determining the mechanisms of resistance to drug ...€¦ · leading fungal pathogen, candida...

223
Determining the mechanisms of resistance to drug combinations and their fitness effects in Candida albicans by Jessica Anne Hill A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of Molecular Genetics University of Toronto © Copyright by Jessica A. Hill 2014

Upload: others

Post on 04-Feb-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • Determining the mechanisms of resistance to drug combinations and their fitness effects in Candida albicans

    by

    Jessica Anne Hill

    A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

    Department of Molecular Genetics University of Toronto

    © Copyright by Jessica A. Hill 2014

  • ii

    Determining the mechanisms of resistance to drug combinations

    and their fitness effects in Candida albicans

    Jessica Anne Hill

    Doctor of Philosophy

    Department of Molecular Genetics University of Toronto

    2014

    Abstract

    The evolution of drug resistance is occurring rapidly and threatening our arsenal of

    antimicrobials. This is a poignant issue with antifungals, whose targets are limited due to

    conservation between fungi and the human host. A promising strategy is to enhance

    effectiveness of our current antifungals using combination therapy. Adding an inhibitor of one

    of the key stress response proteins, Hsp90 or calcineurin, to an antifungal azole renders the

    activity of the azole fungicidal and abrogates azole resistance. However, evidence from clinical

    isolates indicates that resistance to the drug combination can still arise. Determining the

    mechanisms of resistance to these drug combinations and their fitness effects will provide

    information about how resistance can occur and whether it is likely to spread in a population. To

    address what mechanisms of resistance can arise to drug combinations, I evolved strains of the

    leading fungal pathogen, Candida albicans, and its genetically tractable relative Saccharomyces

    cerevisiae, in the presence of an azole and the Hsp90 inhibitor geldanamycin (GdA) or the

    calcineurin inhibitor FK506. The vast majority of strains (286) went extinct, however, I

    identified and validated mechanisms of resistance in 13 strains. A variety of resistance

    mechanisms were found, from well-established resistance mechanisms to novel ones. To

    address the fitness effects of resistance to an azole in combination with GdA or FK506, I took

  • iii

    advantage of these C. albicans in vitro evolved strains and a series of C. albicans clinical

    isolates that transition to Hsp90- and calcineurin-independent azole resistance. I determined that

    resistance to the drug combination is often but not universally costly in the absence of drug and

    in novel, clinically relevant stress environments. Together these results provide a basis for our

    understanding of key parameters for the evolution of resistance to drug combinations: the

    variety of mechanisms that encompass resistance, and their fitness costs.

  • iv

    Acknowledgments

    The successful completion of this thesis would not have been possible without several

    individuals.

    I am indebted to my supervisor, Leah Cowen, under whose guidance I have grown to become a

    better scientist. Leah’s commitment to effective communication, elegant design and

    thoroughness will stay with me beyond this degree, as well as her commitment to supporting her

    students.

    I have had the pleasure of working with Aneil Agrawal, William Navarre and Corey Nislow. As

    a supervisory committee, Aneil, Will and Corey have been consistently supportive and

    insightful and I would like to recognize their influence on this body of work.

    My time in the Cowen lab afforded me the opportunity to meet some incredible scientists and

    people. Throughout the years my lab mates may have changed but they all share a passion for

    science. I will miss the stimulating scientific conversations in the Cowen lab. All the more I will

    miss the special friendships I have formed in this lab, because no one can understand the

    challenges of grad school the way your lab mates can. I would like to especially thank Rebecca

    Shapiro, Sheena Singh, Nicole Robbins, Michelle Leach, Teresa O’Meara, Tanvi Shekhar-

    Guturja, Elizabeth Polvi, Amanda Veri and Cathy Collins.

    Beyond the lab, there have been a number of scientists in my life whose friendship, support and

    curiousity has not been diminished by time or physical separation. I cannot imagine how

    finishing this degree would have looked without their involvement, nor would I want to. I am

    especially grateful to Heather Maughan, Dilara Ally and Aleeza Gerstein for their friendship and

    insights. My non-scientist friends have been instrumental in the completion of this degree. Over

  • v

    a period of time that has included many difficulties in my life, I have gratefully received

    support from Chioma Ume, Andrea Benvenuto and Dave Seevaratnam, among others. I cannot

    adequately thank these people.

    I am eternally grateful for the love and support of my family. My sister, Kelly Hill, has been the

    bedrock of my support and her persistence in life and commitment to high standards has been an

    inspiration. I would also like to thank my brother-in-law, Gabe Plener, for his unwavering

    support. My grandparents, Theresa and Michael Illes, and Margaret and George Hill, had been

    incredibly influential and loving; they showed me that hard work is a requirement for success.

    Finally, I must thank my parents, Anne and Lawrence Hill, for their unconditional love and

    support, and for giving me the freedom to pursue my interests. I dedicate this work to their

    memory.

  • vi

    Table of Contents

    Acknowledgments ........................................................................................................................ iv

    Table of Contents .......................................................................................................................... vi

    List of Tables ................................................................................................................................. x

    List of Figures ............................................................................................................................... xi

    List of Abbreviations .................................................................................................................. xiii

    Chapter 1 Introduction ................................................................................................................... 1

    1 Introduction ............................................................................................................................. 1

    1.1   Evolution of drug resistance in microbes ......................................................................... 1  

    1.2   Candida albicans .............................................................................................................. 3  

    1.3   Antifungal drugs ............................................................................................................... 5  

    1.3.1 Major antifungal drugs and their mode of action ....................................................... 5

    1.3.2 Drug resistance mechanisms 8

    1.4   Strategies for thwarting the evolution of drug resistance ............................................... 15  

    1.4.1 Developing new antifungals ..................................................................................... 15

    1.4.2 Combination therapy ................................................................................................ 15

    1.4.3 Targeting stress response: Hsp90 and calcineurin ................................................... 16

    1.5   Natural variation in resistance to antifungal drugs ......................................................... 20  

    1.5.1 Variation between species ........................................................................................ 20

    1.5.2 Variation between strains ......................................................................................... 23

    1.5.3 Variation within a population .................................................................................. 24

    1.6   Strategies for studying the evolution of drug resistance ................................................. 26  

  • vii

    1.6.1 Examining clinical isolates ...................................................................................... 26

    1.6.2 Using experimental evolution .................................................................................. 31

    1.7   Thesis rationale ............................................................................................................... 36  

    Chapter Two Genetic and Genomic Architecture of the Evolution of Resistance to Antifungal

    Drug Combinations ...................................................................................................................... 39

    2 Genetic and Genomic Architecture of the Evolution of Resistance to Antifungal Drug

    Combinations ............................................................................................................................... 40

    2.1   Introduction .................................................................................................................... 40  

    2.2   Materials and Methods ................................................................................................... 44  

    2.2.1 Strains and culture conditions .................................................................................. 44

    2.2.2 Strain construction ................................................................................................... 50

    2.2.3 Plasmids and oligonucleotides ................................................................................. 56

    2.2.4 Plasmid construction ................................................................................................ 61

    2.2.5 Evolution experiment ............................................................................................... 63

    2.2.6 Minimum inhibitory concentration and checkerboard assays .................................. 69

    2.2.7 Genome sequencing ................................................................................................. 70

    2.3   Results ............................................................................................................................ 73  

    2.3.1 Experimental evolution of C. albicans and S. cerevisiae yields resistance to the

    combination of an azole and an inhibitor of Hsp90 or calcineurin ...................................... 73

    2.3.2 Cross-resistance assays as a strategy to predict distinct mechanisms of resistance . 80

    2.3.3 Mutations in HSP90 confer resistance to azole and geldanamycin in two S.

    cerevisiae lineages and one C. albicans lineage .................................................................. 83

    2.3.4 Mutations in FPR1 confer resistance to azole and FK506 in two S. cerevisiae

    lineages ................................................................................................................................ 89

  • viii

    2.3.5 Whole genome sequencing identifies candidate resistance mutation in additional

    S. cerevisiae evolved lineages ............................................................................................. 93

    2.3.6 Whole genome sequencing identifies extensive aneuploidy in four of the C.

    albicans evolves lineages, and additional candidate resistance mutation in two of the C.

    albicans lineages ................................................................................................................ 100

    2.4   Discussion ..................................................................................................................... 110  

    Chapter 3 Fitness Trade-offs Associated with the Evolution of Resistance to Antifungal Drug

    Combinations ............................................................................................................................. 119

    3 Fitness Trade-offs Associated with the Evolution of Resistance to Antifungal Drug

    Combinations ............................................................................................................................. 119

    3.1   Introduction .................................................................................................................. 119  

    3.2   Materials and Methods ................................................................................................. 122  

    3.2.1 Strains and culture conditions ................................................................................ 122

    3.2.2 Strain construction ................................................................................................. 126

    3.2.3 Plasmids and oligonucleotides ............................................................................... 132

    3.2.4 Drug susceptibility assays ...................................................................................... 139

    3.2.5 In vitro competition experiments ........................................................................... 140

    3.2.6 In vivo competition experiments ............................................................................ 142

    3.2.7 Growth rate kinetics ............................................................................................... 143

    3.2.8 Filamentation assays .............................................................................................. 143

    3.2.9 Cidality spotting assay ........................................................................................... 143

    3.3   Results .......................................................................................................................... 144  

    3.3.1 In vivo and in vitro evolution of resistance to drug combinations ......................... 144

  • ix

    3.3.2 Strains with distinct genetic backgrounds resistant to drug combinations are more

    fit than ancestral strains in drug combinations .................................................................. 148

    3.3.3 The mechanism of resistance in late-stage clinical isolates is largely independent of

    Hsp90 and calcineurin, in contrast to some experimentally evolved strains ..................... 151

    3.3.4 TAC1A736V, UPC2A643V and ERG11R467K confer Hsp90- and calcineurin-independent

    azole resistance in late-stage clinical isolates .................................................................... 155

    3.3.5 Resistance to fluconazole and geldanamycin or fluconazole and FK506 is costly in

    several host relevant environments .................................................................................... 159

    3.3.6 Late-stage clinical isolates are unable to filament in response to geldanamycin due

    to TAC1A736V-regulated efflux of geldanamycin ................................................................ 163

    3.3.7 Relative fitness is variable in a host model system ................................................ 170

    3.4   Discussion ..................................................................................................................... 173  

    Chapter 4 Conclusion, general discussion and future directions ............................................... 179

    4 Conclusion, general discussion and future directions ......................................................... 179

    4.1   Conclusion .................................................................................................................... 179  

    4.2   Discussion ..................................................................................................................... 180  

    4.2.1 Combination therapy in the treatment of disease ................................................... 180

    4.3   Future directions ........................................................................................................... 187  

  • x

    List of Tables

    Table 2.1. Strains used in this study .................................................................................. 45

    Table 2.2. Plasmids used in this study ............................................................................... 57

    Table 2.3. Oligonucleotides used in this study .................................................................. 59

    Table 2.4. Evolution experiment treatments and conditions ............................................. 65

    Table 2.5. Mean coverage whole-genome sequenced strains ............................................ 73

    Table 2.6. High confidence single nucleotide variants (coding and non-coding) identified in S.

    cerevisiae ........................................................................................................... Separate file

    Table 2.7. High confidence single nucleotide variants (coding and non-coding) identified in C.

    albicans .............................................................................................................. Separate file

    Table 2.8. Non-synonymous S. cerevisiae single nucleotide variants ............................... 95

    Table 2.9. Number of high confidence C. albicans single nucleotide variants (SNVs)

    (coding and non-coding) .................................................................................................... 97

    Table 2.10. Non-synonymous C. albicans single nucleotide variants ............................. 108

    Table 3.1. Corresponding strain names between Chapters Two and Three .................... 123

    Table 3.2. Strains used in this study ................................................................................ 123

    Table 3.3. Plasmids used in this study ............................................................................. 133

    Table 3.4. Primers used in this study ............................................................................... 136

    Table 3.5. Stressor concentrations for competition and growth rate assays .................... 141

  • xi

    List of Figures

    Figure 1.1. Mode of action of azoles and echinocandins and resistance mechanisms ...... 10

    Figure 1.2. Studying the evolution of resistance to antifungals ........................................ 28

    Figure 2.1. Dose response matrices were used to select concentrations for the evolution

    experiment ......................................................................................................................... 68

    Figure 2.2. Design and outcome of the experimental evolution of resistance to drug

    combinations ...................................................................................................................... 76

    Figure 2.3. The populations evolved distinct resistance profiles ....................................... 79

    Figure 2.4. Cross-resistance profiles provide a strategy to predict resistance

    mechanisms ....................................................................................................................... 83

    Figure 2.5. Mutations in HSP90 confer resistance to azole and geldanamycin in two S.

    cerevisiae lineages and in one C. albicans lineage ............................................................ 87

    Figure 2.6. Ca-G-10 is slightly resistant to azole and geldanamycin, and slightly cross-resistant

    to azole and 17-AAG, a structural derivative of geldanamycin ........................................ 89

    Figure 2.7. Mutations in FPR1 confer resistance to azole and FK506 in two S. cerevisiae

    lineages .............................................................................................................................. 92

    Figure 2.8. Whole genome sequencing identifies mutations that confer resistance to azole and

    FK506, as well as azole and geldanamycin ....................................................................... 99

  • xii

    Figure 2.9. Six C. albicans lineages evolved with azole and FK506 share the same cross-

    resistance profile, and a mutation in CNA1 and LCB1 confer resistance ........................ 102

    Figure 2.10. Aneuploidies identified in four C. albicans lineages that evolved resistance to the

    combination of azoles and calcineurin inhibitors ............................................................ 106

    Figure 3.1. Lab-passaged strains and late-stage clinical isolates of C. albicans evolved resistance

    to drug combinations ....................................................................................................... 147

    Figure 3.2. Fitness differences between strains are often reflected in differences in growth rates

    ......................................................................................................................................... 150

    Figure 3.3. Fluconazole resistance of clinical isolates is largely Hsp90- and calcineurin-

    independent, in contrast to in vitro evolved strains ......................................................... 154

    Figure 3.4. Together, TAC1A736V, UPC2A643V and ERG11R467K confer resistance to fluconazole in

    combination with geldanamycin or FK506 ..................................................................... 158

    Figure 3.5. Resistance to drug combinations is costly in the absence of drug and in different

    stressful environments ..................................................................................................... 162

    Figure 3.6. Late-stage clinical isolates do not filament in response to geldanamycin .... 166

    Figure 3.7. Genetic background for the in vitro evolved strains has increased susceptibility to

    geldanamycin ................................................................................................................... 168

    Figure 3.8. TAC1A736V abolishes filamentation in late-stage clinical isolates .................. 170

    Figure 3.9. Fitness in a host model is variable ................................................................ 173

  • xiii

    List of Abbreviations

    17-AAG 17-allylamino-17-demethoxygeldanamycin

    ANOVA analysis of variance

    ATP adenosine triphosphate

    CsA cyclosporin A

    DNA deoxyribonucleic acid

    EBI ergosterol biosynthesis inhibitor

    FLP flippase recombination enzyme

    GdA geldanamycin

    GFP green fluorescent protein

    Hsp90 heat shock protein 90

    MIC minimum inhibitory concentration

    MPC mutant prevention concentration

    MSW mutant selection window

    NAT nourseothricin

    OD optical density

    PCR polymerase chain reaction

    NF-AT nuclear factor of activated T cells

    PKC protein kinase C

    RAD radicicol

  • xiv

    RNA ribonucleic acid

    YPD yeast peptone dextrose

  • 1

    Chapter 1 Introduction

    1 Introduction

    1.1 Evolution of drug resistance in microbes

    The evolution of resistance to antimicrobials is a ubiquitous phenomenon. The success of a new

    antimicrobial is ultimately undermined by the evolution of resistance that inevitably follows. For

    example, resistance to penicillin was found in Gram-positive bacteria a mere six years after its

    clinical use was initiated (Barber and Whitehead 1949). Resistance arises rapidly, as evidenced

    by the nearly 1000 beta-lactam antibiotic inactivating beta-lactamases that have been identified, a

    ten-fold increase since 1990 (Davies and Davies 2010).

    Strong selection for resistance is imposed by the widespread deployment of antimicrobials in the

    clinic as well as in environmental settings, such as the treatment of crops or livestock with

    antimicrobials. With the increase in the immunocompromised population due to infections like

    HIV and modern medical breakthroughs like chemotherapy and organ transplantation, the

    population of individuals susceptible to infection has increased. This combination of factors

    makes antimicrobial resistance one of the most serious threats to public health. Furthermore, the

    economic effects of drug resistance are astonishing, as the combined cost of treatment of patients

    with drug-resistant infections, pesticide use on crops for the treatment of drug-resistant pests and

    crop losses is a minimum of $33 billion USD annually (Palumbi 2001).

    Treatment of resistant pathogens is a constant struggle. The arsenal of antimicrobials is

    becoming increasingly limited as resistance to multiple drug classes is emerging. The microbes

    that pose the greatest risk to public health are now resistant to multiple drug classes, such as the

    bacteria Clostridium difficile and drug-resistant Neisseria gonorrhoeae (CDC 2013). The rate at

  • 2

    which new antimicrobials are being produced is vastly outpaced by the rate at which resistance

    emerges. Despite this, fewer new antimicrobials are being developed. Since the 1970’s only one

    antibiotic and one antifungal with novel modes of action have reached the clinic (the

    oxazolidinones and the echinocandins, respectively) (Norrby, Nord et al. 2005). This may be

    partly due to the fact that research and development of new drugs, particularly clinical trials, is

    exceedingly expensive and therefore often pursued foremost by pharmaceutical companies.

    However, pharmaceutical companies are deterred by lower return per treatment of antimicrobials

    relative to developing drugs prescribed for life, such as those used to treat hypertension or

    diabetes. Furthermore, the number of novel drug targets is limited, particularly for fungi who

    share close evolutionary relationships with their host, ultimately causing host toxicity issues.

    Effective stewardship of our antimicrobials is a critical component of the fight against antibiotic

    resistance (Laxminarayan, Duse et al. 2013). Making our current arsenal of antimicrobials more

    effective is therefore of paramount importance.

    In particular, fungal drug resistance is a pressing issue. Antifungal drug resistance is

    tremendously costly, in terms of both human mortality (McNeil, Nash et al. 2001) and financial

    burden (Dasbach, Davies et al. 2000). The most significant challenge in developing new

    antifungals is avoiding host toxicity while effectively killing the pathogen. The close

    evolutionary relationship between fungal pathogens and their human hosts limits the number of

    fungal-specific drug targets available (Baldauf, Roger et al. 2000). Resistance to antifungals

    evolves at a rate faster than new antifungals are being developed. The most commonly used

    classes of antifungals, the azoles and the echinocandins, target the biosynthesis of the fungal-

    specific cell membrane sterol ergosterol and the fungal cell wall, respectively (Cowen 2008).

    These limitations in drug development underscore the importance of determining how drug

    resistance evolves in order to predict and prevent it.

  • 3

    1.2 Candida albicans

    Candida albicans is a clinically relevant organism with which to study the evolution of drug

    resistance. It is an opportunistic pathogen and the leading cause of death due to mycotic

    infection, and the fourth most common cause of nosocomial bloodstream infections (Pfaller and

    Diekema 2007). C. albicans is found as a commensal on 15 – 60% of asymptomatic individuals,

    where is it able to reside in diverse niches such as the oral cavity and the gastrointestinal tract

    (Odds, Gow et al. 2001). In immunocompetent hosts, it causes superficial infections in moist

    areas (typically the groin and toe webs), however, upon impairment of the host immune system,

    C. albicans is able to penetrate tissues to cause serious invasive fungal infections of the blood,

    normally sterile body fluids, and of deep tissue and organs (Horn, Neofytos et al. 2009; Marr

    2010; Pfaller and Diekema 2010). Thus, immunocompromised individuals and those hospitalized

    with serious underlying diseases, such as stem cell transplant recipients, solid organ transplant

    recipients, those undergoing major surgery, those undergoing chemotherapy, those infected with

    HIV, neonatal infants and those of advanced age, are particularly susceptible to invasive C.

    albicans infection (Pfaller and Diekema 2010). Due to advances in medical technology, the

    immunocompromised population has increased, which has lead to a 207% increase in

    bloodstream infections caused by fungal pathogens between 1979 and 2000 in the United States

    (Martin, Mannino et al. 2003).

    A key feature of C. albicans virulence is its ability to undergo morphogenesis. C. albicans is

    polymorphic, existing as yeast or in its pseudohyphal or hyphal filamentous forms. It is thought

    that the ability to penetrate tissues and cause deep-seated infection requires filamentous growth,

    while bloodstream dissemination to target organs requires yeast form growth (Saville, Lazzell et

    al. 2003). Indeed, several studies have indicated that mutants locked in either morphology are

  • 4

    attenuated in virulence (Shapiro, Robbins et al. 2011). The ability to undergo morphogenesis also

    facilitates survival in the host, as macrophage killing by C. albicans is associated with the ability

    to transition between morphologies (McKenzie, Koser et al. 2010). The ability to undergo

    morphogenesis is essential to the lifestyle of C. albicans.

    Candida albicans exhibits a high level of genomic plasticity. Widespread aneuploidy in lab

    strains was discovered by comparative genomic hybridization, and this aneuploidy occurred

    more frequently in strains exposed to stressors like the process of transformation and

    counterselection on 5-fluoro-orotic acid (Selmecki, Bergmann et al. 2005). It has recently been

    shown that exposure to the antifungal fluconazole also induces aneuploidy (Harrison, Hashemi et

    al. 2014). This evidence suggests that aneuploidy can be adaptive, and indeed this has been

    demonstrated by the formation of an isochromosome of the left arm of chromosome 5 in

    response to fluconazole exposure. This isochromosome contained extra copies of a

    transcriptional activator of drug efflux pumps (TAC1) and the drug target itself, ERG11

    (Selmecki, Forche et al. 2006). Furthermore, the stress of passage through a host model system

    induced more large-scale genomic rearrangements than passage in test tubes (Forche, Magee et

    al. 2009). This remarkable genome lability enhances the ability of C. albicans to cope with a

    broad spectrum of stressors.

    Genomic flexibility, lack of meiotic cycle and inability to maintain plasmids can render

    molecular studies with C. albicans challenging (Berman and Sudbery 2002). It is advantageous,

    therefore, to perform complementary experiments with its genetically tractable relative,

    Saccharomyces cerevisiae, with whom it often shares drug resistance phenotypes and underlying

    molecular mechanisms (Cowen and Steinbach 2008). Tools for study with S. cerevisiae are

    greater in number and more developed. S. cerevisiae supports plasmids, has a complete sexual

  • 5

    cycle, and has more efficient homologous recombination. Complementary studies in S.

    cerevisiae can provide insights into drug resistance mechanisms relevant to C. albicans.

    1.3 Antifungal drugs

    1.3.1 Major antifungal drugs and their mode of action

    The number of fungal-specific drug targets is limited due to conservation with the human host.

    The majority of antifungals that are clinically deployed target the fungal-specific sterol

    ergosterol, or its biosynthesis, or the fungal cell wall (Cowen 2008; Shapiro, Robbins et al.

    2011). The utility of these antifungals is compromised by their often fungistatic

    (antiproliferative) nature. Reservoirs of cells tolerant to fungistatic drugs are poised for the

    evolution of resistance. This section will focus on antifungals currently available for the

    treatment of the most prevalent fungal infections.

    1.3.1.1 Azoles

    The most widely deployed class of antifungals, the azoles, are ergosterol biosynthesis inhibitors

    (EBIs). EBIs are a large class of antifungals that function by targeting enzymes involved in the

    biosynthesis of ergosterol, the bulk sterol of fungal membranes that is involved in maintaining

    membrane fluidity and cellular signal. Ergosterol is the fungal equivalent of cholesterol

    (Anderson 2005). Ergosterol is synthesized from squalene; the genes in this pathway encode

    products that act in the following order: Erg1, Erg7, Erg11, Erg24, Erg25, Erg26, Erg27, Erg6,

    Erg2, Erg3, Erg5 and Erg4 (White, Marr et al. 1998).

    Azoles have been clinically deployed for almost 30 years. They inhibit fungal growth by binding

    to the P450 enzyme α-14-lanosterol demethylase (encoded by ERG11), a late-stage enzyme in

    the ergosterol biosynthesis pathway. Inhibition of ERG11 blocks the production of ergosterol and

    leads to the accumulation of methylated sterol intermediates in the cell membrane, which alters

  • 6

    membrane fluidity and inhibits growth (Cowen 2008). Inhibition of ERG11 also reduces the

    overall amount of ergosterol in the cell, thus interfering with vacuolar H+-ATPase activity, which

    is necessary for ion homeostasis and fungal virulence (Zhang, Gamarra et al. 2010).

    Azoles are classified into two groups, the imidazoles and newer triazoles. Imidazoles suffer from

    toxicity and bioavailability issues and are only available in topical, over-the-counter formulations

    for vaginal or skin infections (Ostrosky-Zeichner, Casadevall et al. 2010). Triazoles must be

    prescribed and are available in oral or injection formats, and are used in the treatment of systemic

    fungal infections. The four triazoles currently available clinically are fluconazole, itraconazole,

    posaconazole and voriconazole. Azoles are effective for treating diverse fungi, including species

    of Candida, Cryptococcus, and Aspergillus, as well as dimorphic and dermatophyte fungi

    (Shapiro, Robbins et al. 2011). Posaconazole is also effective against zygomycete fungi

    (Alastruey-Izquierdo, Castelli et al. 2009). However, azoles are fungistatic against yeasts like

    Candida, such that strong selective pressure for resistance is imposed upon tolerant cells

    (Anderson 2005). Clinical resistance in yeasts is a significant clinical issue.

    1.3.1.2 Echinocandins

    The newest class of antifungals is the echinocandins. These drugs disrupt cell wall biosynthesis

    by inhibiting the enzyme β-(1,3)-glucan synthase, the subunits of which are encoded by FKS1

    and FKS2 (Cowen and Steinbach 2008). This results in a loss of cell wall integrity. The

    echinocandins are broadly effective against many fungi including Candida and Aspergillus

    species, but not against Cryptococcus or zygomycetes. Unlike azoles, echinocandins are

    generally fungicidal against yeasts and resistance is not yet widespread, although numerous

    clinical isolates with elevated resistance have been identified (Shapiro, Robbins et al. 2011). The

  • 7

    echinocandins have little to no toxicity to the host, and, because they are fungicidal, they are

    becoming the drug of choice for treatment of invasive fungal infections.

    1.3.1.3 Polyenes

    The polyenes have been used in the clinic for the past 50 years. They bind to ergosterol in fungal

    cell membranes, resulting in the formation of pores that allow small molecules like ions to

    diffuse across the membrane, ultimately resulting in cell death (White, Marr et al. 1998). The

    most commonly deployed polyenes are amphotericin B and nystatin. Amphotericin B is effective

    against Candida, Cryptococcus and Aspergillus infections and was previously considered the

    gold standard for treatment of invasive fungal infections; it is highly insoluble and is typically

    delivered in liposomal formulations to increase solubility when given intravenously. Therapeutic

    use of amphotericin B is complicated by significant liver toxicity, poor pharmacokinetics and

    other side effects like fevers and tremors (Paterson, Seaton et al. 2003). As such, it is not

    typically used to treat chronic fungal infections or for prophylaxis and is now considered a third-

    line defense. Nystatin is also nephrotoxic and highly insoluble; it is typically given only in

    lozenge format to treat oropharyngeal candidiasis. Resistance to polyenes is not widespread, such

    that a recent survey of over 9000 clinical isolates found that 99.8% were sensitive to

    amphotericin B (Pfaller, Espinel-Ingroff et al. 2012). This is in part due to their limited use and

    fungicidal nature, but also due to the high fitness cost resistance incurs (Vincent, Lancaster et al.

    2013). The best-characterized resistance mechanism is concurrent loss of function of ERG11 and

    ERG3, both members of the ergosterol biosynthesis pathway (Sanglard, Ischer et al. 2003).

    1.3.1.4 Other antifungals

    Other than the azoles, commonly used classes of EBIs are the allylamines and morpholines. The

    allylamines, including terbinafine, and the thiocarbamates, including tolnaftate, act by inhibiting

  • 8

    Erg1; they are available in over-the-counter topical preparations. In contrast, morpholines such

    as fenpropimorph are commonly used in agricultural settings to target fungal pathogens of plants

    (Zocco, Van Aarle et al. 2011). Morpholines target two enzymes in the ergosterol biosynthetic

    pathway, C-8 sterol isomerase (encoded by ERG2) and C-14 sterol reductase (encoded by

    ERG24).

    There are several other antifungals that are clinically employed or currently in development. The

    classic antifungal 5-flucytosine inhibits fungal nucleic acid biosynthesis, but its effect is

    fungistatic and resistance develops readily (White, Marr et al. 1998). 5-flucytosine is therefore

    typically only used in combination with other antifungals. Emerging antifungals include

    sordarins and nikkomycin Z. Sordarins are semi-synthetic natural products that had shown

    antifungal activity as early as the 1970’s (Ostrosky-Zeichner, Casadevall et al. 2010). They

    function as antifungals by inhibiting fungal elongation factor 2 during protein biosynthesis. The

    sordarin derivative FR290581 has demonstrated robust antifungal activity against C. albicans

    and is currently in development. Nikkomycin Z interferes with cell wall biosynthesis by

    competitively inhibiting chitin synthases. Phase I clinical trials of nikkomycin Z are currently

    underway (Ostrosky-Zeichner, Casadevall et al. 2010).

    1.3.2 Drug resistance mechanisms

    Acquired resistance to antifungals can occur through several mechanisms. The canonical

    resistance mechanisms are drug target alteration or overexpression, reduction of intracellular

    drug accumulation, and upregulation of stress response pathways (Figure 1.1). These resistance

    mechanisms need not occur in isolation; in fact clinical isolates frequently exhibit several

    mechanisms of resistance. This section focuses on genetically stable resistance mechanisms.

  • 9

  • 10

    Figure 1.1. Mode of action of azoles and echinocandins and resistance mechanisms. A)

    Azoles enter the cell by facilitated diffusion and inhibit Erg11, blocking the production of

    ergosterol and resulting in the accumulation of a toxic sterol intermediate, which results in cell

    membrane stress. Resistance to azoles can arise by: B) upregulation of two classes of efflux

    pumps that remove the drug from the cell; C) mutation or overexpression of ERG11, which

    minimizes the impact of the drug on the target; or D) loss-of-function mutation of ERG3, which

    blocks the accumulation of a toxic sterol that is otherwise produced when Erg11 is inhibited by

    azoles. E) Echinocandins inhibit β-(1,3)-glucan synthase (the catalytic subunit is encoded by

    FKS1 and FKS2 in S. cerevisiae), and thus disrupt cell wall integrity. F) Resistance to

    echinocandins can arise by mutations in FKS1 or FKS2 that minimize the impact of the drug on

    the target.

  • 11

    1.3.2.1 Alteration of drug target

    The most direct mechanism to increase resistance to an antimicrobial is by mutation of the drug

    target that reduces drug binding. One example of this is alteration of ERG11 encoding the azole

    target enzyme in C. albicans. In a well-characterized series of matched susceptible and resistant

    isolates, an R467K amino acid substitution was identified that conferred azole resistance (White

    1997). This mutation reduces the susceptibility of the enzyme to azoles, likely through

    decreasing the azole binding affinity (Lamb, Kelly et al. 2000). Numerous other ERG11

    mutations have been identified within several mutational ‘hot spots’ that result in amino acid

    substitutions that are in or adjacent to the active site. Elevated resistance is often associated with

    subsequent loss of heterozygosity that results in two mutated alleles of the gene.

    Echinocandin resistance also occurs by target alteration. In echinocandin-resistant isolates of C.

    albicans, mutations in the target genes FKS1 and FKS2 (also referred to as GSL1 and GSC1)

    have been found, leading to alteration of the glucan synthase (Munro 2010). Similar mutations in

    FKS1 and FKS2 contributing to echinocandin resistance have been identified in C. glabrata

    (Zimbeck, Iqbal et al. 2010). Elevated echinocandin resistance in C. albicans has been associated

    with loss of heterozygosity of FKS1, leaving two mutant alleles, as with ERG11 and azole

    resistance (Niimi, Monk et al. 2010).

    1.3.2.2 Overexpression of drug target

    Increased expression of the drug target can also confer resistance. For example, overexpression

    of ERG11 has been documented in azole-resistant C. albicans clinical isolates. ERG11

    overexpression can occur by gain-of-function mutations in the transcriptional activator encoded

    by UPC2, leading to overexpression of ergosterol biosynthesis genes including ERG11 (Dunkel,

    Liu et al. 2008; Hoot, Smith et al. 2011). Multiple studies identified mutations in the same C-

  • 12

    terminal region of UPC2 in azole-resistant C. albicans clinical isolates (Dunkel, Liu et al. 2008;

    Heilmann, Schneider et al. 2010; Hoot, Smith et al. 2011). In a clinical isolate containing UPC2

    A643T, loss of heterozygosity occurred rendering the mutation homozygous and further

    elevating resistance. That these amino acid substitutions in diverse isolates arise in a restricted

    region of the protein suggests that this region has an essential regulatory function.

    Recently, two UPC2 homologs were identified in C. glabrata that, like in C. albicans, regulate

    the response to azoles and hypoxia (Nagi, Nakayama et al. 2011). However, UPC2 homologs are

    limited to these and related species. In other ascomycetes and in basidiomycetes such as C.

    neoformans, ergosterol biosynthesis is regulated at the transcriptional level by an SREBP-like

    transcription factor Sre1 (Bien and Espenshade 2010). This protein has functional homology to

    the mammalian cholesterol regulatory transcription factor SREBP. In A. fumigatus, a similar

    SREBP homolog called SrbA regulates sterol biosynthesis (Willger, Puttikamonkul et al. 2008).

    These species also contain homologs of some of the other proteins that regulate the SREBP

    pathway, such as the proteases or chaperones involved in regulating SREBP proteolytic cleavage

    (Bien and Espenshade 2010). In both C. neoformans and A. fumigatus, the SREBP-like proteins

    are required for sterol regulation in response to hypoxia and antifungal drugs, and are also

    implicated in virulence in animal models.

    1.3.2.3 Increased drug efflux

    Reduction of intracellular drug concentration by active drug efflux can also lead to antifungal

    drug resistance, especially for the azoles, and is well studied in C. albicans (White, Marr et al.

    1998). In C. albicans, azole efflux is mediated by overexpression of the ATP-dependent pumps

    Cdr1 and Cdr2, or the major facilitator pump Mdr1. Tac1 is the central transcriptional activator

    of CDR1 and CDR2 expression in C. albicans. TAC1 gain-of-function mutations have been

  • 13

    identified and are often homozygous in CDR1 overexpressing clinical isolates (Coste, Karababa

    et al. 2004; Coste, Turner et al. 2006). Another C. albicans transcription factor, Ndt80 has been

    implicated in CDR1 transcriptional activation (Wang, Yang et al. 2006). Interestingly, CDR2

    overexpression can confer a slight decrease in susceptibility to echinocandins, however, only on

    solid agar media (Schuetzer-Muehlbauer, Willinger et al. 2003; Silver, Oliver et al. 2008). The

    major transcriptional activator of MDR1 expression in C. albicans is Mrr1; MRR1 mutations can

    lead to overexpression of MDR1 (Dunkel, Liu et al. 2008). However, unlike Upc2, neither Tac1

    nor Mrr1 are activated in response to azoles.

    Increased efflux of antifungals is a major mechanism of resistance in other fungal pathogens as

    well. In C. neoformans, overexpression of the efflux pumps encoded by AFR1 and MDR1

    confers a decrease in azole susceptibility (Sionov, Lee et al. 2010). A. fumigatus also encodes a

    major facilitator gene MDR1, but expression of this gene in S. cerevisiae does not impart

    resistance to azoles, but only to the drug cilofungin, which is not used clinically (Morschhauser

    2010). An ABC transporter encoding gene in A. fumigatus, atrF, was shown to be induced at the

    transcriptional level by azoles, and overexpression of this gene was correlated with azole

    resistance (Morschhauser 2010).

    1.3.2.4 Genome plasticity

    Aneuploidy can facilitate increased resistance to antifungals. While aneuploidy itself is not

    known to confer resistance, the increased dosage of specific genes on an extra chromosome can

    contribute to resistance. In C. albicans clinical isolates, evidence of genomic plasticity

    contributing to resistance is well documented. A survey of 70 azole-resistant and susceptible

    isolates revealed that resistant isolates had a preponderance of aneuploidies, the most common of

    which involved chromosome 5, often with duplications of the left chromosome arm.

  • 14

    Recombination events occurring at a common breakpoint in repetitive sequences flanking the

    centromere led to production of an isochromosome with two left arms of chromosome 5, termed

    i(5L) (Selmecki, Forche et al. 2006). Functional analyses confirmed that in i(5L) containing

    strains, azole resistance is due to duplication of both ERG11 and TAC1 alleles that reside on the

    left arm of chromosome 5 (Selmecki, Gerami-Nejad et al. 2008). C. neoformans also adapts to

    azoles by duplication of specific chromosomes, including chromosome 1, which harbors ERG11

    as well as the major azole transporter gene AFR1 (Sionov, Lee et al. 2010).

    1.3.2.5 Upregulation of stress response pathways

    Stress response pathways play a critical role in mediating both basal tolerance and resistance to

    diverse antifungals. For example, the molecular chaperone Hsp90 regulates crucial responses to

    both azoles and echinocandins in diverse fungi (Cowen and Lindquist 2005; Cowen 2008;

    Cowen 2009; Singh, Robbins et al. 2009). In C. albicans, Hsp90 regulates drug resistance by

    stabilizing key regulators of cellular stress responses, including the protein phosphatase

    calcineurin and the terminal mitogen activated protein kinase in the Pkc1 cell wall integrity

    signaling pathway, Mkc1 (Singh, Robbins et al. 2009; LaFayette, Collins et al. 2010). The cell

    wall integrity pathway regulates not only resistance to drugs that target the cell wall, but also

    resistance to drugs that target the cell membrane (LaFayette, Collins et al. 2010). Drugs that

    inhibit Hsp90, calcineurin, or Pkc1 reduce azole and echinocandin resistance of isolates that

    evolved resistance in a human host (Cruz, Goldstein et al. 2002; Singh, Robbins et al. 2009;

    LaFayette, Collins et al. 2010). While upregulation of stress response pathways has not yet been

    identified as a resistance mechanism, it is clear that these signaling pathways are key components

    of the cellular circuitry regulating drug resistance.

  • 15

    1.4 Strategies for thwarting the evolution of drug resistance

    1.4.1 Developing new antifungals

    There are relatively few drugs available to treat fungal infections and resistance to these

    antifungals has had a significant clinical impact, such that the development of new antifungals

    will be essential to keep the upper hand in the battle against these pervasive pathogens.

    Generally, new antifungals have an average of ten years after the start of clinical use before

    resistance becomes common. This necessitates the perpetual identification of new drug targets

    and small molecules to inhibit those targets.

    The development of new antifungals is complicated by the evolutionary relatedness of these

    eukaryotic pathogens to humans. To avoid host toxicity, differences between fungi and humans

    must be exploited to develop new drugs; however, new targets remain somewhat elusive. One set

    of candidates is the fungal-specific zinc cluster transcription factors, many of which regulate

    aspects of fungal drug resistance (such as Upc2, Tac1 and Mrr1). Efforts to identify drugs that

    inhibit these proteins have not yet been successful. The search for new fungal-specific drug

    targets will be an ongoing area of research for the control of fungal infections.

    1.4.2 Combination therapy

    Combination therapies have long been known to be potentially more effective than single drug

    therapy at limiting the development of resistance. Typically, combination therapy employs two

    antifungals that target more than one pathway (White, Marr et al. 1998). The synergistic effect of

    appropriately selected antifungals in combination will more substantially reduce the fungal

    population size, therefore reducing the genetic space over which mutations can arise.

    Furthermore, in order for resistance to occur, mutations would need to arise that permit fungal

    growth in the presence of both drugs.

  • 16

    One example of combination therapy is the use of 5-flucytosine and amphotericin B to treat

    fungal infections. This drug combination is commonly used in the treatment of cryptococcal

    meningitis (Bennett, Dismukes et al. 1979). Resistance to 5-flucytosine develops readily, but its

    effectiveness can be prolonged when used in combination with amphotericin B (White, Marr et

    al. 1998). This combination therapy is complicated, however, due to the nephrotoxicity of

    amphotericin B. Notably, 5-flucytosine can also be used in combination with azole therapy.

    Other combinations that show promise in vitro and in animal models are Hsp90 inhibitors with

    azoles or echinocandins in the treatment of C. albicans or A. fumigatus infections (Cowen, Singh

    et al. 2009). As with Hsp90 inhibitors, calcineurin inhibitors enhance antifungal efficacy and

    block the emergence of drug resistance, motivating the search for non-immunosuppressive

    agents that retain antifungal activity (Steinbach, Cramer et al. 2007).

    1.4.3 Targeting stress response: Hsp90 and calcineurin

    Traditional antifungals target fungal pathways required for proliferation or viability. Recent work

    has focused on inhibiting cellular factors that are required for the evolution of resistance but do

    not directly confer resistance. The molecular chaperone Hsp90 and its client protein calcineurin

    have been shown to be required for resistance or tolerance to echinocandins and azoles in

    Candida and Aspergillus species, as previously described in this chapter. Other regulators of

    fungal drug resistance exist: the protein kinase C (PKC) signaling pathway has been identified as

    regulator of fungal stress response such that with deletion of PKC1 in C. albicans or S.

    cerevisiae, resistance to azoles is abrogated (LaFayette, Collins et al. 2010). Here I focus on the

    regulators relevant to my work, Hsp90 and calcineurin.

  • 17

    1.4.3.1 Hsp90

    Hsp90 is a fascinating, highly abundant, essential protein in eukaryotes. As a molecular

    chaperone, it facilitates the folding of its meta-stable client proteins, helping prevent aggregation

    in the crowded environment of the cytoplasm under basal conditions (Taipale, Jarosz et al. 2010).

    Under proteotoxic stress conditions, such as high temperature, or exposure to heavy metals,

    hypoxia and acidosis, Hsp90 expression levels are increased in an effort to maintain protein

    homeostasis in a hostile environment (Whitesell and Lindquist 2005). Its clients are diverse but

    tend to be regulatory hubs, such as kinases and transcription factors. The factors that determine

    what make a protein an Hsp90 client had remained elusive until recently, when it was determined

    that thermal instability of a protein dictated its association with Hsp90 (Taipale, Krykbaeva et al.

    2012). It has been estimated that Hsp90 interacts with ~10% of the proteome in S. cerevisiae, and

    has comparably extensive interactions in C. albicans (Zhao, Davey et al. 2005; Diezmann,

    Michaut et al. 2012).

    Hsp90 is a highly conserved protein. It is composed of N-terminal, middle and C-terminal

    domains and functions as a homodimer, dimerizing at the C-terminal domain (Whitesell and

    Lindquist 2005). The flexible middle domain is thought to be important for client protein

    recognition, with the assistance of co-chaperones (Meyer, Prodromou et al. 2003). The N-

    terminal domain contains a site that binds to and hydrolyzes ATP as a requirement of Hsp90

    function. This nucleotide-binding pocket contains an unusual Bergerat fold, making it distinct

    from other ATPases (Dutta and Inouye 2000). Several Hsp90 inhibitors, including geldanamycin

    (GdA) and radicicol (RAD), competitively inhibit the binding of ATP to this nucleotide-binding

    pocket, which blocks Hsp90 function (Stebbins, Russo et al. 1997; Whitesell and Lindquist

    2005).

  • 18

    A fascinating aspect of Hsp90’s biology is its role as both a phenotypic capacitor and a

    phenotypic potentiator. Landmark studies in Drosophila melanogaster and Arabidopsis thaliana

    demonstrated that Hsp90 buffered underlying genetic variation that was exposed upon Hsp90

    depletion, demonstrating its role as a capacitor (Rutherford and Lindquist 1998; Queitsch,

    Sangster et al. 2002). Recently, it has been demonstrated that Hsp90 acts as a capacitor for eye

    morphology variation in surface populations of cavefish, whose cave-dwelling conspecifics lose

    their eyes. Raising surface populations of cavefish in cave conditions unmasks underlying

    genetic variation, similar to Hsp90 depletion, suggesting that Hsp90 plays a role in canalizing

    eye morphology (Rohner, Jarosz et al. 2013). Hsp90’s role as a phenotypic potentiator is

    exemplified by resistance to azoles and echinocandins in diverse fungi (Cowen 2013). A

    systematic study assayed the phenotypes upon Hsp90 inhibition of a diverse set of 102 S.

    cerevisiae strains, and found that Hsp90 worked as a phenotypic capacitor as frequently as a

    phenotypic potentiator (Jarosz and Lindquist 2010). In cancer cells, Hsp90’s physiological

    protein stabilizing activity is subverted to chaperone proteins that have become destabilized by

    oncogenic promoting mutations (Whitesell and Lindquist 2005). Hsp90 inhibitors, such as

    derivatives of GdA, are in clinical development as anti-cancer agents.

    Hsp90 inhibitors have recently been tested to assess their utility in combination therapy with

    azoles or echinocandins in the wax moth G. mellonella and mouse models of infection (Cowen,

    Singh et al. 2009). While effective at enhancing the efficacy of antifungals in the wax moth

    larval model of infection, toxicity to the mouse in the context of an acute fungal infection

    compromised the therapeutic utility although in a biofilm rat catheter model, there was no

    toxicity when the drug delivery and fungal infection were localized (Robbins, Uppuluri et al.

    2011). For systemic infections, fungal-specific Hsp90 inhibitors would therefore be an extremely

    valuable tool to prevent the development of resistance in vivo and extend the clinical usefulness

  • 19

    of antifungal drugs. Hsp90 holds great promise as a therapeutic target for extending the clinical

    efficacy of existing drugs to which antifungal resistance has already been documented.

    1.4.3.2 Calcineurin

    Calcineurin is a Ca+-activated, serine/threonine phosphatase that plays a crucial role in cellular

    stress response, from yeast to humans (Fox and Heitman 2002). In diverse fungi, it has roles in

    several critical cellular processes, including drug resistance, morphogenesis, virulence, cell cycle

    progression and ion homeostasis (Shapiro, Robbins et al. 2011). In humans, it is implicated in

    many critical developmental and cellular processes, most notably in immune response by

    dephosphorylation of the cytoplasmic component of the nuclear factor of activated T cells (NF-

    AT), which leads to transcriptional activation of IL-2 and T cell activation (Steinbach, Reedy et

    al. 2007).

    Calcineurin is a heterodimer, composed of a catalytic subunit and a regulatory subunit. Upon the

    release of calcium stores, Ca+-calmodulin binds to the catalytic subunit, which displaces its

    autoinhibitory domain and exposes its active site (Fox and Heitman 2002). The catalytic subunit

    of calcineurin is stabilized by Hsp90 in both C. albicans and S. cerevisiae (Imai and Yahara

    2000; Singh, Robbins et al. 2009). Activity of calcineurin also requires the association of the

    catalytic subunit with the regulatory subunit; mutant regulatory subunits unable to bind to the

    catalytic subunit abolish calcineurin activity (Watanabe, Perrino et al. 1996).

    The immune response controlled by calcineurin has distinguished it as a target for

    immunosuppressive drugs. FK506 and cyclosporin A are calcineurin inhibitors exploited for

    their inhibitory effect on T cell activation in humans. These drugs have a distinctive method of

    calcineurin inhibition: they bind to the peptidyl-prolyl isomerases Fpr1 and Cpr1 (for FK506 and

    CsA, respectively) and these drug-protein complexes bind to the hydrophobic interface between

  • 20

    the catalytic and regulatory subunits (Fox and Heitman 2002). Because neither of these drug-

    protein complexes block the active site, inhibition is thought to occur by occlusion of protein

    substrates. The regions to which FK506-Fpr1 and CsA-Cpr1 bind are distinct but overlapping

    (Huai, Kim et al. 2002). This inhibition of calcineurin by FK506 and CsA is also conserved

    between S. cerevisiae, C. albicans and humans.

    1.5 Natural variation in resistance to antifungal drugs

    Fungal pathogens are comprised of diverse members of the fungal kingdom. Variation in

    response to antifungals exists between both distantly and closely related species, as well as

    between strains within a species and even between cells within a single strain. Pathogens can

    differ in their inherent ability to survive and reproduce despite the detrimental presence of drug,

    often referred to as tolerance. Pathogens can also acquire specific mechanisms that reduce the

    inhibitory effect of the drug, referred to as resistance. High levels of basal tolerance can enable

    the evolution of resistance by maintaining viability in the context of strong selective pressure.

    The vast differences among species and strains in response to specific antifungals highlights the

    importance of selecting the appropriate antifungal to administer during clinical treatment.

    This section will discuss the differences in resistance and tolerance between and within the major

    pathogenic fungal species. Differences in resistance within a species will be covered and special

    phenotypic states conferring elevated resistance will be examined.

    1.5.1 Variation between species

    The major pathogenic fungi, Candida species, Aspergillus fumigatus and Cryptococcus

    neoformans, are very distantly related and vary dramatically in their ability to tolerate and

    acquire resistance to antifungals (Cowen 2008; Shapiro, Robbins et al. 2011). For example,

    fluconazole is one of the most frequently deployed antifungals in treating Candida infections, yet

  • 21

    it is completely ineffective against A. fumigatus (Marichal and Vanden Bossche 1995). While

    much remains to be understood about its intrinsic tolerance, A. fumigatus can be rendered

    susceptible to azoles by deletion of the transcriptional regulator of sterol production, srbA

    (Willger, Puttikamonkul et al. 2008). A. fumigatus is susceptible to other azoles, such as

    itraconazole and posaconazole (Pfaller and Diekema 2010). Tolerance to azoles in A. fumigatus

    can be abrogated by the addition of an inhibitor of Hsp90 or calcineurin (Cowen, Singh et al.

    2009). While C. neoformans is inherently susceptible to azoles, it shows remarkable tolerance to

    the echinocandins (Abruzzo, Griffith et al. 1997; Krishnarao and Galgiani 1997). How C.

    neoformans is able to tolerate echinocandin concentrations that are typically inhibitory is unclear,

    although it is known that C. neoformans β-(1,3)-glucan synthase (the echinocandin target) is

    inhibited by echinocandins (Thompson, Douglas et al. 1999). Echinocandin tolerance in C.

    neoformans is abrogated by inhibition of calcineurin (Del Poeta, Cruz et al. 2000). The

    echinocandins are fungicidal against C. albicans and fungistatic against A. fumigatus (Denning

    2003).

    Variation in tolerance also occurs between closely related fungal pathogens. Candida krusei and

    Candida glabrata, for example, have lower intrinsic susceptibility to fluconazole (Blot, Janssens

    et al. 2006). In C. krusei, this is mainly due to a low affinity for Erg11 (the target of azoles)

    (Lamping, Ranchod et al. 2009). Other azoles, however, can be effective in the treatment of C.

    krusei infection (Hoffman, Ernst et al. 2000). A recent study assaying the epidemiological cutoff

    value of voriconazole for over 16,000 strains of Candida species found reduced susceptibility in

    C. guilliermondii, C. glabrata and C. krusei and greater susceptibility in C. albicans, C.

    tropicalis, C. parapsilosis, C. kefyr and C. lusitaniae (Pfaller, Espinel-Ingroff et al. 2012). Unlike

    other Candida species, C. lusitaniae and C. guilliermondii are intrinsically resistant to

    amphotericin B. Candida species also vary in their sensitivity to echinocandins. C. parapsilosis,

  • 22

    for example, shows greater intrinsic resistance to caspofungin and its incidence in the clinic

    increases with the application of caspofungin to treat candidemia, while incidence of the more

    susceptible C. glabrata and C. tropicalis decrease (Forrest, Weekes et al. 2008; Arendrup,

    Rodriguez-Tudela et al. 2011).

    Variation exists between Aspergillus species in their susceptibility to the polyene amphotericin

    B. Testing hundreds of clinical and environmental Aspergillus isolates revealed higher levels of

    amphotericin B tolerance in Aspergillus flavus and Aspergillus terreus, while A. fumigatus and

    Aspergillus glaucus were more susceptible (Araujo, Pina-Vaz et al. 2007). It has been speculated

    that the intrinsic amphotericin B resistance of A. terreus is due to increased catalase production,

    which limits oxidative damage (Blum, Perkhofer et al. 2008). Additionally, most Aspergillus

    species are susceptible to echinocandins, but there is reduced susceptibility in select A. nidulans

    isolates (Araujo, Pina-Vaz et al. 2007).

    The mechanisms of resistance an organism acquires are influenced by the genetic architecture of

    tolerance. While most fungal pathogens ultimately evolve antifungal resistance by several

    mechanisms, the most prevalent mechanisms may vary between species whose intrinsic

    resistance varies. For example, resistance of A. fumigatus to azoles is often due to alteration in

    the drug target, cyp51A, the A. fumigatus equivalent of ERG11 (Balashov, Gardiner et al. 2005).

    C. neoformans infrequently evolves resistance to azoles, and does so largely by increased azole

    efflux or alteration in the drug target (Venkateswarlu, Taylor et al. 1997; Pfaller, Messer et al.

    2005), or via heteroresistance and aneuploidy for chromosome 1 (Sionov, Lee et al. 2010).

    Similarly, increased azole efflux and target alteration are common mechanisms of azole

    resistance in C. albicans (Shapiro, Robbins et al. 2011). With respect to echinocandins, the most

    frequently encountered resistance mechanism in C. albicans clinical isolates is mutation in the

  • 23

    drug target gene, FKS1 (Park and Morschhauser 2005). However, A. fumigatus echinocandin

    resistance is mostly facilitated through stress response pathways and rarely by FKS1 mutations

    (Gardiner, Souteropoulos et al. 2005).

    Mechanisms of resistance can vary between closely related species as well, like the model yeast

    Saccharomyces cerevisiae and C. glabrata, which is more closely related to S. cerevisiae than to

    other Candida species. S. cerevisiae is able to acquire fluconazole resistance with the addition of

    exogenous ergosterol by importing the sterol (Kuo, Tan et al. 2010). However, C. glabrata lacks

    sterol import genes and typically becomes fluconazole resistant by increased efflux (Kuo, Tan et

    al. 2010).

    1.5.2 Variation between strains

    Natural variation within a species can account for differing levels of tolerance among strains. For

    example, erg3∆-mediated azole resistance in S. cerevisiae varies between strains W303 and

    BY4741/2 in its dependence on downstream effectors of calcineurin, Crz1, Hph1 and Hph2, in a

    manner that is dependent on nutrient signaling (Cowen, Carpenter et al. 2006; Robbins, Collins

    et al. 2010).

    Clinically isolated Candida species have a broad diversity of genotypes. Four major clades of C.

    albicans strains have been identified by multilocus sequence typing and DNA fingerprinting

    (Odds, Bougnoux et al. 2007; MacCallum, Castillo et al. 2009). Nearly all 5-flucytosine-resistant

    samples are members of the first clade, which has acquired a mutation that confers resistance

    (Tavanti, Davidson et al. 2005). Likewise, one C. dubliniensis clade has been associated with 5-

    flucytosine resistance (Al Mosaid, Sullivan et al. 2005). Members of the C. albicans fourth clade

    demonstrate increased tolerance to amphotericin B (Blignaut, Molepo et al. 2005). However, no

    correlation between azole resistance and multilocus genotype was found in a set of C. albicans

  • 24

    clinical isolates, indicating that the azole resistance was too transient to become associated with a

    neutral marker (Cowen, Sirjusingh et al. 1999). Similarly, five C. glabrata clades have been

    identified and the isolates with the highest levels of azole resistance were distributed across

    clades (Dodgson, Pujol et al. 2003). Very little is known about the impact of population structure

    on resistance in the major fungal pathogens besides Candida species.

    1.5.3 Variation within a population

    Variation in antifungal resistance occurs not only between and within fungal species, but also

    within a population of cells. Unlike the predominantly stable variation in resistance found

    between strains and species, variation with a population of cells is often more transient.

    Heteroresistance is an intriguing example of variation in azole resistance in C. neoformans that is

    commonly documented in clinical isolates and in both serotypes A and D, as well as in C. gattii

    (Sionov, Chang et al. 2009). Heteroresistance is defined as the emergence of a subset of azole-

    resistant cells in addition to azole-susceptible cells from a susceptible progenitor. This

    subpopulation gains azole resistance in a stepwise manner and loses this resistance with extended

    passage in azole-free media (Sionov, Chang et al. 2009). Recently, comparative genomic

    hybridization experiments determined that the azole-resistant cells are disomic, most commonly

    for chromosome 1 (Sionov, Lee et al. 2010). The duplicated chromosomes contain genes

    important for resistance, such as the azole target ERG11 and azole transporter AFR1, and azole

    resistance was shown to correlate with the number of disomic chromosomes. Heteroresistance is

    also observed in strains of C. albicans (Marr, Lyons et al. 2001).

    There are several other distinct ways in which variation in resistance within a population has

    been described. Heterogeneous resistance is a broad term for a phenomenon in fungi and

    bacteria, which involves subpopulations of resistant cells that emerge at a low frequency (10-1 to

  • 25

    10-4) from a larger susceptible population; colonies formed by the resistant cells again give rise

    to the same distribution with only a few resistant cells among a larger susceptible population

    (Martinez, Lopez-Ribot et al. 2002). Another related phenomenon is high frequency azole

    resistance in which some strains produce azole-resistant colonies at a frequency that is higher

    than a typical mutation rate (Martinez, Lopez-Ribot et al. 2002). In contrast to heterogeneous

    resistance, the resistant colonies that emerge in the high frequency azole resistance context have

    stable resistance phenotypes and it is thought that altered mitochondrial function might play a

    role in this phenomenon.

    Biofilms are a clinically important example of heterogeneity in drug resistance. Fungal biofilms

    are complex surface-associated communities surrounded by an extracellular matrix (Blankenship

    and Mitchell 2006). Besides having important virulence implications, biofilms demonstrate a

    level of antifungal resistance that is unprecedented in planktonic cells. C. neoformans biofilms,

    for example, have significantly higher resistance to amphotericin B and the echinocandin

    caspofungin than planktonic cells (Martinez and Casadevall 2006). C. albicans biofilms are

    extremely resistant to azoles (d'Enfert 2006). There are several mechanisms that contribute to the

    resistance of C. albicans biofilms, including increased levels of β-(1,3)-glucan (Nett, Lincoln et

    al. 2007), but the most important may be the advent of persister cells. Biofilms offer refuge from

    antifungal drugs and some host immune response, which facilitates the development of persister

    cells that are able to tolerate high drug concentrations. Persisters are a phenotypic variant that is

    only able to arise from biofilms (LaFleur, Kumamoto et al. 2006). Persister cells are frequently

    found in bacterial biofilms; while bacterial persisters are quiescent, it is unclear whether fungal

    persisters are quiescent as well (LaFleur, Kumamoto et al. 2006). Persisters are thought to be a

    major factor affecting the recalcitrance of fungal infection to treatments.

  • 26

    1.6 Strategies for studying the evolution of drug resistance

    1.6.1 Examining clinical isolates

    One of the best ways to study how resistance to antifungals evolves in the human host is to

    examine fungal specimens that have been isolated from an infected individual over time. The

    most clinically relevant mechanisms of resistance can be elucidated and ultimately targeted to

    abrogate resistance. However, exclusively examining clinical isolates can be limiting due to the

    small sample sizes and the inability to control population parameters such as the number of

    generations, effective population size and the genotype of the initial, susceptible strain. This

    makes it challenging to determine what selective pressures directly contribute to the evolution of

    resistance. For these reasons, in vivo evolution of antifungal resistance studies are complemented

    by in vitro experimental evolution counterparts (Figure 1.2), as discussed in a later section. This

    section focuses on studies of the evolution of drug resistance in the human host.

  • 27

    Figure 1.2. Studying the evolution of resistance to antifungals. The evolution of resistance to

    antifungals can be monitored in real time and the mechanisms of resistance can be dissected, in

    both in vitro and in vivo evolved strains. A) Propagating a strain in the presence of antifungal

    selects for cells that have acquired a mutation that permits growth in the presence of the drug

    (grey cells). With continued propagation, additional mutations can accumulate (black cells) that

    confer resistance, abrogate the cost of resistance or have no beneficial effect on their own

    (‘hitchhiking’ mutations). B) Resistance evolves over time in the human host undergoing

  • 28

    antifungal treatment. Serial isolates are taken from the host and their resistance to the

    administered antifungal can be assayed. Frequently, multiple mutations conferring resistance are

    found in clinical isolates (grey and black cells). The image of the human form was generously

    provided by Carolyn Thomas. C) The clonality of evolved strains must be confirmed in order to

    determine how frequently beneficial mutations arose. Clinical isolates can represent several

    strains that the host was infected with. D) Resistance mechanisms can be determined in evolved

    strains by several methods.

  • 29

    1.6.1.1 Population dynamics of fungi in the human host

    When an antifungal-resistant fungal strain arises in an infected patient, it can be due to resistance

    mutations occurring in the initial strain, or due to replacement with a resistant strain or species

    (reviewed in (White, Marr et al. 1998)). Replacement with a resistant species can occur due to

    intrinsic differences in drug susceptibility between related species, and due to the fact that many

    fungi are environmentally ubiquitous. For example, it is well documented that during treatment

    a patient with an initially azole-susceptible C. albicans strain may develop a resistant infection

    due to acquisition of a species that is intrinsically less susceptible to azoles, such as C. glabrata.

    It is also possible that a patient can acquire a resistant strain of the same species, and multiple

    studies have focused on strain replacement in the clinical setting; between 20-33% of cases of

    resistance that developed during treatment were due to strain replacement. To distinguish clonal

    relationships from strain replacement requires the susceptible and resistant isolates from the

    patient and involves typing methods such as restriction-length fragment polymorphism (RFLP)

    analysis or multi-locus sequence typing (MLST), which both take advantage of genotypic

    variability between strains of the same species. Clonal relationships among isolates would

    suggest that resistance emerged in the lineage due to the acquisition of resistance mutations.

    1.6.1.2 Progressive drug resistance

    Matched sets of clinical isolates from fungal infections have informed the field a great deal as to

    how resistance mutations arise during the course of antifungal treatment. One of the best-

    characterized sets of isolates illustrates the concept of progressive drug resistance, or resistance

    that increases over time as sequential isolates accumulate multiple resistance mutations (White,

    Pfaller et al. 1997). This set of 17 isolates of C. albicans was collected over two years from an

    HIV-infected patient who was receiving azole treatment for recurrent oropharyngeal candidiasis.

  • 30

    The initial isolate in the series was highly susceptible to fluconazole treatment, but by the end of

    the two-year period had acquired a level of resistance approximately 200-fold higher than that of

    the original isolate. This set of isolates was extensively characterized to identify the molecular

    mechanisms of azole resistance (reviewed in (White, Marr et al. 1998; Anderson 2005)).

    Initially, there was a selection of a strain variant, as detected by single nucleotide

    polymorphisms. This was followed by overexpression of the efflux pump gene MDR1,

    mutations in the drug target ERG11, the A643V mutation in the transcriptional regulator UPC2

    that resulted in overexpression of ERG11, and ultimately overexpression of the drug efflux pump

    genes CDR1 and CDR2 due to mutations in the transcription factor TAC1 (White 1997; White

    1997; Marr, Lyons et al. 1998; Hoot, Smith et al. 2011). The progressive accumulation of

    resistance mutations in these isolates exemplifies how resistance often develops in a stepwise

    fashion, with high-level resistance due to the combination of multiple mechanisms.

    1.6.1.3 Whole genome sequencing to identify mechanisms of resistance

    Recent advances in sequencing technology allow for whole transcriptome or whole genome

    sequencing approaches to complement studies of antifungal drug resistance. For example, it

    would be feasible to perform whole genome sequencing (DNA-seq) to identify acquired

    molecular mechanisms of drug resistance in clinical isolates, as well as identify novel resistance

    mechanisms if matched sets of resistant and susceptible isolates are available. This approach is

    contingent on matched sets, because of tremendous variation in genotypes among strains that are

    not clonal. Indeed, even amongst clonal isolates from the same patient, there is a baseline level

    of variation in genotype in the fungal population that would complicate determining functionally

    relevant mutations. In addition to genome DNA-seq, RNA-seq technology could enable the

    identification of genes that are specifically overexpressed in drug-resistant isolates and thereby

    illuminate mechanisms of resistance. The most reproducible setting for genome-scale analyses

  • 31

    would be in vitro experimental evolution of drug resistance in the laboratory, where the initial

    population and selection conditions can be carefully controlled and reproduced.

    1.6.1.4 Fitness effects of resistance mutations

    Genetic alterations that confer resistance to an antifungal, such as point mutations in the drug

    target or specific aneuploidies, confer a selective advantage in the presence of the antifungal.

    However, these mutations may cause a fitness defect in the absence of drug. For example, some

    of the mutations that confer azole resistance occur in transcriptional regulators and result in

    altered expression of multiple genes (e.g. UPC2, TAC1, and MRR1 mutations). The pleiotropic

    effects of drug resistance may therefore be costly as discussed in the section below.

    1.6.2 Using experimental evolution

    A powerful method that is gaining prominence in the study of antifungal resistance is

    experimental evolution. Experimental evolution generally refers to a population initiated from a

    single progenitor that is propagated for several generations, such that new traits can evolve and

    be monitored in real-time (Elena and Lenski 2003). This broadly appealing approach is used to

    model theoretical evolutionary questions as well more pragmatic issues in strain development

    (Figure 1.2).

    This section will describe the contribution experimental evolution has made to our understanding

    of how antifungal resistance occurs. Comparisons to more exhaustive studies in bacteria and the

    model yeast S. cerevisiae will also be discussed.

    1.6.2.1 Benefits and limitations of experimental evolution

    In vitro evolution experiments can provide a simple model for how drug resistance is generated

    in the host. The researcher can manipulate parameters and thus more explicitly determine the

  • 32

    factors that give rise to drug resistance. Another benefit is that these experiments are easily

    replicated so the sample size can be much greater than what can be expected from retrieving

    clinical isolates from patients. Experimental evolution can be performed with strains where the

    genome is sequenced, such that mutations conferring resistance can be more readily identified.

    However, experimental evolution cannot capture the complexity of growth in the human host. In

    the human body, organisms experience diverse challenges including variation in

    microenvironments, nutrient limitation, spatial structure and competition with other, unrelated

    pathogens. Despite these limitations, there has been much success in the in vitro evolution of

    lineages that exhibit mechanisms of resistance found in clinical isolates (Cowen, Sanglard et al.

    2000; Scully and Bidochka 2005; MacLean, Hall et al. 2010).

    1.6.2.2 Experimental evolution methodology

    Employing experimental evolution to study the evolution of drug resistance in fungi involves

    propagating cells in the presence of an antifungal, which provides the selective pressure for the

    evolution of resistance. Typically, cells are passaged by serial dilution of stationary phase culture

    to fresh medium containing a concentration of antifungal that typically is inhibitory but not lethal

    (Dunham 2010). It is crucial that the population size at each transfer is great enough that any

    beneficial mutations (conferring resistance) that have occurred are not lost to random drift (Elena

    and Lenski 2003). This process is repeated until a sufficient number of generations have occurred

    for substantial changes to accrue in the population, conferring resistance. Occasionally,

    propagation occurs on solid medium containing antifungal. In this instance, cells are scraped or

    washed off of the plate to ensure a sufficient population size, and then transferred to a new plate.

  • 33

    1.6.2.3 Experimental evolution studies in fungi

    There are three major factors important in the evolution of drug resistance: the mutations

    conferring resistance; the cost of maintaining resistance in terms of reduced reproduction (fitness

    cost); and the nature of mutations that can mitigate this cost of resistance (compensatory

    mutations). These parameters have important treatment implications: the deployment of

    antifungals selects for resistance, if resistance is costly then discontinuing treatment may allow

    susceptible cells to out-compete their resistant counterparts, but if the cost of resistance is

    reduced by compensatory mutations, the resistant population will not be eliminated by selection.

    1.6.2.4 Experimental evolution studies in C. albicans

    The vast majority of experimental evolution studies with fungal pathogens have been performed

    with C. albicans. C. albicans replicate populations propagated for 330 generations in the

    presence of fluconazole evolved resistance to fluconazole, as well as cross-resistance to

    ketoconazole and itraconazole (Cowen, Sanglard et al. 2000). The expression levels of four

    genes known to be important in azole resistance in clinical isolates (ERG11, CDR1, CDR2 and

    MDR1) varied among the populations, indicating the occurrence of different mechanisms of

    resistance. Subsequent studies monitoring the expression of 5000 open reading frames in a subset

    of these evolved lineages identified three different expression profiles that matched expression

    patterns found in clinical isolates (Cowen, Nantel et al. 2002), further underscoring the utility of

    using in vitro evolution as a model for in vivo evolution of drug resistance. Aneuploidy also

    arose frequently in these populations and conferred resistance by increased gene dosage

    (Selmecki, Forche et al. 2006). For example, an isochromosome formed by the duplication of the

    left arm of chromosome 5 increased resistance through elevated expression of two genes:

    ERG11, encoding the drug target, and TAC1, encoding a transcription factor that regulates

    expression of the ABC transporter genes CDR1 and CDR2 (Selmecki, Gerami-Nejad et al. 2008).

  • 34

    Studies examining the fitness effects of resistance mutations in human pathogenic fungi have

    been performed exclusively in C. albicans to date. Fitness was measured by competing evolved

    strains against their progenitor in both selective (with antifungal) and non-selective conditions,

    and by measuring doubling times. Despite a predicted cost of resistance, the fitness of drug-

    resistant populations in non-selective medium was just as often significantly better than the

    progenitor as it was worse (Cowen, Kohn et al. 2001). In some populations, the chromosome 5

    isochromosome provided a fitness advantage in both the presence and absence of fluconazole

    (Selmecki, Dulmage et al. 2009). This suggests that resistance need not be costly, and if it is,

    compensatory mutations occur readily enough to ameliorate this cost. Furthermore, any cost of

    resistance in these lineages was abrogated with additional evolution, indicating compensatory

    mutati