<<

Histoplasma capsulatum: Drugs and Sugars

Dissertation

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate School of The State University

By

Kristie Goughenour

Graduate Program in

The Ohio State University

2020

Dissertation Committee

Dr. Chad A. Rappleye, Advisor

Dr. Stephanie Seveau

Dr. Jesse Kwiek

Dr. Jason Slot

1

Copyrighted by

Kristie Goughenour

2020

2

Abstract

Histoplasma capsulatum is a thermally dimorphic fungal capable of causing clinical symptoms in immunocompetent individuals. It exists as hyphae in the environment but transitions to the phase upon encountering the host, with exposure to mammalian body temperature triggering this phase change. It is endemic to the Ohio and River Valleys in the United States, Latin America (specifically

Brazil, Venezuela, Argentina, and Columbia), and parts of Africa, with limited reports in

China and India and demonstrates a clear medical relevance and healthcare burden.

Antifungal options for Histoplasma are limited due to a lack of fungal-specific targets.

Additionally, most studies do not take into account clinically relevant testing of fungal morphotypes and assume a one-size-fits-all approach to fungal drug testing, contributing to false leads and inaccurate frequencies of resistance. In this thesis, we develop and standardize an appropriate method for Histoplasma susceptibility testing. We show that current CLSI testing methodology is insufficient for testing of Histoplasma and that antifungal susceptibility is often phase-dependent in Histoplasma.

In addition to a need for novel , there is a real need for novel, fungal-specific drug targets. As a result, target identification is an important stage in antifungal ii development, particularly for large-scale screens of compound libraries where the target is completely unknown. In this thesis, we combine a traditional genetics approach and a novel co-fractionation mass spectrometry approach to identify target genes for a novel antifungal compound, 41F5. We generated 41F5-resistant strains and identified 11 potential target genes containing SNPs. In addition, we developed a biochemical approach using size co-fractionation of proteins with unmodified drug via size co-fractionation and compound detection by LC-MS/MS. Drug co-fractionating proteins were identified by proteomics, generating 34 candidate targets.

We are currently validating candidates through overexpression of the candidate genes.

This thesis also investigates virulence factors that could serve as novel fungal-specific drug targets. We work to characterize the O-linked mannosylated proteome in to identify proteins that may contribute to growth at elevated temperatures (38°C). Using a combined bioinformatics approach and a novel lectin-free biotinylation-based O-linked mannosylated purification we are working to identify O-linked mannosylated proteins and determine which contribute to Histoplasma thermotolerance.

We were additionally interested in other fungal-specific characteristics such as chitin as a main structural component of the fungal cell wall. As chitinases have not been comprehensively studied in fungi as a whole, this thesis generates a comprehensive phylogenetic survey of fungal chitinases in addition to characterization of enzymatic activities and gene expression of Histoplasma chitinases. We determined that the

iii previous phylogenetically determined A clade of chitinases was not monophyletic and that the A and C clades need to be recategorized. In Histoplasma, we identify a hyphal phase-specific chitinase, Cts3. We also identify a recent gene duplication event in

Histoplasma (Cts2 and Cts4) with differences in expression and enzymatic activity.

Finally, we show that enzymatic activity of chitinases does not follow previous phylogenetic predictions, indicating that more chitinases need to be studied to identify their roles in individual fungi.

iv

Dedication

This work is dedicated to my mom and dad for all their support over the years. I couldn’t have done it without them. It’s also dedicated to my friends both from grad school and before for all the good times and awesome science conversations over the years.

v

Acknowledgments

I want to begin by thanking and acknowledging by advisor Dr. Chad Rappleye. Your support and willingness to let me try new projects and techniques have been really fundamental to my development as a scientist. I’ve learned so much on how to do research well from being in his lab. In addition, I’d like to thank Jason Slot, for all his invaluable assistance with the phylogenetic work for the chitinases project. I’ve learned so much from it and his mentorship was critical for the project and helping me to develop new skills. I’d also like to thank my committee for their support and advice throughout the years.

I want to acknowledge my amazing colleagues (past and present). Andrew, Qian,

Stephanie, Peter, Amita and Janice. You have provided great scientific advice and amazing companionship throughout the years. I’d also like to acknowledge the other

OSU Microbiology graduate students and SAM. You all have provided a welcoming, helpful and collaborative environment.

vi

Vita

2014…………………………………………………………………….B.A. Microbiology

Ohio Wesleyan University

2014-2015…………………………………………………………...University Fellowship

Ohio State University

2015-2018………………………………………………… Graduate Teaching Associate

Department of Microbiology,

The Ohio State University

2018, 2019……………………………………………...NIH/NIAID T32 Graduate Fellow

CMIB/IDI

The Ohio State University

2020-Pesent………………………………………………… Graduate Research Associate

Department of Microbiology,

The Ohio State University

vii

Publications

Tan CY, Wang F, Anaya-Eugenio G, Gallucci J, Goughenour KD, Rappleye CA, Spjut

R, Carcache de Blanco E, Kinghorn AD, Rakotondraibe LH. 2019. α-Pyrone and Sterol

Constituents of aurantiacobrunneum, a Fungal Associate of the Lichen,

Niebla homalea. Journal of Natural Products. 82 (9), 2529-2536 DOI:

10.1021/acs.jnatprod.9b00340

Garfoot AL, Goughenour KD, Wüthrich M, Rajaram MVS, Schlesinger LS, Klein BS,

Rappleye CA. 2018. O-mannosylation of proteins enables Histoplasma yeast survival at mammalian body temperatures. mBio doi: 10.1128.

Goughenour KD, Rappleye CA. 2017. Antifungal therapeutics for dimorphic fungal . Virulence. 8(2):211-221. doi:10.1080/21505594.2016.1235653.

Goughenour KD, Balada-Llasat J-M, Rappleye CA. 2015. Quantitative microplate- based growth assay for determination of antifungal susceptibility of Histoplasma capsulatum . J Clin Microbiol 53:3286–3295. doi:10.1128/JCM.00795-15.

viii

Fields of Study

Major Field: Microbiology

ix

Table of Contents

Abstract ...... ii Dedication ...... v Acknowledgments ...... vi Vita ...... vii List of Tables ...... xiii List of Figures ...... xiv Chapter 1. Introduction ...... 1 1.1 Histoplasma capsulatum ...... 1 1.2 Antifungal treatment for Histoplasma ...... 3 1.2.1 Current Antifungals avaibile for Clinical use ...... 3 1.2.2 New and Alternative antifungal developments ...... 7 1.3 O-Linked Mannosylation and Chitinases in Histoplasma ...... 10 1.4 Conclusions ...... 11 Chapter 2. Quantitative Microplate-Based Growth Assay for Determination of Antifungal Susceptibility of Histoplasma capsulatum Yeasts ...... 14 2.1 Introduction ...... 14 2.2 Materials and Methods ...... 17 2.2.1 Fungal strains and culture ...... 17 2.2.2 Preparation of Histoplasma inocula ...... 18 2.2.3 Microtiter plate growth ...... 19 2.2.4 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide–based yeast vitality assay...... 19 2.2.5 Resazurin-based yeast vitality assay ...... 20 2.2.6 Antifungal drug susceptibility tests ...... 20 2.3 Results ...... 21 2.3.1 Microtiter-plate based growth of Histoplasma yeasts ...... 21 2.3.2 Optimization of microtiter-based assays with end-point indicators of yeast metabolism ...... 30 2.3.3 Application of the microtiter assay for drug sensitivity tests ...... 33 2.4 Discussion ...... 40

x

Chapter 3. Development of a novel method for target identification of the novel antifungal 41F5 ...... 45 3.1 Introduction ...... 45 3.2. Materials and Methods ...... 47 3.2.1. Yeast strains and growth ...... 47 Table 3.1 Fungal Strains used in this study ...... 48 3.2.2. Generation of Cryptococcus 41F5 resistant mutants ...... 49 3.2.3. Antifungal susceptibility testing ...... 49 3.2.4. Whole genome sequencing of Cryptococcus mutants ...... 50 3.2.5. Detection of 41F5 in cell lysate by HPLC-MS/MS ...... 51 3.2.6. Proteomics analysis ...... 53 3.2.7. Generation of Histoplasma strains overexpressing His-tagged potential target proteins ...... 56 3.3 Results ...... 57 3.3.1 Identification of SNPs from Cryptococcus 41F5 resistant mutants ...... 57 3.3.2 Co-fractionation approach ...... 60 3.3.3 Validation and Optimization of the Co-fractionation approach ...... 63 3.3.4 Identification of 41F5 Co-fractionating Proteins ...... 66 3.3.5 Identification of the target of 41F5 ...... 70 3.4 Discussion ...... 71 Chapter 4. Investigation into the O-linked mannosylated proteome of Histoplasma capsulatum ...... 75 4.1 Introduction ...... 75 4.2 Materials and Methods ...... 76 4.2.1 Histoplasma growth ...... 76 4.2.2. Proteomics analysis ...... 78 4.2.3. Bioinformatics ...... 80 4.2.4 Generation of RNAi Knockdowns ...... 80 4.2.5. Temperature sensitivity assay ...... 83 4.2.6. Determination of the O-linked proteome ...... 83 4.3 Results ...... 84 4.3.1 Mass spectrometry fails to adequately detect O-linked proteins ...... 84 4.3.2 Bioinformatics approach to identification of O-linked mannosylated proteins ...... 87 4.3.3 Biotinylation approach to identification of O-linked mannosylated proteins ...... 91 4.6 Discussion ...... 94 Chapter 5. Diversification of fungal chitinases and their functional differentiation in Histoplasma capsulatum ...... 97 5.1 Introduction ...... 97 5.2 Materials and Methods ...... 100 5.2.1 Phylogenetic Analyses ...... 100 5.2.2 Chitinase gene expression ...... 101 5.2.3 Purification of H. capsulatum chitinases ...... 104 5.2.4 Chitinase activity and specificity determination ...... 106 5.3 Results ...... 107 xi

5.3.1 Diversification of fungal GH18 domains ...... 107 5.3.2 Taxonomic distribution of GH18 chitinases ...... 111 5.3.3 Phylogeny of GH18 ...... 112 5.3.4 Evolution of architecture ...... 114 5.3.5 H. capsulatum chitinases ...... 120 5.4 Discussion ...... 128 5.4.1 Previous chitinase ontologies are largely robust to increased sampling, but the A clade may be polyphyletic...... 128 5.4.2 Some chitinase sub-clades have specific taxonomical biases with potential ecological and evolutionary ramifications...... 129 5.4.3 Chitinases containing LysM domains are distributed among plant pathogenic, insect parasitic and saprotrophic fungi ...... 132 5.4.4 Histoplasma capsulatum chitinase expression differences suggest specific functional roles ...... 133 5.4.5 Histoplasma capsulatum chitinase activities do not follow phylogenetically predicable trends in specificity ...... 135 5.4.6 Evolution of the Histoplasma chitinases indicates a degree of differentiation and multiplicity that is reflected in fungal chitinases in general reflecting a potential division of work for chitinases ...... 136 Chapter 6. Conclusions ...... 139 Bibliography ...... 145

xii

List of Tables

Table 1.1 In vitro antifungal MICs for dimorphic fungal pathogens ...... 5

Table 2.1 Antifungal susceptibility of clinical isolates ...... 37

Table 3.1 Fungal Strains used in this study ...... 48

Table 3.2 Primers used in this study ...... 56

Table 3.3 Antifungal susceptibility of Cryptococcus lines to 41F5 and ...... 58

Table 3.4 SNPs Identified in 41F5-resistant Cryptococcus most likely to contribute to resistance ...... 59

Table 3.5 Presence of 41F5 in different size fractions ...... 66

Table 3.6 Proteins identified as co-fractionating with 41F5 ...... 69

Table 3.7 IC50 of Candidate 41F5 Target Overexpressions ...... 71

Table 4.1 Histoplasma strains used in this study ...... 77

Table 4.2 Primers used in this Study ...... 82

Table 4.3 Predicted O-linked Mannosylated Proteins ...... 89

Table 5.1 Primers used in this study ...... 103

Table 5.2 Histoplasma strains ...... 105

Table 5.3 Bootstrap support values for chitinase sub-clades ...... 110

Table 5.4 Clade distribution of Chitin Binding Domains and Enzymatically Characterized

Members ...... 118 xiii

List of Figures

Figure 2.1 Growth of Histoplasma yeast in microtiter plates ...... 23

Figure 2.2 Correlation of OD595 with McFarland turbidity standards and yeast culture density...... 26

Figure 2.3 Optimization of inocula sizes for maximization of the dynamic range of the microplate-based assay...... 29

Figure 2.4 Optimization of parameters for metabolic reduction of MTT (A, B) or resazurin (C, D) for relative quantitation of Histoplasma yeast cell density...... 32

Figure 2.5 Antifungal dose-response curves for Histoplasma cells tested with 96-well microtiter plate microdilution assays...... 35

Figure 2.6 Antifungal susceptibility profile of Fluconazole-resistant isolates ...... 39

Figure 2.7 Summary of the optimized growth and assay parameters for microdilution testing of Histoplasma yeasts...... 41

Figure 3.1 Schematic for the Co-fractionation approach to 41F5 detection...... 62

Figure 3.2 Validation and optimization of 41F5 detection by UPLC-MS2...... 65

Figure 3.3 Detection of 41F5 in FPLC fractions...... 68

Figure 4.1 Proteomic approach fails to identify O-linked mannosylated proteins...... 86

xiv

Figure 4.2 Bioinformatic determination of candidates and initial temperature sensitivity assay...... 88

Figure 4.3 Biotinylation for purification of O-linked mannosylated proteins...... 93

Figure 5.1 Phylogenetic analysis of fungal GH18 domains...... 109

Figure 5.2 LYSM IQ Tree...... 115

Figure 5.3 Phylogenetic analysis of Hisch88_09649...... 116

Figure 5.4 Diagram of Histoplasma capsulatum chitinase genes...... 122

Figure 5.5 Histoplasma capsulatum chitinase expression...... 125

Figure 5.6 Histoplasma capsulatum chitinases enzymatic activity...... 127

xv

Chapter 1. Introduction

This chapter contains work adapted from Goughenour KD, Rappleye CA. Antifungal therapeutics for dimorphic fungal pathogens. Virulence. 2017;8(2):211–221. doi:10.1080/21505594.2016.1235653

1.1 Histoplasma capsulatum infection

Histoplasma capsulatum is a thermally dimorphic primary pathogen capable of infecting both immunocompromised and immunocompetent individuals. It exists in a hyphal form in the environment but transitions to yeast upon encountering the human host, and this morphological transition is required for virulence (Medoff et al. 1986; Nguyen and Sil

2008). Temperature appears to be the central factor that determines the form/lifestyle of the dimorphic fungi, with exposure to mammalian body temperature triggering adoption of the pathogenic mode (Horwath, Fecher, and Deepe 2015). It is endemic to the Ohio and valleys in the United States, Latin America (specifically Brazil,

Venezuela, Argentina, and Columbia), and parts of Africa, with limited reports in China and India (Edwards et al. 1969; Bahr et al. 2015; Zhao et al. 2001). These areas have been largely defined by clinical case prevalence (Chu et al. 2006) or skin-reactivity tests to Histoplasma antigens (Edwards et al. 1969; Kauffman 2007). Ecological aspects of soils that favor growth of the hyphal forms and/or animal or bird patterns for dispersal are thought to underlie endemicity. Inhalation of conidia initiates infection (Kauffman 2009).

The conidia are produced solely by the environmental mycelia. The small size of these

1 propagules facilitates aerosolization upon environmental disturbance and deposition into alveoli once inhaled. Thus, occupational and recreational activities that contribute to soil and environmental disruption are a primary risk factor for infection (Wheat et al. 2016).

Most infection outbreaks can be traced back to specific events that led to conidia release from soils and their subsequent inhalation (Kauffman 2009).

Within the lungs, conidia transition into yeast cells, the pathogenic phase which survive innate immune defenses. Histoplasma yeasts are internalized by host alveolar macrophages and reside almost exclusively within these host cells. Unlike opportunistic fungal pathogens, Histoplasma is not controlled by the innate axis of the immune system.

Control of the infection requires activation of CD4+ T cells, and consequently individuals lacking aspects of cellular immunity (e.g., HIV, immunosuppression due to tissue or organ transplantation, TNFα blockade, etc.) typically progress to severe and disseminated disease. Elimination of symptoms has been assumed to indicate clearance of the infection, but evidence is now suggesting that at least in some individuals, the infection can enter a latent state which can re-emerge later when the balance between pathogen and host immunity is altered (e.g., immunosuppression of the host) (Ankobiah et al. 1990;

Jain, Evans, and Peterson 2006; van Koeveringe and Brouwer 2010; Nakelchik and

Mangino 2002; Vergidis et al. 2015).

Since inhalation is the route of exposure, is initially a pulmonary disease.

In immunocompetent individuals, mild disease is mostly subclinical, often going undiagnosed. Infection causes varying degrees of pneumonia and influenza-like symptoms. For the majority of individuals, symptoms typically resolve without requiring

2 intervention. Roughly 5% of Histoplasma are estimated to require clinical management (Hage et al. 2015). For individuals inhaling a larger inoculum, or those that have some deficiency in cellular immune response, disease is more severe and the infection typically disseminates to extrapulmonary sites via the hemolymphatic system.

The annual incidence of infections by Histoplsama is likely inaccurate as under-diagnosis and under-reporting of infections is common. Furthermore, most infections are self- limiting without requiring clinical intervention. Nevertheless, reports of 6.1 cases per

100,000 person-years in the Midwest of the United States are seen (Vallabhaneni et al.

2016). The number of clinical cases typically range from 5-30% of infections, although in some outbreaks report 68% with symptomatic infections (Chamany et al. 2004). One survey of hospital records in the United States tallied over 3360 cases of histoplasmosis hospitalizations in one year (Chu et al. 2006). Less than half of these were in immunocompromised individuals.

1.2 Antifungal treatment for Histoplasma

1.2.1 Current Antifungals available for Clinical use

The three major classes of currently avaibile antifungals inclues the polyenes, the and the . The minimum inhibitory concentrations of the currently availvble antifungals for Histoplasma are shown in Table 1.1. Other endemic thermally dimorphic pathogenic fungi are included for comparison. Of important note is the higher MICs to

3 echinocandins, the least cytotoxic antifungal, for the pathogenic phases of endemic thermally dimorphic pathogenic fungi (Table 1.1 and Chapter 2). The reasons underlying this natural resistance of yeasts to the β-glucan synthase inhibitors is currently unknown.The in vitro antifungal susceptibilities for and - drugs have been validated in murine models of dimorphic fungal disease and/or in clinical trials and as a consequence, the Infectious Disease Society of America has released treatment guidelines for infections with Histoplasma (Wheat et al. 2007). Despite its potential for host toxicity, amphotericin B is recommended for severe disseminated disease, with one of the liposomal formulations preferred. For less severe situations, for follow up after amphotericin B, or for possible prophylaxis of highly susceptible individuals (e.g. AIDS),

IDSA guidelines recommend treatment with an oral azole (i.e., ) with monitoring of serum concentrations to ensure sufficient absorption and bioavailability.

To ensure sufficient clearance of the Histoplasma infection, treatments typically involve protracted regimens. For mild disease, treatment durations can range from several months to a year depending on the specific dimorphic . Treatment of disseminated disease and disease in immunocompromised hosts can be a year or longer depending on the antifungal used. Beyond the prolonged treatment times, the potential for latency, rather than clearance of the infection, further complicates antifungal management of

Histoplasma.

4

Table 1.1 In vitro antifungal MICs for dimorphic fungal pathogens

MIC range (µg/mL) Drug class Antifungal Histoplasma Blastomyce Paracoccidioide References s s Polyenes Amphoterici Y: <0.03-2.0 Y:<0.03- Y: 0.06-2.0 Y: 0.25-2.0 (Andreu et al. 2003; Cordeiro et al. 2006; n B M: 0.26-2.5 2.0 M: No data M: 0.03-0.50 Cruz et al. 2013; González et al. 2001; M:No data Goughenour, Balada-Llasat, and Rappleye 2015; Kathuria et al. 2014; Kohler et al. 2000; A M LeMonte et al. 2000; Ren-Kai Li et al. 2000; Nakai et al. 2003; Ramani and Chaturvedi 2007; Sugar and Liu 1996; Zhanel et al. 1997) Y: No data Y: <0.01- Y: <0.01-0.03 Y: (de Aguiar Cordeiro et al. 2006; Andreu et M: 0.17 0.25 M: No data M: 0.03-0.16 al. 2003; Chapman et al. 1998; Cruz et al. M: 0.1-0.4 2013; Ramani and Chaturvedi 2007; Zhanel et al. 1997) Fluconazole Y: 0.25-8.0 Y: 0.06-32 Y: 0.13-0.50 Y: No data (Chapman et al. 1998; Cordeiro et al. 2006; M: 2.0-32 M: 0.06-32 M: No data M: 2.0-64 Cruz et al. 2013; van Duin, Casadevall, and Nosanchuk 2002; González et al. 2001; González et al. 2005; Goughenour, Balada- Llasat, and Rappleye 2015; Kathuria et al. 2014; A M LeMonte et al. 2000; Nakai et al. 2003; Ramani and Chaturvedi 2007; Sugar and Liu 1996; Wheat et al. 2006; Zhanel et al. 1997) Itraconazole Y: <0.01-0.5 Y:<0.01- Y: <0.01-0.06 Y: <0.03-0.50 (Andreu et al. 2003; Chapman et al. 1998; M: 0.03-1.0 0.13 M: No data M: 0.03-1.0 Cordeiro et al. 2006; Cruz et al. 2013; van M: 0.03-4 Duin, Casadevall, and Nosanchuk 2002; González et al. 2005; Kathuria et al. 2014; Ren-Kai Li et al. 2000; Nakai et al. 2003;

5 Ramani and Chaturvedi 2007; Sugar and Liu 1996) Y: 0.03-0.50 Y:<0.03- Y: No data Y: <0.03-2.0 (Cordeiro et al. 2006; González et al. 2005; M: <0.01-2.0 0.25 M: No data M: 0.03-1.0 Kathuria et al. 2014; Ren-Kai Li et al. 2000; M: 0.06-2.0 Ramani and Chaturvedi 2007; Wheat et al. 2006) Y: <0.01- Y: <0.02- Y: No data Y: No data (Espinel-Ingroff 1998; González et al. 2005; 0.50 0.06 M: No data M: 0.06-1.0 Kathuria et al. 2014; Sugar and Liu 1996; M: 0.02-2.0 M: <0.02- Wheat et al. 2006) 2.0 Echinocandin Y: >64 Y: 32-64 Y: >64 Y: No data (Nakai et al. 2003) s M: 0.03-0.06 M: <0.01- M: 4-16 M: 0.02 0.03 Caspofungin Y: 8-32a Y: No data Y: No data Y: No data (Cordeiro et al. 2006; Espinel-Ingroff 1998; M: 0.02-4.0 M: 0.5-8.0 M: No data M: 8-64 González et al. 2001; Goughenour, Balada- Llasat, and Rappleye 2015; Kohler et al. 2000) a two studies reported low MICs for yeast which disagree with the majority of studies based on clinical isolates in India (0.03-1.0 µg/mL) (Kathuria et al. 2014) and for a single laboratory Histoplasma strain (MIC <0.125 µg/mL) (van Duin et al. 2002)

6 1.2.2 New and Alternative antifungal developments

With the known host toxicity risks of current antifungals (Bates et al. 2001; Cheng et al. 1982;

Nett and Andes 2016), and the endogenous resistance of dimorphic fungi to the lower toxicity echinocandins, new and alternative antifungal drug options have been explored. The eukaryotic nature shared by both the dimorphic fungal pathogens and the mammalian host is recognized as a significant obstacle to antifungal drug development. Consequently, many molecules with promising antifungal activity fail due to inadequate selectivity.

Following the goal of reduced host toxicity, some currently clinically used drugs are being evaluated for repurposing as antifungals. This includes as their safety profiles are well-established. For example, drugs targeting the folate pathway have been effective against microbes from bacteria to parasites (Capasso and Supuran 2014), are orally available, and have a moderately safe host toxicity profile. These drugs have already been a means of prophylaxis against Pneumocystis fungal infections in HIV+ individuals. Sulfonamides in combination with dihydrofolate reductase inhibitors (e.g. cotrimoxazole) have activity against Histoplasma yeasts

(8-14 µg/mL and 2-3 µg/mL, respectively) (Brilhante et al. 2010). Using a similar rationale, ciprofloxacin, a fluoroquinolone antibacterial with very low host toxicity, has been tested for antifungal potential. The fluoroquinolones inhibit bacterial type II topoisomerases (e.g., DNA gyrase) and were suggested to potentially also inhibit fungal topoisomerase. While not effective against mycelial forms of Histoplasma, tests against Histoplasma yeasts showed ciprofloxacin has an in vitro MIC of 62.5-250 µg/mL (Brilhante et al. 2013). While not considered a low MIC, the lack of significant host toxicity for ciprofloxacin creates a good selectivity index and raises

7 its potential, at least as an adjunct therapy for fungal infections. Addition of ciprofloxacin to amphotericin B or itraconazole indicated some additive effects could be gained through drug combinations. Anti-tuberculosis drugs show some, albeit low potency, antifungal activity for dimorphic fungal pathogens. When tested in vitro against Histoplasma yeasts, pyrazinamide, isoniazid, and ethambutol had very high MICs (540 µg/mL, 126 µg/mL, and 1240 µg/mL, respectively (Cordeiro et al. 2011)) arguing these antibacterial compounds lack any significant antifungal utility. However, using isoniazid as the core, chemical structure modification improved its antifungal potential; three of nine isoniazid derivatives had significantly lower

MICs against Histoplasma yeasts with one having inhibitory activity at concentrations as low as

7.8 µg/mL (de Aguiar Cordeiro et al. 2014). Antiretroviral protease inhibitors have inhibitory activity against Histoplasma. Saquinavir is active against both filamentous and yeast forms with an MIC around 0.4 µg/mL (Brilhante et al. 2016). Another protease inhibitor, ritonavir, has an

MIC of 1.0 µg/mL against Histoplasma mycelia but is 7-fold more potent against yeast cells. As

HIV+ individuals comprise a significant portion of clinical Histoplasma cases, this raises the possibility that management of HIV through protease inhibitor cocktails could simultaneously provide prophylaxis against infection by dimorphic fungal pathogens. Repurposing of anti- cancer drugs has also been fruitful in the search for new antifungal prospects. AR-12 is a celecoxib-derivative that has been shown to have antimicrobial properties against a wide range microbes including bacteria, parasites, fungi, and even viruses (Baxter et al. 2011; Chiu et al.

2009; Collier et al. 2016; Hoang et al. 2016; Mohr et al. 2015). Early mechanistic studies in cancer cells suggested AR-12 inhibited cellular phosphoinositide-dependent kinase-1 (Pdk1), however this does not appear to be the case in fungi. Recently, it was shown that AR-12 inhibits acetyl-CoA synthetase, an essential in fungi (Koselny et al. 2016). AR-12 has antifungal

8 activity against a broad collection of fungi, including the dimorphic fungal pathogens at concentrations around 4-8 µg/mL (Koselny et al. 2016). In vitro, AR-12 is fungicidal against

Histoplasma yeasts (Rappleye CA, personal communication). These antifungal characteristics combined with AR-12’s safety as established in Phase I trials make AR-12 an attractive antifungal candidate for further development.

There are also some novel drugs in development as specifically antifungal compounds. Farnesol, an alcohol present in some essential oils but is also produced by C. albicans as a quorum sensing molecule, inhibits Histoplasma yeast growth with an average MIC of 0.02 µM (Brilhante et al.

2015) As a caution, tests with Histoplasma yeasts followed the CLSI methodology (M27-A3), which has been shown to be inferior for reliable testing of Histoplasma yeast susceptibility (see chapter 2 (Goughenour, Balada-Llasat, and Rappleye 2015)). The most advanced new antifungal option for clinical development to manage dimorphic fungal pathogens is nikkomycin Z. The polyoxins (from which the nikkomycins are derived) are peptide modified-nuceloside analogs which were originally identified by screening Streptomyces products for antifungal and insecticidal activities (Isono, Asahi, and Suzuki 1969). These compounds were later shown to be inhibitors of chitin synthesis (Endo and Misato 1969). Since chitin synthase is absent from mammalian cells, these compounds have very high selectivity for fungi. Following polyoxin studies on C. albicans, the nikkomycins were demonstrated to have good potency in vitro against dimorphic fungal pathogens, including Histoplasma (Garfoot et al. 2016; Goldberg et al. 2000;

Hector, Zimmer, and Pappagianis 1990). The nikkomycins are orally available, simplifying extension of studies to animal models of mycoses caused by dimorphic fungal pathogens.

Administration of nikkomycin Z in murine models of lethal histoplasmosis resulted in decreased

9 organ fungal burdens (3- to 28-fold) and a concomitant enhancement of mouse survival at doses from 20 to 100 mg/kg per day (Goldberg et al. 2000; Graybill et al. 1998; Hector, Zimmer, and

Pappagianis 1990). However, significant protection from lethal infection was only realized at moderate inocula suggesting nikkomycin antifungal activity may not be sufficiently rapid to curb severe infections. Still the ability to safely administer such doses is a testament to the high selectivity gained by targeting the fungi-specific molecule chitin synthase.

1.3 O-Linked Mannosylation and Chitinases in Histoplasma

Many proteins secreted by the pathogenic yeast phase of Histoplasma but not the avirulent mycelia phase are heavily glycosylated indicating a probable role in virulence (Holbrook et al.

2011). Recently our lab demonstrated that the post-tanslational modification of O-linked mannosylation is essential for Histoplasma virulence (Garfoot et al. 2018). O-linked mannosylation in Histoplasma generates linear mannose chains attached to serines and theonines and involves an initial mannose attachment by the protein mannosyltransferase (Pmt) family of proteins (Pmt1, 2, and 4), with subsequent extension carried out by a single Mnt1 (Garfoot et al.

2018). We demonstrated that strains lacking Pmt1, Pmt2, and Mnt1 were severely attenuated in a mouse model of infection (Garfoot et al. 2018). Cells lacking Pmt2 had reduced detection by the

β-glucan receptor Dectin-1 although cells were still able to enter and survive within phagocytes

(Garfoot et al. 2018). The virulence defect in o-linked deficient cells was linked to a lack of thermotolerance by these mutants leading to yeast death at 38°C (Garfoot et al. 2018).

Thermotolerance in fungi is considered essential for virulence during systemic infections

(Bhabhra and Askew 2005). While previous studies in Histoplasma have shown that blocking the temperature-dependent transition to yeast is essential for virulence, the factors needed to 10 proliferate at these higher temperatures are not well understood (Shen and Rappleye 2017).

Identifying which proteins are o-linked mannosylated (and therefore may be non-functional in the mannosylation-deficient strains) could pinpoint factors Histoplasma needs to survive at high temperatures. Once these features are identified they may serve as fungal-specific drug targets allowing for the inhibition of Histoplasma thermotolerance and therefore death of the fungus.

As previously mentioned, one of the most promising novel antifungal compounds is nikkomycin

Z, which is an inhibitor of chitin synthesis (Endo and Misato 1969). Chitin is a major structral component of the fungal cell wall, ranging from 0.5%-5% in yeasts to ≥20% in filamentous fungi

(Hartl, Zach, and Seidl-Seiboth 2012). Chitinases are glycosyl hydrolases that break the glycosidic bonds between the N-acetylglucosamine monomers which make up chitin. The roles of chitinase are highly varible in fungi and are not particularly well-characterized into patterns.

In Histoplasma, the chitinases have not been invesigated. In other organisms, some chitinases have shown roles in cell wall remodeling during various phases of life and morphologies (Duo-

Chuan 2006; Seidl 2008b). Chitinases and the remodeling of the fungal cell wall may play important roles in Histoplasma survival and morphology and may serve as potiential drug targets.

1.4 Conclusions

Current clinical management of infections by Histoplasma is limited to azole-class antifungal drugs and amphotericin B. While orally available, the azoles are not without host toxicity issues and the treatment course is lengthy for infections by dimorphic fungi. Development of resistance to azoles is not widespread, although treatment failures due to azole resistance have occurred in 11

Histoplsama and other dimorphs (Kriesel et al. 2008; Wang et al. 2005; Wheat et al. 2001).

Unfortunately, the better-tolerated antifungals lack efficacy against the pathogenic- phase of the dimorphic fungal pathogens raising the need for alternative or second-line treatment options (Table 1 and Chapter 2).

While a number of strides have been made in repurposing existing drugs and development of new inhibitors of fungal growth, careful attention must be paid to the unique challenges posed by

Histoplasma. As only the yeast state of Histoplasma is present within the mammalian host, antimicrobial susceptibilities need to be performed with these pathogenic-phase cells, not the mycelia which has led to erroneous conclusions (See Chapter 2). Standard CLSI methodology does not address concerns for specific testing of dimorphic fungi and needs to be optomized for

Histoplasma (See Chapter 2). Additionally, since Histoplasma yeast cells reside within host phagocytes, it is also advisable for in vitro tests to be followed by tests on drug effectiveness on intracellular yeasts, at least during initial drug development stages. Our lab has attempted to take this approach with the indtification of novel anti-Histoplasma compounds (Edwards, Kemski, and Rappleye 2013).

The overall selectivity of antifungal drug candidates is critical for progression of drugs through the development pipeline. Many of the repurposed drugs have relatively high MICs (greater than

100 ug/mL), bringing into question their therapeutic utility. However, if their selectivity is sufficiently high, formulations may be developed to facilitate the high serum and tissue levels required for drug efficacy. Lower MICs have been found for drugs targeting the folate pathway: an isoniazid-hydrazone derivative, antiretroviral protease inhibitors, and the anti-cancer drug

12

AR-12, all of which are expected to be reasonably well-tolerated by the mammalian host. Novel drugs with good in vitro MICs and good selectivity include thioredoxin-reductase inhibitors, an aminothiazole compound, and nikkomycin Z. Since nikkomycin Z targets an enzyme absent from the host, nikkomycin has an excellent basis for high selectivity against fungi. In addition, nikkomycin Z has maintained antifungal effectiveness against multiple dimorphic fungal pathogens in animal models of disease. The identification of additional novel fungal-specific drug targets would be of great benefit to furthering antigungal development. To this end, the identification of novel virulence factors, such as O-linked mannosylated proteins which may confer thermotolerance to Histoplasma during infection, and a better understanding of the basic biology of the organism, such as the role of chitinases, are required to generate new fungal- specific drug targets.

13

Chapter 2: Quantitative Microplate-Based Growth Assay for Determination of Antifungal

Susceptibility of Histoplasma capsulatum Yeasts

This chapter was previously published as Goughenour KD, Balada-Llasat JM, Rappleye CA.

Quantitative Microplate-Based Growth Assay for Determination of Antifungal Susceptibility of

Histoplasma capsulatum Yeasts. J Clin Microbiol. 2015;53(10):3286–3295. doi:10.1128/JCM.00795-15

2.1 Introduction

The increasing incidence of fungal disease necessitates adequate and timely assessment of antifungal susceptibility to guide the selection and implementation of antifungal therapies.

Consequently, the Clinical and Laboratory Standards Institute (CLSI) and the European

Committee on Antimicrobial Susceptibility Testing (EUCAST) have established reference test methods for susceptibility to the major classes of antifungals, which include the polyenes, azoles, and the recently developed echinocandins (Cuenca-Estrella and Rodriguez-Tudela 2010; Pfaller and Diekema 2012). These standards utilize broth macrodilution or microdilution assays for testing yeasts (M27-A3 (Rex and Clinical and Laboratory Standards Institute 2008) and E.DEF

7.1 (Subcommittee on Antifungal Susceptibility Testing (AFST) of the ESCMID European

14

Committee for Antimicrobial Susceptibility Testing (EUCAST) 2008)) and filamentous fungi

(M38-A2 (Reference Method for Broth Dilution Antifungal Susceptibility Testing of

Filamentous Fungi. 2008) and E.DEF 9.1(Subcommittee on Antifungal Susceptibility Testing of the ESCMID European Committee for Antimicrobial Susceptibility Testing 2008)). Procedures, including inoculum preparation, media, assay duration, and endpoint definitions, have now been established for testing some of the more commonly encountered yeast-form pathogenic fungi

(i.e., Candida and Cryptococcus). However, notably missing from the M27-A3 and

E.DEF 7.1 standards for testing of yeasts are methods for testing the yeast forms of the dimorphic fungi.

The dimorphic fungi that cause systemic disease are characterized by distinct yeast and filamentous morphological states (Klein and Tebbets 2007; Maresca and Kobayashi 2000).

Temperature is the principal morphology-determining factor, with yeast forms characterizing infections in mammals (Kauffman 2007). Histoplasma capsulatum is the most common clinically encountered dimorphic fungal pathogen in the United States, with an estimated 10,000 to 20,000 hospitalizations annually (Chu et al. 2006). Histoplasma is acquired by inhalation of mycelia-produced conidia from the environment. In the lung, mammalian body temperature triggers differentiation of the conidia into yeasts instead of filamentous hyphae (Kauffman

2007). Histoplasma yeasts parasitize phagocytic cells, and the yeast-infected phagocytes can facilitate extrapulmonary dissemination of the fungus to cause life-threatening disseminated histoplasmosis, particularly in immunocompromised individuals. In the majority of otherwise healthy individuals, low-dose inoculation results in a self-limiting disease with the onset of cell- mediated immunity. With higher doses, Histoplasma causes acute disease, even in

15 immunocompetent individuals (Chu et al. 2006; Hage, Knox, and Wheat 2012).

Despite the fact that the yeast form is the state most relevant to disease, some in vitro antifungal testing for Histoplasma has been performed only on the filamentous form (Borelli et al. 2008;

Espinel-Ingroff 1998; González 2009). Both yeast- and filamentous-form Histoplasma cells are generally susceptible to polyenes and -class antifungals in vitro (Cheung et al. 1975;

Hage et al. 2011; Nakai et al. 2003). Early studies on the echinocandins suggested that this antifungal class, with very low host toxicity, could be effective against Histoplasma infections

(Espinel-Ingroff 1998; Johnson and Perfect 2003; van Duin, Casadevall, and Nosanchuk 2002).

However, some of these early reports were based on tests with the filamentous form, and subsequent in vitro susceptibility tests employing the yeast form demonstrated that the

Histoplasma yeasts are comparatively resistant to the echinocandins (Hage et al. 2011; Nakai et al. 2003; Kohler et al. 2000). This morphology-based discrepancy in antifungal susceptibility is not limited to echinocandins and highlights the need to use the most appropriate cell type during antifungal testing (Cordeiro et al. 2016). With the increasing importance of monitoring antifungal drug resistance and the acceleration of new antifungal drug discovery efforts, optimized methods for high-throughput antifungal tests on Histoplasma yeasts are essential.

In this study, we standardized a broth microdilution method for assaying Histoplasma yeast growth. This study was motivated by the inadequacy of CLSI-defined yeast culture procedures for the growth of Histoplasma yeasts, the clinical form of Histoplasma. In some previous studies, the CLSI M27-A2 and -A3 standards have been adapted for Histoplasma yeast (Edwards,

Kemski, and Rappleye 2013; Wheat et al. 2001; Wheat et al. 2006; Andreu et al. 2003), but no

16 consensus methods or optimization of assay parameters have been detailed. Here, we optimize the inoculum size and assay duration to maximize the dynamic range of a microdilution assay for

Histoplasma. The assay quantifies yeast growth through spectrophotometric measurement of culture turbidity to provide a more objective conclusion of endpoints than visual inspection allows. To complement culture density measurements, we also determined optimal parameters for the use of colorimetric and fluorescent indicators of yeast cell density and vitality. These methodologies were validated for use in a clinical microbiology laboratory setting by establishing antifungal susceptibilities of multiple clinical isolates from both North American phylogenetic groups of Histoplasma and a Latin American isolate.

2.2 Materials and Methods

2.2.1 Fungal strains and culture

Histoplasma capsulatum strains used were the wild-type North American type 2 (NAm2) strain

G217B (ATCC 26032), the wild-type Panama strain G186A (ATCC 26029), 7 clinical sample isolates from the Ohio State Medical Center, and 3 clinical isolates from MiraVista Labs (Wheat et al. 2006; Wheat, MaWhinney, et al. 1997; Wheat, Marichal, et al. 1997). Isolates were assigned to the North America NAm1 and NAm2 phylogenetic groups (Kasuga et al. 2003) based on PCR-restriction fragment length polymorphism markers (Bohse and Woods 2007).

Histoplasma cells were maintained as yeasts by growth at 37°C on a medium optimized for

Histoplasma yeast growth (Histoplasma macrophage medium [HMM]) supplemented with 25

17

µM FeSO4 and solidified with 0.6% agarose (30) or on a brain heart infusion (BHI) medium

(Becton, Dickenson and Co.) solidified with 0.6% agarose instead of agar. For liquid culture, basal cell culture media were used as the rich growth media. F-12 (Gibco) and RPMI 1604

(Gibco) media were compared, as F-12 is the base composition of HMM medium, and RPMI is more commonly used in clinical laboratories for microdilution assays. Both broth media were buffered to a pH of 7.0 with 25 mM HEPES and were optionally supplemented with 1.5% glucose and/or 0.7 mM cystine for nutritional tests. For growth in microtiter plates (see below), yeasts were added to wells in a total volume of 100 µL. For quantitative platings, serial dilutions of yeast suspensions were plated on solid HMM medium, as this medium supports the robust growth of individual colonies, and incubated at 37°C until colonies developed (8 to 10 days).

2.2.2 Preparation of Histoplasma inocula

Microtiter plate inocula of Histoplasma yeasts were prepared either by suspending yeast colonies in F-12 medium or from broth cultures of Histoplasma yeasts that had been pregrown in culture tubes for 48 h. For pregrowth of Histoplasma yeasts before dilution into microtiter plates, yeast colonies were inoculated into 2.5 ml liquid medium in 16 mm × 100 mm test tubes. Tubes were placed at a 30-degree angle and incubated for 2 days at 37°C with shaking (200 rpm) in a

Multitron orbital shaker (Infors, Inc.). Yeast density was determined directly using hemacytometer counts, estimated by optical density at 595 nm (OD595), or by a turbidity comparison to McFarland standards. For estimation of yeast density by OD595, a standard curve of OD595 versus the yeast cell number was used. For comparison to McFarland standards,

McFarland standards (of 0.5, 1.0, or 1.5) were prepared by mixing defined volumes of 1% BaCl2

18

(1.175% BaCl2 dihydrate) with 1% H2SO4 (50 µl, 100 µl, and 200 µl of BaCl2 in a final volume of 1 mL). Yeast suspensions were adjusted with growth medium to a density equivalent to 1.0

McFarland standard (representing approximately 8 × 106 yeasts/ml) and further diluted to the desired inoculum.

2.2.3 Microtiter plate growth

A total of 50 µL of yeast culture at twice the desired final yeast density was added to wells of a

96-well flat-bottomed microtiter plate. Subsequently, 50 µl of growth medium or medium containing antifungal drugs at twice the desired concentration (typically 2-fold dilutions from 32

µg/mL to 0.03 µg/mL) was added to each well for a final volume of 100 µL. Plates were incubated at 37°C for 5 days with twice-daily agitation (30 s at 1,000 rpm) on a microplate mixer

(Eppendorf) to improve aeration. Culture turbidity was determined by measurement of the absorbance of each well at 595 nm with a Synergy 2 microplate reader (Biotek).

2.2.4 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide–based yeast vitality assay.

Histoplasma yeasts in 100 µL of growth medium were added to 96-well plates and grown as described above for approximately 96 h. A 5× solution of 3-(4,5-dimethylthiazol-2-yl)-2,5- diphenyltetrazolium bromide (MTT) was prepared in growth medium containing 0.5 mM menadione as an electron shuttle. Then, 25 µL of the MTT solution was added to each well that

19 contained Histoplasma yeasts as well as to wells that contained growth medium without

Histoplasma yeasts. Plates were incubated at 37°C to allow for reduction of MTT and sufficient color development. The assay was terminated by the addition to each well of 50 µL of 20% SDS with 10 mM HCl. After at least 2 h of incubation at ambient temperature for reduced MTT extraction, the microtiter plates were centrifuged for 5 min at 2,000 × g to collect yeast cell debris, and 50 µl of the supernatant was transferred to a microtiter plate containing 50 µl of distilled H2O. The extracted MTT was quantified by absorbance readings at 570 nm after correction with absorbance at 690 nm using a Synergy 2 plate reader (BioTek).

2.2.5 Resazurin-based yeast vitality assay

Histoplasma yeasts in 100 µL of growth medium were added to 96-well plates and grown as described above for approximately 96 h. 10 µL of a 10X concentrated solution of resazurin (7- hydroxy-10-oxidophenoxazin-10-ium-3-one) were added to each well, including control wells containing growth media without yeast cells and the microtiter plates incubated at 37°C. Kinetics of resorufin (reduced resazurin; 7-hydroxyphenoxazin-3-one) production was quantified by measurement of fluorescence (530/25 excitation, 590/35 nm emission wavelengths) every 20 min using a Synergy2 plate reader (BioTek).

2.2.6 Antifungal drug susceptibility tests

Antifungals used in clinical practice were applied to Histoplasma yeast and included

20 amphotericin B, fluconazole, and caspofungin. Antifungals were prepared according to the CLSI reference standard M27-A3 (Rex and Clinical and Laboratory Standards Institute 2008). For the

50% inhibitory concentration (IC50) and MIC determinations for Histoplasma yeasts, a 2-fold dilution series of antifungal drugs was prepared, and 50 µL of drug-containing solution was added to the wells of a 96-well microtiter plate containing 50 µL of Histoplasma yeasts at a density of 4×106 to 8×106 yeasts/mL (producing a final yeast density of 2 × 106 to 4 × 106 yeasts/mL). Yeast growth was assessed by measurement of turbidity at 595 nm after 96 to 144 h and the growth was normalized to wells containing no antifungal drug. Endpoint resazurin reduction assays were subsequently employed to confirm OD595 readings. Relative growth at each antifungal drug concentration was used to generate dose-response curves, and the IC50 was determined by nonlinear regression. MIC values were based on the minimum concentration of antifungal drug that prevented development of visual turbidity or lack of resazurin fluorescence compared to growth in the absence of antifungal drug. The morphology of Histoplasma cells following drug treatment was evaluated by phase-contrast microscopy at 100X magnification.

2.3 Results

2.3.1 Microtiter-plate based growth of Histoplasma yeasts

To determine the maximal yeast cell density that could be achieved in microtiter plates, replicate plates (n = 3) were inoculated with 1 × 107, 1 × 106, 1 × 105, and 1 × 104 yeasts/mL of the

G217B strain, a clinical strain representative of North American isolates. Plates were incubated at 37°C for 5 days. Yeast growth and viability were monitored daily by measurement of optical

21 density at 595 nm (OD595) and by plating serial dilutions of the yeast suspension to enumerate

CFU, respectively. Maximal growth in 96-well microtiter plates saturated at approximately 2.5 ×

107 CFU/mL (Fig. 2.1A). Yeast growth reached saturation at 2 to 3 days, 4 days, and 5 days for 1

× 107, 1 × 106, and 1 × 105 yeasts/mL of starting culture density, respectively. Yeast cultures started at 1 × 104 yeasts/mL significantly lagged in growth and reached saturation only after 6 to

7 days of incubation (data not shown). Yeast viability as determined by CFU closely paralleled the optical density of each well at 595 nm (Fig. 2.1B), indicating that optical density was a reliable indicator of yeast cell growth. Importantly, Histoplasma microdilution assays started at both the high and low inoculum levels recommended for yeasts by the CLSI procedures (2.5 ×

103 and 0.5 × 103 yeasts/mL, respectively (Rex and Clinical and Laboratory Standards Institute

2008) resulted in a complete failure of Histoplasma yeast growth (data not shown).

22

Figure 2.1 Growth of Histoplasma yeast in microtiter plates

Histoplasma yeasts were grown in 96-well microtiter plates, and the yeast cell density was measured by CFU (A) or optical density at 595 nm (OD595) (B). Growth media were inoculated with 1 × 107 (circles), 1 × 106 (squares), 1 × 105 (diamonds), and 1 × 104 (triangles) yeasts/mL, and microtiter plates were incubated at 37°C. Plates were shaken (60 s at 1,000 rpm) twice daily to improve aeration. Yeast growth was assessed at 24-hour intervals. (A) For quantitative platings, dilutions of yeast cultures from 96- well microtiter plates were plated on solid HMM medium to enumerate viable CFU. (B) Culture turbidity was measured by OD595 in a plate reader. (C) Effect of supplementation of growth media with glucose (Glc) or cysteine (Cys) on yeast cell 6 density (OD595). Data shown represent growth at 5 days after inoculation with 4 × 10 yeasts/ml in Ham's F-12 (black bars) or RPMI (gray bars) buffered to a pH of 7.0. Data points represent the average ± standard deviation of biological replicate cultures (n = 3).

23

As maximal growth obtained in microtiter plates was less dense than when Histoplasma yeasts were grown in culture tubes, we investigated whether glucose or cysteine was limiting in the

RPMI-based medium. CLSI procedures for growth of yeast-form organisms (i.e., Candida and

Cryptococcus) typically utilize a defined rich medium (e.g., RPMI) buffered to a pH of 7.0 (Rex and Clinical and Laboratory Standards Institute 2008). HMM, a medium optimized for growth of

Histoplasma and macrophages, is based on Ham's F-12 cell culture medium and is supplemented with 1.8% glucose and 0.7 mM cystine (Worsham and Goldman 1988). The increased cystine provides additional organic sulfur, since Histoplasma yeast cells and mycelia differ in their requirements for organic sulfur (Boguslawski and Stetler 1979; McVeigh I, Houston WE 1972;

Pine 1954; Salvin 1949; Stetler and Boguslawski 1979). Although both F-12 and RPMI have cysteine (0.2 mM and 0.4 mM, respectively), we tested if supplementation of F-12 or RPMI with cystine increased the yield of Histoplasma yeasts in microtiter plates. Neither added cystine nor added glucose significantly improved the overall rate or total growth of Histoplasma yeasts; no significant differences in yield were detected by one-way analysis of variance (ANOVA) (n = 3;

Fig. 2.1C). With some strains of Histoplasma, F-12 supported 20% to 80% higher yeast cell densities than RPMI (data not shown). We suspect that trace metals (i.e., Fe2+, Cu2+, Zn2+) present in F-12 but absent from the RPMI formulation may account for the improved growth in the F-12 medium. Thus, F-12 (lacking any supplements) buffered to a pH of 7.0 was selected as the growth medium for microtiter plate-based growth of Histoplasma.

Enumeration of yeasts by hemacytometer counts was used to establish a standard curve of optical density versus yeast cell density and to correlate Histoplasma yeast cell number to McFarland standards. Since spectrophotometers vary in their readings for light scattering due to sample

24 turbidity, the McFarland standards provide a defined optical density reference. McFarland standards of 0.5 to 4.0 were prepared in triplicate, and their optical densities at 595 nm were determined using a plate reader (Fig. 2.2A). In parallel, different yeast culture densities

(determined by hemacytometer counts) were correlated to optical density readings (Fig. 2.2B).

Both linear relationships were subsequently combined to produce a standard curve of the turbidity reference (McFarland standards) and the Histoplasma yeast cell density (Fig. 2.2C). A

McFarland standard of 1.0 was equivalent to 8.4 × 106 Histoplasma yeasts/mL.

25

Figure 2.2 Correlation of OD595 with McFarland turbidity standards and yeast culture density.

Optical density readings (OD595) were correlated to McFarland turbidity standards (A) or yeast cell density (B). (A) McFarland standards were constructed by mixing BaCl2 with 1% H2SO4 and then measuring the optical density. Data points represent individual readings for replicate cultures (n = 3), and the dashed line indicates the computed linear relationship (B) For determination of yeast cell density, dilutions of Histoplasma yeast cultures were counted by hemacytometer to enumerate yeast cells. Data points represent individual Histoplasma cultures. Dashed line indicates the relationship between OD595 and yeast cell density by linear regression. (C) The corresponding OD595 for each McFarland standard was matched to the OD595 standard curve for Histoplasma yeast culture density to yield a linear relationship for McFarland standards and the titer of the yeast culture.

26

To optimize the dynamic range of the growth of Histoplasma yeasts in 96-well microtiter plates, we investigated the parameters of starting yeast cell density and incubation time using cultures of

G217B in triplicate. Starting inocula of less than 1 × 105 yeasts/mL displayed significant lags in growth (Fig. 2.1A) or no growth at all. While inocula of around 1 × 107 yeasts/mL produced maximal growth in a short time (2 to 3 days), these inocula already showed significant optical density in the absence of any growth, thereby decreasing the dynamic range (Fig. 2.3A). Starting concentrations of around 5 × 105 yeasts/mL reached maximal growth at 4 days. Comparison of the OD595 readings at 3 to 4 days (72 to 96 h) with the OD595 readings at the start of the assay showed that initial inocula at 2 × 106 to 4 × 106 yeasts/mL had the largest dynamic range (Fig.

2.3B) and still produced near-maximal growth by 4 days of incubation. A total of 96 h of incubation provided the maximum difference, with an overall dynamic range of 35-fold and 77- fold for 4 × 106 and 2 × 106 yeasts/mL inocula, respectively. As a further test for optimal starting yeast cell density, different inocula of Histoplasma yeasts were grown in microtiter plates in the presence or absence of the antifungal drug amphotericin B. Calculation of the Z′-factor, a statistical measure of the discriminatory power of an assay for comparing growth in the presence or absence of drug (Zhang, Chung, and Oldenburg 1999), showed that, while all starting inocula could discriminate between growth and no growth (Z′ > 0.5), an inoculum of 2 × 106 to 4 × 106 yeasts/mL was optimal. From these results, we established 4 × 106 yeasts/mL as the starting yeast cell density and 96 h for the growth time for the microtiter plate-based OD595 growth assay.

Greater consistency in growth kinetics was found when the initial inoculum was prepared from a liquid culture of Histoplasma yeast pregrown in culture tubes for 2 days than when yeast

27 suspensions were made from colonies collected directly from solid medium. This was not due to transfer of any spent medium, since washing the pregrown cells before inoculation of microtiter plates yielded identical results as simply diluting pregrown cells (data not shown). The initial inoculum for the microdilution assay in 96-well plates may be determined by hemacytometer counts or by the adjustment of a yeast suspension to a McFarland standard of 1.0 followed by a

2-fold dilution of the adjusted suspension.

28

Figure 2.3 Optimization of inocula sizes for maximization of the dynamic range of the microplate-based assay.

(A) Representative growth curves of Histoplasma yeasts initiated at starting densities of 3.2 × 107 (circles), 8.0 × 106 (squares), 2 × 106 (diamonds), or 5 × 105 (triangles) yeast cells/mL. Data points represent average optical density readings ± standard deviation of replicate cultures (n = 3). (B) Determination of the dynamic range of OD595 readings for different inocula after growth at 37°C. The dynamic range was determined as the difference between the initial OD595 reading at the time of 0 h and then at the time of 72 h (diamonds), 87 h (squares), or 96 h (circles). (C) Z′-factor statistic indicating the discriminatory power of the microtiter plate-based assay of Histoplasma yeast growth. Cultures were initiated at 5 × 105 yeasts/mL to 3.2 × 107 yeasts/mL. For computation of the Z′-factor, wells containing cultures grown in the absence of antifungal drug were compared to wells cultured in the presence of 5 µg/ml of amphotericin B to suppress growth. Data points represent the average ± standard deviation of replicate cultures (n = 3).

29

2.3.2 Optimization of microtiter-based assays with end-point indicators of yeast metabolism

Measurement of optical density provides an easy and efficient way to quantify yeast growth over time. Metabolic indicators of cell density, based on metabolism-dependent reduction of colorimetric and fluorescent dyes, provide an alternative measure of yeast cell density as well as provide an indicator of yeast vitality. Two accepted endpoint assays are cellular reduction of the tetrazolium dye MTT to its formazan, which can be monitored by absorption at 590 nm, and reduction of resazurin to fluorescent resorufin. The colorimetric change in both reactions can also be used as a more objective qualitative indicator than visual determination of turbidity. To determine the suitability of these assays for measuring Histoplasma yeast cell growth, we tested the ability of Histoplasma yeasts to reduce these dyes and optimized the dye concentration used and the timing of maximal color/fluorescence development.

To determine the optimal concentration of MTT and resazurin substrates to be used, MTT was added to suspensions of Histoplasma yeasts at final concentrations of 0 to 1.0 mg/mL, and resazurin, at final concentrations of 0 to 200 µM (n = 3 for each). We used yeast cell densities at or above the maximum titer obtained in 96-well microtiter plate growth to ensure that the substrates would not be limiting for yeast grown to saturation density in microtiter plates. A dose-response relationship between the yeast cell number and formazan formation was observed

(Fig. 2.4A). Saturating levels of the MTT substrate for the highest yeast cell densities occurred around 0.4 mg/mL. Thus, 0.5 mg/mL MTT was chosen as the substrate concentration that would not be limiting for the assay with Histoplasma yeasts. To determine the timing for maximal

30 formazan formation, we grew Histoplasma yeasts in microtiter plates for 96 h, added 0.5 mg/mL

MTT, and incubated the plates at 37°C for different time intervals from 0 to 240 min before stopping the reaction. Maximal MTT reduction to its formazan was achieved at 180 min after

MTT addition (Fig. 2.4B). As with the MTT assay, a dose-response relationship between the yeast cell number and resorufin formation from resazurin was observed (Fig. 2.4C). Levels of the resazurin substrate became saturating at 100 µM. Concentrations of resazurin above 100 µM partially inhibited the reduction of resazurin to resorufin (Fig. 2.4C and data not shown).

Maximum fluorescence was achieved by 100 to 120 min of incubation with 100 µM resazurin

(Fig. 2.4D). At 90 min, the reaction was near but not yet at saturation levels; thus, 90 min was designated the optimal time for the resazurin reduction assay.

To determine if the MTT- and resazurin-based assays were suitable to detect inhibition of

Histoplasma growth, the assays were applied to Histoplasma yeasts grown in microtiter plates in the presence of the antifungal drug fluconazole (Fig. 2.4B and 2.4D). This test showed that the

MTT assay could efficiently discriminate between growth and inhibition of growth of

Histoplasma yeasts, with a dynamic range of 7-fold. No additional formazan was formed by incubation times longer than 180 min; however, the background of nongrowing yeasts continued to increase, thereby falsely reporting the relative difference between inhibited and uninhibited yeasts. Comparison of the resazurin assay between Histoplasma yeasts grown in the presence or absence of the antifungal drug fluconazole showed that the resazurin assay also had excellent discrimination between yeast growth and growth inhibition, with a Z′-factor of 0.85 and a dynamic range of 12-fold.

31

Figure 2.4 Optimization of parameters for metabolic reduction of MTT (A, B) or resazurin (C, D) for relative quantitation of Histoplasma yeast cell density.

For optimization of substrate concentration (A, C) Histoplasma yeast suspensions at 1 × 108 (circles), 7 × 107 (squares), 4 × 107 (diamonds), and 1 × 107 (triangles) yeasts/mL were incubated with various amounts of substrate. For optimization of the timing of the development reactions (B, D), Histoplasma yeasts in wells of a 96-well microtiter plate were grown for 96 h in the absence (circles) or presence (squares) of 4 µg/mL of fluconazole (Flc) to suppress yeast growth and then incubated with MTT or resazurin substrates. MTT reduction was monitored by formazan product formation, which was extracted from cells with SDS and quantified by absorbance at 570 nm. Resazurin reduction was monitored by fluorescence of the resorufin product. Histoplasma yeasts were incubated with MTT at final concentrations ranging from 0 to 1.0 mg/mL (A), with 0.5 mg/ml MTT and the formazan product extracted at 30 min intervals for quantification (B), with resazurin at final concentrations ranging from 0 to 200 µM (C), or with 100 µM resazurin and the fluorescence of the resorufin product measured at 20 min intervals (D). Data points represent the average ± standard deviation of replicate cultures (n = 3).

32

2.3.3 Application of the microtiter assay for drug sensitivity tests

To validate the optimized 96-well microtiter plate assay for drug sensitivity studies, we determined the concentrations for 50% inhibition (IC50) and the MICs of antifungal drugs against the G217B strain of Histoplasma capsulatum (Fig. 2.5). Using the OD595 parameters optimized above, yeast growth was determined over a range of antifungal drug concentrations in triplicate cultures. As optical density can quantify partial yeast growth, dose-response curves could be constructed, and the IC50 could be determined from linear regression. MICs were determined by visual inspection of the plate for lack of turbidity. Treatment of Histoplasma yeasts with amphotericin B showed an IC50 of 1.0 ± 0.03 µg/mL (mean ± standard deviation [SD]) and an

MIC of 2.0 µg/mL (Fig. 2.5A). Fluconazole had an IC50 of 0.24 ± 0.01 µg/mL (mean ± SD) and an MIC of 0.5 µg/mL (Fig. 2.5B). For the echinocandin caspofungin, Histoplasma yeasts were characterized by an IC50 of 10.4 ± 0.5 µg/mL (mean ± SD) and an MIC of 32 µg/mL (Fig. 2.5C).

These results are in agreement with established drug sensitivities for Histoplasma by macrodilution assay and modified microdilution assays (Kohler et al. 2000; Wheat et al. 2006; A.

M. LeMonte et al. 2000; R. K. Li et al. 2000). Examination of Histoplasma cells by microscopy at inhibitory and subinhibitory drug concentrations showed no drug-induced morphological conversion to mycelia from the yeast-form inocula (data not shown). Using the microdilution assay, we confirmed that the antifungal susceptibility profile of yeasts differed from mycelia; 20- fold more caspofungin was required to inhibit yeasts than mycelia (Fig. 2.5C). In contrast mycelia were more resistant to fluconazole than were the Histoplasma yeasts (Fig. 2.5B). No significant difference in the susceptibility to amphotericin B was found between the two

Histoplasma morphologies (Fig. 2.5A), consistent with previous observations of amphotericin B

33 effects on yeasts and mycelia (Hage et al. 2011). The variation in antifungal drug effects on the different morphologies underscores the importance of using the clinically relevant form (i.e., yeasts) for antifungal susceptibility profiling of Histoplasma.

34

Figure 2.5 Antifungal dose-response curves for Histoplasma cells tested with 96-well microtiter plate microdilution assays.

Charts depict representative dose-response curves of yeast to the antifungal compounds amphotericin B (A), fluconazole (B), and caspofungin (C) as determined by quantitative growth assay (turbidity at 595 nm). Histoplasma yeasts were inoculated into buffered F-12 media at a density of 2 × 106 yeasts/mL and grown for 96 hours, at which point the OD595 was measured. Yeast cell density was normalized to wells with no antifungal drugs added, and the relative growth was plotted. Data points represent the average ± standard deviation of replicate cultures (n = 3) for each antifungal drug concentration tested. Trend lines indicate the regression line for the yeast-phase dose response which was used to derive the IC50 for each antifungal drug. MICs were determined by visual inspection of wells and identification of wells without turbidity.

35

We applied the optimized microtiter plate-based assay to test the drug sensitivity profile of clinical isolates of Histoplasma using yeast-phase cells. The isolates represent clinical strains from the North America phylogenetic group 1 (NAm1), the more clinically prevalent North

America group 2 (NAm2), and a Latin American isolate (LAm), as well as from a laboratory strain originally from Panama (G186A; Pan). The antifungal susceptibility results of these clinical isolates are presented in Table 2.1. In general, Histoplasma clinical isolates were similar in their susceptibility to amphotericin B but NAm1, LAm, and the Panama isolates had slightly higher natural resistance to fluconazole (2- to 3-fold higher IC50). All isolates were relatively resistant to caspofungin, with MICs above 8 µg/ml. Of note, the NAm1, LAm, and Panama isolates were more resistant than NAm2 clinical isolates to the β-glucan synthase inhibitor caspofungin.

36

Table 2.1 Antifungal susceptibility of clinical isolates Susceptibility datab (µg/mL) for: Amphotericin B Fluconazole Caspofungin Phylogenetic a Isolate group IC50 MIC IC50 MIC IC50 MIC 0.11 ± 10– HC17 NAm1 0.02 0.24 1.88 ± 0.52 2.5 8.9 ± 2.4 20 0.08 ± 0.06– 0.64– 10– HC01 NAm2 0.03 0.16 0.56 ± 0.03 1.28 8.1 ± 2.0 16 0.06 ± 0.04– 0.64– HC10 NAm2 0.04 0.16 0.48 ± 0.11 1.28 5.3 ± 0.9 8–10 0.05 ± 0.04– HC14 NAm2 0.04 0.16 0.59 ± 0.22 1.28 5.5 ± 0.8 8–10 0.07 ± 0.06– 0.32– HC16 NAm2 0.04 0.32 0.25 ± 0.07 0.64 5.6 ± 3.6 8–16 0.06 ± 0.06– 0.64– HC22 NAm2 0.03 0.16 0.81 ± 0.16 1.28 6.4 ± 2.9 8–16 0.06 ± 1.28– 11.6 ± HC30 Lam 0.01 0.12 1.12 ± 0.43 2.56 0.6 20 G186 0.15 ± 0.24– 2.56– 23.4 ± 32– A Pan 0.06 0.64 1.82 ± 0.45 5.12 7.4 40 0.04 ± 12.3 ± IN2 NAm2 0.01 0.12 12.2 ± 0.5 32 0.8 20 0.07 ± 0.06– 145.3 ± 34.4 ± 40– IN14 NAm2 0.01 0.13 14.8 256–512 6.9 80 0.15 ± 0.25– 10.3 ± NY17 RFLP-VI 0.02 0.50 52.1 ± 29.7 128–256 3.5 40 aPhylogenetic group classification as described by Kasuga et al. (Kasuga et al. 2003): NAm1, North American type 1; NAm2, North American type 2; LAm Latin American isolates or, in the case of NY17, based on the RFLP classification scheme of Keath et al. (Keath, Kobayashi, and Medoff 1992). b IC50 expressed as average ± standard deviation (n = 3); MIC expressed as range (low to high; n = 3).

37

As further confirmation of the parameters for the Histoplasma microdilution assay for clinical and laboratory application, we tested three isolates with established resistance to fluconazole.

Clinical isolates IN2, IN14, and NY17 originated from AIDS patients who had failed or relapsed following fluconazole therapy (Wheat et al. 2006; Kasuga et al. 2003; Bohse and Woods 2007).

Compared to typical inhibitory levels of fluconazole for NAm2-class Histoplasma yeasts, we observed 22- to 200-fold increases in the IC50s and MICs of fluconazole for isolates IN2 and

IN14, respectively (Table 2.1 and Fig. 2.5). The NY17 strain also had decreased susceptibility to fluconazole (MICs, 128 to 256 µg/mL). Interestingly, the IN14 isolate also showed approximately 3-fold higher resistance to caspofungin (Fig. 2.5), even though this antifungal was not part of the treatment regimen for the patient from whom it was isolated. Further investigation of the very high fluconazole resistance of the IN14 isolate identified a mutation in the ERG11B gene that resulted in substitution at amino acid 165 (L165R; data not shown).

38

Figure 2.6 Antifungal susceptibility profile of Fluconazole-resistant isolates

Microdilution assays were used to test the susceptibility of fluconazole-resistant isolates IN2 (blue symbols), IN14 (red symbols), and NY14 (green symbols) to fluconazole (A), caspofungin (B), and amphotericin B (C). G217B (black symbols) susceptibility is provided as a wild-type NAm2 strain for comparison. Dose-response curves were computed by non-linear regression and error bars represent standard- deviations among tests (n=3). 39

2.4 Discussion

In this study, we optimized parameters for growing and measuring Histoplasma yeast growth in

96-well microtiter plates. Differences in antifungal susceptibilities of Histoplasma yeasts cells compared to mycelia (Hage et al. 2011; Nakai et al. 2003; Kohler et al. 2000) raise the importance of standardizing methodologies for Histoplasma yeasts, the clinically relevant morphological form. Although CLSI has developed procedures for yeast-form fungal pathogens, these methods do not work for the culture of Histoplasma yeasts. Some studies on Histoplasma have adapted the CLSI macrobroth procedures for dimorphic fungi (Wheat et al. 2001; Wheat et al. 2006; Connolly et al. 1999; R. K. Li et al. 2000), but no optimized methodology for a microdilution assay has been standardized. Our optimized methods are summarized in Fig. 2.7 and differ from the CLSI procedures for yeast by the following: higher inoculum density (2 × 106 to 4 × 106 yeasts/mL compared to 500 to 2,500 yeasts/mL), longer growth time (4 to 6 days compared to 1 to 2 days) due to the much slower growth of Histoplasma, and growth in lower volume of media (100 µL compared to 200 µL) and in flat-bottomed wells (instead of U-bottom wells) to improve oxygen diffusion. Using these parameters, saturated growth densities of approximately 2.5 × 107 Histoplasma yeasts/mL were achieved. We suspect that reduced oxygen availability caused by growth in microtiter plate wells accounts for the lower densities than are typically achieved in culture tubes. Nonetheless, sufficient growth was present to reliably distinguish growth of Histoplasma over the inoculum density. We observed that improved consistency of growth of Histoplasma yeasts in microtiter plates was achieved if the initial inoculum was prepared from pregrown liquid cultures. Nevertheless, preparation of the initial inoculum by suspension of yeast colonies taken directly from solid BHI medium was sufficient for the microdilution assay if the yeasts were obtained from freshly streaked plates (4 to 8 days

40 old).

Figure 2.7 Summary of the optimized growth and assay parameters for microdilution testing of Histoplasma yeasts.

Inoculum size, growth medium, and growth conditions are listed for initiation of the tests in 96-well microtiter plates. Assays that can be used for determination of Histoplasma growth (culture turbidity, MTT reduction, and resazurin assays) provide both visual (qualitative) and spectrophotometric (quantitative) results. Further details of the endpoint assay methodologies are found in the text.

41

Measurement of culture turbidity by optical density at 595 nm provides an easy and efficient means of determining the Histoplasma yeast number, which facilitates high-throughput assays and antifungal susceptibility profiling. Daily OD595 measurements provide a kinetic view of

Histoplasma yeast growth. OD595 readings are less subjective than visual inspection of wells, and, furthermore, spectrophotometric quantitation provides data for partially, but not fully, inhibited growth. Measurement of light scatter due to turbidity differs between spectrophotometers; thus, the creation of standard curves for OD595 versus turbidity using

McFarland standards should be performed for individual plate reader machines to calibrate readings. Starting yeast cell densities of 2 × 106 to 4 × 106 yeasts/mL provide an optimal balance between length of growth period in microtiter plates and dynamic range for growth measurement. An inoculum of 2 × 106 to 4 × 106 yeasts/mL yields initial plate readings (or readings of fully inhibited yeast growth) near zero, and the yeast typically reach saturation in 4 to

5 days (Fig. 2.4), although the growth period is extended to 6 days for some slower growing strains. This results in a dynamic range over 30-fold between measurements of the inoculum and saturated growth.

Metabolic activity indicator dyes can also be used to quantify Histoplasma yeasts grown in microtiter plates. We show that both MTT- and resazurin-based endpoint assays are applicable to the Histoplasma microdilution assay. With both assays, color can be used as a qualitative indicator of yeast cell number, which is less subjective than the visual determination of turbidity.

Spectrophotometric measurement of formazan formation or fluorescence due to the conversion of resazurin to resorufin also provides more objective, quantitative results. The accurate determination of relative growth requires that the measurements for maximum numbers of

42

Histoplasma yeasts do not saturate the assay color or fluorescence development. For MTT, we determined that 0.5 mg/mL MTT was not limiting and that a reaction time of 3 h at 37°C maximized the dynamic range of the assay. For resazurin, 100 µM and a time of 90 min at 37°C were the nonlimiting substrate concentration and optimal assay reaction time before saturation, respectively. The MTT and resazurin assays both are particularly useful for measuring the growth of Histoplasma strains which exhibit clumped rather than dispersed growth in a liquid medium (e.g., strains of the NAm1 class or Latin American isolates), since these assays are based on cellular metabolism rather than on yeast particles which, when aggregated, can lead to variable turbidity readings.

Testing of multiple clinical Histoplasma isolates from diverse geographic locations demonstrates that the optimized microdilution assay is broadly applicable to Histoplasma isolates. Antifungal susceptibility testing of the isolates showed similar susceptibility profiles for fungicidal amphotericin B. NAm1, LAm, and the Panama isolates had slightly increased natural resistance to fluconazole compared with the NAm2 isolates, but all were still below the fluconazole breakpoints defined for Candida yeasts (Clinical and Laboratory Standards Institute 2012;

Fothergill et al. 2014). Yeasts from the LAm and Panama, but not the NAm2, phylogenetic groups are characterized by cell walls that have α-glucan polysaccharides in addition to the β- glucan polysaccharide (Domer 1971; Reiss 1977; Reiss et al. 1977; Edwards, Alore, and

Rappleye 2011; Edwards and Rappleye 2011), and these isolates showed increased resistance to caspofungin compared to the resistance of NAm2 isolates. This correlation between α-glucan and less sensitivity to β-glucan synthase inhibitors is intriguing, yet the mechanism remains undefined at this point. Nonetheless, it indicates that Histoplasma strains should not be

43 considered a homogeneous group with regard to the natural profiles of antifungal susceptibility.

In addition, testing of isolates from patients that had failed or relapsed following fluconazole treatment confirms the acquisition of fluconazole resistance in these isolates and shows the applicability of the microdilution assay to provide accurate information on antifungal susceptibilities to improve clinical management of histoplasmosis.

In conclusion, the optimized parameters for measurement of Histoplasma yeast growth in 96- well microtiter plates, either through optical density measurement or through endpoint colorimetric/fluorescence assays, fill a gap in the CLSI methodology for antifungal susceptibility testing of Histoplasma yeasts. It is hoped that these standardized methods will provide for more reliable profiling of clinical isolates of Histoplasma as well as accelerate high-throughput screening of compounds for new antifungals active against the pathogenic phase of Histoplasma

(Edwards, Kemski, and Rappleye 2013).

44

Chapter 3. Development of a novel method for target identification of the novel antifungal 41F5

3.1 Introduction

41F5 was initially identified as a potent antifungal through the screening of a 3,600-compound library (Edwards, Kemski, and Rappleye 2013). The library contained compounds with structural similarity to purines or purine analog scaffolds and was screened using a high throughput method of monitoring intramacrophage growth, the native environment of Histoplasma throughout the course of infection, by fluorescence of RFP-expressing Histoplasma yeast (Edwards, Kemski, and Rappleye 2013). After initial high-throughput screening at 10µM, the IC50 of compounds that reduced yeast growth by at least 75% were verified and those that demonstrated a dose dependent inhibition and IC50 of less than 5µM underwent secondary host cell toxicity screening

(Edwards, Kemski, and Rappleye 2013). Host cell toxicity screening was performed using LacZ- transgenic P388D1 macrophages with β-galactosidase. Production of β-galactosidase was directly proportional to the number of surviving macrophages and therefore used as a measure of macrophage survival (Edwards, Kemski, and Rappleye 2013). 41F5 was selected as the most promising compound as it has an MIC of 2 µM, an IC50 of 0.87 µM and low host cell toxicity

(Edwards, Kemski, and Rappleye 2013). 41F5 was also tested against several other pathogenic fungi, and was found to be effective against C. neoformans with a MIC of 1.25 µM, but not effective against Blastomyces (MIC>40 µM), Histoplasma’s closest pathogenic relative

(Edwards, Kemski, and Rappleye 2013).

45

Once determined that a compound is a promising novel drug candidate, these compounds are modified to increase commercial and medical benefits. This includes modifying or making derivatives of the original compound in order to increase drug-like characteristics such as potency, solubility, stability, and bioavailability. One common way to determine how to modify drug candidates is Structure-activity relationship (SAR) studies. These are used to determine key structures and functional groups required for activity and how changes to other functional groups on the compound alter chemical properties of that compound. A SAR analysis was determined for 41F5 but was not able to significantly improve overall upon the original structure (Khalil et al. 2015). Thus another way to examine how to modify a novel candidate drug was required. One method used in the development of novel antibiotics is called Rational Drug Design (RDD). This method can be used in combination with screens to identify how candidate drugs interact with target molecules in order to mathematically model how modifications to the compound will change the interaction with the target. This allows for targeted modifications to the compound that increase desired drug-like characteristics without disrupting drug-target interactions.

However, this method requires a target molecule to be known.

41F5 is a prime candidate for RDD once the target is known. As antifungal targets are severally limited (Goughenour and Rappleye 2016), and 41F5 is not thought to act via known antifungal targets (Edwards, Kemski, and Rappleye 2013) identifying a novel target would be a large step forward for the field. RDD would then be able to be used to modify 41F5 to increase its drug- like characteristics and potentially design novel compounds against the same target. We have utilized a dual genetics and novel co-fractionation biochemical approach to identification of the

46 target of 41F5. We generated and whole-genome sequenced 6 41F5-resistant C. neoformans lines to identify single nucleotide polymorphisms (SNPs) in potential gene targets. While no consistent individual target was determined in all strains, some candidates were identified. In addition, a novel co-fractionation approach was developed to identify proteins that co-fractionate with 41F5 as potential targets. Current work is ongoing to validate these potential candidates by determination if increased 41F5 is required to inhibit fungi when the candidate target molecule is overexpressed (Titov and Liu 2012).

3.2. Materials and Methods

3.2.1. Yeast strains and growth

H. capsulatum strains used in this study are listed in the Table 1 and were derived from the

WU15 strain, a uracil auxotroph derivative of G217B. H. capsulatum yeasts were grown in H. capsulatum-macrophage medium (HMM) For growth of uracil auxotrophs, HMM was supplemented with 100 µg/ml uracil. Yeasts were grown with continuous shaking (200 rpm) at

37°C. For growth on solid medium, HMM was solidified with 0.6% agarose and supplemented with 25 µM FeSO4. C. neoformans strains used in this study are listed in the Table 1 and were derived from the H99 wild-type strain. strain S288C was used in this study. Yeasts were grown on YPD plates or RPMI liquid cultures. C. neoformans and S. cerevisiae were grown in liquid at with continuous shaking (200 rpm) at 30°C.

47

Table 3.1 Fungal Strains used in this study

Strain Genotype Other Designation1 Wild type Cryptococcus neoformans H99 serotype A Wild type Cryptococcus neoformans B3501 serotype D S288C Wild type Saccharomyces cerevisiae WT (ATCC204508) G217B Wild type NAm1 isolate WT (ATCC26032) WU15 G217B ura5-42Δ G217B ura5-42Δ zzz::pAM05 (URA5, OSU415 ACS2:6xHIS) ACS2 G217B ura5-42Δ zzz::pAM08 (URA5, OSU416 SAH1:6xHIS) SAH1 G217B ura5-42Δ zzz::pAM01 (URA5, OSU417 WRS1:6xHIS) WRS1 G217B ura5-42Δ zzz::pAM02 (URA5, OSU418 DED81:6xHIS) DED81 G217B ura5-42Δ zzz::pAM03 (URA5, OSU419 ERG10:6xHIS) ERG10 Generated through serial CNG1 41F5 resistant H99 isolate passaging in this study Generated through serial CNK2 41F5 resistant H99 isolate passaging in this study Generated through serial CNL3 41F5 resistant H99 isolate passaging in this study Generated through serile CN12 41F5 resistant H99 isolate passaging in this study Generated through serial CN15 41F5 resistant H99 isolate passaging in this study Generated through serial CN22 41F5 resistant H99 isolate passaging in this study Generated through serile CN Beta Fluconazole and 41F5 resistant H99 isolate passaging in this study 1gene designations: zzz::T-DNA: T-DNA integration at an undetermined chromosomal location URA5: orotate phosphoribosyltransferase ACS2: Acetyl-CoA synthetase SAH1: S-adenosyl-L-homocysteine hydrolase WRS1: Cytoplasmic tryptophanyl-tRNA synthetase DED81: Aspartyl/asparaginyl-tRNA synthetase ERG10: biosynthesis pathway, Acetyl-CoA C-acetyltransferase

48

3.2.2. Generation of Cryptococcus 41F5 resistant mutants

Independent lines of 41F5-resistant Cryptococcus yeasts were generated by serial passage of

Cryptococcus in increasing concentrations of 41F5. 1x105 cells/mL were grown in RPMI for 3 days at 30°C for each passage. As isolates became capable of proliferating in the starting concentrations of 41F5 (0.4 µM), the concentration of the drug was doubled. After 25 passages, isolates were challenged to determine a preliminary MIC for 41F5 (see 3.2.3). Passaged lines capable of growing in >16 µM 41F5 from passaging were selected for secondary testing. Single colonies were isolated and subsequently passaged without 41F5 selection to determine the stability of the resistance. Stably resistant lines were identified and re-tested to determine the

MIC to 41F5. Isolate lines with an MIC >64µM (>50-fold increase in resistance) were also tested for fluconazole resistance to identify those with specific resistance to 41F5 (see 3.2.3). Isolates with 41F5 resistance but fluconazole susceptibility were utilized for whole genome sequencing

(Table 3.3 see 3.2.4).

3.2.3. Antifungal susceptibility testing

In order to identify potential general antifungal resistance mechanisms like drug-efflux pumps, all 41F5 resistant mutants were screened for fluconazole reduced susceptibility to another antifungal (i.e., fluconazole). Fluconazole was prepared according to the CLSI reference standard M27-A3 (Rex and Clinical and Laboratory Standards Institute 2008). For IC50 and MIC determinations for Histoplasma yeasts, in a 96 well plate, a 2-fold dilution series of antifungal

49 drugs was added to Histoplasma yeasts at a density of 2-4x106 yeasts/mL and incubated at 37°C for 96-144 hours. Yeast growth was then assessed via turbidity at 595nm and/or endpoint resazurin reduction assays (Goughenour, Balada-Llasat, and Rappleye 2015). Briefly, 100 µM resazurin (7-hydroxy-10-oxidophenoxazin-10-ium-3-one) was added to each well, including control wells containing growth media without yeast cells, and the microtiter plates were incubated at 37°C. Resorufin (reduced resazurin; 7-hydroxyphenoxazin-3-one) production was quantified by measurement of fluorescence (530/25 excitation, 590/35 nm emission wavelengths) after 90 min using a Synergy2 plate reader (BioTek). Relative growth at each antifungal drug concentration was used to generate dose-response curves, and the IC50 value was determined by non-linear regression. MIC values were based on the minimum concentration of antifungal drug that prevented development of measurable turbidity or lack of resazurin fluorescence compared to growth in the absence of antifungal drug. For IC50 and MIC determinations for Cryptococcus yeasts the method was as above with the following modifications: yeast cells were inoculated at 1x104 cells/mL in RPMI and grown for 48hrs at

30°C, and resorufin was measured 3hrs after adding resazurin.

3.2.4. Whole genome sequencing of Cryptococcus mutants

Cryptococcus DNA from 41F5-specific resistant mutants was extracted as follows. 3mL cultures were grown overnight at 30°C, spun down, and the cell pellet frozen at -80°C overnight. The cell pellet was lyophilized then mechanical breakage with 0.5 µm diameter glass beads was preformed. CTAB buffer was added to samples and thoroughly mixed by vortexing, followed by incubation at 65°C for 30 minutes then cooling in an ice bath for 10min (Pitkin, Panaccione, and

50

Walton 1996). Phenol chloroform extraction was then preformed with the aqueous layer removed for isopropanol DNA precipitation. DNA was washed with 70% ethanol then resuspended in TE buffer. Samples were treated with RNase for 30 min at 37°C, then samples were cleaned using the NucleoSpin gDNA Clean-up kit (Macherey Nagel). Samples were then sent for whole genome sequencing. Multiplex sequencing using the Illumina HiSeq2500 platform was used to obtain genome coverage of 20-25x. SNPs were then identified for each isolate sequenced via comparison to the 41F5-sensitive parental line. SNPs were prioritized as those that result in single amino acid residue changes in coding regions of proteins (known proteins will be prioritized over hypothetical), followed by promoter SNPs.

3.2.5. Detection of 41F5 in cell lysate by HPLC-MS/MS

Histoplasma yeasts were grown to exponential phase (as described above) in 160 mL. 32µL of

25mM 41F5 was added to Histoplasma cultures for 8-12hrs to allow target proteins to bind the antifungal compound. Fungal cell lysates were prepared from the yeast cells by suspension in

PBS with 1X Protease Inhibitor (ingle Use Cocktail EDTA-Free; ThermoScientific) added and mechanical disruption with 0.5 um diameter glass beads. Lysates were kept on ice to prevent degradation. Debris was removed from the cellular lysate by centrifugation (10 minutes at

14,000 x g). Additional lysate preparations were as follows. To determine if native protein was required for 41F5 interaction Histoplasma lysates were heat denatured or degraded with proteinase K. Lysates were heat denatured by incubation at 95°C for 10 min. Lysates were degraded by incubation in 30mM TrisHCl (pH 8) with proteinase K treatment at 37°C overnight.

51

Additionally, Saccharomyces cerevisiae cells were grown with 41F5 and lysate prepared as previously described.

Separation of the lysates and 41F5 for subsequent target identification was performed by ultrafiltration filter co-fractionation followed by size-exclusion chromatography co-fractionation and detection of extracted 41F5 using LC-MS/MS by the procedure that follows (modified based on: Becher et al., 2017; Chan et al., 2012). Initial size fractionation of the Histoplasma capsulatum lysates was first performed using a series of ultrafiltration filter fractionations with molecular weight cut-off (MWCO) filters (VIVASPIN 2 PES Membrane Sartorius). The series of MWCO filters used were 300 kDa, 100kDa, 50kDa, 30kDa, then 10kDa. Subsequent fractionation was preformed by size exclusion chromatography (SEC) on an ÄKTA UPC-900 fast protein liquid chromatography (FPLC) system with P-920 pump and Frac-950 automated fraction collector (Amersham Biosciences) equipped with a Superdex 200 Increase 10/300 GL column (GE Lifesciences) where 40 fractions (500 µl each) were collected at 2 minute intervals in a mobile phase of 10mM phosphate buffered saline (PBS) at pH 7. The fractions were then chloroform extracted and the chloroform was evaporated using a CentriVap centrifugal concentrator (Labconco). Samples were resuspended in 60% acetonitrile:water (HPLC-grade;

Sigma Aldrich) and transferred to autosampler vials for LC-MS/MS detection of 41F5.

For detection of 41F5, Ultra performance liquid chromatography (UPLC) separation was perfomed with a 50 µl injection volume of the reconstituted fractions onto a BEC C18 column

(1.7µm 2.1 x100 mm; Waters) with VanGuard BEC C18 pre-column (1.7µm 2.1 x 5 mm;

Waters) thermostatted at 50 °C using a 5 min isocratic elution with a mobile phase consisting of

52

70% acetonitrile:water with 0.1% formic acid on a Vanquish UPLC equipped with autosampler and column compartment (Thermo). The UPLC was coupled to a TSQ Quantiva mass spectrometer (Thermo) through a heated electrospray ionization source (HESI) with a vaporizer and ion transfer tube temperature of 325 °C, sheath gas flow of 42 arbitrary units, auxiliary gas flow of 12 arbitrary units, sweep gas flow of 1 arbitrary unit, and spray voltage of 3.5 kV.

Detection was performed using multiple reaction monitoring (MRM) mode with a collision gas pressure of 1.5 mTorr following the transitions of 351 m/z to 241, 141, and 83 m/z using collision energies of 23.3 V, 43.6 V and 33.5 V respectively.

For proteomics work the aqueous phase from the chloroform extraction was split and both cold acetone precipitated and methanol precipitated to separate out proteins. The protein pellets were spun down (10 minutes at 14000 x g) and resuspended overnight in water. These were then run on a 10% SDS PAGE gel and slices were submitted for proteomic analysis. In addition, FPLC analysis on lysate was repeated and the same fractions that were directly identified as containing

41F5 were submitted for proteomics analysis without further modification to prevent complications from protein precipitation and re-solubilization as different methodologies can impact the protein composition seen in proteomic analysis and therefore bias results (Zellner et al. 2005).

3.2.6. Proteomics analysis

In gel digestion took place with sequencing grade-modified trypsin (Promega, Madison WI) added to SDS-PAGE gel slices for 6 hours at 37ºC. The peptides were then extracted from the

53 polyacrylamide with 50% acetonitrile and 5% formic acid. Non-gel samples were also trypsin digested overnight at 37 °C. Peptide concentration was determined by nanodrop (A280nm).

Orbitrap Fusion Capillary-liquid chromatography-nanospray tandem mass spectrometry

(Capillary-LC/MS/MS) of protein identification was performed on a Thermo Scientific orbitrap

Fusion mass spectrometer equipped with an EASY-Spray™ Sources operated in positive ion mode. Samples were separated on an easy spray nano column (PepmapTM RSLC, C18 3µ

100A, 75µm X150mm Thermo Scientific) using a 2D RSLC HPLC system from Thermo

Scientific. Each sample was injected into the µ-Precolumn Cartridge (Thermo Scientific) and desalted with 0.1% Formic Acid in water for 5 minutes. Mobile phase A was 0.1% Formic Acid in water and acetonitrile (with 0.1% formic acid) was used as mobile phase B. Flow rate was set at 300nL/min. MS/MS data was acquired with a spray voltage of 1.7 KV and a capillary temperature of 275 °C is used.

The scan sequence of the mass spectrometer was based on the preview mode data dependent

TopSpeed™ method: the analysis was programmed for a full scan recorded between m/z 400 –

1600 and a MS/MS scan to generate product ion spectra to determine amino acid sequence in consecutive scans starting from the most abundant peaks in the spectrum in the next 3 seconds.

To achieve high mass accuracy MS determination, the full scan was performed at FT mode and the resolution was set at 120,000. The AGC Target ion number for FT full scan was set at 2 x

105 ions, maximum ion injection time was set at 50 ms and micro scan number was set at 1.

MSn was performed using ion trap mode to ensure the highest signal intensity of MSn spectra using both CID (for 2+ and 3+ charges) and ETD (for 4+-6+ charges) methods. The AGC Target

54 ion number for ion trap MSn scan was set at 1000 ions, maximum ion injection time was set at

100 ms and micro scan number was set at 1. The CID fragmentation energy was set to 35%.

Dynamic exclusion is enabled with a repeat count of 1 within 60s and a low mass width and high mass width of 10ppm.

Sequence information from the MS/MS data was processed by converting the .raw files into a merged file (.mgf) using MSConvert (ProteoWizard) and were searched using Mascot Daemon by Matrix Science version 2.5.1 (Boston, MA) and the database searched against an in-house generated Histoplasma capsulatum genome database fragmented into 10 kb contigs which contained 5196 total database entries. This database was constructed from the draft assembly of the Histoplasma G186A genome (December 2004 assembly, Washington University Genome

Sequencing Center). The mass accuracy of the precursor ions were set to 10ppm, accidental pick of 1 13C peaks was also included into the search. The fragment mass tolerance was set to 0.5 Da.

Four missed cleavages for the enzyme were permitted. A decoy database was also searched to determine the false discovery rate (FDR) and peptides were filtered according to the FDR. The significance threshold was set at p<0.05 and bold red peptides is required for valid peptide identification. A decoy database was also searched to determine the false discovery rate (FDR) and peptides were filtered at a FDR of 1%. Proteins identified with at least 2 unique peptides were considered and reliable identification. Any modified peptides are manually checked for validation.

55

3.2.7. Generation of Histoplasma strains overexpressing His-tagged potential target proteins

Candidate target genes identified from the co-fractionation studies were amplifed by PCR (Table

3.2) from wild-type H. capsulatum G217B genomic DNA and cloned into the H. capsulatum expression vector pAG38, placing the gene under transcriptional control of the H2B constitutive promoter and fusing the potential target gene to a C-terminal hexahistidine tag. Overexpression plasmids were transformed into H. capsulatum WU15 yeasts by Agrobacterium-mediated transformation (Zemska and Rappleye 2012) and transformants selected by uracil prototrophy.

Transformants were screened by immunoblotting of culture filtrates and cellular lysates for the hexahistidine tag (GnScrpt A00186 to 6xhis).

Table 3.2 Primers used in this study

Primer Primer Sequence 5' to 3' Direction1 ACS2-1 ATGTCTGACGGTCCAGTTACGC Forward ACS2-2 CTTCTTTGAAGCGTGGTAAACT Reverse SAH1-1 ATGTCAACTCACCAAACATTAACC Forward SAH1-2 CATTTCAGCCTTGAAAGGGCCA Reverse DED81-1 ATGGCCGCCACCACTTCAAT Forward DED81-2 TGGTGTACACCGTCCAGTAAAG Reverse WRS1-1 ATGACTACCCCGCCTCCTGG Forward WRS1-2 ATTTTCATCATGACGTTGTTTCG Reverse ERG10-1 ATGCAGTGCTCTCTCAGAACAG Forward ERG10-2 GTCGACCTTTTCCACTCTCTG Reverse 1Direction relative to gene transcription

56

3.3 Results

3.3.1 Identification of SNPs from Cryptococcus 41F5 resistant mutants

Serial passaging of Cryptococcus yeasts was used to generate 41F5-resistant mutants as

Histoplasma yeasts quickly start converting from the yeast phase to hyphae even in sub- inhibitory concentrations of 41F5. The conversion of yeast to mycelia alters the antifungal susceptibility of Histoplasma, preventing it from being a good candidate for serial passaging

(Goughenour, Balada-Llasat, and Rappleye 2015). Using this approach 6 independent lines were generated that were resistant to 41F5 but susceptible to fluconazole (Table 3.3). Fluconazole was chosen to monitor for potential non-specific drug efflux pump-mediated resistance mechanisms because they are a well-described mechanism of azole resistance in Candida (Kanafani and

Perfect 2008). Following isolation of a clonal line, the genomes of these lines were sequenced and single nucleotide polymorphisms (SNPs) were identified. Of the SNPs, particular attention was focused on missense, but not non-sense, mutations in the coding sequence (CDS) (as inhibition of fungal growth by 41F5 suggests the target is essential) and SNPs in regions upstream of genes, which might result in increased gene expression (e.g., increased drug-target production). 12 high priority candidate SNPs were identified (Table 3.4) with 2 of immediate interest (Erg20 in line L3 and a serine/threonine kinase in line K2).

57

Table 3.3 Antifungal susceptibility of Cryptococcus lines to 41F5 and Fluconazole

41F5 Fluc Fluc Fluc Strain (MIC) (MIC) (IC50) (logIC50) WT 2 2.5 1.2 0.1 CNG1 ≥64 5 3.2 0.5 CNK2 ≥64 5 2.6 0.4 CNL3 ≥64 10 4.4 0.6 CN12 ≥64 5 3.0 0.5 CN15 ≥64 2.5 1.7 0.2 CN22 ≥64 5 1.3 0.1 CN Beta ≥64 20 17.0 1.2

41F5 results are reported as Minimum Inhibitory Concentrations (MICs) as resistant lines were not inhibited at any concentration tested and therefore inhibitory concentrations of 50% (IC50) were not able to be calculated. Fluconazole results are reported as MICs and IC50. IC50s are calculated by linear regression from dose response curves. logIC50 is used to measure asymmetrical confidence intervals. Each strain was tested in biological triplicate.

58

Table 3.4 SNPs Identified in 41F5-resistant Cryptococcus most likely to contribute to resistance

Line Locus Encoded Function Mutation CNAG_07784 Promoter CNG1 (-23) hypothetical mutation CNAG_07784 Promoter CNG1 (-30) hypothetical mutation 4 amino acid CNK2 CNAG_03376 hypothetical insertion CNAG_08026 Promoter CNK2 (-404) hypothetical mutation CNK2 CNAG_01905 Ser/Thr kinase K>T CNL3 CNAG_02084 Erg20: farnesyl diphosphate synthase A154V homolog of RPN7 (proteasome CN12 CNAG_06899 regulatory subunit) Y>D hypothetical cytoplasmic protein (non- CN12 CNAG_02079 essential gene) H>Y CN12 CNAG_01699 Histone deacetylase 1/2 V53A 8 amino acid CN15 CNAG_05333 Hypothetical protein insertion Frameshift CN22 CNAG_07938 Hypothetical protein deletion

59

3.3.2 Co-fractionation approach

As a complementary approach to genetics, candidate drug targets were identified through a biochemical approach for proteins that could interact with 41F5. Initial attempts to generate a bioactive, tagged 41F5 compound that could be used for column purification were unsuccessful as the molecule could not be modified at any substituent positions without losing activity (Ishita et al., 2018). We therefore developed an approach to identify cellular proteins that co-fractionate with 41F5 as potential drug targets and UPLC-MS/MS (Ultra-high performance liquid chromatography-mass spectrometry/mass spectrometry) strategy for target identification (Chan et al. 2012; Becher et al. 2017). As 41F5 and target-interacting 41F5 should size fractionate differently (Fig. 3.1) we should be able to use size fractionation to identify which proteins co- fractionate with 41F5. To track 41F5 in cellular fractions, we exploited a mass-spectrometry based procedure to enable high sensitivity and specificity in detection. Briefly, to identify 41F5 interacting proteins, fungi are treated with 41F5 to allow target proteins to bind the antifungal compound and then cellular lysates are prepared. Lysates were fractionated by size (molecular weight cutoff filters then fast protein liquid chromatography (FPLC)) then cholorform extraction of 41F5 was preformed. UPLC-MS/MS was preformed on the chloroform fractions to identify

41F5-positive fractions are identified. The corresponding aqueous fractions along with non- chloroform extracted fractions identified as 41F5 positive are used for proteomic analysis by mass spectrometry to identify proteins that co-fractionate with 41F5. The aqueous fraction from the chloroform extraction provides a matched sample to the 41F5-detected fraction, which is the most accurate way to show 41F5-interacting proteins. However, the level of manipulation of the sample can lead to protein loss and subsequent gaps in the proteomics analysis. Therefore

60 proteomics were also preformed on fractions without further processing to ensure no proteins are lost. These two resulting lists are combined into a single proteomics list for candidate targets.

61

Figure 3.1 Schematic for the Co-fractionation approach to 41F5 detection.

Free 41F5 and target bound 41F5 will size fractionate differently. Some of every fraction is held for proteomic analysis. The rest of the fraction undergoes chloroform extraction of 41F5. The aqueous layer contains precipitated proteins and is submitted for proteomic analysis as well. The chloroform layer is dried and resuspended for HPLC-MS/MS detection of 41F5 in each fraction. Fractions identified as 41F5 positive are used for proteomic analysis.

62

3.3.3 Validation and Optimization of the Co-fractionation approach

To track the 41F5 compound, the UPLC retention time and MS/MS pattern for 41F5 was used.

Using only 41F5, a characteristic MS/MS pattern was then determined with a primary 351m/z

MS1 peak fracturing into three diagnostic MS2 peaks at 241m/z, 141m/z and 83m/z (Fig 3.2 A).

By UPLC (C18 column), the retention time of 41F5 was determined to be 1.90 min. This retention time and MS/MS peak profile was used to identify the 41F5 molecule in lysate fractions.

For initial size fractionation, we used ultrafiltration of lysates through different molecular weight cutoff filters (MWCO). This process separated the lysate into broad size ranges of >300kDa,

300-100kDa, 100-50kDa, 50-30kDa, 30-10kDa, and <10kDa. In the Histoplasma lysate from yeasts incubated without 41F5, we do not detect any 41F5 (Table 3.5). Passage of the 41F5 drug through the size filters in the absence of cells revealed it only in the high molecular weight sample (>300kDa range; Table 3.5). We hypothesize this to be the result of the largely hydrophoblic compound forming large aggregates that cannot pass through the 300 kDa MWCO filter. Lysates prepared from Histoplasma yeasts incubated with 41F5 overnight had 41F5 present in the >300 kDa fraction, but also some 41F5 in the 300-100kDa fraction (Table 3.5).

We interpret this as the 41F5 in the >300 kDa fraction representing free drug that did not bind any target within the yeast cell while the 41F5 present in the 100-300 kDa size was bound to a cellular component that fractionates in that size, bring the drug with it. The molecule to which the 41F5 molecule binds is likely a protein or protein complex between 100-300 kDa as the 41F5 molecule is not carried into the 300-100kDa range if the lysate is treated with heat or Proteinase

63

K before running through the MWCO filters (Table 3.5). The interation is specific to a

Histoplasma protein as incubation of Sachromyces cerevisiae (an organismnot inhibited by

41F5) with the drug and fractionation of cellular lysates prepared from these yeasts did not result in 41F5 being found in the 100-300 kDa fraction.

64

Figure 3.2 Validation and optimization of 41F5 detection by UPLC-MS2.

Relative diagnostic peaks at 241m/z, 141m/z and 83m/z of 41F5 generated from further fractionation of the initial 351m/z compound (A). UPLC-MS2 detection of 41F5 showing a retention time from the UPLC (x-axsis) at 1.90 minutes and the relative detection of 241m/z, 141m/z and 83m/z size molecules (B). Acetonitrile only negative control (left) showing background noise detection and 41F5 (right) detection is a clean strong peak. 65

Table 3.5 Presence of 41F5 in different size fractions

300- Sample <100kDa 100kDa >300kDa

Drug only - - ++

Histoplasma lysate only - - -

Drug + Histoplasma lysate - + ++

Drug + Heat-denatured Histoplasma lysate - - ++

Drug + Proteinase K-Treated Histoplasma lysate - - ++

Drug + S. cerevisiae lysate - - ++ ++ indicates a large peak of 41F5 detected; + indicates a peak detected; - indicates no peak detected

3.3.4 Identification of 41F5 Co-fractionating Proteins

The cellular lysate fraction from the 300-100kDa fraction was further size fractionated by FPLC to reduce the number of potential 41F5-interacting proteins. Each of the resulting fractions was tested for the presence of 41F5 (Fig. 3.3A). 41F5 was seen in the fractions 5-11 as part of the solvent front where molecules that do not separate in the range of the column (molecular weights

(Mr ) from 10,000 to 600,000) can be seen. This is likely free drug that has dissociated from the target over the course of the experiment forming aggregates too large for the column to separate.

66

In the fractions within separation range of the column, one clear 41F5 peak is seen in fraction 15

(Fig. 3.3A and 3.3B). Fraction 15, along with a non 41F5-containg control fraction (fraction 26) to rule out proteins that do not interact or cleanly fractionate, were sent for proteomic analysis by

LC-MS/MS. 194 proteins were identified in the 41F5-containing fraction that had a Quantitative

Value (Normalized Total Spectra) of greater than 10 (Supplemental File 1). The co- fractionating proteins were further prioritized as potential 41F5-interacting proteins as those that would be predicted to be required for viability (Table 3.6). This criterion was used as proteins in very small amounts were unlikely to be able to bind enough drug for detection. In addition as the drug is inhibitory the target is likely to be essential. All potential target genes were annotated by

BLAST analysis to identify homologs in three common fungal organisms S. cerevisiae, C. albicans and A. nidulans and the protein’s predicted essentially determined from analysis in these model genetic organisms.

67

Figure 3.3 Detection of 41F5 in FPLC fractions.

(A) Detection of the characteristic 241 m/z (blue) and 141 m/z (red) MS2 peaks in each collected fraction from UPLC analysis. (B) Protein concentration as determined by absorbance (280nm) in each fraction, with the 41F5 containing peak (15) and negative control peak (26) selected for proteomics indicated by red arrows.

68

Table 3.6 Proteins identified as co-fractionating with 41F5

Histoplasma Gene Homolog ID Annotation Histo217_09243 ACS2 Acetyl-CoA synthetase Histo217_09046 SAH1 S-adenosyl-L-homocysteine hydrolase ATP domain Histo217_03917 WRS1 Cytoplasmic tryptophanyl-tRNA synthetase Histo217_00804 DED81 Aspartyl/asparaginyl-tRNA synthetase Histo217_05335 ERG10 Ergosterol biosynthesis pathway Histo217_00586 GUS1 Glutamyl-tRNA synthetase Histo217_07910 CDC60 Leucyl-tRNA synthetase Histo217_08619 SES1 Seryl-tRNA synthetase Histo217_02253 ALA1 Alanine-tRNA ligase Histo217_03699 KRS1 Lysyl-tRNA synthetase Histo217_07818 THS1 Threonyl-tRNA synthetase Histo217_06011 YNL247W Cysteinyl-tRNA synthetase Histo217_03385 VAS1 Valyl-tRNA synthetase Histo217_07051 ILS1 Isoleucyl-tRNA synthetase Histo217_05386 PYK1 Pyruvate kinase Histo217_08561 GRS1 Glycyl-tRNA synthetase Histo217_05806 PKC1 Protein kinase C Histo217_07968 RSE1 MMS1_N and CPSF_A domain-containing Vesicle coat complex COPII, subunit SEC23 Histo217_02917 SEC23 [Intracellular trafficking and secretion] Histo217_05835 RPN1 26S proteasome regulatory complex component Histo217_01025 GFA1 Glucosamine:fructose-6-phosphate aminotransferase Protein similar to ribosomal protein Histo217_01193 RPL30 L7Ae/L30e/S12e/Gadd45 family protein Histo217_02798 BMH1/2 14-3-3 domain-containing Histo217_04092 ADE13 Adenylosuccinate lyase Histo217_06798 ALD Betaine aldehyde dehydrogenase Histo217_08045 KRE5 UDP-glucose:glycoprotein glucosyltransferase Protein containing domains CDC37_N, CDC37_M, Histo217_02640 CDC37 and CDC37_C Protein containing domains Ube1_repeat1, E1_4HB, Histo217_00835 UBA1 Ube1_repeat2, and GST_C_Omega_like Histo217_04309 RFA1 Replication factor-a protein 1 Histo217_07128 BMH1/2 14-3-3 domain-containing ATPase; HSPA5-like_NBD domain-containing Histo217_03829 KAR2 protein Histo217_04542 RPS9B/A 40S ribosomal protein S9 Histo217_07720 RPS31 Ubiquitin-like (Ubl) domain; Ribosomal protein S27a

69

3.3.5 Identification of the target of 41F5

To determine the target of 41F5 from the candidate proteins from Histoplasma and

Cryptococcus, we will use correlation of decreased susceptibility to the drug upon overexpression of the target protein; overexpression of the target should increase the amount of drug needed to inhibit all of the target protein which will increase the observed MIC (Titov and

Liu 2012). For the Cryptococcus resistant strains we will use Agrobacterium mediated transformation of C. neoformans (B3501) (Fu et al. 2011; Walton, Idnurm, and Heitman 2005).

This work is on-going with plasmid optimization to ensure high levels of expression of the target gene via promoter selection taking place. In Histoplasma Agrobacterium-mediated transformation is being used to introduce transgenes in which the candidate 41F5-interacting proteins are fused to a C-terminal hexahistidine-tag (6xHis) and expression controlled by the

H2B promoter, which drives high level, constitutive expression. The inclusion of the C-terminal tag facilitates confirmation of transgene overexpression by immunoblotting of lysates and affinity purification (Garfoot et al. 2016) of the protein for future biochemical characterization.

Preliminary susceptibility testing of 5 candidate proteins (ACS2, SAH1, WRS1, DED81, and

ERG10) has been preformed using two independent transformants for each gene (Table 3.7).

None show an increased IC50. Work is continuing on the others in the list (Table 3.6). Once a target is identified by an increase in resistance, final validation can occur. For final validation, candidate proteins will be purified from cellular lysates using metal-affinity chromatography (via the 6xHis tag) and testing for direct interaction (via the mass-spectrometry co-fractionation with purified protein) and inhibition of any predicted biochemical function by 41F5. 41F5 derivative

70 compounds from the SAR study that have lost antifungal activity will be used as negative control.

Table 3.7 IC50 of Candidate 41F5 Target Overexpressions

IC50 logIC50 Strain (µM) WT 0.9 -0.02 ACS2 0.7 -0.1 SAH1 1.0 -0.00002 WRS1 0.9 -0.03 DED81 0.8 -0.1 ERG10 0.6 -0.2

Results are calculated by linear regression from dose response curves of two independent transformations of each overexpression strain. logIC50 is used to measure asymmetrical confidence intervals.

3.4 Discussion

Antifungal options for treating fungal disease remain limited, particularly for Histoplasma and the thermal dimorphs (Goughenour and Rappleye 2016; Rauseo et al. n.d.). As , fungi have very similar proteins to those of the mammalian host, thereby limiting fungal-specific targets. In this study we combined a standard genetic approach with a co-fractionation approach to attempt to identify the molecular target of 41F5. We have developed a highly sensitive and specific assay to monitor the presence of 41F5 in biological samples that enabled the biochemical fractionation approach. One advantage of his is we were able to use the 41F5 drug

71 without having to modify the compound in any way, which could alter its activity. A second advantage it holds over radiolabeling the drug to track it is that the approach does not have to be modified for each drug (i.e., by labeling individual atoms during synthesis).

Our study identified several potential target proteins through our genetic approach.

Unfortunately, due to our small sample size, no consistent SNP pattern was found among multiple resistant lines. We initially investigated SNPs as the method of mutation as amino acid changes in the target protein are a direct way to determine if the protein interacts with a drug.

Our candidate list is limited by the identification of several hypothetical proteins, which makes speculation on their essentiality difficult. Of the few point mutations we identified, the ergosterol biosynthesis protein (Erg20) and the serine/threonine kinase are of greatest interest. Ergosterol biosynthesis is the target of the azole class of antifungals, although the azoles target the Erg11 gene (Cowen et al. 2015). Nonetheless, as ergosterol is essential for fungal viability and relatively specific, inhibition of the ergosterol pathway is an attractive target. The serine/threonine kinase is a attractive candidate as the original screen for 41F5 was comprised of purines or purine analog scaffolds therefore it would be reasonable that such as kinases that bind ATP would be potential targets. However, this specific serine/threonine kinase is not predicted to be essential in Cryptococcus questioning whether it is a direct target of 41F5 (Lee et al. 2016). This highlights a problem with the genetic approach to target identification. Sometimes the mutations that generate resistance are not in the target of the drug itself. For example, in

Candida albicans mutations in the transcription factor that regulates Erg11, UPC2, or mutation in the stress response pathway can also lead to azole resistance (Cowen et al. 2015). Therefore, analysis of the results from whole genome sequencing is often difficult to interpret. In addition,

72 chromosomal abnormalities such as copy number can lead to antifungal resistance (Cowen et al.

2015). These are not detected in SNP calling and add another layer of analysis that can be preformed. Currently we are still working on testing the current potential targets from the SNP analysis; however a chromosomal analysis of the mutants may still be preformed on the data from the whole genome sequencing in the future.

This study also developed a co-fractionation method for identifying drug targets through their potential biochemical interaction with the 41F5. The method showed that 41F5 interacts with a protein target in the 100-300kDa size range. Further fractionation and prioritization identified 34 likely protein target candidates. The candidate list contains several interesting proteins. Acetyl coenzyme A (acetyl-CoA) synthetase (ACS2) is the proposed target of the antifungal AR-12

(Koselny et al. 2016) and was therefore a high priority candidate to test in our overexpression system. However, overexpression of ACS2 did not confer any resistance to 41F5. Preliminary overexpression tests of Erg10 however, did not confer any 41F5 resistance. The other clear pattern seen in this proteomic dataset is the high numbers of tRNA synthetases. Although not currently seen in fungi as drug targets, tRNA synthetases inhibitors are promising antiparasitic drugs (Pham et al. 2013). This holds promise that a similar inhibitor could work in fungi as they are both eukaryotic pathogens. While overexpression of the first two tested tRNA synthetases genes WRS1 and DED81 (tryptophan and aspartate respectively) did not show any resistance, there are 10 more that are still being tested. No other large-scale patterns were seen in the proteomics data.

73

Very little correlation was seen between the genetics and co-fractionation approaches. No target gene was identified in both sets and the only pathway overlap was in the ergosterol biosynthesis pathway with Erg20 identified in the Cryptococcus SNP analysis of 41F5-resistant mutants and

Erg10 co-fractionating with 41F5 in Histoplasma lysates. However, it is difficult to predict that

41F5 inhibits two separate enzymes within the same pathway. As the Erg10 gene in

Histoplasma does not confer resistance, this is not likely the target in both organisms. Therefore, there are two possible reasons why we do not see overlap in the data sets: the two organisms do not have the same target or the genetics approach only identified indirect resistant mutations in

Cryptococcus. As to the first reason Histoplasma and Cryptococcus are members of two different phyla of fungi ( and respectively) and therefore have a huge amount of evolutionary divergence between them. To address this concern, the co-fractionation approach could be used to generate a proteomics list for Cryptococcus 41F5 co-fractionating proteins.

Initial attempts at this, however, showed that 41F5 in Cryptococcus lysates does not fractionate in the same size range as it does in Histoplamsa lysates. Alternatively, the genetics approach may have identified indirect factors in Cryptococcus that confer resistance not through changes in the target itself but through other interacting proteins or pathways. Therefore overexpressing the Cryptococcus homolog of the Histoplasma target could lead to confirmation in

Cryptococcus. This highlights why the development of this new method is such a benefit to the field of drug discovery. This method is based on direct interaction of target and drug avoiding off-target resistance mechanisms. It also can be adapted for any organism, allowing for use in target identification in any kind of drug development.

74

Chapter 4. Investigation into the O-linked mannosylated proteome of Histoplasma capsulatum

4.1 Introduction

O-linked glycosylation is one of the two major types of glycosylation found in fungi. It commonly consists of a linear series of mannose residues attached at serines and threonines and is performed by a relatively simple pathway (Loibl and Strahl 2013). The initial mannose attachment is performed by protein mannosyltransferases (PMTs) with subsequent mannose additions by the MNT family of proteins, both of which have been characterized in C. albicans

(Loibl and Strahl 2013; Munro et al. 2005). Mutants of Histoplasma in three PMT proteins

(Pmt1, Pmt2, and Pmt4) and the only identified Histoplasma MNT (Mnt1) were previously generated and examined for their role in virulence (Garfoot et al. 2018). Confirmation of the lack of O-linked mannosylation in the mutants was performed by observing molecular weight shifts in Cfp4, a protein with a mucin-like domain that is predicted to be heavily o-linked mannosylated. The Cfp4 protein of the pmt2 mutant has a reduction in mannose content

(Garfoot et al. 2018), but Sod3, a protein without predicted O-linked mannosylation, is the same molecular weight in mutants and wild type strains (Garfoot et al. 2018). In addition, mass spectrometry was used to directly show that pmt2 mutants had reduced mannosylation on purified Cfp4 proteins (Garfoot et al. 2018). In vivo mouse assays show virulence defects in the pmt1, pmt2, and mnt1 mutants (Garfoot et al. 2018). These results are consistent with reports that in C. albicans the two MNT proteins are necessary for adhesion and virulence (Munro et al.

2005). To explain the virulence defect seen in mice, our lab has worked on characterizing the 75 pmt2 mutant extensively, and have discovered that the mutant is temperature sensitive, which leads to the loss of virulence (Garfoot et al. 2018). This work attempts to expand upon our understanding of O-linked mannosylation in Histoplasma in order to identify mannosylated proteins and determine which proteins play a role in temperature sensitivity and therefore virulence. To do this we performed a preliminary bioinformatics search for O-linked mannosylated proteins, and identified some proteins with mucin-like domains (for example,

Cfp4); however, there is no true consensus sequence for O-linked mannosylation, which significantly hindered this approach. In addition, we are working on developing a novel approach to purification of O-linked mannosylated proteins via biotinylation and then identification by mass spectrometry. Initial attempts and differentiation of O-linked mannosylated proteins by mass spectrometry alone were unsuccessful.

4.2 Materials and Methods

4.2.1 Histoplasma growth

Histoplasma strains used in this study were derived from wild-type strain G217B (ATCC

26032) (Table 4.1) and were grown on HMM (Worsham and Goldman 1988). The uracil auxotrophs of Histoplasma were grown on/in HMM supplemented with 100 ug/mL uracil. For maintenance, cultures were grown at 37 °C under 5% CO2/95% air with shaking at 200 rpm for liquid cultures. Solid media was prepared by addition of 0.6% agarose and 25 µM FeSO4.

76

Table 4.1 Histoplasma strains used in this study Strain Genotype Other Designation1 G217B Wild type NAm1 isolate WT (ATCC26032) WU15 G217B ura5-42Δ OSU194 G217B ura5-42Δ zzz::pAG21 (G418R,GFP) RNAi sentinel background G217B ura5-42Δ zzz::pAG21 (G418R,GFP) OSU379 zzz:pED02 (URA5, gfp-RNAi) GFP-RNAi G217B ura5-42Δ zzz::pAG21 (G418R,GFP) OSU380 zzz:pKG22 (URA5, gfp:CRH1-RNAi) CRH1-RNAi G217B ura5-42Δ zzz::pAG21 (G418R,GFP) OSU381 zzz:pKG16 (URA5, gfp:ENG3-RNAi) ENG3-RNAi G217B ura5-42Δ zzz::pAG21 (G418R,GFP) OSU382 zzz:pKG17 (URA5, gfp:PDI-RNAi) PDI-RNAi G217B ura5-42Δ zzz::pAG21 (G418R,GFP) OSU383 zzz:pED02 (URA5, gfp-RNAi) GFP-RNAi G217B ura5-42Δ zzz::pAG21 (G418R,GFP) OSU384 zzz:pKG23 (URA5, gfp:BGL5-RNAi) BGL5-RNAi G217B ura5-42Δ zzz::pAG21 (G418R,GFP) OSU385 zzz:pKG24 (URA5, gfp:CRR1-RNAi) CRR1-RNAi G217B ura5-42Δ zzz::pAG21 (G418R,GFP) OSU386 zzz:pED02 (URA5, gfp-RNAi) GFP-RNAi G217B ura5-42Δ zzz::pAG21 (G418R,GFP) OSU387 zzz:pKG18 (URA5, gfp:GH-RNAi) GH-RNAi G217B ura5-42Δ zzz::pAG21 (G418R,GFP) OSU388 zzz:pKG19 (URA5, gfp:CTS4-RNAi) CTS4-RNA1 1gene designations: zzz::T-DNA: T-DNA integration at an undetermined chromosomal location URA5: orotate phosphoribosyltransferase apt3: aminoglycoside phosphotransferase (G418 resistance) GFP: green-fluorescence protein CRH1: GH16 fungal CRH1 transglycosylase; predicted role in cell wall architecture, required for the transfer of chitin to 1,6-beta-glucan in the cell wall in S. cerevisiae ENG3: GH16 fungal Lam16A glucanase PDI: Protein Disulfide Isomerase (PDIa) family, YbbN superfamily neg reg of GroEL BGL5: Beta-glucosidase 5, Glycosyl hydrolase family 1 domain CRR1: GH16 fungal CRR1 transglycosylase; specific glycosidase involved in spore wall assembly during sporulation and potentially involved in copper import in S. cerevisiae. GH: Glycosyl hydrolases family 15 (Glucoamylase (glucan1,4-alpha-glucosidase)), C-terminal CBM20 (carbohydrate-binding module, family 20) domain CTS4: glycosyle hydrolase 18 domain; chitinase

77

4.2.2. Proteomics analysis

Supernatants of saturated Histoplasma pmt2 (RNAi) knockdown, mnt1 (RNAi) knockdown and wild type Histoplasma strains in biological triplicate were collected and concentrated 100X.

Samples of the concentrated supernatants were treated overnight with PNGase F to remove N- linked glycosylation (Holbrook et al. 2014). Proteins were separated by electrophoresis 1(0%

SDS-PAGE gel) and gel slices were submitted for trypsin digestion and proteomic analysis.

In gel digestion took place with sequencing grade-modified trypsin (Promega, Madison WI) added to SDS-PAGE gel slices for 6 hours at 37ºC. The peptides were then extracted from the polyacrylamide with 50% acetonitrile and 5% formic acid. Non-gel samples were also trypsin digested overnight at 37 °C. Peptide concentration was determined by nanodrop (A280nm).

Orbitrap Fusion Capillary-liquid chromatography-nanospray tandem mass spectrometry

(Capillary-LC/MS/MS) of protein identification was performed on a Thermo Scientific orbitrap

Fusion mass spectrometer equipped with an EASY-Spray™ Sources operated in positive ion mode. Samples were separated on an easy spray nano column (PepmapTM RSLC, C18 3µ

100A, 75µm X150mm Thermo Scientific) using a 2D RSLC HPLC system from Thermo

Scientific. Each sample was injected into the µ-Precolumn Cartridge (Thermo Scientific) and desalted with 0.1% Formic Acid in water for 5 minutes. Mobile phase A was 0.1% Formic Acid in water and acetonitrile (with 0.1% formic acid) was used as mobile phase B. Flow rate was set at 300nL/min. MS/MS data was acquired with a spray voltage of 1.7 KV and a capillary temperature of 275 °C is used.

78

The scan sequence of the mass spectrometer was based on the preview mode data dependent

TopSpeed™ method: the analysis was programmed for a full scan recorded between m/z 400 –

1600 and a MS/MS scan to generate product ion spectra to determine amino acid sequence in consecutive scans starting from the most abundant peaks in the spectrum in the next 3 seconds.

To achieve high mass accuracy MS determination, the full scan was performed at FT mode and the resolution was set at 120,000. The AGC Target ion number for FT full scan was set at 2 x

105 ions, maximum ion injection time was set at 50 ms and micro scan number was set at 1.

MSn was performed using ion trap mode to ensure the highest signal intensity of MSn spectra using both CID (for 2+ and 3+ charges) and ETD (for 4+-6+ charges) methods. The AGC Target ion number for ion trap MSn scan was set at 1000 ions, maximum ion injection time was set at

100 ms and micro scan number was set at 1. The CID fragmentation energy was set to 35%.

Dynamic exclusion is enabled with a repeat count of 1 within 60s and a low mass width and high mass width of 10ppm.

Sequence information from the MS/MS data was processed by converting the .raw files into a merged file (.mgf) using MSConvert (ProteoWizard) and were searched using Mascot Daemon by Matrix Science version 2.5.1 (Boston, MA) and the database searched against an in-house generated Histoplasma capsulatum genome database fragmented into 10 kb contigs which contained 5196 total database entries. This database was constructed from the draft assembly of the Histoplasma G186A genome (December 2004 assembly, Washington University Genome

Sequencing Center). The mass accuracy of the precursor ions were set to 10ppm, accidental pick of 1 13C peaks was also included into the search. The fragment mass tolerance was set to 0.5 Da.

79

Four missed cleavages for the enzyme were permitted. A decoy database was also searched to determine the false discovery rate (FDR) and peptides were filtered according to the FDR. The significance threshold was set at p<0.05 and bold red peptides is required for valid peptide identification. A decoy database was also searched to determine the false discovery rate (FDR) and peptides were filtered at a FDR of 1%. Proteins identified with at least 2 unique peptides were considered and reliable identification. Any modified peptides are manually checked for validation.

4.2.3. Bioinformatics

Histoplasma predicted proteome was screened using the NetOGlyc (v. 4.0) for identification of regions likely to be O-linked glycosylated. Proteins identified as likely to be glycosylated (at least 1 site with at least >0.5 probability) were then screened for the presence of a secretion signal using the SignalP (v. 5.0) software (Eukaryotes with D scores >0.45 as positive)

(Almagro Armenteros et al. 2019). These proteins were then screened for domain annotations using Conserved Domain (CD) search from NCBI (Marchler-Bauer et al. 2017; Marchler-Bauer et al. 2015; Marchler-Bauer et al. 2011; Marchler-Bauer and Bryant 2004). Some gene annotation was preformed by identification from NCBI (blastp) of homologs in the model organisms Saccharomyces cerevisiae, sp., and .

4.2.4 Generation of RNAi Knockdowns

Histoplasma yeast cells were transformed by Agrobacterium-mediated transformation (Zemska and Rappleye 2012). The RNAi vectors used were created in the gfp-sentinel vector (pED02)

80

(Garfoot et al. 2016) by using 500 to 1,000 bp of the coding regions (Table 4.2). Vectors were transformed into gfp-expressing sentinel strain OSU194 and sentinel green fluorescent protein

(GFP) fluorescence was quantified with a modified gel documentation system (Youseff and

Rappleye 2012) and ImageJ software (Schneider, Rasband, and Eliceiri 2012).

81

Table 4.2 Primers used in this Study

1 Primer Primer Sequence 5' to 3' Direction Target Gene 03868-1 CTCTCTGGTCTTGTCACAAACG Forward CRH1 03868-2 CTTTCCGTGCCTTTCGCTTG Reverse CRH1 03868-3 CTCTCTGGTCTTGTCACAAACG Forward CRH1 03868-4 CTTTCCGTGCCTTTCGCTTG Reverse CRH1 03793-1 TCCAGAAGCTCCGCCTCAAA Forward CTS4 03793-2 CCCGGCCTTCAGGTTGTC Reverse CTS4 03793-3 TCCAGAAGCTCCGCCTCAAA Forward CTS4 03793-4 CCCGGCCTTCAGGTTGTC Reverse CTS4 03564-1 TACAATCCGACGACCTGGACGAG Forward CRR1 03564-2 GTGGACGGCGGTGTGGCTCT Reverse CRR1 03564-3 TACAATCCGACGACCTGGACGAG Forward CRR1 03564-4 GTGGACGGCGGTGTGGCTCT Reverse CRR1 Eng3-1 ATCATCAATACCGCCTTTTGTGG Forward ENG3

Eng3-2 CAATGATGTGTTCCAGCAATGTG Reverse ENG3

Eng3-3 ATCATCAATACCGCCTTTTGTGG Forward ENG3 Eng3-4 CAATGATGTGTTCCAGCAATGTG Reverse ENG3

Bgl5-1 AGAAAACCCCAGCGGAGCAA Forward BGL5 Bgl5-2 GGGTTTTTCCAGAGATCGTTCG Reverse BGL5 Bgl5-3 AGAAAACCCCAGCGGAGCAA Forward BGL5 Bgl5-4 GGGTTTTTCCAGAGATCGTTCG Reverse BGL5 09219-1 TTTGGGAAGAGGTTGACGGCA Forward GH 09219-2 CGCAACAGCCAAATATCCATC Reverse GH 09219-3 TTTGGGAAGAGGTTGACGGCA Forward GH 09219-4 CGCAACAGCCAAATATCCATC Reverse GH 05213-1 CCTCAAACCCCCTCTCAACG Forward PDI 05213-2 ACTTCACCAACATTTAGGACGCC Reverse PDI 05213-3 CCTCAAACCCCCTCTCAACG Forward PDI 05213-4 ACTTCACCAACATTTAGGACGCC Reverse PDI 1Direction relative to gene transcription

82

4.2.5. Temperature sensitivity assay

Histoplasma cultures were grown in liquid culture to exponential phase then diluted to 1x107 yeasts/mL and dotted (10μL) on replicate HMM plates in a four fold dilution series. Plates were incubated at 35°C or 38°C. Plates were observed for visual growth and GFP fluorescence to verify RNAi knockdown was successful after 7 and 10 days respectively.

4.2.6. Determination of the O-linked proteome

Porcine stomach mucin (PSM) was used in combination with 100X concentrated Histoplasma supernatant to develop this method. 2 mg protein (amount used regardless of protein source) was treated to remove N-linked glycans with PNGase F (New England Biolabs) incubation at 37 °C overnight after heat denaturing (95 °C for 10 min). Treated supernatant was desalted (Sephadex-

G25 resin) then oxidized with 20 mM sodium periodate (Thermo fisher) in 1M sodium acetate for 30 min on ice in the dark. Samples were exchanged into phosphate buffered saline (PBS) then 5mM biotin was added (EZ-Link Alkoxyamine-PEG-Biotin; Fisher Scientific) for 2 hours while rotating at room temperature. Samples were desalted again to remove excess biotin.

Samples were purified using a streptavidin (Thermo fisher) resin to bind biotinylated proteins.

Protein was eluted via 8M guanidine-HCl (pH 1.5) or addition of 5mM biotin to the column.

Proteins were separated for visualization under reducing conditions by 10% SDS-polyacrylamide gel electrophoresis and transferred to nitrocellulose membranes or directly added to nitrocellulose membranes. Biotinylated proteins were detected with streptavidin horseradish peroxidase (HRP)-conjugated anti-monoclonal antibodies to biotin (Peroxidase Labeled

83

Strepavidin Kirkegaard &Perry Laboratories) and visualized with HRP chemiluminescent substrate (Millipore).

4.3 Results

4.3.1 Mass spectrometry fails to adequately detect O-linked proteins

Originally our approach for identification of O-linked mannosylated proteins attempted to exploit the size differences seen in previous gel-shift assays (Garfoot et al. 2018). Our approach is outlined in Figure 4.1A. The mass to charge ratios (m/z) of peptides in the pmt2 and mnt1 spectra were planned to be compared. Peptides shifted by factors of 162m/z (the size of a single mannose) in the mnt1 mutant, which is not capable of further mannose attachment after the initial mannose, would be identified as mannosylated. Cfp4 was chosen as a known O-mannosylated protein control to verify our spectra analysis. As most proteins are predicted to contain multiple serines and threonines that are mannosylated, we planned to identify these fragments in the pmt2 spectra, which should not have any mannosylation, but not in the mnt2 spectra, as the additional multiple mannosylation will make their m/z ratios difficult to predict in proteomics data. In order to prevent loss of mannosylation by ionization during mass spectrometry, ionization via electron transfer dissociation (ETD) was used in addition to analysis using collision-induced dissociation

(CID). ETD is considered the best way to generate a high amount of peptide backbone fragmentation, which is required for peptide sequence determination while minimizing fragmentation of post-translational modifications on peptide residues (Kim and Pandey 2012;

Mechref 2012; Mikesh et al. 2006).

84

Unfortunately, this approach was not successful. We ran proteomics samples of supernatants of

Histoplasma pmt2 (RNAi) knockdown, mnt1 (RNAi) knockdown and wild-type Histoplasma.

However, even for the control protein Cfp4, the proteomics analysis was not able to identify the peptide that contains the mucin-domain in any sample (Fig. 4.1B top). The same was true for another protein, Eng1, which has also been shown in gel-shifts to be O-linked glycosylated

(unpublished by Andrew Garfoot) (Fig. 4.1B bottom). In addition, we also saw that the Eng1 peptides that were not predicted to be mannosylated (amino acids 1-57) were also not always detected. Because of this, all missing peptides could not be accurately identified or considered as being O-linked mannosylated. As we did not observe complete coverage of non-glycosylated peptides, our comparative identification approach could not work.

85

Figure 4.1 Proteomic approach fails to identify O-linked mannosylated proteins.

(A) Schematic of the comparative proteomic approach where supernatants from the mannosylated mnt1 and non-mannosylated pmt2 will be collected, trypsin digested then LC-MS/MS used for identification of peptides. Differences in spectra between the two mutants will be identified as O-linked mannosylated. (B) Detection of peptides of two known O-linked mannosylated proteins Cfp4 (top) and Eng1 (bottom). Yellow indicates detection via mass spectrometry, and peptides from trypsin digestion that are predicted to be regions of mannosylation are indicated in red boxes. Green residues are predicted modification sites (non-O-linked glycosylation).

86

4.3.2 Bioinformatics approach to identification of O-linked mannosylated proteins

As another approach this study tried to identify O-linked mannosylated proteins, mucin-like domain containing proteins were identified by bioinformatics. While there is not a true consensus sequence for O-linked mannosylation, software exists that will determine the probability that a serine or threonine residue is O-linked glycosylated based on artificial neural networks trained on the O-GLYCBASE. We took all predicted proteins from Histoplasma and using prediction software (NetOGlyc) identified proteins likely to be O-linked glycosylated. We then checked these proteins for predicted secretion signals. Proteins containing secretion signals were then analyzed for domain annotation. This resulted in 51 high-priority proteins to investigate (Fig.

4.2A and Table 4.3). Priority was given to proteins with large numbers of predicted sites as well as known homologs from Aspergillus spp. and Saccharomyces cerevisiae. As the attenuated virulence of O-linked mannosylation-deficient strains was found to be due to temperature sensitivity (Garfoot et al. 2018), temperature sensitivity was identified as a efficient phenotype to screen for as a virulence proxy. To test if these proteins identified via the bioinformatic approach are required for growth at elevated temperatures, two independent RNAi knockdowns of selected genes were generated and tested for growth at 35°C and 38°C with the pmt2 deficient mutant serving as a positive control for sensitivity to higher temperatures. Initially one Cts4 RNAi line showed reduced growth at 38°C but this phenotype was not observed with a second independent line (Fig. 4.2B). More independent RNAi knockdowns of Cts4 were generated (3) and none were inhibited in growth at 38°C. Therefore the initial temperature sensitive Cts4 strain is not due to the silencing of CTS4 and is more likely a result of the specific location of the T-DNA insertion into the genome.

87

Figure 4.2 Bioinformatic determination of candidates and initial temperature sensitivity assay.

(A) Reductive path taken by bioinformatics for determination of the final list of potential O-linked mannosylated proteins. (B) RNAi knockdown strains, pmt2 knockout temperature sensitive control, and WT strains grown on HMM plates at 35°C (left) and 38°C (right).

88

Table 4.3 Predicted O-linked Mannosylated Proteins

Protein Histoplasma Name Protein ID Domains Bgl5 Histo217_03475 Glycosyl hydrolase family 1 Crh Histo217_03868 GH16_fungal_CRH1_transglycosylase, Crh Crr1 Histo217_03564 GH16_fungal_CRH1_transglycosylase-related to Crr1 Cts1 Histo217_04563 GH18_chitinase; GPI anchored Cts4 Histo217_03793 GH18_chitinase Glycosyl hydrolase family 81; Family of eukaryotic beta-1,3- Eng1 Histo217_07625 glucanases. ENG1 Eng3 Histo217_02903 GH16_fungal_Lam16A_glucanase; DUF4045 Pectate lyase superfamily protein/glycosyl hydrolase family 28 (related), Peptidases_S8_S53, Herpes BLLF1 glycoprotein Exg6 Histo217_03934 superfamily Pectate_lyase 3 superfamily, glyco_hydro_28 superfamily (cell Exg8 Histo217_00486 wall metabolism) HYP- Glycosyl hydrolases family 15; CBM20_glucoamylase Glycosyl (Glucoamylase (glucan1,4-alpha-glucosidase), C-terminal hydrolase Histo217_09219 CBM20 (carbohydrate-binding module, family 20) domain.) HYP- Protein Disulfide Isomerase (PDIa) family, YbbN superfamily PDI Histo217_05213 neg reg of GroEL CFEM (eight cysteine-containing domain present in fungal extracellular membrane proteins), UBQ superfamily (ubiquitin Histo217_02157 homologues) Histo217_03482 Fasciclin; Fas1/BglH3 Histo217_08993 NOX_Duox_like_FAD_NADP, Ferric_reductase like Histo217_08263 Pepsin-like aspartic proteases Maintenance of mitochondrial morphology protein 1 Histo217_06703 superfamily, SARAF superfamily -ER? Histo217_03445 GPI-anchored protein Histo217_00951 Epstein-Barr virus nuclear antigen 3 (EBNA-3) superfamily Saccharomyces cerevisiae YHR202W and related proteins, N- terminal metallophosphatase domain; YHR202W is an uncharacterized Saccharomyces cerevisiae UshA-like protein with two domains, an N-terminal metallophosphatase domain Histo217_06409 and a C-terminal nucleotidase domain. Protein of unknown function (DUF1620), PQQ-dependent Histo217_05977 dehydrogenases and related proteins Peptidases_S8_Protein_convertases_Kexins_Furin-like, Histo217_01703 Proprotein convertase P-domain Histo217_02065 Multi-glycosylated core protein 24 (MGC-24), sialomucin WSC domain; This domain may be involved in carbohydrate Histo217_07821 binding. Protein of unknown function (DUF4448) Histo217_04605 CFEM; DNA_pol3_delta2 superfamily

89

Histo217_08411 Atrophin-1 superfamily Histo217_00207 ATP-synt_S1 superfamily Histo217_02788 CFEM Histo217_05959 WSC superfamily Peptidase_M14NE-CP-C_like superfamily; WW domain; The WW domain is a protein module with two highly conserved Histo217_04031 tryptophans that binds proline-rich peptide motifs in vitro Histo217_00567 Atrophin-1 superfamily; CFEM PDI family, MPD1-like subfamily; composed of eukaryotic proteins similar to Saccharomyces cerevisiae MPD1, also Histo217_02906 contains Aspergillus niger prpA Histo217_03498 EamA-like transporter family; Atrophin-1 superfamily Histo217_07183 Atrophin-1 superfamily Histo217_07169 Atrophin-1 superfamily; PAT1 superfamily Histo217_05175 CFEM Histo217_00721 Cupredoxin STKc_IRE1 (Catalytic domain of the Serine/Threonine kinase, Inositol-requiring protein 1), The Luminal domain, a dimerization domain, of the Serine/Threonine protein kinase, Inositol-requiring protein 1; Ribonuclease 2-5A; This domain is Histo217_06551 an endoribonuclease. Histo217_03333 GPI-anchored Histo217_00587 DUF 4965/1793/5127 Histo217_00193 p450 superfamily Histo217_01411 CYYR1 superfamily-cysreine and tyrosine rich Histo217_03075 GPI-anchored Histo217_03382 Atrophin-1 superfamily Histo217_08008 Atrophin-1 superfamily Histo217_04553 Gdt1; LGT superfamily HYOU1-like_NBD (Saccharomyces cerevisiae Lhs1p (also Histo217_06954 known as Cer1p, SsI1)) Histo217_04892 Atrophin-1 superfamily Histo217_08154 Phosphatidylethanolamine-Binding Protein (PEBP) domain Histo217_06883 Abhydrolase superfamily RSN1_7TM superfamily (Calcium-dependent channel-seven transmembrane domain); seven transmembrane domains Histo217_08878 (Extracellular tail, of 10TM putative phosphate transporter) Domain annotations obtained from CDD. Gene names obtained from homology to Saccharomyces cerevisiae and Aspergillus spp.

90

4.3.3 Biotinylation approach to identification of O-linked mannosylated proteins

As our initial experimental and bioinformatic approaches to identification of O-linked mannosylated proteins did not identify proteins required for thermotolerance, we next tried a biochemical approach to identification of O-linked mannosylated proteins. Histoplasma O-linked glycosylation is comprised of repeating mannose (i.e. mannosylation) residues, which is not congruent with common lectin purification methods (Cummings et al. 2015). Therefore we used a lectin-independent method based on biotinylation of mannose to capture mannosylated proteins. Our novel approach is illustrated in Figure 4.3A. Briefly, N-linked glycosylation is removed from a protein sample with a specific enzyme (PNGaseF). Then we oxidized terminal mannose residues on any remaing O-linked mannosylated proteins. This generates a carbonyl reactive group (-CHO) that can be linked to biotin (via a reactive alkoxyamine group conjugated to biotin via a polyethylene glycol (PEG) spacer arm) resulting in a terminal biotin that can be captured by affinity chromatography with streptavidin resulting in purification of proteins with

O-linked mannosylation. Once selectively purified, we proposed to identify these proteins with mass spectrometry proteomics. We have been working on developing and validating this method. Initial attempts to biotinylate O-linked glycosylation in both Histoplasma supernatants and Porcine Stomach Mucin (PSM), a heavily O-linked glycosylated protein that was used for methodology development and optimization, failed to detect attached biotin. As our approach is heavily dependent on chemical reactions, we were concerned that the NP-40-containing denaturing buffer used during the PNGaseF treatment to remove N-linked glycosylation was inhibiting subsequent steps. When we removed the NP-40 from the PNGaseF treatment (heat denatured proteins before treatment as a replacement) we managed to detect the presence of biotin via western blot, but not when NP-40 was present despite the same oxidation and

91 biotinylation treatments (Fig. 4.3B). However, this verified that biotinylation of these proteins was possible. We do not believe that the NP-40 inhibited oxidation as Pro-Q Emerald 300

Glycoprotein staining of standard PNGase F treatment indicated the presence of oxidized carbohydrate groups (data not shown) We were also concerned with excess biotin potentially inhibiting binding of biotinylated proteins to a streptavidin column via competitive inhibition.

We did try to reduce the amount of biotin in the biotinylation step but were unable to obtain labeling of protein when the biotin concentration was lower than 25mM. Therefore, we have instead focused on extensive desalting following the biotinylation process (Fig. 4.3C). Currently we have managed to get biotinylated proteins to bind a streptavidin column as seen by the reduction of biotin signal seen by western analysis in the flow through (FT) fractions compared to the pre-column samples (Pre) (Fig. 4.3D). Proteins are not bound with 100% efficiency, but that is not unexpected as we cannot accurately estimate the number of biotin-binding sites on proteins and therefore available to bind the streptavidin resin. We are currently working on capture of biotinylated proteins using streptavidin-agarose beads. Attempts using 8M guanidine-

HCL (pH 1.5) and biotin competition to elute (Fig. 4.3D) have so far been unsuccessful. Work is continuing to optimize this portion of the protocol, with future attempts aimed at using trypsin to liberate peptides associated with O-mannosylated proteins, which will then be available for mass spectrometry identification.

92

Figure 4.3 Biotinylation for purification of O-linked mannosylated proteins.

(A) Schematic for the biotinylation of O-linked mannosylated proteins for purification and identification. (B) Immunoblot of biotin (via streptavidin-HRP visualization) detection in Histoplasma supernatant (No treatment), Histoplasma supernatant after PNGase F treatment, oxidation and biotinylation (DB) and Histoplasma supernatant after PNGase F treatment without NP-40, oxidation and biotinylation (No DB). (C) Detection of biotinylated PSM from 5 different biotinylation procedures on an immunoblot transferred from a SDS-PAGE gel. Increasing concentrations of biotin during biotinylation (left to right, maximum 5mM). (D) Detection of biotinylated PSM during two elution procedures on an immunoblot transferred from a SDS-PAGE gel. Samples are: protein not run through the column (Pre), flow-through from a streptavidin agarose column (FT), flow-through during wash, and elution by 8M guanidine hydrochloride (GuHCl) pH 1.5 (right) or by addition of 5mM biotin (left).

93

4.6 Discussion

This study attempts to characterize the O-linked proteome in Histoplasma in order to identify O- linked mannosylated proteins that are necessary for thermotolerance and virulence. Our initial approach of using mass spectrometry to compare the proteomes of mnt1 and pmt2 deficient strains to identify mannosylated peptides failed. This was due to the limitations of the proteomics analysis. The analysis relied on identification peaks detected via predicted m/z ratios. For identification of proteins, whole protein sequences are run through a series of predictive algorithms in order to predict the m/z ratio of every peptide that could be generated. While some degree of post-translational modification can be accounted for, it again has to be a specific m/z ratio (size) difference. In this analysis, we were not able to predict exactly how many additional mannoses should be expected for each peptide. This was due to a variety of factors: not having a consensus sequence, mannosylation potentially not being uniform on every individual protein of a gene, and potential variability in which potential sites were mannosylated. In addition, for most mass spectrometry analyses, a general idea of the size of peptide expected is necessary as most peak collections occur in specified m/z ranges. With the increased mass from the mannose residues it is likely many peptides were outside the range of detection used for data collection. In addition, we failed to detect predicted O-linked mannosylated peptides in the pmt2 mutant strain, which should have reduced mannosylation. This is most likely due to promiscuity of the other

PMT proteins to target proteins. Our lab has observed residual O-linked glycosylation of specific proteins (e.g. Cfp4) in mannosylation deficient strains (Garfoot et al. 2018). Due to the fact the slight changes in mass negate our ability to detect the mannosylated peptide, this residual activity

94 is likely the reason we failed to detect any of the control peptides. Previous attempts to create double knockouts in Histoplasma have not been viable (Andrew Garfoot, Unpublished data).

We also took a bioinformatics approach to identifying O-mannosylated proteins. Due to the lack of a specific consensus sequence, we decided to focus our bioinformatics approach on identifying a pool of prioritized targets to screen using thermotolerance as a phenotype. While we successfully generated a reasonable list of potentially mannosylated proteins to test, there were no obviously thermotolerance-related proteins. In addition, the software may miss some mannosylated proteins, especially those with reduced numbers of potential O-linked sites or falsely selects proteins with potential sites but which are not actually glycosylated. However, we can continue to test genes from our bioinformatic list for heat sensitivity.

Due to the limitations of a bioinformatic approach for O-linked glycosylation, and our initial failure at an experimental dataset, we worked to develop a novel method to purify O-linked glycosylated proteins in Histoplasma. While several studies have used lectins to probe glycosylation, this was not a viable approach for studying Histoplasma O-linked glycosylation.

Jacalin is a lectin that has been used in some studies to purify O-linked glycosylated proteins for mass spectrometry proteomic analysis (Durham and Regnier 2006; Hortin and Trimpe 1990).

However, Jacalin binds N-acetylgalactosamine, not mannose. As Histoplasma’s O-linked glycosylation consists of mannoses (as seen in mass spectrometry of purified Cfp4), this is not as viable approach for our purposes (Garfoot et al. 2018). The common ligand for mannose residues, Concanavalin A (Con A) is actually used to identify N-linked glycosylation

(Cummings et al. 2015). Our lab did attempt a pilot test using ConA to be thorough, but did not

95 observe binding of O- mannosylated proteins (Andrew Garfoot, unpublished results). Therefore we needed to develop a novel O-linked glycosylation-specific method.

The approach we have been developing is based on the attachment of biotin specifically to mannose and the exploitation of the well-developed streptavidin-biotin purification protocols.

We have successfully managed to biotinylate O-linked mannosylated proteins. However, the elution of proteins from the streptavidin resin has proven difficult. Traditional guanidine hydrochloride elution has failed, and we were unable to obtain a cleavable biotin that will react with carbonyl groups as it has been discontinued. We are attempting a trypsin-proteolysis strategy to release peptides from the captured biotinylated O-linked mannosylated proteins.

Another alternative method we could attempt is to use click chemistry. Some studies have managed to use azido-N-Acetylgalactosamine (GalNAc) to incorporate a reactive azido group into O-linked glycosylation during protein post-translational modification, then use alkyne- biotins to subsequently label the O-linked glycosylated proteins (Boyce et al. 2011; Dube et al.

2006). These alkyne-biotins are cleavable via the linker without breaking the streptavidin-biotin bond and therefore are much easier to release, which would solve our elution problem. However, these studies were done in mammalian cells (Boyce et al. 2011; Dube et al. 2006). We will first have to establish if the azido-sugar precursor molecules can (1) enter Histoplasma and (2) be incorporated into glycosylated proteins without modification (e.g., loss of the azido group). It does however; offer another avenue for further work.

96

Chapter 5: Diversification of fungal chitinases and their functional differentiation in Histoplasma

capsulatum

5.1 Introduction

Chitin is a 1, 4- β-linked N-acetyl-D-glucosamine (GlcNAc) polymer. As the second most abundant biopolymer after cellulose (Tharanathan and Kittur 2003), chitin and its deacetylated derivative chitosan are abundant sources of organic carbon and nitrogen that have many potential industrial uses, from biomedical to agricultural to water engineering (Ravi Kumar 2000; Zargar,

Asghari, and Dashti 2015). Consequently, there is a great interest in identifying enzymes that efficiently hydrolyze chitin into more soluble mono and oligomers of GlcNAc. Chitin degrading enzymes also have potential applications in the breakdown of the chitinous structures of fungal and arthropod agricultural pests (Hamid et al. 2013).

Chitinases (E.C 3.2.2.14) are glycosyl hydrolases that are found in a wide range of plants, bacteria, and fungi. For plants and bacteria, which lack chitin, chitinases play roles in defense.

Chitin is an important structural component of the fungal cell wall, ranging from 0.5%-5% in yeasts to ≥20% in filamentous fungi (Hartl, Zach, and Seidl-Seiboth 2012). Correspondingly, yeasts possess relatively few chitinases (e.g., two in Saccharomyces cerevisiae) compared to the

10-20 chitinases encoded in filamentous fungal genomes (Seidl 2008a). The expansion of

97 chitinases in filamentous fungi has prompted investigation of chitinase roles in formation of hyphal structures (Seidl 2008a). While some studies have shown chitinases whose expression is specifically induced during hyphal formation (Gruber, Kubicek, and Seidl-Seiboth 2011; Takaya et al. 1998), many others are not (Duo-Chuan 2006; Gruber et al. 2011; Seidl 2008a) suggesting alternative roles for diverse chitinases. For example, there is evidence that specific chitinases facilitate mycoparasitism of other fungi by Trichoderma (Boer et al. 2007; Cruz et al. 1992;

Seidl et al. 2005).

Fungal chitinases are characterized by the presence of the gycoside hydrolase 18 family (GH18)

(Seidl 2008a) domain. Additional common domains include: an N-terminal signal peptide region, a serine/theronine-rich region, one or more chitin-binding domains (CBDs) and extension in the C-terminal region, although none of these are conserved in all fungal chitinases or required for a protein to be considered a chitinase (Duo-Chuan 2006). The LysM domain is a domain that allows for binding to polysaccharides such as peptidoglycan and chitin (i.e. a CBD) that is particularly important in chitinases and fungi. A three-clade classification system for fungal chitinases (clades A, B, and C) emerged from several investigations into the diversity of GH18- domain enzymes among fungi (Karlsson and Stenlid 2008; Karlsson and Stenlid 2009; Seidl et al. 2005). Clade A chitinases contain the GH18 catalytic domain but were reported to have no

CBD(Seidl 2008a). Clade A enzymes are secreted enzymes between 40-50 kDa and are the most well-studied of the three clades (Seidl 2008a). Clade B chitinases are variable in size and in the presence and number of CBDs. Clade C chitinases are similar to clade A, but have been distinguished by their typically larger size (140-170 kDa) and the presence of multiple CBDs and

LysM motifs that are particularly characteristic of this clade (Seidl 2008a; Seidl et al. 2005)

98

(Karlsson and Stenlid 2008; Karlsson and Stenlid 2009; Seidl et al. 2005). Clade C chitinases are the least well characterized of all the clades, containing very few members that have been included in previous phylogenetic studies, none of which have been functionally characterized

(Duo-Chuan 2006; Seidl 2008a). However, this three clade classification is based on analyses with a limited number of fungal genomes, and it remains to be determined if it is robust to more in-depth sampling (Karlsson and Stenlid 2008; Karlsson and Stenlid 2009; Seidl et al.

2005).

While many fungal chitinases have been transcriptionally profiled, the characterization of fungal chitinase enzymatic specificity is particularly limited, leaving broad assumptions about clade- specific functions untested. For example, Clades A and C are class V (fungal/bacterial) chitinases, which are generally assumed to be exochitinases based on modeling of the conserved binding groove as deep and tunnel shaped (Duo-Chuan 2006; Hartl, Zach, and Seidl-Seiboth

2012; Seidl 2008a). Clade B corresponds to class III (fungal/plant) chitinases and are therefore predicted to be endochitinases due to the modeling of their binding grooves as shallow and open

(Duo-Chuan 2006; Hartl, Zach, and Seidl-Seiboth 2012; Seidl 2008a). While some of these assumptions are supported by the activities of specific chitinases, for example those of

Trichoderma harzianum CHIT33 and CHIT42 (Boer et al. 2007; Lienemann et al. 2009), these chitinases also have multiple types of activity and further examples need to be studied (Hartl,

Zach, and Seidl-Seiboth 2012). The inconsistent use of diverse chitin substrates (e.g., crustacean chitin, or fungal chitin), further confounds the accuracy of assumptions about functional diversity within and among clades, and while there are reports of complete transcriptional/deletion studies and sub-clade transcriptional or deletion studies (Alcazar-Fuoli et al. 2011; Dünkler et al. 2005;

99

Gruber et al. 2011; Gruber, Kubicek, and Seidl-Seiboth 2011), enzymatic activity studies are not as comprehensive.

In this chapter, we identified the GH18 domain proteins encoded in 367 published fungal genomes to provide a complete analysis of the distribution and evolution of GH18 proteins. In addition, we provide initial expression analysis and enzymatic characterization of the chitinase enzymes produced by the thermally , Histoplasma capsulatum. This organism has distinct and tightly controlled morphologies (mycelia or yeasts) each of which has a specific lifestyle (environmental saprobe or pathogen of mammals, respectively) potentially allowing for specific roles for different chitinases. These chitinases provide new enzymatic examples from each of the major clades, including enzymatic activity for the first characterized clade C chitinase.

5.2 Materials and Methods

5.2.1 Phylogenetic Analyses

Putative chitinases were retrieved from 367 fungal genomes, including 6 Histoplasma genomes

(Supplemental File 2) obtained from the Joint Genome Institute (JGI) website. Proteomes were independently mined for glycosyl hydrolase 18 (GH18) containing proteins (El-Gebali et al.

2019) using hidden Markov models (HMM search) (Eddy 2009). A CDD search was then performed on proteins containing GH18 domains to identify additional carbohydrate binding domains (Marchler-Bauer et al. 2017) and LysM domains. Alignments of the GH18 domains and

LysM domains were performed by HMMalign which uses a Viterbi algorithm to align each

100 sequence to the given HMM (Eddy 2009). A dynamic programming algorithm, HMMalign identifies residues in the sequence that match to consensus columns in the original model and displays residues that do not correspond as insertions in the alignment and missing residues as gap characters (HmmerAlign n.d.). GH18 domain sequences were removed if they were missing alignment between positions 87-238. Poorly aligned characters were removed using trimAl (v.

1.4) (Capella-Gutiérrez, Silla-Martínez, and Gabaldón 2009) with a gap threshold of 0.01.

Phylogenetic analysis of the GH18 domains was performed using maximum-likelihood methods using VT+F+G4 (GH18) or WAG+I+G4 (LysM) model (automatically determined) in IQ-TREE

(Nguyen et al. 2015). This tree was used to determine the sequences in the AC clades and the B clade. These sequences were then split into separate phylogenetic analyses again using maximum-likelihood methods (model WAG+G4 for AC tree and PMB+F+I+G4 for 4B tree as automatically determined) implemented in IQ-TREE (Nguyen et al. 2015). Statistical support for all IQ-TREEs was assessed by Ultrafast bootstrap analysis using 1,000 replicates (Minh,

Nguyen, and von Haeseler 2013). An ultrafast bootstrap value of ≥ 95 was considered a strongly supported branch. In addition, maximum-likelihood methods were implemented in RAxML (v.

SSE3) (Stamatakis 2014), 100 alternative runs on 100 distinct trees, bootstrapping (100) and best scoring ML in 1 run and GAMMA models autodetermined (WAG for AC and PMB for B).

Statistical support for RAxML trees was assessed by rapid bootstrapping, where nodes receiving

≥60 percent of bootstraps were considered supported.

5.2.2 Chitinase gene expression

101

H. capsulatum CTS gene transcription was determined by quantitative RT-PCR (qRT-PCR). The

WU15 strain of Histoplasma, a uracil auxotroph of the North America 2 clade was grown on

HMM (Worsham and Goldman 1988) or SDA (40 g/L dextrose and 10 g/L peptone) media supplemented with 100 ug/mL uracil and solidified with 0.6% agarose. For tests of chitin induction of CTS gene transcription, colloidal chitin (prepared from shrimp chitin by the method of (Rodriguez-Kabana et al. 1983)) was added to media (1.2% final concentration). Media were inoculated with Histoplasma cells and incubated at 25 degrees (mycelial culture) for 5 weeks or

37 degrees (yeasts culture) for 7 days. Histoplasma cells were scraped from the solid media and collected by centrifugation (5 minutes at 2000 x g). RNA was isolated from fungal cells by mechanical disruption with 0.5 mm glass beads, extraction with RiboZol (Amresco) and purified using an affinity column (Direct-zol RNA MiniPrep Plus kit; Research Products International).

Following DNA removal with DNase (Invitrogen) RNA was reversed transcribed (Maxima reverse transcriptase; Thermo Scientific) primed with random pentadecamers. Quantitative PCR was carried out using CTS gene specific primer pairs (Table 5.1) with SYBR green-based visualization of product amplification (Bioline). Changes in CTR transcript levels relative to

Actin (ACT1) and Ribosomal Protein S15 (RPS15) were determined using the ΔΔCt method(Schmittgen and Livak 2008) after normalization of cycle thresholds to ACT1 and RPS15 mRNA levels. RNAs were prepared from 3 biological replicates for each media/condition.

Samples with transcripts below the detectable limit were set to -12.00 for analysis purposes.

102

Table 5.1 Primers used in this study

Primer Use Primer Sequence 5' to 3' Direction1 CTS1-8 qPCR TGTTGTTGTTGACGGATTTGATT Forward CTS2-7 qPCR GGCAAAGATCGCAGATTCC Forward CTS3-5 qPCR CGCAAACGTCAAACCAGATT Forward CTS4-3 qPCR CAAGTTCTATCTAACCGTCGC Forward CTS5-1 qPCR ATGCGACTTGTGGACCTCTC Forward CTS6-7 qPCR CTTCTCACCGCTCAGGGATG Forward CTS7-1 qPCR ACGGCGGGAGGGGAAAGGGT Forward CTS8-5 qPCR CGAGGACTGTGGAAGGGAGG Forward CTS1-5 expression ATGGCTCTCAAAAGCCTTC Forward

CTS2-5 expression ATGAAACAGCTCCCCCGCTGC Forward

CTS3-5 expression ATGTCCACCTACACCGTTGCCA Forward

CTS4 expression ATGGGCCCCTTTTTCCTCTACA Forward

CTS5 expression ATGCGTTTGTTTAGAGTCCCC Forward

CTS6 expression ATGAAGACCCTGAATGGCAT Forward

CTS7 expression ATGTATCCTTTTCAGAGATCTCCC Forward

CTS8 expression ATGAAAAAACTCCACGCCCA Forward CTS1 qPCR TCAGGGGGACATTGCGGA Reverse CTS2 qPCR TTGGTGGAGAAAGGAGTTGAAT Reverse CTS3 qPCR CCGATGGAGAGGAGGGTT Reverse CTS4 qPCR TGGATTTTGGATCGGATTTG Reverse CTS5 qPCR TGCTCCTTGTCGGTCATCCC Reverse CTS6 qPCR CACAGCCATTGATTAACTCC Reverse CTS7 qPCR CTCCGTTCCCACCTCCGCAT Reverse CTS8 qPCR GGACTGCGGCTGAGATTGTT Reverse CTS1 expression ACTGCCAGTGTACATCGGT Reverse

CTS2 expression GTGTTGTGTCGTACCATTGGGA Reverse

CTS3 expression CTGAAATCCCTTCTTCATATTCTCAT Reverse

CTS4 expression CTTCTCCGCAAACCCGGC Reverse

103

CTS5 expression AGCTGTCCCCTTATATCCCTC Reverse

CTS6 expression ATGCTTTTGGAGTCGGTTCA Reverse

CTS7 expression TCCTCCTCTTCTCCTCTTACCC Reverse

CTS8 expression CTTCACAATCCCAATCTCTTTC Reverse 1Direction relative to gene transcription

5.2.3 Purification of H. capsulatum chitinases

Chitinases were purified from transgenic H. capsulatum yeasts overexpressing each protein.

Chitinase-encoding genes were amplified by PCR (Table 5.1) from wild-type H. capsulatum

G217B genomic DNA and cloned into the H. capsulatum expression vector pAG38 placing the gene under transcriptional control of the H2B constitutive promoter and fusing the chitinase to a

C-terminal hexahistidine tag. For Cts1, sequences encoding the putative GPI-attachment site

(nucleotides 2141-2215) were removed. Overexpression plasmids were transformed into H. capsulatum WU15 yeasts by Agrobacterium-mediated transformation (Zemska and Rappleye

2012) and transformants selected by uracil prototrophy. Transformants were screened by immunoblotting of culture filtrates and cellular lysates for the hexahistidine tag (GnScrpt

A00186 to 6xhis). Chitinase-expressing transformats were grown in liquid HMM until stationary phase and the yeasts separated from the culture supernatant by centrifugation (5 minutes at 2000 x g). For Cts1, Cts2, Cts4, Cts5, Cts6, Cts8, culture filtrates were prepared by filtration of the supernatant (0.45 um pore; Millipore), concentration 100 fold by ultrafiltration (10 kDa pore), and the proteins exchanged into phosphate-buffered saline (PBS). For Cts3 and Cts7, lysates were prepared from the yeast cells by suspension in PBS and mechanical breakage with 0.5 um

104 diameter glass beads. Debris was removed from the cellular lysate by centifiguation (10 minutes at 12000 x g). Hexahistidine-tagged chitinase proteins were purified from the concentrated culture filtrates or cellular lysates by metal affinity chromatography (HisPur Co2+ Resin, Thermo

Fisher Scientific) and the elution exchanged into PBS by ultrafiltration. Resultant protein concentrations were determined using a Bradford assay (Sigma-Aldrich) and purity of the protein preparation determined by SDS-PAGE followed by silver staining. Purified proteins were stored at −20 °C in 50% glycerol.

Table 5.2 Histoplasma strains

Other Strain Genotype Designation1 WT G217B Wild type NAm1 isolate (ATCC26032) WU15 G217B ura5-42Δ G217B ura5-42Δ zzz::pKG27 (URA5, OSU395 CTS1ΔGPI:6xHIS) CTS1 G217B ura5-42Δ zzz::pAM11 (URA5, OSU402 CTS2:6xHIS) CTS2 G217B ura5-42Δ zzz::pJW05 (URA5, OSU403 CTS3:6xHIS) CTS3 G217B ura5-42Δ zzz::pKG26 (URA5, OSU399 CTS4:6xHIS) CTS4 G217B ura5-42Δ zzz::pJW06 (URA5, OSU404 CTS5:6xHIS) CTS5 G217B ura5-42Δ zzz::pJW08 (URA5, OSU405 CTS6:6xHIS) CTS6 G217B ura5-42Δ zzz::pJW09 (URA5, OSU407 CTS7:6xHIS) CTS7 G217B ura5-42Δ zzz::pJW07 (URA5, OSU406 CTS8:6xHIS) CTS8 1gene designations: zzz::T-DNA: T-DNA integration at an undetermined chromosomal location URA5: orotate phosphoribosyltransferase CTS: chitinase (GH18 domain)

105

5.2.4 Chitinase activity and specificity determination

Chitinase enzymatic activities were determined via the fluorimetric chitinases assay kit (Sigma-

Aldrich-CS1030 - Chitinase Assay Kit, Fluorimetric). Three artificial substrate mimics, 4-

Methylumbelliferyl N,N’-diacetyl- β-D-chitobioside, 4-Methylumbelliferyl N-acetyl- β-D- glucosaminide and 4-Methylumbelliferyl β-D-N,N’,N’’-triacetylchitotriose were used to determine the activity of each chitinases with regards to chitobiosidase, exochitinase, and endochitinase activity, respectively. Hydrolysis of each substrate releases 4-methylbelliferone, the fluorescence of which was measured using a plate reader (360 nm excitation and 450 nm emission; BioTek Synergy 2). For chitinase activity reactions, varying amounts of purified chitinases were added to 0.5mg/mL substrate in reaction buffer (see kit protocol) and incubated at 37°C. Reactions were monitored by endpoint fluorescence. Data is reported as nanograms of

4-methylumbelliferone released per minute per nanogram of enzyme based on a standard curve of 4-methylumbelliferone. Activity assays were performed in triplicate. Significant differences in expression (P<0.05) were determined by ANOVA and pairwise differences determined by

Tukey’s post-hoc test.

106

5.3 Results

5.3.1 Diversification of fungal GH18 domains

We identified 3,888 GH18 domain-containing proteins in 373 publicly available fungal genomes using a HMM search (Eddy 2009). 494 of these proteins contain the LysM chitin binding domain considered to be characteristic of the C clade chitinases in fungi (Seidl 2008a; Seidl et al. 2005).

1,250 CBDs (ChtBD1, ChtBD3, COG3979, CBM_1 and CBM_19) were also identified in the

GH18 domain containing proteins (Marchler-Bauer et al. 2017).

An alignment of the GH18 domain was 1,840-characters long; however the end quarter of the alignment is not well-conserved and actually missing particularly in the B clade gene members

(even functionally validated chitinases (Dünkler et al. 2005; Hurtado-Guerrero and van Aalten

2007)). Maximum likelihood phylogenetic analysis using IQ-Tree identified the previously reported distinct AC and B clades, although these were not well-supported (70% of rapid bootstraps (RB)) (Fig, 5.1A). Clades and sub-clades, determined by identifying the most recent common ancestor of chitinases previously categorized as sub-clade members (Karlsson and

Stenlid 2008; Karlsson and Stenlid 2009; Seidl et al. 2005), were generally recovered. Further expansion of these clades recovered more strongly supported nodes without encroaching on other clades (Fig. 5.1). In order to retain more alignable characters and reduce error introduced by homoplasy for finer resolution within clades and sub-clades, we separately analyzed sequences on either half of a bifurcation between B and AC clades in the complete tree (Figs. 5.1B, 5.1C,

Table 5.3). We did not recover a B-I clade in the complete tree, and no B sub-clades received strong bootstrap support (Fig. 5.1A, Table 5.3). However, in the IQ-tree restricted to the B clade,

B-I (98% RB) B-IV (99% RB) and B-V (97% RB) clades were supported, while clades B-II and 107

B-III were not supported with the current strict sub-clade calling, however, slight expansions of the sub-clades would generate supported sub-clades consistent with previous analyses (Fig. 5.1B and Table 5.3). While a RAxML analysis of the B clade recovered these B sub-clades none were supported, suggesting the finer scale topology is not always robust to differences in methodology

(Table 5.1 and Supplemental File 3). For the A clades, we recovered distinct, yet unsupported A-

IV (58% RB) and A-V (50% RB) clades after reclassifying one Trichoderma reesei gene (Chi18-

5) from an A-V to an A-IV (Fig. 5.1A, and Table 5.3). The A-V sub-clade was not monophyletic in the AC only IQ-Tree (Fig. 5.1C). A-II (100% RB) and A-III (100% RB) clades were supported in all analyses (Fig. 5.1A, Fig. 5.1C and Table 5.3). The A-III sub-clade is also supported (99% RB, AC IQ-Tree) as earlier diverging from the A-IV, A-V, A-II and C clades, as is the A-IV and-V sub-clades (98%RB) from the A-II and C clades (Fig. 5.1A, Fig. 5.1C and

Table 5.3). Sub-clade C-I and was supported (99% RB) while sub-clade C-II was not (89% RB)

(Fig. 5.1A and Table 5.3). However, in the individual trees C-II is either supported (76%RB in

RAxML) or would be supported by including sister sequences (Fig. 5.1A and Fig. 5.1C).

Interestingly, in our analysis clade C (both C-I and C-II) groups closely and with strong support with A-II (97% RB All GH-18 IQ-Tree).

108

Figure 5.1 Phylogenetic analysis of fungal GH18 domains.

(A) Phylogenetic analysis of all GH18 domains with clades indicated by color (A is magenta, B is Cyan, and C is green) and sub clades indicated and collapsed (left). By each clade is a pie chart with the normalized percentage of each chitinase in the clade by early diverging fungi (blue), Basidiomycota (red) and Ascomycota (green). A heat map of the location of domain (GH18, CBD, or LysM) per genome by taxonomical class and GH-18 sub-clade with darker color indicating more domains/genome for each class (see Supplemental File 4). A simple cartoon of taxonomic tree is included below for reference. (B) Phylogenetic analysis of clade B fungal GH18 domains with sub clades indicated and collapsed. (C) Phylogenetic analysis of clades A and C fungal GH18 domains with clades indicated by color (A is magenta and C is green) and sub clades indicated and collapsed. (A) is rooted along the B vs AC clade split that has been widely accepted in the literature (Karlsson and Stenlid 2008, 2009; Seidl et al. 2005) and (B) and (C) are rooted as in the complete tree. Trees were built using maximum-likelihood methods and branch support assessed by 1000 Ultrafast bootstrap replicates (values shown) in IQ-tree. As sub-clade A-V is polyphyletic in AC IT Tree (C), the tree nodes from previously identified chitinases are indicated with red stars.

109

Table 5.3 Bootstrap support values for chitinase sub-clades

Clade All GH18 IQ Tree A/C or B IQ Tree A/C or B RAxML A-II 100 100 79 A-III 100 100 83 A-IV 58 67 54 A-V 50 NA 6 C-I 99 91 68 C-II 98 80 76 B-I NA 98 10 B-II 69 73 47 B-III 60 38 17 B-IV 84 99 1 B-V 82 97 54

110

5.3.2 Taxonomic distribution of GH18 chitinases

Each sub-clade hosts a very different assortment of taxa (Fig. 5.1A). In terms of numbers and diversity among the sub-clades the Ascomycota have the most members, particularly

Leotiomycetes, , , and . For individual orders,

Malasseziales, , Monoblepharidales, Neocallimastigales, and Rozella have the fewest chitinases per genome (1 or 0.5) and Auriculariales, Basidiobolales, Geastrales, and Xylariales have the most (>17). B sub-clades-I, -II, and -IV are almost entirely composed of Ascomycota.

The B-IV sub-clade consists entirely of Saccharomycetales, while the B-I and B-II sub-clades are made of non-Saccharomycetales, Ascomycota. B-I consists of , ,

Dothideomycetes, Eurotiomycetes, Leotiomycetes, Sordariomycetes, and Xylonomycetes, with multiple orders of Dothideomycetes, Eurotiomycetes and Sordariomycetes represented. B-II contains multiple orders of Eurotiomycetes, Dothideomycetes, Leotiomycetes, and

Sordariomycetes. Sub-clade B-V also contains a large number of Basidiomycota, however several members from Ascomycota and a single early diverging fungus (Basidiobolus meristosporus) are also represented. B-III is almost entirely Basidiomycota with the exception of sequences from the early diverging Mucoromycetes. Variation in taxa is also seen in the A sub- clades with A-II and A-IV containing mostly Ascomycota with isolated Basidiomycota

(Tremellales in A-II and Pucciniales in A-IV). The A-II sub-clade consists of Eurotiomycetes,

Sordariomycetes, Leotiomycetes, and Dothideomycetes although within the Dothideomycetes the order Capnodiales and the were not represented. For the

A-IV sub-clade only the Dothideales, Orbiliales, and Lecanorales orders did not have members represented. A-III is composed of both Basidiomycota and early diverging fungal chitinases.

111

Malasseziales was the only Basidiomycota with no A-III chitinase while in the early diverging fungi Glomerales, Blastocladiales, and Rozella also lacked members. The A-V sub-clade contains Ascomycota, Basidiomycota and early diverging fungi. All Ascomycota were represented in this clade and all Basidiomycota were represented. For the early diverging fungi only Rozella, , Glomerales, and Basidiobolales have A-V chitinases. The C-I sub- clade mostly contains Ascomycota with a small assortment of sequences from 6 orders of

Agaricomycetes while the C-II sub-clade mostly contains both Ascomycota and limited early diverging fungi (Kickxellales, , and Blastocladiales).

5.3.3 Phylogeny of GH18

Much of the chitinase phylogeny is consistent with vertical inheritance. More derived branches of the chitinase phylogeny frequently track the species phylogeny, particularly in the A sub- clades (A-V, A-IV, A-III, and A-II) and the B-V sub-clade. The B-V sub-clade contains distinct

Agaricomycetes, Eurotiomycetes, and Dothideomycetes groups, and a single species of

Leotiomycetes within a Sordariomycetes group. The A-V sub-clade contains a distinct

Saccharomycetes clade in addition to multiple Agaricomycetes, Sordariomycetes, and

Eurotiomycetes clades. The A-IV sub-clade contains distinct clades containing Sordariomycetes,

Eurotiomycetes, or Dothideomycetes. The A-III sub-clade contains a clade of Sordariomycetes within a clade otherwise composed of Agaricomycetes. The A-II sub-clade has large separate

Sordariomycetes and Dothideomycetes groups. Despite grouping of related species, the topology of GH-18 phylogenies is complex, suggesting domain and gene duplications are common, and specific instances of horizontal gene transfer are supported (Supplemental File 5). Some sub- clades in particular (e.g. the C sub-clades) show few taxa-specific clades.

112

Two poorly-placed clades of mainly fungi were re-analyzed using NCBI as a target database, which more confidently placed these sequences among bacteria. One of these genes, a non-specific B clade protein also containing a signal secretion peptide (e.g. XP_006673222.1

Cordyceps_militaris, Supplemental File 5) appears to have been acquired from bacteria of an unknown lineage to specifically insect-pathogenic Hypocreales fungi. A much larger group of

Hypocreales and some other possess a chitinase supported to have been transferred from Actinobacteria to Hypocreales (e.g. XP_006670951.1 Cordyceps militaris,

Supplemental File 5). Another group of insect-associated Hypocreales chitinases from the A clade (close to but not included in the A-IIIs; e.g. KOM21536.1 Ophiocordyceps unilateralis), is dispersed in a clade of early diverging fungal chitinases that is expanded among the arthropod- associated (Supplemental File 5). Outside Hypocreales, the arthropod associated Basidiobolus meristosporus (Basme2_176417; B-V sub-clade) appears to have acquired a chitinase from Agaricomycetes that is most similar to sequences in Auriculariales

(Supplemental File 5). Another inter-phylum transfer appears to have occurred from

Pezizomycotina to Panaeolus (Pancy2_12872, Agaricales), and the nearest sequence in

Uncinocarpus reesei (Uncre1_6593) associates this event with the dung decay niche

(Supplemental File 5).

113

5.3.4 Evolution of domain architecture

Multiple GH-18 domains were rare with only 11 proteins containing multiple GH18 domains (all contain 2 GH-18 domains except Psicy2_12260 which contained 3). These were all in members of Basidiomycota and most were Agaricales (9) with single proteins from Polyporales and

Pucciniales. They were in the A-III (6), B-III (3) and B-V (2) sub-clades, and this did not appear linked to individual taxa. Interestingly only 5 of these GH18 domains in proteins with multiple

GH18s were predicted to be active (all in the B sub-clades).

LysM domains occur in 4 subclades, but are primarily restricted to the C-II sub-clade (Table

5.4). C-II sub-clade members also include the previously described LysM containing class III

Hce2 (Homologs of C. fulvum Ecp2) effector proteins (Stergiopoulos et al. 2012). The A-V sub- clade includes a LysM domain monophyletic group of most remaining LysM-containing chitinases, and this LysM domain appears to have been independently acquired from a bacterial spore assembly protein (Fig. 5.2 and Fig. 5.3) from the GH18 domain. Almost all other LysM- containing chitinases, which include the C-I, remaining A-V (Talma12_7383) and B-I (except

Spoth2_113450), share a different recent LysM common ancestor. There is evidence of additional LysM duplications of various ages that have resulted in phylogenetic diversity among

LysM domains in the same protein (Fig. 5.2).

114

Figure 5.2 LYSM IQ Tree.

Phylogenetic analysis of all LYSM domains with A-V sub-clade monophyletic group indicated in red.

115

LysM peptidoglycan-binding domain-containing protein [Clostridium sp. W14A] SafA/ExsA family spore coat assembly protein [Firmicutes bacterium] SafA/ExsA family spore coat assembly protein [ humiferrea] SafA/ExsA family spore coat assembly protein [Moorella mulderi] MULTISPECIES: SafA/ExsA family spore coat assembly protein [unclassified Moorella (in: Bacteria)] TPA: peptidoglycan-binding protein [Firmicutes bacterium] TPA: peptidoglycan-binding protein LysM [Firmicutes bacterium] MULTISPECIES: SafA/ExsA family spore coat assembly protein [unclassified Thermoanaerobacter] SafA/ExsA family spore coat assembly protein [Thermoanaerobacter thermohydrosulfuricus] SafA/ExsA family spore coat assembly protein [Thermoanaerobacter thermohydrosulfuricus] SafA/ExsA family spore coat assembly protein [Thermoanaerobacter kivui] MULTISPECIES: SafA/ExsA family spore coat assembly protein [Thermoanaerobacter] SafA/ExsA family spore coat assembly protein [Thermoanaerobacter ethanolicus] spore coat assembly protein SafA [Thermoanaerobacter thermocopriae] TPA: peptidoglycan-binding protein [Thermoanaerobacter sp.] SafA/ExsA family spore coat assembly protein [Thermoanaerobacter wiegelii] MULTISPECIES: SafA/ExsA family spore coat assembly protein [Thermoanaerobacter] SafA/ExsA family spore coat assembly protein [Thermoanaerobacter italicus] SafA/ExsA family spore coat assembly protein [Caldanaerobius polysaccharolyticus] SafA/ExsA family spore coat assembly protein [Caldanaerobius fijiensis] TPA: hypothetical protein [Ruminococcaceae bacterium] hypothetical protein BGN88_08100 [Clostridiales bacterium 43-6] hypothetical protein CVU91_10990 [Firmicutes bacterium HGW-Firmicutes-16] SafA/ExsA family spore coat assembly protein [Bacillus aquimaris] SafA/ExsA family spore coat assembly protein [Bacillus taxi] SafA/ExsA family spore coat assembly protein [Bacillus circulans] SafA/ExsA family spore coat assembly protein [Bacillus pumilus] SafA/ExsA family spore coat assembly protein [Papillibacter cinnamivorans] LysM peptidoglycan-binding domain-containing protein [Acetatifactor muris] LysM domain protein [Acetatifactor muris] LysM peptidoglycan-binding domain-containing protein [Tumebacillus permanentifrigoris] LysM peptidoglycan-binding domain-containing protein [Paenibacillus anaericanus] endochitinase 1 [Aspergillus steynii IBT 23096] chitinase [Aspergillus sclerotioniger CBS 115572] glycoside hydrolase family 18 protein [Aspergillus carbonarius ITEM 5010] endochitinase 1 precursor [Aspergillus ibericus CBS 121593] chitinase [Aspergillus piperis CBS 112811] endochitinase 1 [Aspergillus eucalypticola CBS 122712] glycoside hydrolase family 18 protein [Aspergillus luchuensis CBS 106.47] chitinase [Aspergillus niger] chitinase [Aspergillus costaricaensis CBS 115574] chitinase [Aspergillus luchuensis] chitinase [Aspergillus neoniger CBS 115656] endochitinase 1 [Aspergillus phoenicis ATCC 13157] endochitinase 1 [Aspergillus niger CBS 513.88] glycoside hydrolase superfamily [Aspergillus welwitschiae] endochitinase 1 [Aspergillus lacticoffeatus CBS 101883] endochitinase 1 [Aspergillus awamori] hypothetical protein ASPBRDRAFT_127465 [Aspergillus brasiliensis CBS 101740] TFIIH p62 subunit, N-terminal domain family protein [Aspergillus niger] endochitinase 1 [Aspergillus sclerotiicarbonarius CBS 121057] endochitinase 1 [Aspergillus heteromorphus CBS 117.55] chitinase [Aspergillus ellipticus CBS 707.79] endochitinase [Aspergillus japonicus CBS 114.51] hypothetical protein ASPACDRAFT_22298 [Aspergillus aculeatus ATCC 16872] chitinase [Aspergillus aculeatinus CBS 121060] chitinase [Aspergillus fijiensis CBS 313.89] chitinase [Aspergillus brunneoviolaceus CBS 621.78] endochitinase 1 [Aspergillus saccharolyticus JOP 1030-1] hypothetical protein ATNIH1004_010616 [Aspergillus tanneri] SafA/ExsA family spore coat assembly protein [Ottowia sp. GY511] LysM domain-containing protein [Streptomyces sp. T1317-0309] MULTISPECIES: LysM peptidoglycan-binding domain-containing protein [Dyella] Phage-related lysozyme (muramidase), GH24 family [Dyella marensis] hypothetical protein A1O9_12367 [Exophiala aquamarina CBS 119918] endochitinase 1 [ H538.4] chitinase 6 [Coccidioides immitis RS] chitinase 1 [Coccidioides immitis RMSCC 2394] endochitinase 1 [ reesii 1704] hypothetical protein AJ80_07171 [Polytolypa hystricis UAMH7299] hypothetical protein AJ80_09553 [Polytolypa hystricis UAMH7299] LysM peptidoglycan-binding domain-containing protein [Tumebacillus avium] SafA/ExsA family spore coat assembly protein [Tumebacillus sp. BK434] LysM peptidoglycan-binding domain-containing protein [Acidibacillus ferrooxidans] hypothetical protein AYJ22_15055 [Acidibacillus ferrooxidans] endochitinase 1 [ canis CBS 113480] chitinase 1 [Nannizzia gypsea CBS 118893] hypothetical protein H113_05964 [ rubrum MR1459] hypothetical protein ARB_04752 [Trichophyton benhamiae CBS 112371] endochitinase [Trichophyton equinum CBS 127.97] endochitinase [ CBS 118892] endochitinase [Trichophyton violaceum] hypothetical protein H107_06057 [Trichophyton rubrum CBS 202.88] endochitinase [ CBS 112818] hypothetical protein H101_07620 [Trichophyton interdigitale H6] endochitinase [Trichophyton mentagrophytes] TPA: putative Class V chitinase [Trichophyton benhamiae CBS 112371] class V chitinase, putative [ HKI 0517] chitinase [Emmonsia crescens] hypothetical protein ACJ72_08582 [Emmonsia sp. CAC-2015a] hypothetical protein AJ79_10228 [Helicocarpus griseus UAMH5409] hypothetical protein AJ78_06737 [Emergomyces pasteurianus Ep9510] hypothetical protein ACJ73_09978 [Blastomyces percursus] chitinase [ ATCC 26199] chitinase [Blastomyces dermatitidis ATCC 18188] chitinase [Blastomyces gilchristii SLH14081] chitinase [Blastomyces parvus] endochitinase 1 precursor [Histoplasma capsulatum NAm1] chitinase [Histoplasma capsulatum H143] 0.1 Hiscah88_9646 chitinase [Histoplasma capsulatum G186AR]

Figure 5.3 Phylogenetic analysis of Hisch88_09649.

Phylogenetic analysis the closest 1000 most closely related sequences identified by NCBI to Hisch88_09649 (highlighted in yeallow). Hisch88_09649 is used as a representative member of the A-V clade LYSMs. Fungal strains are indicated with green dots while bacteria are indicated with tan dots.

116

Chitin binding domains are spread across early diverging fungi, Ascomycota and Basidiomycota.

In early diverging fungi, Zoopagomycota have relatively large numbers of CBDs particularly in

Basidiobolomycetes, and to a lesser degree and Kickxellomycotina. In

Ascomycota CBDs are widely dispersed through the classes and orders. In Basidiomycota CBDs are mostly restricted to Agaricomycetes. The earliest diverging B sub-clade, B-V has a low percentage of chitinases with CBDs (Table 5.4), while there are more of these domains in other sub-clades particularly the B-III clade. In the AC superclade, there is a low percentage of A-Vs with CBDs, while the most are in C-I and C-II (Table 5.4).

117

Table 5.1 Clade distribution of Chitin Binding Domains and Enzymatically Characterized Members Clade CBD LysM Characterized Members containing Gene Organism Activity Reference (Total members proteins in clade) (% in clade) A-II 0 1 (not in Cful Aspergillus niger Exochitinase (van Munster et al. 2012) clade H. capsulatum (181) itself) Cts5 exo/ (this study) endochitinase H. capsulatum chitobiosidase Cts8 Exochitinase (this study) A-III (470) 0 0 A-IV (181) 0 0 Cts7 H. capsulatum no activity (this study) A-V (757) 115 23 ChiB1 Aspergillus Chitobiosidase (Jaques et al. 2003) (15.21%) fumigatus

Ech42/Chit42 endochitinase (Haran et al. 1995) Trichoderma chitobiosidase harzianum CiX1/Cts1 endochitinase (Fukamizo et al. 2001) Cts3 Coccioidies immitis endochitinase (this study) chitobiosidase Cts2 H. capsulatum endochitinase (this study) chitobiosidase Cts4 H. capsulatum exo/endo and (this study) chitobiosidase H. capsulatum

118

C-I (563) 417 6 Cts6 H. capsulatum exo/ (this study) (77.94%) endochitinase chitobiosidase C-II (297) 273 238 (92.23%) B-I (253) 45 (17.86%) 5 ChiA1 Aspergillus Endochitinase (Rush et al. 2010) fumigatus

Cts1 H. capsulatum Endochitinase (this study)

Ech30 Trichoderma Endochitinase (Hoell et al. 2005) atroviride B-II (160) 54 (33.96%) 0 Chit33 Trichoderma endochitinase (Boer et al. 2007; Haran et harzianum chitobiosidase al. 1995) B-III (208) 185 (89.37%) 0 B-IV (56) 23 1 (not in CTS1 Sachromyces Endochitinase (Hurtado-Guerrero and (41.82%) clade cerevisiae van Aalten 2007; Kuranda itself) and Robbins 1991)

(Selvaggini et al. 2004) CHIT2&3 Candidia Endochitinases albicans (Colussi, Specht, and KlCts1p Endochitinase Taron 2005) Kluyveromyces lactis B-V (442) 12 (2.72%) 0 “B” Bbchit1 Beauveria Endochitinase (Fang et al. 2005) bassiana

119

5.3.5 H. capsulatum chitinases

5.3.5.1 Diversification and taxonomic distribution

The H. capsulatum genome encodes eight chitinase (Cts) enzymes (Fig. 5.4), which are widely distributed in the GH-18 phylogeny and are focused on in this study. Six Histoplasma genomes were used for analysis (H. capsulatum G186AR, H. capsulatum G217B, H. capsulatum H143, H. capsulatum H88, H. capsulatum NAM1, and H. capsulatum TMU). While the genomes generally contain the same chitinases, there are slight differences in clade distribution. H. capsulatum H143 and H88 do not contain a C1 chitinase while G217B has two that are 100% identical (may be a result of assembly error). H. capsulatum TMU lacks the conserved A-IV chitinase, while H. capsulatum H143 is also missing the otherwise conserved B-I chitinase. Cts1 is the only H. capsulatum sequence in the B-I clade. Cts6 is the sole C clade protein (CI) and also had the larger size characteristic of this sub-clade. The remaining Histoplasma chitinases

(Cts2-5, Cts7-8) are all members of clade A. Cts2, Cts3, and Cts4 are all members of A-V, Cts5 and Cts8 both belong to A-II, and Cts7 is a member of A-IV. Cts4 appears to be a very recent duplication of Cts2 as the closely related fungus Blastomyces dermatitidis has only a single Cts protein that is orthologous to both H. capsulatum’s Cts2 and Cts4 proteins. There are no A-III chitinases in H. capsulatum.

Histoplasma is the only in that contains chitinases in the A-II sub-clade, while other Onygenales tend to contain B-II sub-clade members, which Histoplasma lacks. The C

120 clade (C-I and –II) is expanded in multiple other Onygenales members. This is most obvious in

Microsporum canis with 9 C-I and 4 C-II members. However, upon closer examination, only 3 of each sub-clade are predicted to be chitin degrading as they lack any of the active site residues.

Among Eurotiomycetes, Onygenales and have similar distributions of chitinases among the sub-clades except as noted above. Chaetothyriales differ in that they contain A-III sub-clade chitinases but lack the B-II and C clade chitinases found in the other Eurotiomycetes.

121

Figure 5.4 Diagram of Histoplasma capsulatum chitinase genes.

Chitinase name, gene name and clade are indicated (left) along with domain structure indicated. Active sites and indicated with a blue star. Sub-clades are indicated (A is magenta, B is Cyan, and C is green).

5.3.5.2 Domain architecture and finer motifs

The architectural diversity of Histoplasma chitinases is consistent with the variability observed across the fungal . In particular, Histoplasma chitinases look similar to those of other

Onygenales. For example other Onygenales have Cts3-like proteins with matching architecture

(contain LysM but no signal peptide). Most Onygenales only have 1-2 CBDs in C clade proteins. 122

Microsporum canis has greatly expanded numbers of CBDs and LysM domains corresponding to their expanded C clade, which is unusual for the Onygenales chitinase architecture. At the individual protein level, domain architecture is variable in the A clade. Cts2-4 are all in the A-V sub-clade. Of these, Cts2 and 4 have similar domain architecture consistent with a recent duplication, with a secretion signal at the N-terminus and serine/threonine rich regions, which are common sites for O-linked glycosylation (Loibl and Strahl 2013). Cts3 contrasts with Cts2 and 4 by lacking the secretion signal and containing the only LysM domain in Histoplasma. In sub-clade A-IV, Cts7 lacks a secretion signal and contains a serine/threonine rich region. This is also the only Histoplasma chitinase that lacks conserved aspartate residues predicted to comprise the active site (Hartl et al. 2012). In sub-clade A-II while Cts5 and Cts8 both contain a secretion signal, only Cts5 also contains a CBD. The C clade chitinase, Cts6 (C-I) also contains the secretion signal at the N-terminus and serine/threonine rich regions. It also contains two CBDs.

The B clade chitinase in Histoplasma, Cts1 (B-I), has a GPI anchor region in addition to a secretion signal suggesting this enzyme will be anchored to the cell surface of H. capsulatum cells. It also contains a serine/threonine rich region.

5.3.5.3 Expression of H. capsulatum chitinases

To provide insight into the physiological roles of diverse chitinases in Histoplasma, the expression of each was determined under different environmental and nutritional conditions.

Most chitinase-encoding genes (CTS) were expressed at constant, but low levels across most conditions (Fig. 5.5). Surprisingly for a thermally-controlled dimorphic fungus, only CTS3 (A-

V) was specifically up regulated in H. capsulatum filamentous cells (11 fold higher in mycelia compared to yeasts; Fig. 5.5). CTS2 (A-V) was expressed at higher levels overall indicating that

123

Cts2 could be a general growth chitinase or the major functional chitinases under the tested conditions. CTS4 (A-V) and CTS8 (A-II) were expressed at low but consistent levels. CTS1 (B-

I) and CTS6 (C-I) had low, but highly variable levels of expression. CTS7 (A-IV), which lacks the active site D-X-X-D-X-D-X-E residues, and CTS5 (A-II) was at an undetectable level under every condition tested. Thus, with the exception of CTS3, expression studies did not reveal specific environmental conditions for H. capsulatum chitinase expression or a yeast-phase specific chitinase.

124

Figure 5.5 Histoplasma capsulatum chitinase expression.

Expression of CTS genes based on qRT-PCR. Expression was calculated relative to the constitutive expression of the ACT1 and RPS15 genes. H. capsulatum yeast (magenta) or mycelia (blue) were grown in HMM or SDA with or without colloidal chitin (1.2%). Transcript levels were quantified by qRT-PCR and calculated relative to ACT1 and RPS15 transcript levels. Data represent the average expression ± standard deviation of results from biological replicates (n = 3). Sub-clades are indicated (A is magenta, B is Cyan, and C is green).

125

5.3.5.4 Enzymatic activities of H. capsulatum chitinases

To determine if the different clades represented by the H. capsulatum chitinases correspond to different enzyme activities, all eight H. capsulatum chitinases were purified and tested for chitin degradation profiles. Three artificial substrate mimics were used to determine the specificity of each purified chitinase enzyme: exochitinase, endochitinase, and chitobiosidase (hydrolysis of chitobiose, a deacetylated glucosamine dimer to approximate chitosan) activity. The B-clade chitinase Cts1 exhibited endochitinase activity specifically (Fig. 5.6), consistent with the activity suggested by the

B-clade enzyme structure (Hartl, Zach, and Seidl-Seiboth 2012). In contrast, the A-II clade protein, Cts8, only had exochitinase activity (Fig. 5.6). The other A-II clade protein, Cts5 had barely detectable activity on any substrate despite equivalent amounts of protein used. Cts2, Cts3, and Cts4, all closely related proteins from the A-IV/V group exhibited good endochitinase and chitobiosidase activity (Fig. 5.6). Cts4 also showed exochitinase activity suggesting neofunctionalization after duplication from Cts2. Cts7, the member of the A-IV group lacked any detectable activity, consistent with the lack of active site residues in the Cts7 protein (Fig. 5.6). Cts6, H. capsulatum’s only C-subclade protein showed good activities for all three specificities, especially endochitinase activity

(Fig. 5.6).

126

Figure 5.6 Histoplasma capsulatum chitinases enzymatic activity.

Chitinase enzymatic activity rates were calculated as nanograms 4-methylumbelliferone released per minute per nanogram enzyme. Chitobiosidase activity (blue), exochitinase activity (black) and endochitinase activity (red) is differentiated. Data represent the average expression ± standard deviation of results from biological replicates (n = 3). Bars indicate a p value of >0.05. Sub-clades are indicated (A is magenta, B is Cyan, and C is green).

127

5.4 Discussion

5.4.1 Previous chitinase ontologies are largely robust to increased sampling, but the A clade may be polyphyletic.

Our analyses identify the existence of two major classes of chitinases defined by an ancient divergence between B and AC seen in previous phylogenetic analyses. We also recapitulate the previous sub-clades with varying degrees of support (Table 5.3).

However, we find that there is not a single A clade, because C clade chitinases are consistently more closely related to A-II chitinases. The divergence of the A-IIIs and the close clustering of the A-II sub-clade with the C clade suggest the definition of A and C clades could use revision. Since A chitinases are not monophyletic, either C chitinases should be subsumed into a larger A clade, or alternately, the As could be revisioned as multiple new clades informed by gene architecture and history. Previous work has supported both an independent C clade (Alcazar-Fuoli et al. 2011; Karlsson and Stenlid

2008; Seidl et al. 2005) and a C + A-II clade (Karlsson and Stenlid 2009) that is consistent with our analyses. However, only the present analysis strongly supports the divergence of the A-III clade from the rest of the A sub-clades and the Cs, perhaps because the limited sample of A-III chitinases in previous analyses (< 10 members) did not reflect overall A-III diversity. Our analysis suggests additional fungal chitinases should be characterized in order to accurately describe functional diversity among chitinases. Additionally the location of the LysM domain has been used to distinguish the

128

C clade from the A clade (Gruber, Vaaje-Kolstad, et al. 2011; Gruber and Seidl-Seiboth

2012); however, in our analyses, LysMs are restricted to the CII sub-clade and some A-V chitinases, including the Histoplasma Cts3 (Table 5.4). Additionally, the A clade was generally thought to be lacking in CBDs (Hartl et al. 2012), however CBDs are widespread in the A-II sub-clade (Table 5.4), further supporting the need to redefine the

A and C clades.

5.4.2 Some chitinase sub-clades have specific taxonomical biases with potential ecological and evolutionary ramifications.

The B vs AC split appears to be an early divergence. For the Bs (Class III) chitinases, the

B-V sub-clade is most likely closest to the ancestral with members found in early diverging fungi, Basidiomycota, and Ascomycota. In other analyses, B-Vs are the first to diverge after the fungal Bs split from bacterial chitinase (Karlsson and Stenlid 2009). B-I,

-II, and -IV contain mainly Ascomycota and are likely taxa-specific B chitinases. The B-

IV sub-clade is strictly limited to the Saccharomycetales further strengthing the ides this is a taxa-specific chitinase. Looking at the family level does not reveal any additional information, as members are spread across but not entirely encompassing several families. The B-III sub-clade is lacking in Ascomycota and may represent the early diverging fungi and Basidiomycota taxa-specific B chitinases. Interestingly other studies report the B-III sub-clade associating with plant (class III) chitinases (Karlsson and

Stenlid 2009). For the AC (Class V) chitinases, A-IV and A-III are the most widespread

129 found in with members found in early diverging fungi, Basidiomycota, and Ascomycota.

Interestingly, the A-III sub-clade is greatly reduced in Ascomycota, as compared to early diverging fungi and Basidiomycota, suggesting lineage specific retention of this sub- clade. In other studies, the AIII- along with the A-II sub-clades cluster closely with an novel clade of bacterial (class V) chitinases called group V (Karlsson and Stenlid 2009).

This group contains members of flavobacteria and gammaproteobacteria. As flavobacteria are environmental bacteria and many members of gammaproteobacteria are environmental bacteria and include some plant pathogens, it is possible these chitinases have functional roles in saprotrophic nutrition or plant pathogenicity. In general, the

Ascomycota are the most expanded in different chitinase sub-clades particularly

Leotiomycetes, Sordariomycetes, Eurotiomycetes, and Dothideomycetes.

Agaricomycetes (Basidiomycota) have the highest diversity of chitinases, with members in multiple sub-clades. Early diverging fungal genomes are limited to a small number of chitinases. Interestingly even though Rozella allomycis lacks chitin (Jones et al. 2011), it contains a single A-V chitinase, which may play a role in its obligate endoparasitism of

Chytridiomycota or .

The clear taxonomical patterns for certain sub-classes of chitinases do indicate they may be phylogenetically restricted and potentially have specific functions. The clearest example of this is the B-IV sub-clade, which is a Saccharomycetales specific chitinase.

Interestingly for the B-IV sub-clade, there are several characterized members and this clade is relatively well functionally studied (Table 5.4). However, most of these functions

130 relate to cell division and morphology (Colussi, Specht, and Taron 2005; Hurtado-

Guerrero and van Aalten 2007; Kuranda and Robbins 1991; Selvaggini et al. 2004). This is interesting as cell division and morphology are rather fundamentally conserved actives in fungi, making it unusual that this function appears to be done by members in a novel

Saccharomycetales-specific sub-clade. Thus the question arises: are the B clade chitinases all relating to cell division and morphology with the B-IVs the

Saccharomycetales representative, or is this neofunctionalization in the

Saccharomycetales. Other B sub-clade members are not functionally well-characterized making inferences difficult. The other Ascomycota-specific sub-clades A-III and B-III are not as well-studied and are lacking characterized members making additional insight into how the diversification into the Ascomycota chitinase diversification difficult.

Unfortunately the ecology of fungi at the level of resolution we see for the clades is so multifactorial that strict ecological rationales for these differences is difficult to hypothesize. Basidiobolales and Auriculariales have the highest number of chitinases per genome with over 20 each, indicating a major ecological function of chitinases is saphrotrophism. In agreement with this we see the fewest chitinases (≤1) in fungi that are not common in the environment; for example, endoparasitic Rozella, symbiontic

Neocallimastigomycota (digestive tracts in larger herbivores) and Glomerales

(mycorrhizal), and the skin fungi . In addition we see some HGT events that indicate specific ecological roles of chitinases. Although it was not feasible to perform a comprehensive analysis of molecular evolution events among fungal chitinases in this

131 dataset, there is evidence for a significant role of horizontal gene transfer, and the suggestion of associated ecological roles of transferred genes (Supplemental File 5). For example, insect-pathogenic and associated fungi were seen to have HGT events and expansions. As insect exoskeletons are a major source of chitin (Tharanathan and Kittur

2003), expansion and acquisition of chitinases would provide an evolutionary advantage to fungi that feed on them, either through parasitism or saprotrophism.

5.4.3 Chitinases containing LysM domains are distributed among plant pathogenic, insect parasitic and saprotrophic fungi

LsyM domains have been shown to bind chitin and are thought to do the same in fungi although carbohydrate-binding specificities, are not well-described and their role in saprotrophic fungi is not well understood (Gruber et al. 2011).In our analysis the saprotrophic and have increased numbers of

LysM domains supporting their role in saprotrophic nutrition. The largest amount of

LysM domains per genome in our analysis is seen in the . This family contains many strophic members again supporting a role of LysM domains in saprotrophic nutrition. In addition, many species in this family include parasite of plants, insects, or amphibians, indicating an additional role for them in parasitizing organisms with large chitin content. This pattern of increased LysM domains in parasitic and strophic families is also seen in , Eurotiomycetes, ,

Ustilaginomycetes, Mucoromycotina, and . LysM domains have also

132 been shown to work with plant pathogens to bind chitin to prevent host plant immunity activation (Bolton et al. 2008; de Jonge and Thomma 2009). In our analysis we see heavy increases of the LysMs and corresponding C-II clade in some classes that generally contain plant pathogens; Microbotryomycetes, , Dothideomycetes, and

Leotiomycetes. This supports a role of LysM containing chitinases in adaption to plant pathogenesis.

5.4.4 Histoplasma capsulatum chitinase expression differences suggest specific functional roles

H. capsulatum expression data allowed for an initial look at how multiple chitinases could potentially play different roles. Cts3 and Cts1 are most likely morphological in function. Cts1 is secreted and contains a GPI-anchor indicating that it likely localizes to the cell wall. This makes it likely it has a roll in cell wall remodeling or modification, such as in cell division or morphological changes. Cts1 had a low but generally consistent level of expression. As transcriptional data was not collected on synchronized stages of division or morphology it is possible the low expression is a result of a mix of cells in high and low expression at different stages of replication. More investigation into Cts1 is required. Cts3 was the only chitinase that was induced in the hyphal form. As Cts3 is not secreted, it is likely that this chitinases plays a role in cell division or septa formation the mycelial form. It was surprising that more chitinases in Histoplasma did not appear to be

133 regulated by the phase transition and that no chitinases were specifically induced in the yeast phase. For example, chitinases are expanded in those fungi with filamentous growth and are few in yeast-growing fungi like Saccharomyces cerevisiae with the hypothesis being that hyphal growth requires multiple different chitinases (Karlsson and Stenlid

2008). H. capsulatum has a very strict dimorphic regulation (Edwards et al. 2013) controlled by the RYP transcription factors (Webster and Sil 2008) and other common fungal transcription factors, such as the APSES family, have very limited function

(Longo et al. 2018). Therefore any chitinases necessary for hyphal growth should be upregulated in the mycelia stage. The lack of phase specific transition, with all but CTS3 showing roughly the same expression in the yeast and mycelia states, indicates that if the expansion of chitinases in H. capsulatum is related to the hyphal form they must play more specific roles than general growth that were not identified under these conditions.

Cts2 and Cts4 appear to be a recent duplication event and appear to have functionally diverged. CTS2 has consistently higher expression (e.g. up to 21 fold higher in the yeast

SDA CHIT condition) than CTS4. This indicates that under the conditions tested Cts2 may be the general functional chitinase while Cts4 is either simply less used or used under specific unknown conditions. As the only Onygenales members containing an A-II member, the Cts5 and Cts8 genes are of particular interest. However, in our current transcriptional profile we were not able to determine a condition under which they were induced. They may have highly specific condition for induction that have yet to be identified. CTS7’s low to no detectable transcription under the conditions tested. In conjunction with the lack of functional activity, Cts7 may no longer serve a role in

134

Histoplasma. CTS6 expression was too variable to determine what specifically was controlling expression, however it does appear to be under some sort of regulation.

5.4.5 Histoplasma capsulatum chitinase activities do not follow phylogenetically predicable trends in specificity

Chitinase activity of the eight Histoplasma chitinases highlighted the differential enzymatic activity of multiple chitinases and how that variation applies to clade distinctions (Table 5.4). We again see the divergence of function with the duplication event of Cts2 and Cts4. Cts2 has endo- and chitobiosidase activity while Cts4 has some exochitinase activity in addition to endo- and chitobiosidase activity. Cts3 also has endo- and chitobiosidase activity. This is interesting as the three A-V clade members all show strong endo- and chitobiosidase activity. When compared to some other characterized A-

V clade members (Table 5.4) we see a similar pattern. This is not consistent with the prediction that A clade members should have exochitinase activity (Duo-Chuan 2006;

Hartl, Zach, and Seidl-Seiboth 2012; Seidl 2008a). However, one of our A-II members,

Cts8 does have specific exochitinase activity as do other described members (Table 5.4).

The A sub-clades may have functional divergence, which be reflected in the potential polyphyletic phylogony of the As. The other A-II Histoplasma chitinase, Cts5, has very minimal activity (no one substrate is statically preferred over another), which along with the lack of expression suggests that this chitinase may no longer have a functional role.

Alternatively, it may have a highly specialized role and may be optimized for activity

135 under specified conditions other than what we tested. For example, Cts1 in

Saccharomyces cerevisiae shows an unusually low pH preference for activity (Hurtado-

Guerrero and van Aalten 2007). Notably, the A-III is lacking in enzymatically- characterized members (Table 5.4). As the A sub-clades are more variable in enzymatic activity than expected and this sub-clade is so divergent from the other As, these members should be further studied to get any understanding of their activity. This study is also the first report of a characterized C clade member, Cts6, which has both strong endo- and exochitinase activity (in addition to some chitobiosidase activity) arguing against the prediction that chitinases in the C subgroup are only exochitinases. The B clade functional data is much more consistent with predictions, with characterized members of multiple clades having endochitinase activity (Table 5.4) as predicted (Duo-

Chuan 2006; Hartl, Zach, and Seidl-Seiboth 2012; Seidl 2008a). However, sub-clades B-

III and B-V have not been extensively investigated.

5.4.6 Evolution of the Histoplasma chitinases indicates a degree of differentiation and multiplicity that is reflected in fungal chitinases in general reflecting a potential division of work for chitinases

There is variation in chitinase composition among Histoplasma isolates. H. capsulatum

H143 and H88 were missing CI-clade members indicating this is not ancestral for

Histoplasma., while H. capsulatum H143 is also missing the conserved B-I sub-clade member. In addition, H. capsulatum TMU lacks the conserved A-IV member, which is

136 not functional in H. capsulatum G217B or predicted to be functional in several

Onygenales. H. capsulatum TMU may have lost this chitinase. All contain the Cts3 homolog supporting the idea that this is the conserved hyphal chitinase in Histoplasma.

One of the most interesting features of Histoplasma chitinase evolution is the presence of two functional chitinases in the A-II sub-clade. A-II sub-clade members are not widespread in fungal . They are almost exclusively found in the

(sub-division of the filamentous Pezizomycotina). Members are found in Eurotiomycetes,

Sordariomycetes Dothideomycetes and Leotiomycetes. In the Eurotiomycetes Eurotiales,

Onygenales, and Chaetothyriales contain A-II sub-clade members. However,

Histoplasma is the only Onygenales to contain this sub-clade. That indicates these chitinases are ancestral to the Leotiomyceta, however they have been largely lost in the

Onygenales with the exception of Histoplasma. Therefore these may have a specific function that is still necessary in Histoplasma, although what it could be is still unknown.

Cts1 was the B-I clade member. Cts6 is the sole C clade protein (CI) and also had the larger size characteristic of this subgroup. The remaining Histoplasma chitinases are all members of clade A. Cts5 and Cts8 both belong to the A-II subclade. Cts7 is a member of the A-IV sub-clade. Cts2, Cts3, and Cts4 are all members of the A-V group, the location of most of the chitinase genes in Histoplasma. Cts4 appears to be a very recent duplication of Cts2 as the closely related fungus Blastomyces dermatitidis has only a single Cts protein that is orthologous to both H. capsulatum’s Cts2 and Cts4 proteins. H.

137 capsulatum lacks an A-III chitinase. Placing Histoplasma in context of the other

Onygenales analyzed, Histoplasma was the only genus that contained A-II sub-clade members. The other Onygenales tended to contain B-II sub-clade members that

Histoplasma lacks. In addition the C clade (C-I and –II) are expanded in multiple other

Onygenales members. This is most obvious in Microsporum canis with 9 C-I and 4 C-II members. However, upon closer examination, only 3 of each sub-clade are predicted to be active (contain the active site residues) indicating that this larger clade representation may not be as functionally extreme as otherwise indicated. Looking at the Histoplasma chitinases in terms of other members of Eurotiomycetes, the Onygenales and Eurituales show matching patterns of chitinases distribution in the sub-clades. Chaetothyriales differ in that they contain A-III sub-clade chitinases members but are lacking B-II or C clade members seen in the other Eurotiomycetes. When comparing the Eurotiomycetes to other

Pezizomyotina, is rather divergent. It is the only one to contain a B-IV member, and is the only order to be missing a B-V or C-I members from the

Pezizomyotina. Saccharomycetes are limited to the B-IV, C-II, A-IV, and A-V sub- clades, which is less diversity seen in other Pezizomyotina. Sordariomycetes and

Eurotiomycetes had identical patterns of sub-clade distribution as A-III members are missing in all others. Leotiomyceta members are where A-II (excluding Xylonomycetes),

B-I, and C-II (excluding Xylonomycetes) sub-clade members are seen. The A-IV clade is missing in Orbiliomycetes while C-I is missing in Xylonomycetes and Pezizomycetes.

The A-V sub-clade is conserved in all Pezizomyotina.

138

Chapter 6 Conclusions

Antifungal options for Histoplasma are currently limited. This is largely due to the lack of fungal-specific targets. As a result, many novel antifungals in the drug development pipeline are repurposed drugs (Goughenour and Rappleye 2016). In addition, most studies do not take into account clinically relevant testing of fungal morphotypes and assume a one-size-fits-all approach to fungal drug testing, which leads to false leads and inaccurate frequencies of resistance. In chapter 2, we develop and standardize an appropriate method for Histoplasma antifungal susceptibility testing. We show that current CLSI testing methodology calls for too low an initial inoculum and a too-short incubation time to see proper growth of even uninhibited yeast cells. We also show that the antifungal susceptibility is often phase-dependent in Histoplasma. While amphotericin B susceptibility shows no phase dependence, yeast cells are 10 times more susceptible to fluconazole than mycelia. The most important clinically relevant example of this is that hyphal cells of Histoplasma are susceptible to the echinocandin,

Caspofungin (IC50 is 0.4 µg/mL) while the yeasts are not (IC50 is 10.4 µg/mL). This results in the echinocandins being unusable as a therapeutic option. It also highlights the importance of testing the clinically relevant phase (i.e. yeast) in early screening. In addition, we developed and optimized two colorimetric assays (MTT and resazurin) for 139 future high-throughput screens of compounds. Antifungal and more generally antimicrobial development is an area of great concern for human health. As antifungals in particular are limited and effective targets even more so, many see large scale high- throughput screens as promising ways to move the drug development pathway forward.

However, as we have shown, improper testing methodology can be more detrimental to the field than not testing at all. False positives waste resources and false negatives can lead to the abandonment of promising compounds, both of which are much more likely with faulty methodology. There is also a need to take into account the biology of the organism you are testing in these screens and tailor the assays to fit the organism tested.

This is particularly important in fungi where a relatively small number of species display a wide variety of biological differences.

In addition to a need for novel antifungals, there is a real need for novel, fungal-specific targets. As a result, target identification is an important stage in antifungal development, particularly for large-scale screens of compound libraries, where the target is completely unknown. In chapter 3, we work on identifying the target of 41F5, a novel antifungal compound previously identified by our lab (Edwards, Kemski, and Rappleye 2013). We use a two-pronged approach to target identification: a traditional genetics approach and a novel co-fractionation mass spectrometry approach. We generated 6 independent lines of

Cryptococcus neoformans, which are specifically resistant to 41F5. Whole genome sequencing of these 6 lines identified SNPs, which were prioritized into 11 potential target genes. We are currently working on validating these candidates through

140 overexpression studies. In addition, we developed a novel approach to target identification using an untagged size co-fractionation approach. Cells are grown with the drug, and the resulting lysate is size-fractionated. These fractions are then screened for the drug by LC-MS/MS and co-fractionating proteins are identified by proteomics. Using this method, we identified and prioritized 34 candidate targets. We have started verification of these targets using a similar overexpression system and have screened 5 candidates to date. So far none have conferred resistance, but we will continue to screen the remaining candidate targets in the future. In addition, the only common pathway between the two methods was the ergosterol biosynthesis pathway, however overexpression of Erg10 in Histoplasma did not confer resistance. This is not completely unexpected as Histoplasma and Cryptococcus are members of two different phyla of fungi (Ascomycota and Basidiomycota, respectively) and therefore may have different targets due to their being extremely evolutionarily divergent. To address this concern, if overexpressing the Cryptococcus homolog of the Histoplasma 41F5 target fails to confer resistance, the co-fractionation approach could be used to generate a proteomics list for

Cryptococcus 41F5 co-fractionating proteins to identify an independent target. This highlights a major strength of the new method. This method can be adapted for any organism, allowing for use in protein target identification in any kind of drug development. This is particularly important in drug development in fungi. There are large verities of medicinal chemistry and pharmacology techniques that can be used to improve compounds. However, the biological side is limiting in that without fungal targets to

141 inhibit we cannot expand our therapeutic options just attempt to keep up with the pathogens.

In addition, we wanted to investigate virulence factors that could serve as novel fungal- specific drug targets. Of particular focus for our lab is o-linked mannosylation as it confers thermotolerance to Histoplasma, which is required for virulence (Garfoot et al.

2018). In chapter 4, we attempt to identify which proteins are o-linked mannosylated and therefore may contribute to thermotolerance. We show that direct identification of o- linked mannosylation via mass spectrometry proteomics is not a viable approach. We subsequently take a bioinformatics approach and identify 51 prioritized genes to investigate. We have started screening these candidates using temperature sensitivity as a phenotype of knockdowns and have screened 7 genes. None of these have yet been temperature sensitive, however screening is ongoing. In addition, we have been developing a biotinylation protocol for o-linked mannosylated proteins to enable specific purification. While biotinylation has been successful, we are currently working on elution of proteins from the streptavidin resin. We are attempting a trypsin elution method to free all non-mannosylated peptides for identification by proteomics. In addition, we are investigating the possibility of utilizing an azido-N-Acetylgalactosamine (GalNAc) to incorporate a reactive azido group into glycosylation sites, then using alkyne-biotins to subsequently label the o-linked glycosylated proteins which can then be easily eluted via cleavage of the linker (Boyce et al. 2011; Dube et al. 2006). We are working to establish that the azido-sugar precursor molecules can enter Histoplasma and be incorporated. O-

142 linked glycosylation is a major posttranslational modification in fungi that has been shown to be required for virulence in C. albicans, C. neoformans, and H. capsulatum largely contributed to their role in the fungal cell wall (Leach and Brown 2012; Garfoot et al. 2018). However, there is little mechanistic analysis of what specific proteins are involved or any additional role they may play. For example, we see large amounts of glycosyl hydrolases in our bioinformatics O-linked glycosylation data set. We know that some of these glycosyl hydrolases (Eng1) are needed for cell wall remodeling (Garfoot et al. 2016). While most investigations into O-linked mannosylation in fungi focuses on structural and proteolytic protection (Lommel and Strahl 2009), O-linked glycosylation can have effects on enzyme activity (Goettig 2016). There is much work that needs to been done to define the mechanism of O-linked glycosylation related virulence and investigate how specific modified proteins are involved, and how modification impacts these proteins.

In the course of investigating o-mannosylated proteins, we noticed that several chitinases proteins are predicted to be o-linked glycosylated. Although they do not appear to be related to thermotolerance, we were interested in investigating them further as chitin is a unique structural component of the fungal cell wall which can be exploited for the development of fungal-specific drugs such as nikkomycin Z, which inhibits chitin synthesis (Endo and Misato 1969). As chitinases have not been comprehensively studied in fungi as a whole, in chapter 5 we decided to undertake a comprehensive phylogenetic survey of fungal chitinases in addition to characterization of enzymatic activities and

143 gene expression of Histoplasma chitinases. We determined that the previous phylogenetically determined A clade of chitinases was not monophyletic and that the A and C clades need to be recategorized. We identified clear taxonomical patterns in chitinase phylogeny. We saw increased LysM domain-containing chitinases in saprotrophic, insect parasitic, and plant pathogenic containing taxa, suggesting a role for this domain in those functions. In Histoplasma, we identify a hyphal phase-specific chitinase, Cts3. We also identify a recent gene duplication event in Histoplasma (Cts2 and Cts4) with differences in expression and enzymatic activity. Finally, we show that enzymatic activity of chitinases does not follow previous phylogenetic predictions, indicating that more chitinases need to be studied to identify their roles in individual fungi. This chapter highlights the need in the field to pair phylogenetic analyses with experimental characterization. The phylogenetic results offer a large-scale look at the diversity of chitinases in fungi. However, enzymatic characterizations and insight into specific functions of individual chitinases in fungi with multiple chitinases are limited to a small number of fungi that only account for a subset of the phylogenetic diversity we see. Using phylogenetic results to identify gaps in experimental data can provide new insight into our understanding of what we should be looking at.

144

Bibliography

de Aguiar Cordeiro, Rossana, Raimunda Sâmia Nogueira Brilhante, Marcos Fábio Gadelha Rocha, et al. 2006 In Vitro Inhibitory Effect of Antituberculosis Drugs on Clinical and Environmental Strains of . The Journal of Antimicrobial Chemotherapy 58(3): 575–579. de Aguiar Cordeiro, Rossana, Francisca Jakelyne de Farias Marques, Rebecca de Aguiar Cordeiro, et al. 2014 Synthesis and Antifungal Activity in Vitro of Isoniazid Derivatives against Histoplasma Capsulatum Var. Capsulatum. Antimicrobial Agents and Chemotherapy 58(5): 2504–2511.

Alcazar-Fuoli, Laura, Cécile Clavaud, Claude Lamarre, et al. 2011 Functional Analysis of the Fungal/Plant Class Chitinase Family in Aspergillus Fumigatus. Fungal Genetics and Biology 48(4): 418–429.

Almagro Armenteros, José Juan, Konstantinos D. Tsirigos, Casper Kaae Sønderby, et al. 2019 SignalP 5.0 Improves Signal Peptide Predictions Using Deep Neural Networks. Nature Biotechnology 37(4): 420–423.

Andreu, Carlos Manuel Fernández, Agner Martínez León, Yamilé Echemendía Medina, et al. 2003 [In Vitro Sensitivity of Histoplasma Capsulatum Var. Capsulatum to Amphotericin B, Ketoconazole, Itroconazole and Fluconazole]. Revista Cubana de Medicina Tropical 55(2): 76–82.

Ankobiah, W A, K Vaidya, S Powell, et al. 1990 Disseminated Histoplasmosis in AIDS. Clinicopathologic Features in Seven Patients from a Non-Endemic Area. State Journal of Medicine 90(5): 234– 238.

Bahr, Nathan C., Spinello Antinori, L. Joseph Wheat, and George A. Sarosi 2015 Histoplasmosis Infections Worldwide: Thinking Outside of the Valley. Current Tropical Medicine Reports 2(2): 70–80. 145

Bates, D W, L Su, D T Yu, et al. 2001 Mortality and Costs of Acute Renal Failure Associated with Amphotericin B Therapy. Clinical Infectious Diseases 32(5): 686–693.

Baxter, Bonnie K., Louis DiDone, Duana Ogu, Stanford Schor, and Damian J. Krysan 2011 Identification, in Vitro Activity and Mode of Action of Phosphoinositide- Dependent-1 Kinase Inhibitors as Antifungal Molecules. ACS Chemical Biology 6(5): 502–510.

Becher, François, Joseph Ciccolini, Diane-Charlotte Imbs, et al. 2017 A Simple and Rapid LC-MS/MS Method for Therapeutic Drug Monitoring of Cetuximab: A GPCO-UNICANCER Proof of Concept Study in Head-and-Neck Cancer Patients. Scientific Reports 7(1): 2714.

Bhabhra, R., and D. S. Askew 2005 Thermotolerance and Virulence of Aspergillus Fumigatus: Role of the Fungal Nucleolus. Medical 43 Suppl 1: S87-93.

Boer, Harry, Helena Simolin, Sylvain Cottaz, Hans Söderlund, and Anu Koivula 2007 Heterologous Expression and Site-Directed Mutagenesis Studies of Two Trichoderma Harzianum Chitinases, Chit33 and Chit42, in Escherichia Coli. Protein Expression and Purification 51(2): 216–226.

Boguslawski, George, and Dean A. Stetler 1979 Aspects of Physiology of Histoplasma Capsulatum (A Review). Mycopathologia 67(1): 17–24.

Bohse, Megan L., and Jon P. Woods 2007 Expression and Interstrain Variability of the YPS3 Gene of Histoplasma Capsulatum. Eukaryotic Cell 6(4): 609–615.

Bolton, Melvin D., H. Peter van Esse, Jack H. Vossen, et al. 2008 The Novel Cladosporium Fulvum Lysin Motif Effector Ecp6 Is a Virulence Factor with Orthologues in Other Fungal Species. Molecular Microbiology 69(1): 119– 136.

Borelli, C., M. Schaller, M. Niewerth, et al. 2008 Modes of Action of the New Arylguanidine Abafungin beyond Interference with Ergosterol Biosynthesis and in Vitro Activity against Medically Important Fungi. Chemotherapy 54(4): 245–259.

Boyce, Michael, Isaac S. Carrico, Anjali S. Ganguli, et al. 2011 Metabolic Cross-Talk Allows Labeling of O-Linked Beta-N-Acetylglucosamine- 146

Modified Proteins via the N-Acetylgalactosamine Salvage Pathway. Proceedings of the National Academy of Sciences of the United States of America 108(8): 3141–3146.

Brilhante, R. S. N., E. P. Caetano, J. J. C. Sidrim, et al. 2013 Ciprofloxacin Shows Synergism with Classical Antifungals against Histoplasma Capsulatum Var. Capsulatum and Coccidioides Posadasii. Mycoses 56(3): 397–401.

Brilhante, Raimunda Sâmia Nogueira, Érica Pacheco Caetano, Giovanna Barbosa Riello, et al. 2016 Antiretroviral Drugs Saquinavir and Ritonavir Reduce Inhibitory Concentration Values of Itraconazole against Histoplasma Capsulatum Strains in Vitro. The Brazilian Journal of Infectious Diseases 20(2): 155–159.

Brilhante, Raimunda Sâmia Nogueira, Maria Auxiliadora Bezerra Fechine, Rossana de Aguiar Cordeiro, et al. 2010 In Vitro Effect of Sulfamethoxazole-Trimethoprim against Histoplasma Capsulatum Var. Capsulatum. Antimicrobial Agents and Chemotherapy 54(9): 3978– 3979.

Brilhante, Raimunda Sâmia Nogueira, Rita Amanda Chaves de Lima, Francisca Jakelyne de Farias Marques, et al. 2015 Histoplasma Capsulatum in Planktonic and Biofilm Forms: In Vitro Susceptibility to Amphotericin B, Itraconazole and Farnesol. Journal of Medical Microbiology 64(Pt 4): 394–399.

Capasso, Clemente, and Claudiu T. Supuran 2014 Sulfa and Trimethoprim-like Drugs - Antimetabolites Acting as Carbonic Anhydrase, Dihydropteroate Synthase and Dihydrofolate Reductase Inhibitors. Journal of Enzyme Inhibition and Medicinal Chemistry 29(3): 379–387.

Capella-Gutiérrez, Salvador, José M. Silla-Martínez, and Toni Gabaldón 2009 TrimAl: A Tool for Automated Alignment Trimming in Large-Scale Phylogenetic Analyses. Bioinformatics 25(15): 1972–1973.

Chamany, Shadi, Sara A Mirza, John W Fleming, et al. 2004 A Large Histoplasmosis Outbreak among High School Students in , 2001. The Pediatric Infectious Disease Journal 23(10): 909–914.

Chan, Janet N. Y., Dajana Vuckovic, Lekha Sleno, et al. 2012 Target Identification by Chromatographic Co-Elution: Monitoring of Drug- Protein Interactions without Immobilization or Chemical Derivatization. Molecular & Cellular Proteomics : MCP 11(7). https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3394955/.

147

Chapman, S W, P D Rogers, M G Rinaldi, and D C Sullivan 1998 Susceptibilities of Clinical and Laboratory Isolates of Blastomyces Dermatitidis to Ketoconazole, Itraconazole, and Fluconazole. Antimicrobial Agents and Chemotherapy 42(4): 978–980.

Cheng, J T, R T Witty, R R Robinson, and W E Yarger 1982 Amphotericin B Nephrotoxicity: Increased Renal Resistance and Tubule Permeability. Kidney International 22(6): 626–633.

Cheung, S. C., G. Medoff, D. Schlessinger, and G. S. Kobayashi 1975 Response of Yeast and Mycelial Phases of Histoplasma Capsulatum to Amphotericin B and Actinomycin D. Antimicrobial Agents and Chemotherapy 8(4): 498–503.

Chiu, Hao-Chieh, Samuel K. Kulp, Shilpa Soni, et al. 2009 Eradication of Intracellular Salmonella Enterica Serovar Typhimurium with a Small-Molecule, Host Cell-Directed Agent. Antimicrobial Agents and Chemotherapy 53(12): 5236–5244.

Chu, Jaclyn H., Chris Feudtner, Kateri Heydon, Thomas J. Walsh, and Theoklis E. Zaoutis 2006 Hospitalizations for Endemic Mycoses: A Population-Based National Study. Clinical Infectious Diseases: An Official Publication of the Infectious Diseases Society of America 42(6): 822–825.

Clinical and Laboratory Standards Institute 2012 Reference Method for Broth Dilution Antifungal Susceptibility Testing of Yeasts; Informational Supplement M27-S4 32. Clinical and Laboratory Standards Institute, Wayne, PA.

Collier, M. A., K. J. Peine, S. Gautam, et al. 2016 Host-Mediated Leishmania Donovani Treatment Using AR-12 Encapsulated in Acetalated Dextran Microparticles. International Journal of Pharmaceutics 499(1–2): 186–194.

Colussi, Paul A., Charles A. Specht, and Christopher H. Taron 2005 Characterization of a Nucleus-Encoded Chitinase from the Yeast Kluyveromyces Lactis. Applied and Environmental Microbiology 71(6): 2862–2869.

Connolly, P., J. Wheat, C. Schnizlein-Bick, et al. 1999 Comparison of a New Triazole Antifungal Agent, Schering 56592, with Itraconazole and Amphotericin B for Treatment of Histoplasmosis in Immunocompetent Mice. Antimicrobial Agents and Chemotherapy 43(2): 322–328. 148

Cordeiro, R A, R S N Brilhante, M F G Rocha, et al. 2006 In Vitro Activities of Caspofungin, Amphotericin B and Azoles against Coccidioides Posadasii Strains from Northeast, Brazil. Mycopathologia 161(1): 21–26.

Cordeiro, Rossana de Aguiar, Francisca Jakelyne de Farias Marques, Raimunda Sâmia Nogueira Brilhante, et al. 2011 Synergistic Effect of Antituberculosis Drugs and Azoles in Vitro against Histoplasma Capsulatum Var. Capsulatum. Antimicrobial Agents and Chemotherapy 55(9): 4482–4484.

Cordeiro, Rossana de Aguiar, Charlline Vládia Silva de Melo, Francisca Jakelyne de Farias Marques, et al. 2016 Synthesis and in Vitro Antifungal Activity of Isoniazid-Derived Hydrazones against Coccidioides Posadasii. Microbial Pathogenesis 98: 1–5.

Cowen, Leah E., Dominique Sanglard, Susan J. Howard, P. David Rogers, and David S. Perlin 2015 Mechanisms of Antifungal Drug Resistance. Cold Spring Harbor Perspectives in Medicine 5(7). https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4484955/, accessed January 16, 2020.

Cruz, Jesus, Antonio Hidalgo-Gallego, Jose M. Lora, et al. 1992 Isolation and Characterization of Three Chitinases from Trichoderma Harzianum. European Journal of Biochemistry 206(3): 859–867.

Cruz, R C, S M C Werneck, C S Oliveira, et al. 2013 Influence of Different Media, Incubation Times, and Temperatures for Determining the MICs of Seven Antifungal Agents against Paracoccidioides Brasiliensis by Microdilution. Journal of Clinical Microbiology 51(2): 436–443.

Cuenca-Estrella, Manuel, and Juan Luis Rodriguez-Tudela 2010 The Current Role of the Reference Procedures by CLSI and EUCAST in the Detection of Resistance to Antifungal Agents in Vitro. Expert Review of Anti-Infective Therapy 8(3): 267–276.

Cummings, Richard D., Alan G. Darvill, Marilynn E. Etzler, and Michael G. Hahn 2015 Glycan-Recognizing Probes as Tools. In Essentials of Glycobiology. 3rd edition. Ajit Varki, Richard D. Cummings, Jeffrey D. Esko, et al., eds. Cold Spring Harbor (NY): Cold Spring Harbor Laboratory Press. http://www.ncbi.nlm.nih.gov/books/NBK453096/, accessed January 19, 2020.

Domer, Judith E.

149

1971 Monosaccharide and Chitin Content of Cell Walls of Histoplasma Capsulatum and Blastomyces Dermatitidis. Journal of Bacteriology 107(3): 870–877.

Dube, Danielle H., Jennifer A. Prescher, Chi N. Quang, and Carolyn R. Bertozzi 2006 Probing Mucin-Type O-Linked Glycosylation in Living Animals. Proceedings of the National Academy of Sciences of the United States of America 103(13): 4819–4824. van Duin, David, Arturo Casadevall, and Joshua D Nosanchuk 2002 Melanization of Cryptococcus Neoformans and Histoplasma Capsulatum Reduces Their Susceptibilities to Amphotericin B and Caspofungin. Antimicrobial Agents and Chemotherapy 46(11): 3394–3400.

Dünkler, Alexander, Andrea Walther, Charles A. Specht, and Jürgen Wendland 2005 Candida Albicans CHT3 Encodes the Functional Homolog of the Cts1 Chitinase of Saccharomyces Cerevisiae. Fungal Genetics and Biology 42(11): 935–947.

Duo-Chuan, Li 2006 Review of Fungal Chitinases. Mycopathologia 161(6): 345–360.

Durham, Malaika, and Fred E. Regnier 2006 Targeted Glycoproteomics: Serial Lectin Affinity Chromatography in the Selection of O-Glycosylation Sites on Proteins from the Human Blood Proteome. Journal of Chromatography. A 1132(1–2): 165–173.

Eddy, Sean R. 2009 A New Generation of Homology Search Tools Based on Probabilistic Inference. Genome Informatics. International Conference on Genome Informatics 23(1): 205–211.

Edwards, Jessica A., Elizabeth A. Alore, and Chad A. Rappleye 2011 The Yeast-Phase Virulence Requirement for α-Glucan Synthase Differs among Histoplasma Capsulatum Chemotypes. Eukaryotic Cell 10(1): 87–97.

Edwards, Jessica A., Chenxi Chen, Megan M. Kemski, et al. 2013 Histoplasma Yeast and Mycelial Transcriptomes Reveal Pathogenic-Phase and Lineage-Specific Gene Expression Profiles. BMC Genomics 14: 695.

Edwards, Jessica A., Megan M. Kemski, and Chad A. Rappleye 2013 Identification of an Aminothiazole with Antifungal Activity against Intracellular Histoplasma Capsulatum. Antimicrobial Agents and Chemotherapy 57(9): 4349–4359.

Edwards, Jessica A., and Chad A. Rappleye 2011 Histoplasma Mechanisms of Pathogenesis--One Portfolio Doesn’t Fit All. FEMS Microbiology Letters 324(1): 1–9.

150

Edwards, L. B., F. A. Acquaviva, V. T. Livesay, F. W. Cross, and C. E. Palmer 1969 An Atlas of Sensitivity to Tuberculin, PPD-B, and Histoplasmin in the United States. The American Review of Respiratory Disease 99(4): Suppl:1-132.

El-Gebali, Sara, Jaina Mistry, Alex Bateman, et al. 2019 The Pfam Protein Families Database in 2019. Nucleic Acids Research 47(D1): D427–D432.

Endo, A., and T. Misato 1969 Polyoxin D, a Competitive Inhibitor of UDP-N-Acetylglucosamine: Chitin N- Acetylglucosaminyltransferase in Neurospora Crassa. Biochemical and Biophysical Research Communications 37(4): 718–722.

Espinel-Ingroff, A 1998 Comparison of In Vitro Activities of the New Triazole SCH56592 and the Echinocandins MK-0991 (L-743,872) and LY303366 against Opportunistic Filamentous and Dimorphic Fungi and Yeasts. Journal of Clinical Microbiology 36(10): 2950–2956.

Fang, Weiguo, Bo Leng, Yuehua Xiao, et al. 2005 Cloning of Beauveria Bassiana Chitinase Gene Bbchit1 and Its Application To Improve Fungal Strain Virulence. Applied and Environmental Microbiology 71(1): 363– 370.

Fothergill, Annette W., Deanna A. Sutton, Dora I. McCarthy, and Nathan P. Wiederhold 2014 Impact of New Antifungal Breakpoints on Antifungal Resistance in Candida Species. Journal of Clinical Microbiology 52(3): 994–997.

Fu, J., C. Mares, A. Lizcano, Y. Liu, and B. L. Wickes 2011 Insertional Mutagenesis Combined with an Inducible Filamentation Phenotype Reveals a Conserved STE50 Homologue in Cryptococcus Neoformans That Is Required for Monokaryotic Fruiting and Sexual Reproduction. Molecular Microbiology 79(4): 990–1007.

Fukamizo, Tamo, Chiye Sasaki, Elisabeth Schelp, Kara Bortone, and Jon D. Robertus 2001 Kinetic Properties of Chitinase-1 from the Fungal Pathogen Coccidioides Immitis. Biochemistry 40(8): 2448–2454.

Garfoot, Andrew L., Kristie D. Goughenour, Marcel Wüthrich, et al. 2018 O-Mannosylation of Proteins Enables Histoplasma Yeast Survival at Mammalian Body Temperatures. MBio 9(1). https://mbio.asm.org/content/9/1/e02121-17, accessed January 16, 2020.

Garfoot, Andrew L., Qian Shen, Marcel Wüthrich, Bruce S. Klein, and Chad A. Rappleye 151

2016 The Eng1 β-Glucanase Enhances Histoplasma Virulence by Reducing β-Glucan Exposure. MBio 7(2): e01388-15.

Goettig, Peter 2016 Effects of Glycosylation on the Enzymatic Activity and Mechanisms of Proteases. International Journal of Molecular Sciences 17(12). https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5187769/, accessed January 29, 2020.

Goldberg, J, P Connolly, C Schnizlein-Bick, et al. 2000 Comparison of Nikkomycin Z with Amphotericin B and Itraconazole for Treatment of Histoplasmosis in a Murine Model. Antimicrobial Agents and Chemotherapy 44(6): 1624–1629.

González, G M, R Tijerina, L K Najvar, et al. 2001 Correlation between Antifungal Susceptibilities of Coccidioides Immitis in Vitro and Antifungal Treatment with Caspofungin in a Mouse Model. Antimicrobial Agents and Chemotherapy 45(6): 1854–1859.

González, Gloria M. 2009 In Vitro Activities of Isavuconazole against Opportunistic Filamentous and Dimorphic Fungi. Medical Mycology 47(1): 71–76.

González, Gloria M, Annette W Fothergill, Deanna A Sutton, Michael G Rinaldi, and David Loebenberg 2005 In Vitro Activities of New and Established Triazoles against Opportunistic Filamentous and Dimorphic Fungi. Medical Mycology 43(3): 281–284.

Goughenour, Kristie D., Joan-Miquel Balada-Llasat, and Chad A. Rappleye 2015 Quantitative Microplate-Based Growth Assay for Determination of Antifungal Susceptibility of Histoplasma Capsulatum Yeasts. Journal of Clinical Microbiology 53(10): 3286–3295.

Goughenour, Kristie D., and Chad A. Rappleye 2016 Antifungal Therapeutics for Dimorphic Fungal Pathogens. Virulence: 1–11.

Graybill, J R, L K Najvar, R Bocanegra, R F Hector, and M F Luther 1998 Efficacy of Nikkomycin Z in the Treatment of Murine Histoplasmosis. Antimicrobial Agents and Chemotherapy 42(9): 2371–2374.

Gruber, Sabine, Christian P. Kubicek, and Verena Seidl-Seiboth 2011 Differential Regulation of Orthologous Chitinase Genes in Mycoparasitic Trichoderma Species. Applied and Environmental Microbiology 77(20): 7217–7226.

Gruber, Sabine, Gustav Vaaje-Kolstad, Fabiola Matarese, et al. 152

2011 Analysis of Subgroup C of Fungal Chitinases Containing Chitin-Binding and LysM Modules in the Mycoparasite Trichoderma Atroviride. Glycobiology 21(1): 122– 133.

Hage, Chadi A, Marwan M Azar, Nathan Bahr, James Loyd, and L Joseph Wheat 2015 Histoplasmosis: Up-to-Date Evidence-Based Approach to Diagnosis and Management. Seminars in Respiratory and Critical Care Medicine 36(5): 729–745.

Hage, Chadi A., Patricia Connolly, Daniel Horan, et al. 2011 Investigation of the Efficacy of Micafungin in the Treatment of Histoplasmosis Using Two North American Strains of Histoplasma Capsulatum. Antimicrobial Agents and Chemotherapy 55(9): 4447–4450.

Hage, Chadi A, Kenneth S Knox, and Lawrence J Wheat 2012 Endemic Mycoses: Overlooked Causes of Community Acquired Pneumonia. Respiratory Medicine 106(6): 769–776.

Hamid, Rifat, Minhaj A. Khan, Mahboob Ahmad, et al. 2013 Chitinases: An Update. Journal of Pharmacy & Bioallied Sciences 5(1): 21–29.

Haran, Shoshan, Hedva Schickler, Amos Oppenheim, and Ilan Chet 1995 New Components of the Chitinolytic System of Trichoderma Harzianum. Mycological Research 99(4): 441–446.

Hartl, Lukas, Simone Zach, and Verena Seidl-Seiboth 2012 Fungal Chitinases: Diversity, Mechanistic Properties and Biotechnological Potential. Applied Microbiology and Biotechnology 93(2): 533–543.

Hector, R. F., B. L. Zimmer, and D. Pappagianis 1990 Evaluation of Nikkomycins X and Z in Murine Models of , Histoplasmosis, and . Antimicrobial Agents and Chemotherapy 34(4): 587–593.

HmmerAlign N.d. http://www.biology.wustl.edu/gcg/hmmeralign.html, accessed December 26, 2019.

Hoang, Ky V., Heather Curry, Michael A. Collier, et al. 2016 Needle-Free Delivery of Acetalated Dextran-Encapsulated AR-12 Protects Mice from Francisella Tularensis Lethal Challenge. Antimicrobial Agents and Chemotherapy 60(4): 2052–2062.

Hoell, Ingunn A., Sonja S. Klemsdal, Gustav Vaaje-Kolstad, Svein J. Horn, and Vincent G. H. Eijsink 153

2005 Overexpression and Characterization of a Novel Chitinase from Trichoderma Atroviride Strain P1. Biochimica Et Biophysica Acta 1748(2): 180–190.

Holbrook, Eric D., Jessica A. Edwards, Brian H. Youseff, and Chad A. Rappleye 2011 Definition of the Extracellular Proteome of Pathogenic-Phase Histoplasma Capsulatum. Journal of Proteome Research 10(4): 1929–1943.

Holbrook, Eric D., Megan M. Kemski, Sarah M. Richer, L. Joseph Wheat, and Chad A. Rappleye 2014 Glycosylation and Immunoreactivity of the Histoplasma Capsulatum Cfp4 Yeast- Phase Exoantigen. Infection and Immunity 82(10): 4414–4425.

Hortin, G. L., and B. L. Trimpe 1990 Lectin Affinity Chromatography of Proteins Bearing O-Linked Oligosaccharides: Application of Jacalin-Agarose. Analytical Biochemistry 188(2): 271–277.

Horwath, Michael C, Roger A Fecher, and George S Deepe 2015 Histoplasma Capsulatum, Lung Infection and Immunity. Future Microbiology 10: 967–975.

Hurtado-Guerrero, Ramon, and Daan M. F. van Aalten 2007 Structure of Saccharomyces Cerevisiae Chitinase 1 and Screening-Based Discovery of Potent Inhibitors. Chemistry & Biology 14(5): 589–599.

Ishita, Keisuke, Stavros Stefanopoulos, Ahmed Khalil, et al. 2018 Synthesis and Biological Evaluation of Aminothiazoles against Histoplasma Capsulatum and Cryptococcus Neoformans. Bioorganic & Medicinal Chemistry 26(9): 2251–2261.

Isono, K., K. Asahi, and S. Suzuki 1969 Studies on Polyoxins, Antifungal Antibiotics. 13. The Structure of Polyoxins. Journal of the American Chemical Society 91(26): 7490–7505.

Jain, Vipul V, Timothy Evans, and Michael W Peterson 2006 Reactivation Histoplasmosis after Treatment with Anti-Tumor Necrosis Factor Alpha in a Patient from a Nonendemic Area. Respiratory Medicine 100(7): 1291–1293.

Jaques, Alex K., Tamo Fukamizo, Diana Hall, et al. 2003 Disruption of the Gene Encoding the ChiB1 Chitinase of Aspergillus Fumigatus and Characterization of a Recombinant Gene Product. Microbiology, 149(10): 2931– 2939.

Johnson, Melissa D., and John R. Perfect

154

2003 Caspofungin: First Approved Agent in a New Class of Antifungals. Expert Opinion on Pharmacotherapy 4(5): 807–823.

Jones, Meredith D.M., Thomas A. Richards, David L. Hawksworth, and David Bass 2011 Validation and Justification of the Phylum Name Cryptomycota Phyl. Nov. IMA Fungus : The Global Mycological Journal 2(2): 173–175. de Jonge, Ronnie, and Bart P. H. J. Thomma 2009 Fungal LysM Effectors: Extinguishers of Host Immunity? Trends in Microbiology 17(4): 151–157.

Kanafani, Zeina A., and John R. Perfect 2008 Resistance to Antifungal Agents: Mechanisms and Clinical Impact. Clinical Infectious Diseases 46(1): 120–128.

Karlsson, Magnus, and Jan Stenlid 2008 Comparative Evolutionary Histories of the Fungal Chitinase Gene Family Reveal Non-Random Size Expansions and Contractions Due to Adaptive Natural Selection. Evolutionary Bioinformatics Online 4: 47–60. 2009 Evolution of Family 18 Glycoside Hydrolases: Diversity, Domain Structures and Phylogenetic Relationships. Journal of Molecular Microbiology and Biotechnology 16(3– 4): 208–223.

Kasuga, Takao, Thomas J. White, Gina Koenig, et al. 2003 Phylogeography of the Fungal Pathogen Histoplasma Capsulatum. Molecular Ecology 12(12): 3383–3401.

Kathuria, Shallu, Pradeep K Singh, Jacques F Meis, and Anuradha Chowdhary 2014 In Vitro Antifungal Susceptibility Profile and Correlation of Mycelial and Yeast Forms of Molecularly Characterized Histoplasma Capsulatum Strains from India. Antimicrobial Agents and Chemotherapy 58(9): 5613–5616.

Kauffman, Carol A. 2007 Histoplasmosis: A Clinical and Laboratory Update. Clinical Microbiology Reviews 20(1): 115–132.

Kauffman, Carol A 2009 Histoplasmosis. Clinics in Chest Medicine 30(2): 217–25, v.

Keath, E J, G S Kobayashi, and G Medoff 1992 Typing of Histoplasma Capsulatum by Restriction Fragment Length Polymorphisms in a Nuclear Gene. Journal of Clinical Microbiology 30(8): 2104–2107.

Khalil, Ahmed, Jessica A. Edwards, Chad A. Rappleye, and Werner Tjarks 155

2015 Design, Synthesis, and Biological Evaluation of Aminothiazole Derivatives against the Fungal Pathogens Histoplasma Capsulatum and Cryptococcus Neoformans. Bioorganic & Medicinal Chemistry 23(3): 532–547.

Kim, Min-Sik, and Akhilesh Pandey 2012 Electron Transfer Dissociation Mass Spectrometry in Proteomics. Proteomics 12(0): 530–542.

Klein, Bruce S., and Brad Tebbets 2007 Dimorphism and Virulence in Fungi. Current Opinion in Microbiology 10(4): 314–319. van Koeveringe, M P, and R E Brouwer 2010 Histoplasma Capsulatum Reactivation with Haemophagocytic Syndrome in a Patient with Chronic Lymphocytic Leukaemia. The Netherlands Journal of Medicine 68(12): 418–421.

Kohler, S, L J Wheat, P Connolly, et al. 2000 Comparison of the Echinocandin Caspofungin with Amphotericin B for Treatment of Histoplasmosis Following Pulmonary Challenge in a Murine Model. Antimicrobial Agents and Chemotherapy 44(7): 1850–1854.

Koselny, Kristy, Julianne Green, Louis DiDone, et al. 2016 The Celecoxib Derivative AR-12 Has Broad-Spectrum Antifungal Activity In Vitro and Improves the Activity of Fluconazole in a Murine Model of . Antimicrobial Agents and Chemotherapy 60(12): 7115–7127.

Kriesel, John D, Deanna A Sutton, Susan Schulman, Annette W Fothergill, and Michael G Rinaldi 2008 Persistent Pulmonary Infection with an Azole-Resistant Coccidioides Species. Medical Mycology 46(6): 607–610.

Kuranda, M. J., and P. W. Robbins 1991 Chitinase Is Required for Cell Separation during Growth of Saccharomyces Cerevisiae. Journal of Biological Chemistry 266(29): 19758–19767.

Leach, Michelle D., and Alistair J. P. Brown 2012 Posttranslational Modifications of Proteins in the Pathobiology of Medically Relevant Fungi. Eukaryotic Cell 11(2): 98–108.

Lee, Kyung-Tae, Yee-Seul So, Dong-Hoon Yang, et al. 2016 Systematic Functional Analysis of Kinases in the Fungal Pathogen Cryptococcus Neoformans. Nature Communications 7. https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5052723/, accessed January 16, 2020. 156

LeMonte, A M, K E Washum, M L Smedema, et al. 2000 Amphotericin B Combined with Itraconazole or Fluconazole for Treatment of Histoplasmosis. The Journal of Infectious Diseases 182(2): 545–550.

LeMonte, A. M., K. E. Washum, M. L. Smedema, et al. 2000 Amphotericin B Combined with Itraconazole or Fluconazole for Treatment of Histoplasmosis. The Journal of Infectious Diseases 182(2): 545–550.

Li, R. K., M. A. Ciblak, N. Nordoff, et al. 2000 In Vitro Activities of Voriconazole, Itraconazole, and Amphotericin B against Blastomyces Dermatitidis, Coccidioides Immitis, and Histoplasma Capsulatum. Antimicrobial Agents and Chemotherapy 44(6): 1734–1736.

Li, Ren-Kai, Meral A. Ciblak, Nicole Nordoff, et al. 2000 In Vitro Activities of Voriconazole, Itraconazole, and Amphotericin B against Blastomyces Dermatitidis,Coccidioides Immitis, and Histoplasma Capsulatum. Antimicrobial Agents and Chemotherapy 44(6): 1734–1736.

Lienemann, Michael, Harry Boer, Arja Paananen, Sylvain Cottaz, and Anu Koivula 2009 Toward Understanding of Carbohydrate Binding and Substrate Specificity of a Glycosyl Hydrolase 18 Family (GH-18) Chitinase from Trichoderma Harzianum. Glycobiology 19(10): 1116–1126.

Loibl, Martin, and Sabine Strahl 2013 Protein O-Mannosylation: What We Have Learned from Baker’s Yeast. Biochimica et Biophysica Acta (BBA) - Molecular Cell Research 1833(11). Functional and Structural Diversity of the Endoplasmic Reticulum: 2438–2446.

Lommel, Mark, and Sabine Strahl 2009 Protein O-Mannosylation: Conserved from Bacteria to . Glycobiology 19(8): 816–828.

Longo, Larissa V. G., Stephanie C. Ray, Rosana Puccia, and Chad A. Rappleye 2018 Characterization of the APSES-Family Transcriptional Regulators of Histoplasma Capsulatum. FEMS Yeast Research 18(8).

Marchler-Bauer, Aron, Yu Bo, Lianyi Han, et al. 2017 CDD/SPARCLE: Functional Classification of Proteins via Subfamily Domain Architectures. Nucleic Acids Research 45(D1): D200–D203.

Marchler-Bauer, Aron, and Stephen H. Bryant 2004 CD-Search: Protein Domain Annotations on the Fly. Nucleic Acids Research 32(Web Server issue): W327-331. 157

Marchler-Bauer, Aron, Myra K. Derbyshire, Noreen R. Gonzales, et al. 2015 CDD: NCBI’s Conserved Domain Database. Nucleic Acids Research 43(Database issue): D222-226.

Marchler-Bauer, Aron, Shennan Lu, John B. Anderson, et al. 2011 CDD: A Conserved Domain Database for the Functional Annotation of Proteins. Nucleic Acids Research 39(Database issue): D225-229.

Maresca, B., and G. S. Kobayashi 2000 Dimorphism in Histoplasma Capsulatum and Blastomyces Dermatitidis. Contributions to Microbiology 5: 201–216.

McVeigh I, Houston WE 1972 Factors Affecting Mycelial to Yeast Phase Conversion and Growth of the Yeast Phase of Histoplasma Capsulatum. Mycopathol Mycol Appl(47): 135–151.

Mechref, Yehia 2012 Use of CID/ETD Mass Spectrometry to Analyze Glycopeptides. Current Protocols in Protein Science / Editorial Board, John E. Coligan ... [et Al.] 0 12: Unit- 12.1111.

Medoff, G., M. Sacco, B. Maresca, et al. 1986 Irreversible Block of the Mycelial-to-Yeast Phase Transition of Histoplasma Capsulatum. Science (New York, N.Y.) 231(4737): 476–479.

Mikesh, Leann M, Beatrix Ueberheide, An Chi, et al. 2006 The Utility of ETD Mass Spectrometry in Proteomic Analysis. Biochimica et Biophysica Acta 1764(12): 1811–1822.

Minh, Bui Quang, Minh Anh Thi Nguyen, and Arndt von Haeseler 2013 Ultrafast Approximation for Phylogenetic Bootstrap. Molecular Biology and Evolution 30(5): 1188–1195.

Mohr, Emma L., Laura K. McMullan, Michael K. Lo, et al. 2015 Inhibitors of Cellular Kinases with Broad-Spectrum Antiviral Activity for Hemorrhagic Fever Viruses. Antiviral Research 120: 40–47.

Munro, Carol A., Steven Bates, Ed T. Buurman, et al. 2005 Mnt1p and Mnt2p of Candida Albicans Are Partially Redundant α-1,2- Mannosyltransferases That Participate in O-Linked Mannosylation and Are Required for Adhesion and Virulence. Journal of Biological Chemistry 280(2): 1051–1060. van Munster, Jolanda M., Rachel M. van der Kaaij, Lubbert Dijkhuizen, and Marc J. E. 158

C. van der Maarel 2012 Biochemical Characterization of Aspergillus Niger CfcI, a Glycoside Hydrolase Family 18 Chitinase That Releases Monomers during Substrate Hydrolysis. Microbiology 158(8): 2168–2179.

Nakai, Toru, Jun Uno, Fumiaki Ikeda, et al. 2003 In Vitro Antifungal Activity of Micafungin (FK463) against Dimorphic Fungi: Comparison of Yeast-Like and Mycelial Forms. Antimicrobial Agents and Chemotherapy 47(4): 1376–1381.

Nakelchik, Masha, and Julie E Mangino 2002 Reactivation of Histoplasmosis after Treatment with Infliximab. The American Journal of Medicine 112(1): 78.

Nett, Jeniel E., and David R. Andes 2016 Antifungal Agents: Spectrum of Activity, Pharmacology, and Clinical Indications. Infectious Disease Clinics of North America 30(1): 51–83.

Nguyen, Lam-Tung, Heiko A. Schmidt, Arndt von Haeseler, and Bui Quang Minh 2015 IQ-TREE: A Fast and Effective Stochastic Algorithm for Estimating Maximum- Likelihood Phylogenies. Molecular Biology and Evolution 32(1): 268–274.

Nguyen, Van Q., and Anita Sil 2008 Temperature-Induced Switch to the Pathogenic Yeast Form of Histoplasma Capsulatum Requires Ryp1, a Conserved Transcriptional Regulator. Proceedings of the National Academy of Sciences of the United States of America 105(12): 4880–4885.

Pfaller, M. A., and D. J. Diekema 2012 Progress in Antifungal Susceptibility Testing of Candida Spp. by Use of Clinical and Laboratory Standards Institute Broth Microdilution Methods, 2010 to 2012. Journal of Clinical Microbiology 50(9): 2846–2856.

Pham, James S., Karen L. Dawson, Katherine E. Jackson, et al. 2013 Aminoacyl-TRNA Synthetases as Drug Targets in Eukaryotic Parasites. International Journal for Parasitology: Drugs and Drug Resistance 4(1): 1–13.

Pine, Leo 1954 STUDIES ON THE GROWTH OF HISTOPLASMA CAPSULATUM I. Journal of Bacteriology 68(6): 671–679.

Pitkin, J. W., D. G. Panaccione, and J. D. Walton 1996 A Putative Cyclic Peptide Efflux Pump Encoded by the TOXA Gene of the Plant- Carbonum. Microbiology (Reading, England) 142 ( Pt 6): 1557–1565. 159

Ramani, Rama, and Vishnu Chaturvedi 2007 Antifungal Susceptibility Profiles of Coccidioides Immitis and Coccidioides Posadasii from Endemic and Non-Endemic Areas. Mycopathologia 163(6): 315–319.

Rauseo, Adriana M., Ariella Coler-Reilly, Lindsey Larson, and Andrej Spec N.d. Hope on the Horizon: Novel Fungal Treatments in Development. Open Forum Infectious Diseases. https://academic.oup.com/ofid/advance- article/doi/10.1093/ofid/ofaa016/5700872, accessed January 16, 2020.

Ravi Kumar, Majeti N. V 2000 A Review of Chitin and Chitosan Applications. Reactive and Functional Polymers 46(1): 1–27.

Reference Method for Broth Dilution Antifungal Susceptibility Testing of Filamentous Fungi. 2008. Clinical and Laboratory Standards Institute.

Reiss, E. 1977 Serial Enzymatic Hydrolysis of Cell Walls of Two Serotypes of Yeast-Form Histoplasma Capsulatum with Alpha(1 Leads to 3)-Glucanase, Beta(1 Leads to 3)- Glucanase, Pronase, and Chitinase. Infection and Immunity 16(1): 181–188.

Reiss, E., Sara E. Miller, W. Kaplan, and L. Kaufman 1977 Antigenic, Chemical, and Structural Properties of Cell Walls of Histoplasma Capsulatum Yeast-Form Chemotypes 1 and 2 After Serial Enzymatic Hydrolysis. Infection and Immunity 16(2): 690–700.

Rex, John H., and Clinical and Laboratory Standards Institute, eds. 2008 Reference Method for Broth Dilution Antifungal Susceptibility Testing of Yeasts: Approved Standard. 3rd ed. Documents / Clinical and Laboratory Standards Institute, 28,14. Wayne, PA: National Committee for Clinical Laboratory Standards.

Rodriguez-Kabana, R., G. Godoy, G. Morgan-Jones, and R. A. Shelby 1983 The Determination of Soil Chitinase Activity: Conditions for Assay and Ecological Studies. Plant and Soil 75(1): 95–106.

Rush, Christina L., Alexander W. Schüttelkopf, Ramon Hurtado-Guerrero, et al. 2010 Natural Product–Guided Discovery of a Fungal Chitinase Inhibitor. Chemistry & Biology 17(12): 1275–1281.

Salvin, S. B. 1949 Cysteine and Related Compounds in the Growth of the Yeast like Phase of Histoplasma Capsulatum. The Journal of Infectious Diseases 84(3): 275–283. 160

Schmittgen, Thomas D., and Kenneth J. Livak 2008 Analyzing Real-Time PCR Data by the Comparative C(T) Method. Nature Protocols 3(6): 1101–1108.

Schneider, Caroline A., Wayne S. Rasband, and Kevin W. Eliceiri 2012 NIH Image to ImageJ: 25 Years of Image Analysis. Nature Methods 9(7): 671– 675.

Seidl, Verena 2008a Chitinases of Filamentous Fungi: A Large Group of Diverse Proteins with Multiple Physiological Functions. Fungal Biology Reviews 22(1): 36–42. 2008b Chitinases of Filamentous Fungi: A Large Group of Diverse Proteins with Multiple Physiological Functions. Fungal Biology Reviews 22(1): 36–42.

Seidl, Verena, Birgit Huemer, Bernhard Seiboth, and Christian P. Kubicek 2005 A Complete Survey of Trichoderma Chitinases Reveals Three Distinct Subgroups of Family 18 Chitinases: Trichoderma Chitinases. FEBS Journal 272(22): 5923–5939.

Selvaggini, Serena, Carol A. Munro, Serge Paschoud, Dominique Sanglard, and Neil A. R. Gow 2004 Independent Regulation of Chitin Synthase and Chitinase Activity in Candida Albicans and Saccharomyces Cerevisiae. Microbiology, 150(4): 921–928.

Shen, Qian, and Chad A. Rappleye 2017 Differentiation of the Fungus Histoplasma Capsulatum into a Pathogen of Phagocytes. Current Opinion in Microbiology 40: 1–7.

Stamatakis, Alexandros 2014 RAxML Version 8: A Tool for Phylogenetic Analysis and Post-Analysis of Large Phylogenies. Bioinformatics 30(9): 1312–1313.

Stergiopoulos, Ioannis, Yiannis A. I. Kourmpetis, Jason C. Slot, et al. 2012 In Silico Characterization and Molecular Evolutionary Analysis of a Novel Superfamily of Fungal Effector Proteins. Molecular Biology and Evolution 29(11): 3371–3384.

Stetler, D. A., and G. Boguslawski 1979 Cysteine Biosynthesis in a Fungus, Histoplasma Capsulatum. Sabouraudia 17(1): 23–34.

Subcommittee on Antifungal Susceptibility Testing of the ESCMID European Committee for Antimicrobial Susceptibility Testing 2008 EUCAST Technical Note on the Method for the Determination of Broth Dilution 161

Minimum Inhibitory Concentrations of Antifungal Agents for Conidia-Forming Moulds. Clinical Microbiology and Infection: The Official Publication of the European Society of Clinical Microbiology and Infectious Diseases 14(10): 982–984.

Subcommittee on Antifungal Susceptibility Testing (AFST) of the ESCMID European Committee for Antimicrobial Susceptibility Testing (EUCAST) 2008 EUCAST Definitive Document EDef 7.1: Method for the Determination of Broth Dilution MICs of Antifungal Agents for Fermentative Yeasts. Clinical Microbiology and Infection: The Official Publication of the European Society of Clinical Microbiology and Infectious Diseases 14(4): 398–405.

Sugar, A M, and X P Liu 1996 In Vitro and in Vivo Activities of SCH 56592 against Blastomyces Dermatitidis. Antimicrobial Agents and Chemotherapy 40(5): 1314–1316.

Takaya, Naoki, Daisuke Yamazaki, Hiroyuki Horiuchi, Akinori Ohta, and Masamichi Takagi 1998 Cloning and Characterization of a Chitinase-Encoding Gene (ChiA) from Aspergillus Nidulans, Disruption of Which Decreases Germination Frequency and Hyphal Growth. Bioscience, Biotechnology, and Biochemistry 62(1): 60–65.

Tharanathan, Rudrapatnam N., and Farooqahmed S. Kittur 2003 Chitin--the Undisputed Biomolecule of Great Potential. Critical Reviews in Food Science and Nutrition 43(1): 61–87.

Titov, Denis V., and Jun O. Liu 2012 Identification and Validation of Protein Targets of Bioactive Small Molecules. Bioorganic & Medicinal Chemistry 20(6): 1902–1909.

Vallabhaneni, Snigdha, Rajal K. Mody, Tiffany Walker, and Tom Chiller 2016 The Global Burden of Fungal Diseases. Infectious Disease Clinics of North America 30(1): 1–11.

Vergidis, Paschalis, Robin K Avery, L Joseph Wheat, et al. 2015 Histoplasmosis Complicating Tumor Necrosis Factor-\alpha Blocker Therapy: A Retrospective Analysis of 98 Cases. Clinical Infectious Diseases 61(3): 409–417.

Walton, Felicia J., Alexander Idnurm, and Joseph Heitman 2005 Novel Gene Functions Required for Melanization of the Human Pathogen Cryptococcus Neoformans. Molecular Microbiology 57(5): 1381–1396.

Wang, Cheng-Yi, Jih-Shuin Jerng, Jen-Chung Ko, et al. 2005 Disseminated Coccidioidomycosis. Emerging Infectious Diseases 11(1): 177–179.

162

Webster, Rachael Hanby, and Anita Sil 2008 Conserved Factors Ryp2 and Ryp3 Control Cell Morphology and Infectious Spore Formation in the Fungal Pathogen Histoplasma Capsulatum. Proceedings of the National Academy of Sciences of the United States of America 105(38): 14573–14578.

Wheat, J., P. Marichal, H. Vanden Bossche, A. Le Monte, and P. Connolly 1997 Hypothesis on the Mechanism of Resistance to Fluconazole in Histoplasma Capsulatum. Antimicrobial Agents and Chemotherapy 41(2): 410–414.

Wheat, J., S. MaWhinney, R. Hafner, et al. 1997 Treatment of Histoplasmosis with Fluconazole in Patients with Acquired Immunodeficiency Syndrome. National Institute of Allergy and Infectious Diseases Acquired Immunodeficiency Syndrome Clinical Trials Group and Mycoses Study Group. The American Journal of Medicine 103(3): 223–232.

Wheat, L J, P Connolly, M Smedema, et al. 2001 Emergence of Resistance to Fluconazole as a Cause of Failure during Treatment of Histoplasmosis in Patients with Acquired Immunodeficiency Disease Syndrome. Clinical Infectious Diseases 33(11): 1910–1913.

Wheat, L Joseph, Patricia Connolly, Melinda Smedema, et al. 2006 Activity of Newer Triazoles against Histoplasma Capsulatum from Patients with AIDS Who Failed Fluconazole. The Journal of Antimicrobial Chemotherapy 57(6): 1235–1239.

Wheat, L. Joseph, Alison G. Freifeld, Martin B. Kleiman, et al. 2007 Clinical Practice Guidelines for the Management of Patients with Histoplasmosis: 2007 Update by the Infectious Diseases Society of America. Clinical Infectious Diseases 45(7): 807–825.

Wheat, Lawrence J., Marwan M. Azar, Nathan C. Bahr, et al. 2016 Histoplasmosis. Infectious Disease Clinics of North America 30(1): 207–227.

Worsham, P. L., and W. E. Goldman 1988 Quantitative Plating of Histoplasma Capsulatum without Addition of Conditioned Medium or Siderophores. Journal of Medical and Veterinary Mycology: Bi-Monthly Publication of the International Society for Human and Animal Mycology 26(3): 137– 143.

Youseff, Brian H., and Chad A. Rappleye 2012 RNAi-Based Gene Silencing Using a GFP Sentinel System in Histoplasma Capsulatum. In Host-Fungus Interactions: Methods and Protocols. Alexandra C. Brand and Donna M. MacCallum, eds. Pp. 151–164. Methods in Molecular Biology. Totowa, 163

NJ: Humana Press. https://doi.org/10.1007/978-1-61779-539-8_10, accessed January 19, 2020.

Zargar, Vida, Morteza Asghari, and Amir Dashti 2015 A Review on Chitin and Chitosan Polymers: Structure, Chemistry, Solubility, Derivatives, and Applications. ChemBioEng Reviews 2(3): 204–226.

Zellner, Maria, Wolfgang Winkler, Hubert Hayden, et al. 2005 Quantitative Validation of Different Protein Precipitation Methods in Proteome Analysis of Blood Platelets. ELECTROPHORESIS 26(12): 2481–2489.

Zemska, Olga, and Chad A. Rappleye 2012 Agrobacterium-Mediated Insertional Mutagenesis in Histoplasma Capsulatum. Methods in Molecular Biology (Clifton, N.J.) 845: 51–66.

Zhanel, G G, J A Karlowsky, G A Harding, et al. 1997 In Vitro Activity of a New Semisynthetic Echinocandin, LY-303366, against Systemic Isolates of Candida Species, Cryptococcus Neoformans, Blastomyces Dermatitidis, and Aspergillus Species. Antimicrobial Agents and Chemotherapy 41(4): 863–865.

Zhang, null, null Chung, and null Oldenburg 1999 A Simple Statistical Parameter for Use in Evaluation and Validation of High Throughput Screening Assays. Journal of Biomolecular Screening 4(2): 67–73.

Zhao, B., X. Xia, J. Yin, et al. 2001 Epidemiological Investigation of Histoplasma Capsulatum Infection in China. Chinese Medical Journal 114(7): 743–746.

164