br />

with in vitro antifungal susceptibility. J. Clin. Microbiol. 48: 2350-2356.<br />

Badali, H., M.J. Najafzadeh, M. van Esbroeck et al. 2010. The clinical spectrum of Exophiala<br />

jeanselmei, with a case report and in vitro antifungal susceptibility of the species. Med.<br />

Mycol. 48: 318-327.<br />

Badali, H., J. Chander, S. Bansal et al. 2010. First autochthonous case of Rhinocladiella<br />

mackenziei cerebral abscess outside the Middle East. J. Clin. Microbiol. 48: 646-649.<br />

Badali, H., S.A. Yazdanparast, A. Bonifaz. et al. 2013. Veronaea botryosa: molecular<br />

identification with amplified fragment length polymorphism (AFLP) and in vitro antifungal<br />

susceptibility. Mycopathologia 175: 505-513.<br />

Badali, H., S. Khodavaisy, H. Fakhim et al. 2015. In vitro susceptibility profiles of eight antifungal<br />

drugs against clinical and environmental strains of Phaeoacremonium. Antimicrob. Agents<br />

Chemother. 59: 7818-7822.<br />

Badenoch, R.R., C.L. Halliday, D.H. Ellis et al. 2006. Ulocladium atrum Keratitis. J. Clin.<br />

Microbiol. 44: 1190-1193.<br />

Bagyalakshmi, R., K.L. Therese, S. Prasanna et al. 2008. Newer emerging pathogens of<br />

ocular non-sporulating molds (NSM) identified by polymerase chain reaction (PCR)-based<br />

DNA sequencing technique targeting internal transcribed spacer (ITS) region. Curr. Eye<br />

Res. 33: 139-147.<br />

Balajee, S.A., J. Gribskov, M. Brandt et al. 2005a. Mistaken identity: Neosartorya<br />

pseudofischeri and its anamorph masquerading as Aspergillus fumigatus. J. Clin. Microbiol.<br />

43: 5996–5999.


Descriptions of Medical Fungi 243<br />

REFERENCES<br />

Balajee, S.A., J.L. Gribskov, E. Hanley et al. 2005b. Aspergillus lentulus sp. nov., a new<br />

sibling species of A. fumigatus. Eukaryotic Cell 4: 625-632.<br />

Balajee, S.A., D. Nickle, J. Varga et al. 2006. Molecular studies reveal frequent misidentification<br />

of Aspergillus fumigatus by morphotyping. Eukaryotic Cell 5: 1705-1712.<br />

Balajee, S.A., J. Houbraken, P.E. Verweij et al. 2007. Aspergillus species identification in the<br />

clinical setting. Stud. Mycol. 59: 39-46.<br />

Balajee, S.A., A.M. Borman, M.E. Brandt, et al. 2009. Sequence-based identification of<br />

Aspergillus, Fusarium, and Mucorales species in the clinical mycology laboratory: where<br />

are we and where should we go from here? J. Clin. Microbiol. 47: 877-884.<br />

Barnett, J.A., R.W. Payne and D. Yarrow. 1983. Yeasts: characteristics and identification.<br />

Cambridge University Press, London, UK.<br />

Barron, G.L. 1968. The genera of hyphomycetes from soil. Williams & Wilkins Co. Balitmore,<br />

USA.<br />

Barron, M.A., D. A. Sutton, R. Veve et al. 2003. Invasive mycotic infections caused by<br />

Chaetomium perlucidum, a new agent of cerebral phaeohyphomycosis. J. Clin. Microbial.<br />

41: 5302-5307.<br />

Barros, M.B., R. de Almeida Paes and A.O. Schubach. 2011. Sporothrix schenckii and<br />

sporotrichosis. Clin. Microbiol. Rev. 24: 633-654.<br />

Barrs, V.R., T.M. van Doorn, J. Houbraken et al. 2013. Aspergillus felis sp. nov., an emerging<br />

agent of invasive aspergillosis in humans, cats, and dogs. PLoS One. 14;8(6):e64871.<br />

Beguin, H., N. Pyck, M. Hendrickx et al. 2012. The taxonomic status of Trichophyton<br />

quinckeanum and T. interdigitale revisited: a multigene phylogenetic approach. Medical<br />

Mycology 50: 871-882.<br />

Bensch, K., J.Z. Groenewald, J. Dijksterhuis et al. 2010. Species and ecological diversity<br />

within the Cladosporium cladosporioides complex (Davidiellaceae, Capnodiales). Stud.<br />

Mycol. 67: 1-94.<br />

Bensch, K., U. Braun, J.Z. Groenewald et al. 2012. The genus Cladosporium. Stud. Mycol.<br />

72: 1-401.<br />

Beyda, N.D., S.H. Chuang, M.J. Alam et al. 2013. Treatment of Candida famata bloodstream<br />

infections: case series and review of the literature. Antimicrob. Chemother. 68: 438-443.<br />

Bialek, R., A.C. Cirera, T. Herrmann et al. 2003. Nested PCR assays for detection of<br />

Blastomyces dermatitidis DNA in paraffin-embedded canine tissue. J. Clin. Microbiol. 41:<br />

205-208.<br />

Binnicker, M.J., A.S. Popa, J. Catania et al. 2011. Meningeal coccidioidomycosis diagnosed<br />

by real-time polymerase chain reaction analysis of cerebrospinal fluid. Mycopathologia<br />

171: 285-289.<br />

Bishop, J.A., N. Chase, S.S. Magill et al. 2008. Candida bracarensis detected among isolates<br />

of Candida glabrata by peptide nucleic acid fluorescence in situ hybridization: susceptibility<br />

data and documentation of presumed infection. J. Clin. Microbiol. 46: 443-446.<br />

Boekhout, T., E. Guého, P. Mayser and A. Velegraki (eds). 2010. Malassezia and the Skin.<br />

Science and Clinical Practice. Springer, Heidelberg, 319 pp.<br />

Booth, C. 1966. The genus Cylindrocarpon. Mycol. Pap. 104:1-56.<br />

Booth, C. 1971. The genus Fusarium. Commonwealth Mycological Institute, Kew, Surrey,<br />

England.<br />

Booth, C. 1977. Fusarium: laboratory guide to the identification of the major species.<br />

Commonwealth Mycological Institute, Kew, Surrey, England.<br />

Borman, A.M., R. Petch, C.J. Linton et al. 2008. Candida nivariensis, an emerging pathogenic<br />

fungus with multidrug resistance to antifungal agents. J. Clin. Microbiol. 46: 933-938.


244<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Borman, A.M., C.J. Linton, D. Oliver et al. 2009. Pyrosequencing analysis of 20 nucleotides<br />

of internal transcribed spacer 2 discriminates Candida parapsilosis, Candida metapsilosis,<br />

and Candida orthopsilosis. J. Clin. Microbiol. 47: 2307-2310.<br />

Brenier-Pinchart, M.P., H. Pelloux, B. Lebeau et al. 1999. Towards a molecular diagnosis of<br />

invasive aspergillosis? A review of the literature. J. Mycol. Méd. 9: 16-23.<br />

Brilhantea, R.S.N., M.A.B. Fechinea, J.R.L. Mesquita et al. 2012. Histoplasmosis in<br />

HIV-positive patients in Ceará, Brazil: clinical-laboratory aspects and in vitro antifungal<br />

susceptibility of Histoplasma capsulatum isolates. Trans. R. Soc. Trop. Med. Hyg. 106:<br />

484-488.<br />

Brillowska-Dabrowska, A., E. Michałek, D.M. Saunte et al. 2013. PCR test for Microsporum<br />

canis identification. Med. Mycol. 51: 576-579.<br />

Brown, E.M., L.R. McTaggart, S.X. Zhang et al. 2013. Phylogenetic analysis reveals a cryptic<br />

species Blastomyces gilchristii, sp. nov. within the human pathogenic fungus Blastomyces<br />

dermatitidis. PloS One 8: e59237.<br />

Buchta, V. and M. Otcenasek. 1988. Geotrichum candidum - an opportunistic agent of mycotic<br />

diseases. Mycoses. 31: 363-370.<br />

Burges, G.E., C.T. Walls and J.C. Maize. 1987. Subcutaneous phaeohyphomycosis caused<br />

by Exserohilum rostratum in an immunocompetent host. Arch. Dermatol. 123: 1346-1350.<br />

Burgess, L.W. and C.M. Liddell. 1983. Laboratory manual for Fusarium research. Fusarium<br />

Research Laboratory, Department of Plant Pathology and Agricultural Entomology. The<br />

University of Sydney.<br />

Burgess J.W., W.R. Schwan and T.J. Volk. 2006. PCR-based detection of DNA from the<br />

human pathogen Blastomyces dermatitidis from natural soil samples. Med. Mycol. 44: 741-<br />

748.<br />

Buzina, W., D. Lang-Loidolt, H. Braun et al. 2001. Development of molecular methods for<br />

identification of Schizophyllum commune from clinical samples. J. Clin. Microbiol. 39:<br />

2391-2396.<br />

Cabanes, F.J., S. Vega, and G. Castellá. 2011. Malassezia cuniculi sp. nov., a novel yeast<br />

species isolated from rabbit skin. Med. Mycol. 49: 40-48.<br />

Cafarchia, C., R.B. Gasser, L.A. Figueredo et al. 2011. Advances in the identification of<br />

Malassezia. Mol. Cell Probes 25: 1-7.<br />

Cafarchia, C., R. Iatta, M.S. Latrofa et al. 2013. Molecular epidemiology, phylogeny and<br />

evolution of dermatophytes. Infect. Genet. Evol. 20: 336-351.<br />

Calderaro, A., F. Motta, S. Montecchini et al. 2014. Identification of dermatophyte species<br />

after implementation of the in-house MALDI-T<strong>OF</strong> MS database. Int. J. Mol. Sci. 15: 16012-<br />

16024.<br />

Campbell, C.K. and M.D. Smith. 1982. Conidiogenesis in Petriellidium boydii (Pseudallescheria<br />

boydii). Mycopathologia 78: 145-150.<br />

Cano, J. and J. Guarro. 1990. The genus Aphanoascus. Mycol. Res. 94: 355-377.<br />

Cano, J., M. Sagués, E. Barrio et al. 2002. Molecular taxonomy of Aphanoascus and<br />

description of two new species from soil. Stud. Mycol. 47: 153-164.<br />

Cano, J., J. Guarro and J. Gene. 2004. Molecular and morphological identification of<br />

Colletotrichum species of clinical interest. J. Clin. Microbiol. 42: 2450-2454.<br />

Cantón, E., J. Pemán, C. Iniguez et al. 2013. Epidemiological cutoff values for fluconazole,<br />

itraconazole, posaconazole, and voriconazole for six Candida species as determined by<br />

the colorimetric sensititre YeastOne method. J. Clin. Microbiol. 51: 2691-2695.<br />

Cantón, E., J. Pemán, D. Hervás et al. 2012. Comparison of three statistical methods<br />

for establishing tentative wild-type population and epidemiological cutoff values for<br />

echinocandins, amphotericin B, flucytosine, and six Candida species as determined by the<br />

colorimetric Sensititre YeastOne method. J. Clin. Microbiol. 50: 3921-3926.<br />

Carmichael, J.W. 1962. Chrysosporium and some aleuriosporic hyphomycetes. Can. J. Bot.<br />

40: 1137-1173.


Descriptions of Medical Fungi 245<br />

REFERENCES<br />

Casadevall, A. and J.R. Perfect. 1998. Cryptococcus neoformans. ASM Press USA.<br />

Cassagne, C., S. Ranque, A. Normand et al. 2011. Mould routine identification in the clinical<br />

laboratory by matrix-assisted laser desorption ionization time-of- flight mass spectrometry.<br />

PLoS ONE 6(12): e28425.<br />

Catanzaro, A. 1985. Coccidiomycosis. In Fungal Diseases of the Lung, eds G.A. Sarosi and<br />

S.F. Davies. Grune and Stratton Inc.<br />

Cavalier-Smith, T. 1998. A revised six-kingdom system of life. Biol. Rev. Canm. Philos. Soc.<br />

73: 203-266.<br />

Cendejas-Bueno, E., A. Kolecka, A. Alastruey-Izquierdo et al. 2012. Reclassification<br />

of the Candida haemulonii complex as Candida haemulonii (C. haemulonii group I), C.<br />

duobushaemulonii sp. nov. (C. haemulonii group II), and C. haemulonii var. vulnera var.<br />

nov.: three multiresistant human pathogenic yeasts. J. Clin. Microbiol. 50: 3641-51.<br />

Chagas-Neto, T.C., G.M. Chaves and A.L. Colombo. 2008. Update on the genus Trichosporon.<br />

Mycopathologia 166: 121-132.<br />

Chakrabarti, A., M.R. Shivaprakash, I. Curfs-Breuker et al. 2010. Apophysomyces elegans:<br />

epidemiology, amplified fragment length polymorphism typing, and in vitro antifungal<br />

susceptibility pattern. J. Clin. Microbiol. 48: 4580-4585.<br />

Chakrabarti, A., A. Ghosh, G.S. Prasad et al. 2003. Apophysomyces elegans: an emerging<br />

zygomycete in India. J. Clin. Microbiol. 41: 783-788.<br />

Chakrabarti, A., H. Kaur, S.M. Rudramurth et al. 2016. Brain abscess due to Cladophialophora<br />

bantiana: a review of 124 cases. Med.Mycol. 54: 111-119.<br />

Chandler, F.W., W. Kaplan and L. Ajello. 1980. A colour atlas and textbook of the histopathology<br />

of mycotic diseases. Wolfe Medical Publications Ltd.<br />

Chapman, S.W., W.E. Dismukes, L.A. Proia et al. 2008. Clinical practice guidelines for the<br />

management of blastomycosis: 2008 update by the Infectious Diseases Society of America.<br />

Clin. Infect. Dis. 46:1801-1812.<br />

Chemaly, R.F., J.W. Tomford, G.S. Hall et al. 2001. Rapid diagnosis of Histoplasma<br />

capsulatum endocarditis using the AccuProbe on an excised valve. J. Clin. Microbiol. 39:<br />

2640-2641.<br />

Chen, C.A., D. Ellis, T.C. Sorrell et al. 2011. Trichophyton. Chapter 44, Molecular Detection<br />

of Human Fungal Pathogens. Ed: Dongyou Liu. CRC Press.<br />

Chen, S.C., M.A. Slavin, C.H. Heath et al. 2012. Clinical manifestations of Cryptococcus gattii<br />

infection: determinants of neurological sequelae and death. Clin. Infecti. Dis. 55: 789-98.<br />

Chowdhary A, H.S. Randhawa, S.N. Gaur et al. 2013a. Schizophyllum commune as an<br />

emerging fungal pathogen: a review and report of two cases. Mycoses 56: 1-10.<br />

Chowdhary, A., S. Kathuria, P.K. Singh et al. 2013b. Molecular Characterization and in vitro<br />

Antifungal Susceptibility Profile of Schizophyllum commune, an Emerging Basidiomycete<br />

in Bronchopulmonary Mycoses. Antimicrobial Agents Chemother. 57: 2845-2848.<br />

Chowdhary, A., S. Khaturia, K. Agarwal et al. 2014a. Recognizing filamentous basidiomycetes<br />

as agents of human disease: A review. Med. Mycol. 52: 782-797.<br />

Chowdhary, A., K. Agarwal, S. Kathuria et al. 2014b. Allergic bronchopulmonary mycosis<br />

due to fungi other than Aspergillus: a global overview. Crit. Rev. Microbiol. 40: 30-48.<br />

CLSI Interpretive criteria for identification of bacteria and fungi by DNA target sequencing<br />

(MM18-A). 2008. Wayne, PA.<br />

CLSI Reference method for broth dilution antifungal susceptibility testing of yeasts: third edition<br />

(M27-A3). 2008. Wayne, PA.<br />

CLSI Reference method for broth dilution antifungal susceptibility testing of yeasts: fourth<br />

informational supplement (M27-S4). 2012. Wayne, PA.<br />

CLSI Reference method for broth dilution antifungal susceptibility testing of filamentous fungi:<br />

second edition (M38-A2). 2008. Wayne, PA.


246<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Colombo, A., A.C.B. Padovan and G.M. Chaves. 2011. Current Knowledge of Trichosporon<br />

spp. and Trichosporonosis. Clin. Microbiol. Rev. 24: 682-700.<br />

Cooney, D.H. and R. Emerson. 1964. Thermophilic fungi. W.H. Freeman & Co.<br />

Cooter, R.T., I.S. Lim, D.H. Ellis et. al. 1990. Burn wound zygomycosis caused by<br />

Apophysomyces elegans. J.Clin. Microbiol. 28: 2151-2153.<br />

Coriglione, G., G. Stella, L. Gafa et al. 1990. Neosartorya fischeri var fischeri (Wehmer)<br />

Malloch and Cain 1972 (anamorph: Aspergillus fischerianus Samson and Gams 1985) as<br />

a cause of mycotic keratitis. Eur. J. Epidemiol. 6: 382-385.<br />

Correia, A., P. Sampaio, S. James et al. 2006. Candida bracarensis, sp. nov., a novel<br />

anamorphic yeast species phenotypically similar to Candida glabrata. Int. J. Syst. Evol.<br />

Microbiol. 56: 313-317.<br />

Cortez, K.J., E. Roilides, F. Quiroz-Telles et al. 2008. Infections Caused by Scedosporium<br />

spp. Clin. Microbiol. Reviews. 21: 157-197.<br />

Crous, P.W., B. Slippers, M.J. Wingfield et al. 2006. Phylogenetic lineages in the<br />

Botryosphaeriaceae. Stud. Mycol. 55: 235-253.<br />

Crous, P.W., U. Braun, K. Schubert et al. 2007. Delimiting Cladosporium from morphologically<br />

similar genera. Stud. Mycol. 58: 33-56.<br />

Davis, S.R., D.H. Ellis, P. Goldwater et. al. 1994. First human culture-proven Australian case<br />

of entomophthoromycosis caused by Basidiobolus ranarum. J. Med. Vet. Mycol. 32: 225-<br />

230.<br />

da Cunha, K.C., D.A. Sutton, A.W. Fothergill et al. 2012a. Diversity of Bipolaris species<br />

in clinical samples in the United States and their antifungal susceptibility profiles. J. Clin.<br />

Microbiol. 50: 4061-4066.<br />

da Cunha, K.C., D.A. Sutton, J. Gene et al. 2012b. Molecular identification and in vitro<br />

response to antifungal drugs of clinical isolates of Exserohilum. Antimicrob. Agents.<br />

Chemother. 56: 4951-4954.<br />

da Cunha, K.C., D.A. Sutton, A.W. Fothergill et al. 2013. In vitro antifungal susceptibility and<br />

molecular identity of 99 clinical isolates of the opportunistic fungal genus Curvularia. Diagn.<br />

Microbiol. Infect. Dis. 76: 168-174.<br />

da Cunha, K.C., D.A. Sutton, J. Gene et al. 2014. Pithomyces species (Montagnulaceae)<br />

from clinical specimens: identification and antifungal susceptibility profiles. Med. Mycol.<br />

52: 748-757.<br />

de Beer, Z.W., D. Begerow, R. Bauer et al. 2006. Phylogeny of the Quambalariacea fam.<br />

nov., including important Eucalyptus pathogens in South Africa and Australia. Stud. Mycol.<br />

55: 289-298.<br />

de Gruyter, J. M.M. Aveskamp, J.H.C. Woudenberg et al. 2009. Molecular phylogeny of<br />

Phoma and allied anamorph genera: towards a reclassification of the Phoma complex.<br />

Mycol. Res. 113: 508-519.<br />

de Hoog, G.S. 1972. The genera Beauvaria, Isaria, Tritrachium and Acrodontium Gen. Nov.<br />

Stud. Mycol., Centraalbureau voor Schimmelcultures, Baarn. 1: 1-41.<br />

de Hoog, G.S. 1977. Rhinocladiella and allied genera. Stud. Mycol., Centraalbureau voor<br />

Schimmelcultures, Baarn. 15: 1-140.<br />

de Hoog, G.S. 1983. On the potentially pathogenic dematiaceous Hyphomycetes. In: D.H.<br />

Howard (ed). The fungi pathogenic to humans and animals. A: 149-216.<br />

de Hoog, G.S. 1985. The taxonomic structure of Exophiala. in Fungi pathogenic for humans<br />

and animals. Part B: Pathogenicity and detection: II. (ed. D. Howard). Marcel Dekker Inc.<br />

de Hoog, G.S., D. Adelmann, A.O.A. Ahmed et al. 2004. Phylogeny and typification of<br />

Madurella mycetomatis, with a comparison of other agents of eumycetoma. Mycoses 47:<br />

121-130.


Descriptions of Medical Fungi 247<br />

REFERENCES<br />

de Hoog, G.S., D. Attili, V.A. Vicente et al. 2004. Molecular ecology and pathogenic potential<br />

of Fonsecaea species. Med. Mycol. 42: 405-416.<br />

de Hoog, G.S., A.D. van Diepeningen, el-S. Mahgoub et al. 2012. New species of Madurella,<br />

causative agents of black-grain mycetoma. J. Clin. Microbiol. 50: 988-994.<br />

de Hoog, G.S., E. Gueho, F. Masclaux et. al. 1995. Nutritional physiology and taxonomy of<br />

human-pathogenic Cladosporium-Xylohypha species. J. Med. Vet. Mycol. 33: 339-347.<br />

de Hoog, G.S., J. Guarro, J. Gene and M.J. Figueras. 2000. Atlas of Clinical Fungi (second<br />

edition). Centraalbureau voor Schimmelcultures, Utrecht, The Netherlands.<br />

de Hoog, G.S., J. Guarro, J. Gene and M.J. Figueras. 2015. Atlas of Clinical Fungi (Version<br />

4.1.2). Centraalbureau voor Schimmelcultures, Utrecht, The Netherlands.<br />

de Hoog, G.S. and E.J. Hermanides-Nijhof. 1977. The black yeasts and allied hyphomycetes.<br />

Stud. Mycol. No. 15. Centraalbureau voor Schimmelcultures, The Netherlands.<br />

de Hoog, G.S., and R. Horré. 2002. Molecular taxonomy of the Alternaria and Ulocladium<br />

species described from humans and their identification in the routine laboratory. Mycoses<br />

45: 259-276.<br />

de Hoog, G.S., A.S. Nishikaku, G. Fernandez Zeppenfeldt et al. 2007. Molecular analysis<br />

and pathogenicity of the Cladophialophora carrionii complex, with the description of a novel<br />

species. Stud. Mycol. 58: 219-234.<br />

de Hoog, G.S., A.H. Rantio-Lehtimaki and M.TH. Smith. 1985. Blastobotryis; Sporothrix<br />

and Trichosporiella; generic delimitation, new species, and a Stephanoascus teleomorph.<br />

Antontie van Leeuwenhoek. 51: 79-109.<br />

de Hoog, G.S. and M.T. Smith. 2004. Ribosomal gene phylogeny and species delimitation in<br />

Geotrichum and its teleomorphs. Stud. Mycol. 50: 489-515.<br />

de Hoog, G.S. and M.T. Smith. 2011a. Geotrichum Link: Fries (1832), p 1279-1286 In Kurtzman<br />

CP, Fell JW, Boekhout T (ed), The yeasts: a taxonomic study. Elsevier, Amsterdam, the<br />

Netherlands.<br />

de Hoog, G.S. and M.T. Smith. 2011b. Saprochaete Coker & Shanor ex D.T.S. Wagner &<br />

Dawes (1970), p 1317–1330. In Kurtzman CP, Fell JW, Boekhout T (ed), The yeasts: a<br />

taxonomic study. Elsevier, Amsterdam, the Netherlands.<br />

de Hoog, G.S. and M.T. Smith. 2011c. Magnusiomyces Zender (1977), p 565–574 In Kurtzman<br />

CP, Fell JW, Boekhout T (ed), The yeasts: a taxonomic study. Elsevier, Amsterdam, the<br />

Netherlands.<br />

de Hoog, G.S., M.T. Smith and E. Guého. 1986. A revision of the genus Geotrichum and its<br />

teleomorphs. Stud. Mycol. 29: 1-131.<br />

de Hoog, G.S., V. Vincent, R.B. Caligiorne et. al. 2003. Species diversity and polymorphism<br />

in the Exophiala spinifera clade containing opportunistic black yeast-like fungi. J. Clin.<br />

Microbiol. 41: 4767-4778.<br />

de Hoog, G.S. and G.A. de Vries. 1973. Two new species of Sporothrix and their relation to<br />

Blastobotrys nivea. Antonie Van Leeuwenhoek. 39: 515-520.<br />

de Hoog, G.S., X.O. Weenink and A.H.G. Gerrits van den Ende. 1999. Taxonomy of the<br />

Phialophora verrucosa complex with the description of two new species. Stud. Mycol. 43:<br />

107-122.<br />

de Hoog, G.S., J.S. Zeng, M.J. Harrak et al.. 2006. Exophiala xenobiotica sp. nov., an<br />

opportunistic black yeast inhabiting environments rich in hydro carbons. Antonie Van<br />

Leeuwenhoek 90: 257-268.<br />

de Hoog G.S., K. Dukik, M. Monod et al. 2016. Towards a noval multilocus phylogenetic<br />

taxonomy for dermatophytes. Mycopathologia DOI 10.1007/s11046-016-0073-9.<br />

Desjardins, C.A., M.D. Champion, J.W. Holder et al. 2011. Comparative genomic analysis of<br />

human fungal pathogens causing paracoccidioidomycosis. PLoS Genetics DOI: 10.1371/<br />

journal.pgen.1002345.


248<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Desnos-Ollivier, M., S. Bretagne, F. Dromer et al. 2006. Molecular identification of blackgrain<br />

mycetoma agents. J. Clin. Microbiol. 44: 3517-3523.<br />

Desnos-Ollivier, M., C. Blanc, D. Garcia-Hermoso et al. 2014. Misidentification of<br />

Saprochaete clavata as Magnusiomyces capitatus in clinical isolates: utility of internal<br />

transcribed spacer sequencing and matrix-assisted laser desorption ionization-time of flight<br />

mass spectrometry and importance of reliable databases. J. Clin. Microbiol. 52: 2196-98.<br />

Diekema, D.J., B. Petroelje, S.A. Messer et al. 2005. Activities of available and investigational<br />

antifungal agents against Rhodotorula species. J. Clin. Microbiol. 43: 476-478.<br />

Diekema, D.J., S.A. Messer, L.B. Boyken et al. 2009. In vitro activity of seven systemically<br />

active antifungal agents against a large global collection of rare Candida species as<br />

determined by CLSI broth microdilution methods. J. Clin. Microbiol. 47: 3170-3177.<br />

Dixon, D.M. and A. Polak-Wyss. 1991. The medically important dematiaceous fungi and their<br />

identification. Mycoses. 34: 1-18.<br />

Dolatabadi, S., G. Walther, A.H.G. Gerrits van den Ende et al. 2014. Diversity and delimitation<br />

of Rhizopus microsporus. Fung. Divers. 64: 145-163.<br />

Domsch, K.H., W. Gams and T.H. Anderson. 1980. Compendium of soil fungi. Academic<br />

Press.<br />

Domsch, K.H., W. Gams and T.H. Anderson. 2007. Compendium of soil fungi. Second<br />

Edition, IHW-Verlag, Germany.<br />

Duarte A.P.M., F.C. Pagnocca, N.C. Baron et al. 2013. In vitro susceptibility of environmental<br />

isolates of Exophiala dermatitidis to five antifungal drugs. Mycopathologia 175: 455-461.<br />

Duboc de Almeida, G.M., S. Figueiredo Costa, M. Melhem et al. 2008. Rhodotorula spp.<br />

isolated from blood cultures: clinical and microbiological aspects. Med. Mycol. 46: 547-556.<br />

Durie, E.B. and D. Frey. 1957. A new species of Trichophyton from New South Wales.<br />

Mycologia 49: 401-411.<br />

Dworzack, D.L., A.S. Pollock, G.L. Hodges et al. 1978. Zygomycosis of the maxillary sinus<br />

and palate caused by Basidiobolus haptosporus. Arch. Intern. Med. 138: 1274-1276<br />

El Feghaly, R.E., D.A. Sutton, E.H. Thompson et al. 2012. Graphium basitruncatum fungemia<br />

in an immunosuppressed child post stem-cell transplantation. Med. Mycol. Case Rep. 1:<br />

35-38.<br />

Elías, N.A., M.L. Cuestas, M. Sandoval et al. 2012. Rapid identification of Histoplasma<br />

capsulatum directly from cultures by multiplex PCR. Mycopathologia 174: 451-456.<br />

Ellis, D.H. 1981. Ascocarp morphology and terminal hair ornamentation in thermophilic<br />

Chaetomium species. Mycologia. 73: 755-773.<br />

Ellis, D.H. 2005a. Subcutaneous Zygomycetes -Entomophthoromycosis. Chapter 17. In<br />

Topley and Wilson’s Microbiology and Microbial Infections: Medical Mycology, 10th edition,<br />

Hodder Arnold London. pp 347-355.<br />

Ellis, D.H. 2005b. Systemic Zygomycetes -Mucormycosis. Chapter 33. In Topley and Wilson’s<br />

Microbiology and Microbial Infections: Medical Mycology, 10th edition, Hodder Arnold<br />

London. pp 659-686.<br />

Ellis, D.H. and G. Kaminski 1984. Laboratory identification of Saksenaea vasiformis: a rare<br />

cause of zygomycosis in Australia. Sabouraudia: J. Med. Vet. Mycol. 23: 137-140.<br />

Ellis, D.H. and P.J. Keane. 1981. Thermophilic fungi isolated from some Australian soils. Aust.<br />

J. Bot. 29: 689-704.<br />

Ellis, J.J. 1985. Species and varieties in Rhizopus arrhizus - Rhizopus oryzae group as<br />

indicated by their DNA complementarity. Mycologia. 77: 243-247.<br />

Ellis, J.J. 1986. Species and varieties in the Rhizopus microsporus group as indicated by their<br />

DNA complementarity. Mycologia. 78: 508-510.<br />

Ellis, J.J., and L. Ajello. 1982. An unusual source of Aphophysomyces elegans and a method<br />

of stimulating sporulation of Saksenaea vasiformis. Mycologia 74: 144-145.


Descriptions of Medical Fungi 249<br />

REFERENCES<br />

Ellis, J.J. and C.W. Hesseltine. 1966. Two new families of Mucorales. Mycologia. 66: 87-95.<br />

Ellis, J.J. and C.W. Hesseltine. 1965. The genus Absidia: globose spored species. Mycologia.<br />

57: 222-235.<br />

Ellis, J.J. and C.W. Hesseltine. 1966. Species of Absidia with ovoid sporangiospores. II.<br />

Sabouraudia. 5: 59-77.<br />

Ellis, M.B. 1971. Dematiaceous Hyphomycetes. Commonwealth Mycological Institute, Kew,<br />

Surrey, England.<br />

Ellis, M.B. 1976. More Dematiaceous Hyphomycetes. Commonwealth Mycological Institute,<br />

Kew, Surrey, England.<br />

Emmons, C.W. and C.H. Bridges. 1977. Entomophthora coronata, the etiologic agent of a<br />

phycomycosis of horses. Mycologia. 53: 307-312.<br />

Enache-Angoulvant, A. and C. Hennequin. 2005. Invasive Saccharomyces infection: a<br />

comprehensive review. Clin. Infect. Dis. 41: 1559-1568.<br />

Erhard, M., U.C. Hipler, A. Burmester et al. 2008. Identification of dermatophyte species<br />

causing onychomycosis and tinea pedis by MALDI-T<strong>OF</strong> mass spectrometry. Exp Dermatol.<br />

17: 356-361.<br />

Espinel-Ingroff, A., K. Boyle and D.J. Sheehan. 2001. In vitro antifungal activities of<br />

voriconazole and reference agents as determined by NCCLS methods: review of the<br />

literature. Mycopathologia 150: 101-115.<br />

Espinel-Ingroff, A. 2003. In vitro antifungal activities of anidulafungin and micrafungin, licensed<br />

agents and the investigational triazole posaconazole as determined by NCCLS methods<br />

for 12,052 fungal isolates: review of the literature. Rev. Iberoam. Micol. 20: 121-136.<br />

Espinel-Ingroff, A., M.A. Pfaller, B. Bustamante et al. 2014. Multilaboratory study of<br />

epidemiological cutoff values for detection of resistance in eight Candida species to<br />

fluconazole, posaconazole, and voriconazole. Antimicrob. Agents and Chemother. 58:<br />

2006-2012.<br />

Espinel-Ingroff, A., A. Chakrabarti, S. Chowdhary et al. 2015a. Multicenter evaluation<br />

of MIC distributions for epidemiologic cutoff value definition to detect amphotericin B,<br />

posaconazole, and itraconazole resistance among the most clinically relevant species of<br />

Mucorales. Antimicrob. Agents. Chemother. 59: 1745-1750.<br />

Espinel-Ingroff, A., M. Alvarez-Fernandez, E. Cantón et al. 2015b. Multicenter study of<br />

epidemiological cutoff values and detection of resistance in Candida spp. to anidulafungin,<br />

caspofungin, and micafungin using the Sensititre YeastOne colorimetric Method. Antimicrob.<br />

Agents Chemother. 59: 6725-6732.<br />

Espinel-Ingroff, A. and S.E. Kidd. 2015. Current trends in the prevalence of Cryptococcus<br />

gattii in the United States and Canada. Infect. Drug Resist. 8: 89-97.<br />

Estrada-Bárcenas, D.A., T. Vite-Garín, H. Navarro-Barranco et al. 2014. Genetic diversity<br />

of Histoplasma and Sporothrix complexes based on sequences of their ITS1-5.8S-ITS2<br />

regions from the BOLD System. Rev. Iberoam. Micol. 31: 90-94.<br />

Fischer, J. B., and J. Kane. 1971. The detection of contamination in Trichophyton rubrum and<br />

Trichophyton mentagrophytes. Mycopathol. Mycol. Appl. 43: 169-180.<br />

Fisher, M.C., G.L. Koenig, T.J. White and J.W. Taylor. 2002. Molecular and pheno typic<br />

description of Coccidioides posadasii sp. nov., previously recognised as the non-California<br />

population of Coccidioides immitis. Mycologia 94: 73-84.<br />

Formoso, A., D. Heidrich, C.R. Felix et al. 2015. Enzymatic activity and susceptibility to<br />

antifungal agents of Brazilian environmental isolates of Hortaea werneckii. Mycopathologia<br />

180: 345-352.<br />

Fothergill, A.W., M.G. Rinaldi and D.A. Sutton. 2009. Antifungal susceptibility testing of<br />

Exophiala spp.: a head-to-head comparison of amphotericin B, itraconazole, posaconazole<br />

and voriconazole. Medical Mycology 47 (Special Issue): 41-43.


250<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Frankel, D.H. and J.W. Rippon. 1989. Hendersonula toruloidea infection in man.<br />

Mycopathologia 105: 175-186.<br />

Franzot, S.P., I.R. Salkin and A. Casadevall. 1999. Cryptococcus neoformans var. grubii:<br />

separate varietal status for Cryptococcus neoformans serotype A isolates. J. Clin. Microbiol.<br />

37: 838-840.<br />

Gaitanis, G., P. Magiatis, M. Hantschke et al. 2012. The Malassezia genus in skin and<br />

systemic diseases. Clin. Microbiol. Rev. 25: 106–141.<br />

Galgiani, J.N., N.M. Ampel, J.E. Blair et al. 2005. Coccidioidomycosis. Clin. Infect. Dis. 41:<br />

1217-1223.<br />

Gams, W. 1971. Cephalosporium-artige Schimmelpilze (Hyphomycetes). G. Fisher, Stuttgart,<br />

p.262.<br />

Gams, W., M. Christensen, A.H. Onions et al. 1985. Infrageneric taxa of Aspergillus. In:<br />

Advances in Penicillium and Aspergillus Systematics. Samson RA, Pitt JI, eds. New York:<br />

Plenum Press: 55-62.<br />

Garcia-Hermoso, D., D. Hoinard et al. 2009. Molecular and phenotypic evaluation of<br />

Lichtheimia corymbifera (formerly Absidia corymbifera) complex isolates associated with<br />

human mucormycosis: rehabilitation of L. ramosa. J. Clin. Microbiol. 47: 3862-3870.<br />

Garcia-Ruiz, J.C., L. Lopez-Soria, I. Olazabal et al. 2013. Invasive infections caused by<br />

Saprochaete capitata in patients with haematological malignancies: report of five cases<br />

and review of the antifungal therapy. Rev. Iberoam. Micol. 30: 248-255.<br />

Geiser, D.M., M.A. Klich, J.C. Frisvad et al. 2007. The current status of species recognition<br />

and identification in Aspergillus. Stud. Mycol. 59: 1-10.<br />

Geiser, D.M., T. Aoki, C.W. Bacon et al. 2013. One fungus, one name: defining the genus<br />

Fusarium in a scientifically robust way that preserves longstanding use. Phytopathology.<br />

103: 400-408.<br />

George, R.B. and R.L. Penn. 1986. Histoplasmosis. In Fungal diseases of the Lung. eds<br />

Sarosi, G.A. and S.F. Davies. Grune and Stratton Inc.<br />

Gerrits van den Ende, A.H.G. and G.S. de Hoog. 1999. Variability and molecular diagnostics<br />

of the neurotropic species Cladophialophora bantiana. Stud. Mycol. 43: 151-162.<br />

Ghikas, D.V., V.N. Kouvelis and M.A. Typas. 2010. Phylogenetic and biogeographic<br />

implications inferred by mitochondrial intergenic region analyses and ITS1-5.8S-ITS2<br />

of the entomopathogenic fungi Beauveria bassiana and B. brongniartii. BMC Microbiol.<br />

10:174. doi: 10.1186/1471-2180-10-174.<br />

Gilgado, F., J. Cano, J. Gene et al. 2005. Molecular phylogeny of the Pseudallescheria boydii<br />

species complex: proposal of two new species. J. Clin. Microbiol. 43: 4930-4942.<br />

Gilgado, F., J. Gene, J. Cano et al. 2010. Heterothallism in Scedosporium apiospermum and<br />

description of its teleomorph Pseudallescheria apiosperma sp. nov. Med. Mycol. 48: 122-8.<br />

Giraldo, A., J. Gené, D.A. Sutton et al. 2014. Phylogenetic circumscription of Arthrographis<br />

(Eremomycetaceae, Dothideomycetes). Persoonia 32: 102 -114.<br />

Giraldo, A., D.A. Sutton, K. Samerpitak et al. 2014. Occurrence of Ochroconis and Verruconis<br />

species in clinical specimens from the United States. J. Clin. Microbiol. 52: 4189-4201.<br />

Glenn, A., C.W. Bacon, R. Price et al. 1996. Molecular phylogeny of Acremonium and its<br />

taxonomic implications. Mycologia 88:369-383.<br />

Goldschmied-Reouven, A., A. Shvoron, M. Topaz et al. 1989. Saksenaea vasiformis<br />

infection in a burn wound. J. Med. Vet. Mycol. 27: 427-429.<br />

Gomez-Lopez, A., A. Alastruey-Izquierdo, D. Rodrıguez et al. 2008. Prevalence and<br />

susceptibility profile of Candida metapsilosis and Candida orthopsilosis: results from<br />

population-based surveillance of candidemia in Spain. Antimicrob. Agents Chemother. 52:<br />

1506-1509.


Descriptions of Medical Fungi 251<br />

REFERENCES<br />

Gonzalez, G.M., A.W. Fothergill, D.A. Sutton et al. 2005. In vitro activities of new and<br />

established triazoles against opportunistic filamentous and dimorphic fungi. Med. Mycol.<br />

43: 281-284.<br />

Goodman, N.L. and M.G. Rinaldi. 1991. Agents of zygomycosis. In Balows, A., Hau sler, W.J.,<br />

Herrmann, K.L. et al. (eds.), Manual Clinical Microbiology 5th edition. American Society for<br />

Microbiology Washington DC.<br />

Gramaje D, L. Mostert, J.Z. Groenewald et al. 2015. Phaeoacremonium: from esca disease<br />

to phaeohyphomycosis. Fungal Biology 119: 759-783.<br />

Gräser, Y., M. El Fari, W. Presber et al. 1998. Identification of common dermatophytes<br />

(Trichophyton, Microsporum, Epidermophyton) using polymerase chain reactions. Br. J.<br />

Derm. 138: 576-582.<br />

Gräser, Y., M. El Fari, R. Vilgalys et al. 1999a. Phylogeny and taxonomy of the family<br />

Arthrodermataceae (dermatophytes) using sequence analysis of the ribosomal ITS region.<br />

Med. Mycol. 37: 105-114.<br />

Gräser, Y., J. Kühnisch and W. Presber. 1999b. Molecular markers reveal exclusively clonal<br />

reproduction in Trichophyton rubrum. J. Clin. Microbiol. 37: 3713-3717.<br />

Gräser, Y., A.F.A. Kuijpers, M. El Fari et al. 2000a. Molecular and concentional taxonomy of<br />

the Microsporum canis complex. Med. Mycol. 38: 143-153.<br />

Gräser, Y., A.F.A. Kuijpers, W. Presber et al. 2000b. Molecular taxonomy of the Trichophyton<br />

rubrum complex. J. Clin. Microbiol. 38: 3329-3336.<br />

Gräser, Y., S. de Hoog and R.C. Summerbell. 2006. Dermatophytes: recognizing species of<br />

clonal fungi. Med. Mycol. 44: 199-209.<br />

Gräser, Y., J. Scott, and R. Summerbell. 2008. The new species concept in dermatophytes -<br />

a polyphasic approach. Mycopathologia 166: 239-256.<br />

Greer, D.L. and L. Friedman. 1966. Studies on the genus Basidiobolus with reclas sification of<br />

the species pathogenic for man. Sabouraudia. 4: 231-241.<br />

Guarro, J. 2013. Fusariosis, a complex infection caused by a high diversity of fungal species<br />

refractory to treatment. Eur. J. Clin. Microbiol. Infect. Dis. 32: 1491-1500.<br />

Guarro, J., A.S. Kantarcioglus, R. Horre et al. 2006. Scedosporium apiospermum: changing<br />

clinical spectrum of a therapy-refractory opportunist. Medical Mycology 44: 295-327.<br />

Guarro, J., D.K. Mendiratta, H. Sequeira et al. 2007. Acrophialophora fusispora: an emerging<br />

agent of human mycoses. A report of 3 new clinical cases. Diagn. Microbiol. Infect. Dis. 59:<br />

85-88.<br />

Guarro, J., J. Chander, E. Álvarez, et al. 2011. Apophysomyces variabilis infections in<br />

humans. Emerg. Infect. Dis. 17: 134-135.<br />

Gueho, E.S. 1979. Dexoyribonucleic acid base composition and taxonomy in the genus<br />

Geotrichum Link. Antonie van Leeuwenhoek. 45: 199-210.<br />

Gueho, E. and G.S. de Hoog. 1991. Taxonomy of the medical species of Pseudallescheria<br />

and Scedosporium. J. Mycol. Med. 118: 3-9.<br />

Gueho, E., M.Th. Smith, G.S. de Hoog et al. 1992. Contributions to a revision of the genus<br />

Trichosporon. Antonie van Leeuwenhoek. 61: 289-316.<br />

Gueho, E., G. Midgley and J. Guillot. 1996. The genus Malassezia with description of four<br />

new species. Antonie Van Leeuwenhoek. 69: 337-55.<br />

Guillot J. and E. Gueho. 1995. The diversity of Malassezia yeasts confirmed by rRNA<br />

sequence and nuclear DNA comparisons. Antonie Van Leeuwenhoek. 67: 297-314.<br />

Guillot J., E. Gueho, M. Lesourd et al. 1996. Identification of Malassezia species. J. Mycol.<br />

Med. 6: 103-110.<br />

Guillot J., M. Deville, M. Berthelemy et al. 2000. A single PCR-restriction endonuclease<br />

analysis for rapid identification of Malassezia species. Lett. Appl. Microbiol. 31: 400-403.


252<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Guitard, J., A. Angoulvant, V. Letscher-Bru et al. 2013. Invasive infections due to Candida<br />

norvegensis and Candida inconspicua: report of 12 cases and review of the literature.<br />

Medical Mycology 51: 795-799.<br />

Guo, L.N., M. Xiao, F. Kong et al. 2011. Three-locus identification, genotyping, and antifungal<br />

susceptibilities of medically important Trichosporon species from China. J. Clin. Microbiol.<br />

49: 3805-3811.<br />

Gupta, A.K., C.B. Horgan-Bell and R.C. Summerbell. 1998. Onychomycosis associated<br />

with Onychocola canadensis: ten case reports and a review of the literature. J. A. Acad.<br />

Dermatol. 39: 410-407.<br />

Hageage, G.J. and B.J. Harrington. 1984. Use of calcofluor white in clinical mycology.<br />

Laboratory Medicine 15: 109-112.<br />

Hall, M.R., L.M. Brumble, M.A. Mayes et al. 2013. Cutaneous Microsphaeropsis arundinis<br />

infection initially interpreted as squamous cell carcinoma. Int. J. Dermatol. 52: 84-86.<br />

Halliday, C., S.E. Kidd, T.C. Sorrell and S. C-A. Chen. 2015. Molecular diagnostic methods<br />

for invasive fungal disease: the horizon draws nearer? Pathology 47: 257-269.<br />

Harrington, B.J. and G.J. Hageage. 2003. Calcofluor White: A Review of its Uses and<br />

Applications in Clinical Mycology and Parasitology. Laboratory Medicine 34: 361-367.<br />

Heath, C.H., M. A. Slavin, T.C. Sorrell et al. 2009. Population-based surveillance for<br />

scedosporiosis in Australia: epidemiology, disease manifestations and emergence of<br />

Scedosporium aurantiacum infection. Clin. Microbiol. Infect. 15: 689-693.<br />

Hedayati, M.T., A.C. Pasqualotto, P.A. Warn et al. 2007. Aspergillus flavus: human pathogen,<br />

allergen and mycotoxin producer. Microbiology 153: 1677-1692.<br />

Hegedus, D.D. and G.G. Khachatourians. 1996. Identification and differentiation of the<br />

entomopathogenic fungus Beauveria bassiana using polymerase chain reaction and<br />

single-strand conformation polymorphism analysis. J. Invertebr. Pathol. 67: 289-299.<br />

Henrich, T.J., F.M. Marty, D.A. Milner et al. 2009. Disseminated Geotrichum candidum<br />

infection in a patient with relapsed acute myelogenous leukemia following allogeneic stem<br />

cell transplantation and review of the literature. Transpl. Infect. Dis. 11: 458-462.<br />

Hermanides-Nijhof, E.J. 1977. Aureobasidium and allied genera. Stud. Mycol. 15: 141-177.<br />

Hesseltine, C.W. and J.J. Ellis. 1964a. The genus Absidia: Gongronella and cylindricalspored<br />

species of Absidia. Mycologia. 56: 568-601.<br />

Hesseltine, C.W. and J.J. Ellis. 1964b. An interesting case of Mucor, M. ramosissimus.<br />

Sabouraudia. 3: 151-154.<br />

Hesseltine, C.W. and J.J. Ellis. 1966. Species of Absidia with ovoid sporangiospores. I.<br />

Mycologia. 58: 173-194.<br />

Hoffman, K., S. Discher and K. Voigt. 2007. Revision of the genus Absidia (Mucorales,<br />

Zygomycetes) based on physiological, phylogenetic, and morphological characters,<br />

thermotolerant Absidia spp. form a coherent group, Mycocladiaceae fam. nov. Mycol. Res.<br />

111: 1169-1183.<br />

Hoffmann, K., G. Walther, and K. Voigt. 2009. Mycocladus vs. Lichtheimia, a correction<br />

(Lichtheimiaceae fam. nov., Mucorales, Mucoromycotina). Mycol. Res. 113: 277-278.<br />

Hohl, P.E., H.P. Holley, E. Prevost et al. 1983. Infections due to Wangiella dermatitidis in<br />

humans: Report of the first documented case from the United States and a review of the<br />

literature. Reviews of Infectious Diseases. 5: 854-864.<br />

Holland, J. 1997. Emerging zygomycosis of humans: Saksenaea vasiformis and<br />

Apophysomyces elegans. Curr. Top. Med. Mycol. 8: 27-34.<br />

Holländer, H., W. Keilig, J. Bauer, E. Rothemund. 1984. A reliable fluorescent stain for fungi<br />

in tissue sections and clinical specimens. Mycopathologia. 88: 131-134.<br />

Horré, R., G.S. de Hoog, C. Kluczny et al. 1999. rDNA diversity and physiology of Ochroconis<br />

and Scolebasidium species reported from humans and other vertebrates. Stud. Mycol. 43:<br />

194-205.


Descriptions of Medical Fungi 253<br />

REFERENCES<br />

Huguenin, A., A. Lorot and D. Zachar. 2015. Matrix-assisted laser desorption ionizationtime<br />

of flight identification of Schizophyllum commune: perspectives on the review by<br />

Chowdhary et al. Medical Mycology 53: 896-897.<br />

Imai, T., A. Sano, Y. Mikami et al. 2000. A new PCR primer for the identification of<br />

Paracoccidioides brasiliensis on rRNA sequences coding the internal transcribed spacers<br />

(ITS) and 5.8S regions. Med. Mycol. 38: 323-326.<br />

Inderbitzin, P., R.M. Davis, R.M. Bostock, K.V. Subbarao. 2013. Identification and<br />

differentiation of Verticillium species and V. longisporum lineages by simplex and multiplex<br />

PCR assays. PLoS ONE 8(6): e65990.<br />

Irinyi, L., C. Serena, D. Garcia-Hermoso et al. 2015. International Society of Human and<br />

Animal Mycology (ISHAM)-ITS reference DNA barcoding database - the quality controlled<br />

standard tool for routine identification of human and animal pathogenic fungi. Med. Mycol.<br />

53: 313-37.<br />

Irokanulo, E.A.O., C.O. Akueshi and A.A. Makinde. 1994. Differentiation of Cryptococcus<br />

neoformans serotypes A and D using creatinine dextrose bromothymol blue thymine<br />

medium. Br. J. Biomed. Sci. 51: 100-103.<br />

Jackson, L., S.A. Klotz and R.E. Normand. 1996. A pseudoepidemic of Sporothrix cyanescens<br />

pneumonia occurring during renovation of a bronchoscopy suite. J. Med. Vet. Mycol. 28:<br />

455-459.<br />

Jarv, H., J. Lehtmaa, R.C. Summerbell et al. 2004. Isolation of Neosartorya pseudofischeri<br />

from blood: first hint of pulmonary aspergillosis. J. Clin. Microbiol. 42: 925-928.<br />

Jong, S.C. and F.M. Dugan. 2003. Zygomycetes: The Order Entomophthorales. In Howard,<br />

D.H. (ed.), Pathogenic Fungi in Humans and Animals. 2 nd edition, Marcel Dekker Inc., New<br />

York, pp 127-139.<br />

Kanj, S.S., S.S. Amr and G.D. Roberts. 2001. Ramichloridium mackenziei brain abscess:<br />

report of two cases and review of the literature. Med. Mycol. 39: 97-102.<br />

Kaltseis, J., J. Rainer and G.S. de Hoog. 2009. Ecology of Pseudallescheria and Scedosporium<br />

species in human-dominated and natural environments and their distribution in clinical<br />

samples. Med. Mycol. 47: 398-405.<br />

Kane, J., R. Summerbell, L. Sigler et al. 1997. Laboratory handbook of dermatophytes. Star<br />

Publishing Co. Belmont, CA. USA.<br />

Kaneko, T., K. Makimura, M. Abe et al. 2007. Revised Culture-Based System for Identification<br />

of Malassezia Species. J. Clin. Microbiol. 45: 3737-3742.<br />

Kaplan, W. 1977. Protothecosis and infections caused by morphologically similar green algae.<br />

The black and white yeasts. Proceedings of the Fourth International Conference on the<br />

Mycoses. Scientific Publication No. 356. Pan American Health Organization. Washington<br />

D.C. USA.<br />

Kathuria, S., P.K. Singh, J.F. Meis et al. 2014. In Vitro antifungal susceptibility profile<br />

and correlation of mycelial and yeast forms of molecularly characterized Histoplasma<br />

capsulatum strains from India. Antimicrob. Agents and Chemother. 58: 5613-5616.<br />

Katragkou, A., Z.D. Pana, D.S. Perlin et al. 2014. Exserohilum infections: review of 48 cases<br />

before the 2012 United States outbreak. Med. Mycol. 52: 376-386.<br />

Kaufman, L. and P.G. Standard. 1987. Specific and rapid identification of medically important<br />

fungi by exoantigen detection. Ann. Rev. Microbiol. 41: 209-225.<br />

Khan, Z.U., S.J. Lamdhade, M. Johny et al. 2002. Additional case of Ramichloridium<br />

mackenziei cerebral phaeohyphomycosis from the Middle East. Med Mycol. 40: 429-433.<br />

Khan, Z., J. Gené, S. Ahmad et al. 2013. Coniochaeta polymorpha, a new species from<br />

endotracheal aspirate of a preterm neonate, and transfer of Lecythophora species to<br />

Coniochaeta. Antonie van Leeuwenhoek 104: 243-252.


254<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Kidd, S.E., Y Chow, S. Mak et al. 2007. Characterization of environmental sources of the<br />

human and animal pathogen Cryptococcus gattii in British Columbia, Canada, and the<br />

Pacific Northwest of the United States. Appl. Environ. Microbiol. 73: 1433-1443.<br />

King, D.S. 1983. Entomophthorales. In: Howard DH, ed. Fungi pathogenic for humans and<br />

animals. Part A Biology. Marcel Dekker Inc. New York pp 61-73.<br />

Kirk, P., P. Cannon, J. Stalpers and D. Minter. 2008. Dictionary of the Fungi. CABI 784 pp.<br />

Klich, M.A. 2002. Identification of common Aspergillus species. Centraalbureau voor<br />

Schimmelcultures, The Netherlands.<br />

Kluger, E.K., P.K. Della Torre, P. Martin et al. 2004. Concurrent Fusarium chlamydosporium<br />

and Microsphaeropsis arundinis infections in a cat. J. Fel. Med. Surg. 6: 271-277.<br />

Kolecka, A., K. Khayhan, M. Groenewald et al. 2013. MALDI-T<strong>OF</strong> MS identification of<br />

medically relevant species of arthroconidial yeasts. J. Clin. Microbiol. 51: 2491-2500.<br />

Kolecka, A., K. Khayhan, M. Arabatzis et al. 2014. Efficient identification of Malassezia<br />

yeasts by matrix-assisted laser desorption ionization-time of flight mass spectrometry<br />

(MALDI-T<strong>OF</strong> MS). Br. J. Dermatol. 170: 332-341.<br />

Kreger-van Rij, N.J.W. (ed.). 1984. The yeasts, a taxonomic study, 3 rd edition. Elsevier Sci.<br />

Publ., Amsterdam, 1082 pp.<br />

Krockenberger, M.B., P. Martin, C. Halliday et al. 2010. Localised Microsphaeropsis arundinis<br />

infection of the subcutis of a cat. J. Fel. Med. Surg. 12: 231-236.<br />

Kuan, C.S., S.M. Yew, Y.F. Toh et al. 2015. Identification and characterization of a rare fungus,<br />

Quambalaria cyanescens, isolated from the peritoneal fluid of a patient after nocturnal<br />

intermittent peritoneal dialysis. PLoS One 10(12):e0145932.<br />

Kurtzman and J.W. Fell. 1998. The Yeasts: a taxonomic study. 4 th Edition. Elsevier Science<br />

Publishers B.V. Amsterdam.<br />

Kurtzman C.P. 2011. Lodderomyces van der Walt (1971). Chapter 44 In Kurtzman CP, Fell JW,<br />

Boekhout T (ed), The yeasts: a taxonomic study. Elsevier, Amsterdam, the Netherlands.<br />

Kurtzman, C.P., J.W. Fell and T. Boekhout. 2011. The Yeasts, a Taxonomic Study. 5 th Edition<br />

Elsevier B.V.<br />

Kwon-Chung, K.J., I. Polacheck and J.E. Bennett. (1982). Improved diagnostic medium<br />

for separation of Cryptococcus neoformans var. neoformans (Serotypes A and D) and<br />

Cryptococcus neoformans var. gattii (Serotypes B and C). J. Clin. Microbiol. 15: 535-537.<br />

Kwon-Chung, K.J., T. Boekhout, J. Fell and M. Diaz. 2002. Proposal to conserve the<br />

name Cryptococcus gattii against C. hondurianus and C. bacillisporus (Basidiomycota,<br />

Hymenomycetes, Tremellomycetidae). Taxon 51: 804-806.<br />

Kwon-Chung, K.J. and J.W. Bennett. 1992. Medical Mycology. Lea & Febiger, Philadelphia,<br />

861pp.<br />

Lachance, M-A., T. Boekhout, G. Scorzetti et al. 2011. Candida Berhout (1923). Chapter<br />

90 in The Yeasts, a Taxonomic Study, 5 th edition eds Kurtzman, C.P., J.W. Fell and T.<br />

Boekhout, Elsevier B.V. pages 987-1278.<br />

Lackner, M. and G.S. de Hoog. 2011. Parascedosporium and its relatives: phylogeny and<br />

ecological trends. IMA Fungus 2: 39-48.<br />

Lackner, M., G.S. de Hoog, P.E. Verweij et al. 2012a. Species-specific anti-fungal susceptibility<br />

patterns of Scedosporium and Pseudallescheria species. Antimicrob. Agents Chemother.<br />

56: 2635-2642.<br />

Lackner, M., M.J. Najafzadeh, J. Sun et al. 2012b. Rapid identification of Pseudallescheria<br />

and Scedosporium strains using Rolling Circle Amplification. Appl. Environm. Microbiol. 78:<br />

126-133.<br />

Lackner, M., G.S. de Hoog, L. Yang et al. 2014a. Proposed nomenclature for Pseudallescheria,<br />

Scedosporium and related genera. Fungal Div. 67: 1-10.


Descriptions of Medical Fungi 255<br />

REFERENCES<br />

Lackner, M., F. Hagen, J.F. Meis et al. 2014b. Susceptibility and diversity in therapy-refractory<br />

genus Scedosporium. Antimicro. Agents Chemother. 58: 5877-5885.<br />

Lass-Florl, C. and A. Mayr. 2007. Human protothecosis. Clin. Microbiol. Rev. 20: 230-242.<br />

Lau, A.F., S.K. Drake, L.B. Calhoun et al. 2013. Development of a clinically comprehensive<br />

database and a simple procedure for identification of moulds from solid media by Matrix-<br />

Assisted Laser Desorption Ionization - Time of Flight Mass Spectrometry. J. Clin. Microbiol.<br />

51: 828-834.<br />

Lawrence, R.M., W.T. Snodgrass, G.W. Reichel et. al. 1986. Systemic zygomycosis caused<br />

by Apophysomyces elegans. J. Med. Vet. Mycol. 24: 57-65.<br />

Lee, S. and R.T. Hanlin. 1999. Phylogenetic relationships of Chaetomium and similar genera<br />

based on ribosomal DNA sequences. Mycologia 91: 434-442.<br />

Lennon, P.A., C.R. Cooper, Jr., I.F. Salkin et al. 1994. Ribosomal DNA internal transcribed<br />

spacer analysis supports synonomy of Scedosporium inflatum and Lomentospora<br />

prolificans. J . Clin. Microbiol. 32: 2413-2416.<br />

Li, J., J. Xu, F. Bai. 2006. Candida pseudorugosa sp. nov., a novel yeast species from sputum.<br />

J. Clin. Microbiol. 44: 4486-4490.<br />

Liu, J.-K., R. Phookamsak, M. Doilom et al. 2012. Towards a natural classification of<br />

Botryosphaeriales. Fung. Div. 57: 149-210.<br />

L’Ollivier, C., C. Cassagne, A.C. Normand et al. 2013. A MALDI-T<strong>OF</strong> MS procedure for<br />

clinical dermatophyte species identification in the routine laboratory. Med. Mycol. 51: 713-<br />

270.<br />

Lockhart, S.R., S.A. Messer, M.A. Pfaller et al. 2008. Lodderomyces elongisporus<br />

masquerading as Candida parapsilosis as a cause of bloodstream infections. J. Clin.<br />

Microbiol. 46 : 374-376.<br />

Lockhart, S.R., N. Iqbal, C.B. Bolden et al. 2012. Epidemiologic cutoff values for triazole<br />

drugs in Cryptococcus gattii: correlation of molecular type and in vitro susceptibility. Diagn.<br />

Microbiol. Infect. Dis. 73: 144-148.<br />

Lonial, S., L. Williams, G. Carum et al. 1997. Neosartorya fischeri: an invasive fungal<br />

pathogen in an allogeneic bone narrow transplant patient. Bone Marrow Transpl. 19: 753-<br />

755.<br />

Lu, Q., A.H.G. Gerrits van den Ende, J.M.J.E. Bakkers et al. 2011. Identification of<br />

Pseudallescheria and Scedosporium Species by Three Molecular Methods. J. Clin.<br />

Microbiol. 49: 960-967.<br />

Lu, X.L., M.J. Najafzadeh, Y.P. Ran et al. 2013. Taxonomy and epidemiology Mucor irregularis,<br />

agent of chronic cutaneous mucormycosis. Persoonia 30: 48-56.<br />

Lu, X.-l., Z.-h. Liu, Y.-n. Shen et al. 2009. Primary cutaneous zygomycosis caused by<br />

Rhizomucor variabilis: a new endemic zygomycosis? A case report and review of 6 cases<br />

reported from China. Clin. Infect. Dis. 49: e39-e49.<br />

Luangsa-ard, J., J. Houbraken, T. van Doorn et al. 2011. Purpureocillium, a new genus for<br />

the medically important Paecilomyces lilacinus. FEMS Microbiol. Lett. 321: 141-149.<br />

Lunn, J.A. and W.A. Shipton. 1983. Re-evaluation of taxonomic criteria in Cunninghamella.<br />

Trans. Br. Mycol. Soc. 81: 303-312.<br />

Luttrell, E.S. 1978. Biosystematics of Helminthosporium: impact on agriculture. In<br />

Biosystematics in Agriculture. eds. J.A. Romberger et al. Allanheld, Osmon & Co., N.J.<br />

USA.<br />

Lyratzopoulos, G., M. Ellis, R. Nerringer et al. 2002. Invasive infection due to Penicillium<br />

species other than P. marneffei. J. Infect. 45: 184-207.<br />

Machouart, M., P. Menir, R. Helenon et al. 2012. Scytalidium and scytalidiosis: what’s new in<br />

2012? J. Mycol. Méd. 23: 40-46.


256<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Mackenzie, D.W.R., W. Loeffler, A. Mantovani et al. 1986. Guidelines for the prevention,<br />

preservation and control of dermatophytoses in man and animals. World Health Organization.<br />

Madrid, H., M. Ruíz-Cendoya, J. Cano et al. 2009. Genotyping and in vitro antifungal<br />

susceptibility of Neoscytalidium dimidiatum isolates from different origins. Int. J. Antimicrob.<br />

Agents. 34: 351-354.<br />

Madrid, H., K.C. da Cunha, J. Gene et al. 2014. Novel Curvularia species from clinical specimens.<br />

Persoonia 33: 48–6.<br />

Malloch, D. and R.F. Cain. 1972. The Trichocomataceae: Ascomycetes with Aspergillus,<br />

Paecilomyces, and Penicillium imperfect states. Can. J. Bot. 50: 2613-2628.<br />

Malloch, D. and I.F. Salkin. (1984). A new species of Scedosporium associated with<br />

osteomyelidatis in humans. Mycotaxon. 21: 247-255.<br />

Manamgoda, D.S., L. Cai, E.H.C McKenzie et al. 2012. A phylogenetic and taxonomic reevaluation<br />

of the Bipolaris - Cochliobolus - Curvularia complex. Fungal Diversity 56: 131-<br />

144.<br />

Manamgoda, D.S., A.Y. Rossman, L.A. Castlebury et al. 2014. The genus Bipolaris. Stud.<br />

Mycol. 79: 221-288.<br />

Marimón, R., J. Cano, J. Gené, D.A. et al. 2007. Sporothrix brasiliensis, S. globosa, and S.<br />

mexicana, three new Sporothrix species of clinical interest. J. Clin. Microbiol. 45: 3198-<br />

3206.<br />

Marimon, R., C. Serena, J. Gene et al. 2008. In vitro antifungal susceptibilities of five species<br />

of Sporothrix. Antimicrob. Agents Chemother. 52: 732-734.<br />

Matsumoto, T., A.A. Padhye and L. Ajello. 1987. Medical significance of the so-called black<br />

yeasts. Eur. J. Epidemiol. 3: 87-95.<br />

Matsumoto, T., A.A. Padhye, L. Ajello et al. 1984. Critical review of human isolates of<br />

Wangiella dermatitidis. Mycologia. 76: 232-249.<br />

McCullough, M.J., K.V. Clemons, J.H. McCusker et al. 1998. Intergenic transcribed spacer<br />

PCR ribotyping for differentiation of Saccharomyces species and interspecific hybrids. J.<br />

Clin. Microbiol. 36: 1035-1038.<br />

McGinnis, M.R. 1978a. Human pathogenic species of Exophiala, Phialophora, and Wangiella.<br />

In the black and white yeasts. Proceedings of the fourth international conference on the<br />

mycoses. 1978. Scientific Publication No. 356. PAHO. Washington D.C. USA. pp. 37-59.<br />

McGinnis, M.R. 1978b. Taxonomy of Exophiala jeanselmei. Mycopathologia. 65: 79-87.<br />

McGinnis, M.R. 1980. Laboratory handbook of medical mycology. Academic Press.<br />

McGinnis, M.R. and D. Borelli. 1981. Cladosporium bantianum and its synonym Cladosporium<br />

trichoides. Mycotaxon. 13: 127-136.<br />

McGinnis, M.R., W.A. Schell and J. Carson. 1985. Phaeoannellomyces and the<br />

Phaeococcomycetaceae, new dematiaceous blastomycete taxa. J. Med. Vet. Mycol. 23:<br />

179-188.<br />

McGinnis, M.R., D. Borelli, A.A. Padhye and L. Ajello. 1986a. Reclassification of Cladosporium<br />

bantiana in the genus Xylohypha. J. Clin. Microbiol. 23: 1148-1151.<br />

McGinnis, M.R., M.G. Rinaldi and R.E. Winn. 1986b. Emerging agents of Phaeohyphomycosis:<br />

pathogenic species of Bipolaris and Exserohilum. J. Clin. Microbiol. 24: 250-259.<br />

McGinnis, M.R. and A.A. Padhye. 1977. Exophiala jeanselmei, a new combination for<br />

Phialophora jeanselmei. Mycotaxon. 5: 341-352.<br />

McGinnis, M.R., A.A. Padhye and L. Ajello. 1982. Pseudallescheria Negroni et Fischer, 1943<br />

and its later synonym Petriellidium Malloch, 1970. Mycotaxon 9: 94-102.<br />

McGinnis, M.R., L. Pasarell, D.A. Sutton et al. 1997. In vitro evaluation of voriconazole<br />

against some clinically important fungi. Antimicrob. Agents Chemother. 41: 1832-1834.<br />

McGinnis, M.R. and L. Pasarell. 1998a. In vitro testing of susceptibilities of filamentous<br />

ascomycetes to voriconazole, itraconazole, and amphotericin B, with consideration of<br />

phylogenetic implications. J. Clin. Microbiol. 36: 2353-2355.


Descriptions of Medical Fungi 257<br />

REFERENCES<br />

McGinnis, M.R. and L. Pasarell. 1998b. In vitro evaluation of terbinafine and itraconazole<br />

against dematiaceous fungi. Medical Mycology. 36: 243-246.<br />

McTaggart, L., S.E. Richardson, C. Seah et al. 2013. Rapid identification of Cryptococcus<br />

neoformans var. grubii, C. neoformans var. neoformans, and C. gattii by use of rapid<br />

biochemical tests, differential media, and DNA sequencing. J. Clin. Microbiol. 49: 2522-<br />

2527.<br />

Michel, J., D. Maubon, D.A. Varoquaux et al. 2015. Schizophyllum commune: an emergent<br />

or misdiagnosed fungal pathogen in rhinology? Med. Mycol., 2015, 00, 1-9, doi: 10.1093/<br />

mmy/myv084.<br />

Millner, P.D. 1975. Radial growth responses to temperature by 58 Chaetomium species, and<br />

some taxonomic relationships. Mycologia 69: 492-502.<br />

Miranda, K.C., C.R. de Araujo, C.R. Costa et al. 2007. Antifungal activities of azole agents<br />

against the Malassezia species. Int. J. Antimicrob. Agents. 29: 281-284.<br />

Miranda-Zapico, I., E. Eraso, J.L. Hernández-Almaraz et al. 2011. Prevalence and antifungal<br />

susceptibility patterns of new cryptic species inside the species complexes Candida<br />

parapsilosis and Candida glabrata among blood isolates from a Spanish tertiary hospital.<br />

J. Antimicrob. Chemother. 66: 2315-2322.<br />

Mirhendi, H., K. Makimura, G.S. de Hoog et al. 2015. Translation elongation factor 1-α gene<br />

as a potential taxonomic and identification marker in dermatophytes. Med. Mycol. 53: 215-<br />

224.<br />

Misra, P.C., K.J. Srivastava and K. Latas. 1979. Apophysomyces, a new genus of the<br />

Mucorales. Mycotaxon. 8: 377-382.<br />

Mochizuki, T., K. Anzawa, Y. Sakata et al. 2013. Simple identification of Trichophyton tonsurans<br />

by chlamydospore-like structures produced in culture media. J. Dermatol. 40:.1027-1032.<br />

Mok, W.Y. 1982. Nature and identification of Exophiala werneckii. J. Clin. Microbiol. 16: 976-<br />

978.<br />

Monheit, J.E., D.F. Cowan, and D.G. Moore. 1984. Rapid detection of fungi in tissues using<br />

calcofluor white and fluorescence microscopy. Arch. Pathol. Lab. Med. 108: 616-618.<br />

Montel, E., P.D. Bridge and B.C. Sutton. 1991. An integrated approach to Phoma systematics.<br />

Mycopathologia 115: 89-103.<br />

Moore, M.K. 1986. Hendersonula toruloidea and Scytalidium hyalinum infections in London,<br />

England. J. Med. Vet. Mycol. 24: 219-230.<br />

Morjaria, S., C. Otto, A. Moreira et al. 2015. Ribosomal RNA gene sequencing for early<br />

diagnosis of Blastomyces dermatitidis infection. Int. J. Infect. Dis. 37: 122-124.<br />

Morton, F.J. and G. Smith. 1963. The genera Scopulariopsis Bainier, Microascus Zukal, and<br />

Doratomyces Corda. Mycological Papers, No. 86. Commonwealth Mycological Institute,<br />

Kew, London.<br />

Mostert, L., J.Z. Groenewald, R.C. Summerbell et al. 2006. Taxonomy and pathology of<br />

Togninia (Diaporthales) and its Phaeoacremonium anamorphs. Stud. Mycol. 54: 1-115.<br />

Mostert, L., J.Z. Groenewald, R.C. Summerbell et al. 2005. Species of Phaeoacremonium<br />

associated with infections in humans and environmental reservoirs in infected woody<br />

plants. J. Clin. Microbiol. 43: 1752-1767.<br />

Najafzadeh, M.J., C. Gueidan, H. Badali et al. 2009. Genetic diversity and species delimitation<br />

in the opportunistic genus Fonsecaea. Med. Mycol. 47: 17-25.<br />

Najafzadeh, M.J., H. Badali, M.T. Illnait-Zaragozi et al. 2010a. In vitro activities of eight<br />

antifungal drugs against 55 clinical isolates of Fonsecaea spp. J. Clin. Microbiol. 54: 1636-<br />

1638.


258<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Najafzadeh, M.J., J. Sun, V. Vicente et al. 2010b. Fonsecaea nubica sp. nov, a new agent<br />

of human chromoblastomycosis revealed using molecular data. Med. Mycol. 48: 800-806.<br />

Najafzadeh, M.J., D.A. Sutton, M. S. Keisari et al. 2014. In Vitro Activities of Eight Antifungal<br />

Drugs against 104 Environmental and Clinical Isolates of Aureobasidium pullulans.<br />

Antimicrob. Agents and Chemother. 58: 5629-5631.<br />

Nakamura, Y., R. Kano, T. Mural et al. 2000. Susceptibility testing of Malassezia species<br />

using the urea broth microdilution method. Antimicrob. Agents Chemother. 44: 2185-2186.<br />

Nenoff, P., M. Erhhar, J. C. Simon et al. 2013. MALDI-T<strong>OF</strong> mass spectrometry - a rapid<br />

method for the identification of dermatophyte species. Med. Mycol. 51: 17–24.<br />

Ng, K.P., T.S. Soo-Hoo, S.L. Na et al. 2005. The mycological and molecular study of Hortaea<br />

werneckii isolated from blood and splenic abscess. Mycopathologia 159: 495-500.<br />

Nishimura, K. and M. Miyaji. 1983. Studies on the phylogenesis of pathogenic “black yeasts”.<br />

Mycopathologia 81: 135-144.<br />

Nobrega de Almeida J., L.B. de Souza, A.L. Motta et al. 2014. Evaluation of the MALDI-T<strong>OF</strong><br />

VITEK MS system for the identification of Candida parapsilosis, C. orthopsilosis and C.<br />

metapsilosis from bloodstream infections. J. Microbiol. Methods. 105: 105-108.<br />

Nottebrock, H., H.J. Scholer and M. Wall. 1974. Taxonomy and identification of mucormycosis<br />

causing fungi. 1. Synonymity of Absidia ramosa with A. corymbifera. Sabouraudia 12: 64-<br />

74.<br />

Nucci, M. and E. Anaissie. 2007. Fusarium Infections in Immunocompromised Patients. Clin.<br />

Microbiol. Rev. 20: 695-704.<br />

O’Donnell, K.L. 1979. Zygomycetes in culture. Palfrey Contributions in Botany 2. University<br />

of Georgia. pp 257.<br />

O’Donnell, K., D.A. Sutton, A. Fothergill et al. 2008. Molecular phylogenetic diversity,<br />

multilocus haplotype nomenclature, and in vitro antifungal resistance within the Fusarium<br />

solani species complex. J. Clin. Microbiol. 46: 2477-2490.<br />

O’Donnell, K., C. Gueidan, S. Sink et al. 2009a. A two-locus DNA sequence database for<br />

typing plant and human pathogens within the Fusarium oxysporum species complex.<br />

Fungal Genetics and Biology 46: 936-948.<br />

O’Donnell, K., D.A. Sutton, M.G. Rinaldi et al. 2009b. Novel multilocus sequence typing<br />

scheme reveals high genetic diversity of human pathogenic members of the Fusarium<br />

incarnatum-F. equiseti and F. chlamydosporum species complexes within the United<br />

States. J. Clin. Microbiol. 47: 3851-3861.<br />

O’Donnell, K., T.J. Ward, V.A.R.G. Robert et al. 2015. DNA sequence-based identification of<br />

Fusarium: Current status and future directions. Phytoparasitica 43: 583-595.<br />

Ohori, A., S. Endo, A. Sano et al. 2006. Rapid identification of Ochroconis gallopava by a<br />

loop-mediated isothermal amplification (LAMP) method. Vet. Microbiol. 114: 359-365.<br />

Okada, G., T. Kirisits, G.W. Louis-Seize et al. 2000. Epitypification of Graphium penicillioides<br />

Corda, with comments on the phylogeny and taxonomy of Graphium-like synnematous<br />

fungi. Stud. Mycol. 45: 169-188.<br />

Oliveira, D.C., P.G. Lopes, T.B. Spader et al. 2011. Antifungal susceptibilities of Sporothrix<br />

albicans, S. brasiliensis, and S. luriei of the S. schenckii complex identified in Brazil. J Clin<br />

Microbiol. 49: 3047-3049.<br />

Oliveira, M.M.E., R. Almeida-Paesa, M.C. Gutierrez-Galhardob et al. 2014. Molecular<br />

identification of the Sporothrix schenckii complex. Rev Iberoam Micol. 31: 2-6.<br />

Onions, A.H.S., D. Allsopp and H.O.W. Eggins. 1981. Smith’s introduction to industrial<br />

mycology. Edward Arnold.


Descriptions of Medical Fungi 259<br />

REFERENCES<br />

Packeu, A., M. Hendrickx, H. Beguin et al. 2013. Identification of the Trichophyton<br />

mentagrophytes complex species using MALDI-T<strong>OF</strong> mass spectrometry. Med. Mycol. 51:<br />

580-585.<br />

Packeu, A., A. De Bel, C. l’Ollivier et al. 2014. Fast and accurate identification of dermatophytes<br />

by matrix-assisted laser desorption ionization-time of flight mass spectrometry: validation<br />

in the clinical laboratory. J. Clin. Microbiol. 52: 3440-3443.<br />

Padhye, A.A. and J.W. Carmichael. 1972. Arthroderma insingulare sp. nov. another<br />

Gymnoascaceous state of the Trichophyton terrestre complex. Sabouraudia 10: 47-51.<br />

Padhye, A.A., and L. Ajello 1988. Simple method of inducing sporulation by Apophysomyces<br />

elegans and Saksenaea vasiformis. J. Clin. Microbiol. 26: 1861-1863.<br />

Padhye, A.A., G. Koshi, V. Anandi et. al. 1988. First case of subcutaneous zygomycosis<br />

caused by Saksenaea vasiformis in India. Diagn. Microbiol. Infect. Dis. 9: 69-77.<br />

Padhye, A.A., G. Smith, D. McLaughlin et al. 1992. Comparative evaluation of a<br />

chemiluminescent DNA probe and exoantigen test for rapid identification of Histoplasma<br />

capsulatum. J. Clin. Microbiol. 30: 3108-3111.<br />

Padhye, A.A., J.H. Godfrey, F.W. Chandler et al. 1994a. Osteomyelitis caused by Neosartorya<br />

pseudofischeri. J. Clin. Microbiol. 32: 2832-2836.<br />

Padhye, A.A., G. Smith, P.G. Standard et al. 1994b. Comparative evaluation of<br />

chemiluminescent DNA probe assays and exoantigen tests for rapid identification of<br />

Blastomyces dermatitidis and Coccidioides immitis. J. Clin. Microbiol. 32: 867-870.<br />

Paredes, K., D.A. Sutton, J. Cano. et al. 2012. Molecular identification and antifungal<br />

susceptibility testing of clinical isolates of the Candida rugosa species complex and<br />

proposal of the new species Candida neorugosa. J. Clin. Microbiol. 50: 2397-2403.<br />

Pastor, F.J. and J. Guarro. 2008. Alternaria infections: laboratory diagnosis and relevant<br />

clinical features. Clin. Micribiol. Infect. 14: 734-746.<br />

Perdomo, H., D.A. Sutton, D. Garcia et al. 2011a. Spectrum of clinically relevant Acremonium<br />

species in the United States. J. Clin. Microbiol. 49: 243-256.<br />

Perdomo, H., D.A. Sutton, D. García et al. 2011b. Molecular and phenotypic characterization<br />

of Phialemonium and Lecythophora isolates from clinical samples. J. Clin. Microbiol. 49:<br />

1209-1216.<br />

Pendle, S., K. Weeks, M. Priest et al. 2004. Phaehyphomycotic soft tissue infections caused<br />

by the Coelomycetous fungus Microsphaeropsis arundis. J. Clin. Microbiol. 42: 5315-5319.<br />

Perdomo, H., J. Cano, J. Gené et al. 2013. Polyphasic analysis of Purpureocillium lilacinum<br />

isolates from different origins and proposal of the new species Purpureocillium lavendulum.<br />

Mycologia 105: 151-161.<br />

Peterson, S.W. 2000. Phylogenetic relationships in Aspergillus based on rDNA sequence<br />

analysis. In R.A. Samson & J.I. Pitt (eds): Integration of Modern Taxonomic Methods for<br />

Penicillium and Aspergillus Classification, pp. 323-355.<br />

Peterson, S.W. 2008. Phylogenetic analysis of Aspergillus species using DNA sequences<br />

from four loci. Mycologia 100: 205-226.<br />

Pfaller, M.A., and D.J. Diekema. 2010. Epidemiology of invasive mycoses in North America.<br />

Crit. Rev. Microbiol. 36: 1-53.<br />

Pfaller, M.A., M. Castanheira, D.J. Diekema et al. 2011. Wild-type MIC distributions and<br />

epidemiologic cutoff values for fluconazole, posaconazole, and voriconazole when<br />

testing Cryptococcus neoformans as determined by the CLSI broth microdilution method.<br />

Diagnostic Microbiology and Infectious Disease 71: 252-259.<br />

Pfaller, M.A., and D.J. Diekema. 2012. Progress in antifungal susceptibility testing of Candida<br />

spp. by use of Clinical and Laboratory Standards Institute broth microdilution methods,<br />

2010 to 2012. J. Clin. Microbiol. 50: 2846-2856.


260<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Pfaller, M.A., S.A. Messer, L.N. Woosley et al. 2013. Echinocandin and Triazole antifungal<br />

susceptibility profiles for clinical opportunistic yeast and mold isolates collected from 2010<br />

to 2011: application of new CLSI clinical breakpoints and epidemiological cutoff values<br />

for characterization of geographic and temporal trends of antifungal resistance. J. Clin.<br />

Microbiol. 51: 2571-2581.<br />

Pfaller, M.A., P.R. Rhomberg, S.A. Messer et al. 2015. Isavuconazole, micafungin, and<br />

8 comparator antifungal agents’ susceptibility profiles for common and uncommon<br />

opportunistic fungi collected in 2013: temporal analysis of antifungal drug resistance using<br />

CLSI species-specific clinical breakpoints and proposed epidemiological cutoff values.<br />

Diagnostic Microbiology and Infectious Disease 82: 303-313.<br />

Phillips, A.J., A. Alves, J. Abdollahzadeh et al. 2013. The Botryosphaeriaceae: genera and<br />

species known from culture. Stud. Mycol. 76: 51-167.<br />

Pitt, J.I. 1979. The genus Penicillium and its teleomorphic states Eupenicillium and Talaromyces.<br />

Academic Press.<br />

Pore, R.S. 1985. Prototheca taxonomy. Mycopathologia 129: 129-139.<br />

Pritchard, R.C., D.B. Muir, K.H. Archer et al. 1986. Subcutaneous zygomycosis due to<br />

Saksenaea vasiformis in an infant. Med. J. Aust. 145: 630-631.<br />

Pujol, I., C. Aguilar, J. Gene, J. Guarro. 2000. In vitro antifungal susceptibility of Alternaria<br />

spp. and Ulocladium spp. J. Antimicrob. Chemother. 46: 337.<br />

Punithalingam, E. 1979. Sphaeropsidales in culture from humans. Nova Hedwigia. 31: 119-<br />

158.<br />

Pryor, B.M. and R.L. Gilbertson. 2000. Molecular phylogenetic relationships amongst<br />

Alternaria species and related fungi based upon analysis of nuclear ITS and mt SSU rDNA<br />

sequences. Mycol. Res. 104: 1312-1321.<br />

Rainer, J. and G.S. de Hoog. 2006. Molecular taxonomy and ecology of Pseudallescheria,<br />

Petriella and Scedosporium prolificans (Microascaceae) containing opportunistic agents<br />

on humans. Mycol. Res. 110: 151-160.<br />

Ramani, R. and V. Chaturvedi. 2007. Antifungal susceptibility profiles of Coccidioides immitis<br />

and Coccidioides posadasii from endemic and non-endemic areas. Mycopathologia 163:<br />

315-319.<br />

Ramirez, C. 1982. Manual and atlas of the Penicillia. Elsevier Biomedical Press.<br />

Ramos, L.S., M.H.G. Figueiredo-Carvalho, L.S. Barbedo et al. 2015. Candida haemulonii<br />

complex: species identification and antifungal susceptibility profiles of clinical isolates from<br />

Brazil. Antimicrob. Chemother. 70: 111-115.<br />

Raper, K.B. and D.I. Fennell. 1965. The genus Aspergillus. William & Wilkins Co., Baltimore.<br />

Raper, K.B. and C.H. Thom. 1949. A manual of the penicillia. William & Wilkins Co., Baltimore.<br />

Rebell, G. and D. Taplin. 1970. The Dermatophytes. 2nd. revised edition. University of Miami<br />

Press, Coral Gables, Florida. USA.<br />

Rehner, S.A. and E. Buckley 2005. A Beauveria phylogeny inferred from nuclear ITS and EF1-<br />

alpha sequences: evidence for cryptic diversification and links to Cordyceps teleomorphs.<br />

Mycologia 97: 84-98.<br />

Reppas, G., T. Gottlieb, M. Krockenberger et al. 2015. Microsphaeropsis arundinis an<br />

emerging cause of phaeohyphomycosis in cats and people. Microbiol. Australia 36: 74-78.<br />

Revankar, S.G. and D. A. Sutton. 2010. Melanized Fungi in Human Disease. Clin. Microbiol.<br />

Rev. 23: 884-928.<br />

Riddel R.W. 1950. Permanent stained mycological preparations obtained by slide culture.<br />

Mycologia 42: 265-270.<br />

Rippon, J.W. 1988. Medical Mycology. 3rd Edition. W.B. Saunders Co.<br />

Rippon, J.W., P.M. Arnow, R.A. Larson et al. 1985. “Golden tongue” syndrome caused by<br />

Ramichloridium schulzeri. Arch. Dermatol. 121: 892-894.


Descriptions of Medical Fungi 261<br />

REFERENCES<br />

Rodriguez-Tudela, J.L., T.M. Diaz-Guerra, E. Mellado et al. 2005. Susceptibility patterns<br />

and molecular identification of Trichosporon species. Antimicrob. Agents Chemother. 49:<br />

4026-4034.<br />

Rodriguez-Tudela, J.L., J. Berenguer, J. Guarro et al. 2009. Epidemiology and outcome of<br />

Scedosporium prolificans infection, a review of 162 cases. Med. Mycol. 47: 359-370.<br />

Rodrigues, A.M., G.S. de Hoog, D. de Cássia Pires et al. 2014. Genetic diversity and<br />

antifungal susceptibility profiles in causative agents of sporotrichosis. BMC Infect. Dis.<br />

14:219. doi: 10.1186/1471-2334-14-219.<br />

Romeo, O., F. Scordino and G. Criseo. 2011. New insight into molecular phylogeny and<br />

epidemiology of Sporothrix schenckii species complex based on calmodulin encoding<br />

gene analysis of Italian isolates. Mycopathologia 172: 179-86.<br />

Sabatelli, F., R. Patel, P.A. Mann et al. 2006. In vitro activities of posaconazole, fluconazole.<br />

itraconazole, voriconazole, and amphotericin B against a large col lection of clinically<br />

important moulds and yeasts. Antimicrob. Agents Chemother. 50: 2009-2015.<br />

Saksena, S.B. 1953. A new genus of Mucorales. Mycologia 45: 426-436<br />

Salah, H., A.M.S. Al-Hatmi, B. Theelen et al. 2015. Phylogenetic diversity of human pathogenic<br />

Fusarium and emergence of uncommon virulent species. J. Infect. 71: 658-666.<br />

Salkin, I.F., M.R. McGinnis, M.J. Dykstra and M.G. Rinaldi. 1988. Scedosporium inflatum,<br />

an emerging pathogen. J. Clin. Microbiol. 26: 498-503.<br />

Samerpitak, K., E. van der Linde, H.J. Choi et al. 2014. Taxonomy of Ochroconis, genus<br />

including opportunistic pathogens on humans and animals. Fung. Div. 65: 89-126.<br />

Samson, R.A. 1969. Revision of the genus Cunninghamella (Fungi, Mucorales). Proceedings<br />

of the Koninklijke Nederlandse Akademie van Wetenschappen, ser. C, 72: 322-335.<br />

Samson, R.A. 1974. Paecilomyces and some allied hyphomycetes. Stud. Mycol. No. 6. Baarn,<br />

The Netherlands.<br />

Samson, R.A., 1979. A compilation of the Aspergilli described since 1965. Stud. Mycol. 18:<br />

1-40.<br />

Samson, R.A., E.S. Hoekstra, J.C. Frisvad and O. Filtenborg. 1995. Introduction to foodborne<br />

fungi. Centraalbureau voor Schimmelcultures, P.O.Box 273, 3740 AG BAARN, The<br />

Netherlands.<br />

Samson, R.A. and J.I. Pitt (eds). 1990. Modern concepts in Penicillum and Aspergillus<br />

classification. Plenum Press, New York, USA.<br />

Samson, R.A. and J.I. Pitt (eds). 2000. Integration of Modern Taxonomic Methods for<br />

Penicillium and Aspergillus Classification. Harwood, Amsterdam, 510 pp.<br />

Samson, R.A., S. Hong, S.W. Peterson et al. 2007 Polyphasic taxonomy of Aspergillus<br />

section Fumigati and its teleomorph Neosartorya. Stud. Mycol. 59: 147-203.<br />

Samson, R.A., J. Varga and J.C. Frisvad (eds). 2011a. Taxonomic studies on the genus<br />

Aspergillus. Stud. Mycol. 69: 1-97.<br />

Samson, R.A., N. Yilmaz, J. Houbraken et al. 2011b. Phylogeny and nomenclature of the<br />

genus Talaromyces and taxa accommodated in Penicillium subgenus Biverticillium. Stud.<br />

Mycol. 70: 159-183.<br />

Samson, R.A., C.M. Visagie, J. Houbraken et al. 2014. Phylogeny, identification and<br />

nomenclature of the genus Aspergillus. Stud. Mycol. 78: 141-173.<br />

Sandoval-Denis, M., J. Gené, D.A. Sttuon et al. 2015. Acrophialophora, a poorly known<br />

fungus with clinical significance. J. Clin. Microbiol. 53: 1549-55.<br />

Sandoval-Denis, M., D.A. Sutton, A.W. Fothergill et al. 2013. Scopulariopsis, a poorly<br />

known opportunistic fungus: spectrum of species in clinical samples and in vitro responses<br />

to antifungal drugs. J. Clin. Microbiol. 51: 3937-3943.


262<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Sandoval-Denis, M., A. Giraldo, D.A. Sutton et al. 2014a. In vitro antifungal susceptibility of<br />

clinical isolates of Arthrographis kalrae, a poorly known opportunistic fungus. Mycoses 57:<br />

247-248.<br />

Sandoval-Denis, M., D.A. Sutton, J.F. Cano-Lira et al. 2014b. Phylogeny of the clinically<br />

relevant species of the emerging fungus Trichoderma and their antifungal susceptibilities.<br />

J. Clin. Microbiol. 52: 2112-2125.<br />

Sandoval-Denis, M., D.A. Sutton, A. Martin-Vicente et al. 2015. Cladosporium species<br />

recovered from clinical samples in the United States. J. Clin. Microbiol. 53: 2990-3000.<br />

Scalarone, G.M., A.M. Legendre, K.A. Clark et al. 1992. Evaluation of a commercial DNA<br />

probe assay for the identification of clinical isolates of Blastomyces dermatitidis from dogs.<br />

J. Med. Vet. Mycol. 30: 43-49.<br />

Scheel, C.M., Y. Zhou, R.C. Theodoro et al. 2014. Development of a loop-mediated isothermal<br />

amplification method for detection of Histoplasma capsulatum DNA in clinical samples. J.<br />

Clin. Microbiol. 52: 483-488.<br />

Schell, W.A., M.R. McGinnis and D. Borelli. 1983. Rhinocladiella aquaspora a new<br />

combination for Acrotheca aquaspersa. Mycotaxon 17: 341-348.<br />

Schipper, M.A.A. 1976. On Mucor circinelloides, Mucor racemosus and related species. Stud.<br />

Mycol. 12: 1-40.<br />

Schipper, M.A.A. 1978. 1. On certain species of Mucor with a key to all accepted species.<br />

2. On the genera Rhizomucor and Parasitella. Stud. Mycol. No.17. Centraalbureau voor<br />

Schimmelcultures, Baarn, The Netherlands.<br />

Schipper, M.A.A. 1984. A revision of the genus Rhizopus 1. The Rhizopus stolonifer group<br />

and Rhizopus oryzae. Stud. Mycol. 25: 1-19.<br />

Schipper, M.A.A. and J.A. Stalpers. 1984. A revision of the genus Rhizopus II. The Rhizopus<br />

microsporus group. Stud. Mycol. 25: 30-34.<br />

Schipper, M.A.A. and J.A. Stalpers. 2003. Zygomycetes: The Order Mucorales. In Howard,<br />

D.H. (ed.), Pathogenic Fungi in Humans and Animals. 2 nd edition, Marcel Dekker Inc., New<br />

York, 67-125.<br />

Schipper, M.A.A., M.M. Maslen, G.G. Hogg et al. 1996. Human infection by Rhizopus<br />

azygosporus and the occurrence of azygospores in Zygomycetes. J. Med. Vet. Mycol. 34:<br />

199-203.<br />

Schmidt, G., L. Calanni, M. Iacono, and R. Negroni. 2000. Cerinosterus cyanescens<br />

fungemia: report of a case. Proc. 14th ISHAM Congr., B. Aires, p. 272.<br />

Scholer, H.J., E. Müller and M.A.A. Schipper. 1983. Mucorales. In: Howard DH, ed. Fungi<br />

pathogenic for humans and animals, Part A Biology. Marcel Dekker Inc New York, pp 9-59.<br />

Schroers, H-J. K. O’Donnell, S.C. Lamprecht et al. 2009. Taxonomy and phylogeny of the<br />

Fusarium dimerum species group. Mycologia 101: 44-70.<br />

Schrödl, W., T. Heydel, V.U. Schwartze et al. 2012. Direct analysis and identification of<br />

pathogenic Lichtheimia species by matrix-assisted laser desorption ionization-time of flight<br />

analyzer-mediated mass spectrometry. J. Clin. Microbiol. 50: 419-427.<br />

Schubert, K., J.Z. Groenewald, U. Braun et al. 2007. Biodiversity in the Cladosporium<br />

herbarum complex (Davidiellaceae, Capnodiales), with standardization of methods for<br />

Cladosporium taxonomy and diagnostics. Stud Mycol. 58:105–156.<br />

Serena, C., M. Ortoneda, J. Capilla et al. 2003. In Vitro Activities of New Antifungal Agents<br />

against Chaetomium spp and Inoculum Standardization. Antimicrobial. Agents Chemoth.<br />

47: 3161-3164.<br />

Seth, H.K. 1970. A monograph of the genus Chaetomium. Nova Hedwigia 37:1-134.<br />

Seyedmousavi, S., K. Samerpitak, A.J.M.M. Rijs et al. 2014. Antifungal Susceptibility<br />

Patterns of Opportunistic Fungi in the Genera Verruconis and Ochroconis. Antimicro.<br />

Agents Chemoth. 58: 3285-3292.


Descriptions of Medical Fungi 263<br />

REFERENCES<br />

Sfakianakis, A., K. Krasagakis, M. Stefanidou et al. 2007. Invasive cutaneous infection with<br />

Geotrichum candidum - sequential treatment with amphotericin B and voriconazole. Med<br />

Mycol. 45: 81-84.<br />

Shipton, W.A. and P. Zahari. 1987. Sporulation media for Basidiobolus species. J. Med. Vet.<br />

Mycol. 25: 323-327.<br />

Sidamonidze, K., M.K. Peck, M. Perez et al. 2012. Real-time PCR assay for identification of<br />

Blastomyces dermatitidis in culture and in tissue. J. Clin. Microbiol. 50: 1783-1786.<br />

Sigler, L., S.P. Abbott and A.J. Woodgyer. 1994. New records of nail and skin infec tion due<br />

to Onychocola canadensis and description of its teleomorph Arachnomyces nodosetosus<br />

sp. nov. J. Med. Vet. Mycol. 32: 275-285.<br />

Sigler, L. and J.W. Carmichael. 1976. Taxonomy of Malbranchea and some other<br />

hyphomycetes with arthroconidia. Mycotaxon 4: 349-488.<br />

Sigler, L. and H. Congly. 1990. Toenail infection caused by Onychocola canadensis gen. et.<br />

sp. nov. J. Med. Vet. Mycol. 28: 405-417.<br />

Sigler, L., L.M. de la Maza, G. Tan et al. 1995. Diagnostic difficulties caused by a nonclamped<br />

Schizophyllum commune isolate in a case of fungus ball of the lung. J. Clin. Microbiol. 33:<br />

1979-1983.<br />

Silveira, C.P., J.M. Torres-Rodrıguez, E. Alvarado-Ramırez et al. 2009. MICs and minimum<br />

fungicidal concentrations of amphotericin B, itraconazole, posaconazole and terbinafine in<br />

Sporothrix schenckii. J. Med. Microbiol. 58: 1607-1610.<br />

Simmons, E.G. 1967. Typification of Alternaria, Stemphylium and Ulocladium. Mycologia 59:<br />

67-92.<br />

Simmons, E.G. 2007. Alternaria, an Identification Manual. CBS Biodiv. Ser. 6: 1-775.<br />

Simpson, J.A. 2000. Quambalaria, a new genus of eucalypt pathogens. Australasian<br />

Mycologist 19: 57-62.<br />

Sitterlé, E., S. Giraud, J. Leto et al. 2014. Matrix-assisted laser desorption ionizationtime<br />

of flight mass spectrometry for fast and accurate identification of Pseudallescheria/<br />

Scedosporium species. Clinical Microbiology and Infection 20: 929-935.<br />

Sivanesan, A. 1987. Graminicolous species of Bipolaris, Curvularia, Drechslera, Exserohilum<br />

and their teleomorphs. Mycological Paper No. 158. CAB International, U.K.<br />

Skora, M., A.B. Macura and M. Bulanda. 2014. In vitro antifungal susceptibility of Scopulariopsis<br />

brevicaulis isolates. Medical Mycology 52: 723-727.<br />

Sleiman, S., C. Halliday, A.F. Lau et al. 2015. Species identification of filamentous fungi directly<br />

from solid culture media using MALDI-T<strong>OF</strong> MS. Poster ISHAM Congress, Melbourne.<br />

Spiliopoulou, A., E.D. Anastassiou, and M. Christofidou. 2012. Rhodotorula fungemia of an<br />

intensive care unit patient and review of published cases. Mycopathologia 174: 301-309.<br />

Staib F. 1987. Cryptococcus in AIDS Mycological Diagnostic and Epidemiological Observations.<br />

Aids Forshung (AIFO) 2: 363-382.<br />

Stchigel, A.M., D.A. Sutton, J.F. Cano-Lira et al. 2014. Phylogeny of chrysosporia infecting<br />

reptiles: proposal of the new family Nannizziopsiaceae and five new species. Persoonia<br />

31: 86-100.<br />

Steele, T., G.W. Kaminski and D. Hansman. 1977. A case of coccidioidomycosis in Australia.<br />

Med. J. Aust 1: 968-969.<br />

Steinbach, W.J., J.-P. Latgé and D.A. Stevens (eds). 2005. Advances against aspergillosis.<br />

Med. Mycol. 43, Suppl. 1: S1-S319.<br />

Sorrell, T. C. 2001. Cryptococcus neoformans variety gattii. Med. Mycol. 39: 155-168.<br />

Strinivasan, M.C. and M.J. Thirumalachar. 1965. Basidiobolus species pathogenic for man.<br />

Sabouraudia 4: 32-34.<br />

Sugar, A.M. and X.P. Liu. 1996. In vitro and in vivo activities of SCH 56592 against Blastomyces<br />

dermatitidis. Antimicrob. Agents Chemother. 40: 1314-1316.


264<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Sugita, T. 2011. Trichosporon Behrend (1890). Chapter 161 in The Yeasts, a Taxonomic Study,<br />

5 th edition eds Kurtzman, C.P., J.W. Fell and T. Boekhout, Elsevier B.V. pages 2015-2061.<br />

Sugiura, Y. and M. Hironaga. 2010. Arthrographis kalrae, a rare causal agent of onychomycosis,<br />

and its occurrence in natural and commercially available soils. Med. Mycol. 48: 384-389.<br />

Summerbell, R.C., S.A. Rosenthal, and J. Kane. 1988. Rapid method for differentiation of<br />

Trichophyton rubrum, Trichophyton mentagrophytes, and related dermatophyte species. J.<br />

Clin. Microbiol. 26: 2279-2282.<br />

Summerbell, R.C., L. de Repentigny, C. Chartrand et al. 1992. Graft-related endocarditis<br />

caused by Neosartorya fischeri var. spinosa. J. Clin. Microbiol. 30: 1580-1582.<br />

Summerbell, R.C., C. Gueidan, H.J Schroers et al. 2011. Acremonium phylogenetic overview<br />

and revision of Gliomastix, Trichothecium and Sarocladium. Stud. Mycol. 68: 139-162.<br />

Sun, Q.N., A.W. Fothergill, D.I. McCarthy et al. 2002. In vivo activities of posacona zole,<br />

itraconazole, voriconazole, amphotericin B, and fluconazole against 37 clini cal isolates of<br />

zygomycetes. Antimicrob. Agents Chemother. 46: 1581-1582.<br />

Sutton, B.C. 1980. The Coelomycetes, fungi imperfecti with pycnidia, acervuli and stromata.<br />

Commonwealth Mycology Institute, Kew, London.<br />

Sutton, B.C. and B.J. Dyko. 1989. Revision of Hendersonula. Mycol. Res. 93: 466-488.<br />

Sutton, D.A., M. Slifkin, R. Yakulis, and M. G. Rinaldi. 1998. U.S. case report of cerebral<br />

phaeohyphomycosis caused by Ramichloridium obovoideum (R. mackenziei): criteria for<br />

identification, therapy, and review of other known dematiaceous neurotropic taxa. J. Clin.<br />

Microbiol. 36: 708-715.<br />

Taj-Aldeen, S.J., M. Almaslamani, A. Alkhalf et al. 2010. Cerebral phaeohyphomycosis due<br />

to Rhinocladiella mackenziei (formerly Ramichloridium mackenziei): a taxonomic update<br />

and review of the literature. Med. Mycol. 48: 546-556.<br />

Tambini, R., C. Farina, R. Fiocchi et al. 1996. Possible pathogenic role for Sporothrix<br />

cyanescens isolated from a lung lesion in a heart transplant patient. J. Med. Vet. Mycol.<br />

34: 195-198.<br />

Tavanti, A., A.D. Davidson, N.A. Gow et al. 2005. Candida orthopsilosis and Candida<br />

metapsilosis spp. nov. to replace Candida parapsilosis groups II and III. J. Clin. Microbiol.<br />

43: 284-292.<br />

Teixeira, M.M., R.C. Theodoro, F.F. Oliveira et al. 2014. Paracoccidioides lutzii sp. nov.:<br />

biological and clinical implications. Med. Mycol. 52: 19-28.<br />

Theodoro, R.C., M. Teixeira, M.S. Felipe et al. 2012. Genus Paracoccidioides: species<br />

recognition and biogeographic aspects. PLoS One 7: e37694.<br />

Tintelnot, K., G.S. de Hoog, E. Antweiler et al. 2007. Taxonomic and diagnostic markers for<br />

identification of Coccidioides immitis and Coccidioides posadasii. Med. Mycol. 45: 385-393.<br />

Tortorano, A.M., M. Richardson, E. Roilides et al. 2014. ESCMID & ECMM joint guidelines<br />

on diagnosis and management of hyalohyphomycosis: Fusarium spp, Scedosporium spp,<br />

and others. Clin. Microbiol. Infect. 20 Suppl 3: 27-46.<br />

Tragiannidis, A., G. Bisping, G. Koehler et al. 2010. Minireview: Malassezia infections in<br />

immunocompromised patients. Mycoses 53: 187-195.<br />

Tuon, F.F. and S.F. Costa. 2008. Rhodotorula infection. A systematic review of 128 cases from<br />

literature. Rev. Iberoam. Micol. 25: 135-140.<br />

Turnidge, J., G. Kahlmeter, G. Kronvall. 2006. Statistical characterization of bacterial wildtype<br />

MIC value distributions and determination of epidemiological cutoff values. Clin.<br />

Microbiol. Infect. 12: 418-425.<br />

Ueno, R., N. Hanagata, N. Urano et al. 2005. Molecular phylogeny and phenotypic variation<br />

in the heterotrophic green algal genus Prototheca. J. Phycol. 41: 1268-1280.<br />

Vanbreusegham, R., CH. de Vroey and M. Takashio. 1978. Practical guide to medical and<br />

veterinary mycology. Mason Publishing USA, Inc.


Descriptions of Medical Fungi 265<br />

REFERENCES<br />

Vanden Bossche, H., D.W.R. Mackenzie and G. Cauwenbergh (eds.) 1988. Aspergillus and<br />

Aspergillosis. Plenum, New York, 322 pp.<br />

van Diepeningen, A.D., B. Brankovics, J. Iltes et al. 2015. Diagnosis of Fusarium infections:<br />

approaches to identification by the clinical mycology laboratory. Curr. Fungal Infect. Rep.<br />

9: 135-143.<br />

Van Oorschot, C.A.N. 1980. A revision of Chrysosporium and allied genera. Stud. Mycol.<br />

No.20. Centraalbureau voor Schimmelcultures, Baarn, The Netherlands.<br />

Varga, J., J.C. Frisvad and R.A. Samson et al. 2011. Two new aflatoxin producing species,<br />

and an overview of Aspergillus section Flavi. Stud. Mycol. 69: 57-80.<br />

Vaux, S., A. Criscuolo, M. Desnos-Ollivier et al. 2014. Multicenter outbreak of infections by<br />

Saprochaete clavata, an unrecognized opportunistic fungal pathogen. mBio 5(6):e02309-<br />

14. doi:10.1128/mBio.02309-14.<br />

Velegraki, A., E.C. Alexopoulos, S. Kritikou et. al. 2004. Use of fatty acid RPMI 1640 media<br />

for testing susceptibilities of eight Malassezia species to the new triazole posaconazole<br />

and six established antifungal agents by a modified NCCLS M27-A2 microdilution method<br />

and Etest. J. Clin. Microbiol. 42: 3589-3593.<br />

Vidal, P., M. de los Vinuesa, J.M. Sánchez-Puelles et al. 2000. Phylogeny of chrysosporia<br />

infecting reptiles: proposal of the new family Nannizziopsiaceae and five new species.<br />

Revta Iberoam. Micol. 17: 24-31.<br />

Vijaykrishna, D., L. Mostert, R. Jeewon et al. 2004. Pleurostomophora, an anamorph<br />

of Pleurostoma (Calosphaeriales), a new anamorph genus morphologically similar to<br />

Phialophora. Stud. Mycol. 50: 387-395.<br />

Vilela, R., S.M. Silva, F. Riet-Correa et al. 2010. Morphologic and phylogenetic characterization<br />

of Conidiobolus lamprauges recovered from infected sheep. J. Clin. Microbiol. 48: 427-432.<br />

Visagie, C.M., J. Houbraken, J.C. Frisvad et al. 2014. Identification and nomenclature of the<br />

genus Penicillium. Stud. Mycol. 78: 343-371.<br />

Vitale, R.G. and G.S. de Hoog. 2002. Molecular diversity, new species and antifungal<br />

susceptibilities in the Exophiala spinifera clade. Med. Mycol. 40: 545-556.<br />

Voigt, K., E. Cigelnik and K. O’Donnell, K. 1999. Phylogeny and PCR identification of<br />

clinically important zygomycetes based on nuclear ribosomal-DNA sequence data. J. Clin.<br />

Microbiol. 37: 3957-3964.<br />

von Arx, J.A., J. Guarro and M.J. Figueras. 1986. The ascomycete genus Chaetomium.<br />

Beih. Nova Hedwigia 84: 162 pp.<br />

Walther, G., J. Pawlowska, A. Alastruey-Izquierd et al. 2012. DNA barcoding in Mucorales:<br />

an inventory of biodiversity. Persoonia 30: 11-47.<br />

Wahyuningsih, R., I.N. SahBandar, B. Theelen et al. 2008. Candida nivariensis isolated from<br />

an Indonesian human immunodeficiency virus-infected patient suffering from oropharyngeal<br />

candidiasis. J. Clin. Microbiol. 46: 388-391.<br />

Wang, P.-H., S.-W. Wang and Y.-T. Wang. 1999. Phylogenetic relationships among the sections<br />

of form-genus Aspergillus and their teleomorphs inferred from ITS II rDNA sequences.<br />

Chin. Agric. Chem. Soc. 37: 470-480.<br />

Wang, X., Y-F. Fu, R-Y. Wang et al. 2014. Identification of clinically relevant fungi and Prototheca<br />

species by rRNA gene sequencing and multilocus PCR coupled with electrospray ionization<br />

mass spectrometry. PLoS ONE 9(5): e98110.<br />

Weitzman, I. 1984. The case for Cunninghamella elegans, C. bertholletiae and C. echinulata<br />

as separate species. Trans. Br. Mycol. Soc. 83: 527-528.<br />

Weitzman, I., M.R. McGinnis, A.A. Padhye and L. Ajello. 1986. The genus Arthro derma and<br />

its later synonym Nannizzia. Mycotaxon. 25: 505-505.


266<br />

Descriptions of Medical Fungi<br />

REFERENCES<br />

Weitzman, I. and M.Y. Crist. 1980. Studies with clinical isolates of Cunninghamella. II.<br />

Physiological and morphological studies. Mycologia 72: 661-669.<br />

Wieden, M.A., K.K. Steinbronn, A.A. Padhye et al. 1985. Zygomycosis caused by<br />

Apophysomyces elegans. J. Clin. Microbiol. 22: 522-526.<br />

Wilson, C.M., E.J. O’Rourke, M.R. McGinnis et al. 1990. Scedosporium inflatum: Clinical<br />

spectrum of a newly recognised pathogen. J. Infect. Dis. 161: 102-107.<br />

Won, E.J., J.H. Shin, S.C. Lim et al. 2012. Molecular identification of Schizophyllum commune<br />

as a cause of allergic fungal sinusitis. Ann. Lab. Med. 32: 375-379.<br />

Woudenberg, J.H.C., J.Z. Groenewald, M. Binder, and P.W. Crous. 2013. Alternaria<br />

redefined. Stud. Mycol. 75: 171-212.<br />

Xi, L., J. Sun, C. Lu et al. 2009b. Molecular diversity of Fonsecaea (Chaetothyriales) causing<br />

chromoblastomycosis in southern China. Med. Mycol. 47: 27-33.<br />

Xiao, M., L.N. Guo, F. Kong et al. 2013. Practical identification of eight medically important<br />

Trichosporon species by reverse line blot hybridization (RLB) assay and rolling circle<br />

amplification (RCA). Med. Mycol. 51: 300-308.<br />

Yanagihara, M., M. Kawasaki, H. Ishizaki et al. 2010. Tiny keratotic brown lesions on the<br />

interdigital web between the toes of a healthy man caused by Curvularia species infection<br />

and a review of cutaneous Curvularia infections. Mycoscience 51: 224-233.<br />

Yilmaz, N., C.M. Visagie, J. Houbraken et al. 2014. Polyphasic taxonomy of the genus<br />

Talaromyces. Stud. Mycol. 78: 175-341.<br />

Yogo, N., L. Shapiro and K.M. Erlandson. 2014. Sepedonium intra-abdominal infection: a<br />

case report and review of an emerging fungal infection. J. Antimicrob. Chemother. 69:<br />

2583-1585.<br />

Yu, J., G. Walther, A.D. van Diepeningen et al. 2015. DNA barcoding of clinically relevant<br />

Cunninghamella species. Med. Mycol. 53: 99-106.<br />

Yuan, G.F. and S.C. Jong. 1984. A new obligate azygosporic species of Rhizopus. Mycotaxon.<br />

20: 397-400.<br />

Zalar, P., G.S. de Hoog, H.J. Schroers et al. 2007. Phylogeny and ecology of the ubiquitous<br />

saprobe Cladosporium sphaerospermum, with descriptions of seven new species from<br />

hypersaline environments. Stud. Mycol. 58: 157-183.<br />

Zheng, R.-y. and C.-q. Chen. 2001. A monograph of Cunninghamella. Mycotaxon 80: 1-76.<br />

Zeng, J., D.A. Sutton, A.W. Fothergill et al. 2007. Spectrum of clinically relevant Exophiala<br />

species in the U.S.A. J. Clin. Microbiol. 45: 3713-3720.<br />

Zeng, J. and G.S. de Hoog. 2008. Exophiala spinifera and its allies: diagnostics from<br />

morphology to DNA barcoding. Med. Mycol. 46: 193-208.<br />

Zeng., J., P. Feng, Gerrits van den Ende et al. 2014. Multilocus analysis of the Exophiala<br />

jeanselmei clade containing black yeasts involved in opportunistic disease in humans.<br />

Fungal Diversity 65: 3-16.<br />

Zhang, Y., F. Liu, W. Wu, L. Cai. 2015a. A phylogenetic assessment and taxonomic revision<br />

of the thermotolerant hyphomycete genera Acrophialophora and Taifanglania. Mycologia<br />

107: 768-79.<br />

Zhang, Y., F. Hagen, B. Stielow et al. 2015b. Phylogeography and evolutionary patterns in<br />

Sporothrix spanning more than 14,000 human and animal case reports. Persoonia 35:<br />

1-20.<br />

Zhao, Y., R. Petraitiene, T.J. Walsh et al. 2013. A real-time PCR assay for rapid detection<br />

and quantification of Exserohilum rostratum, a causative pathogen of fungal meningitis<br />

associated with injection of contaminated methylprednisolone. J. Clin. Microbiol. 51: 1034-<br />

1036.<br />

Zycha, H., R. Siepmann and G. Linnemann. 1969. Mucorales, eine Beschreibung aller<br />

Gattungen und Arten dieser Pilzgruppe. Cramer Lehre, 355p.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!