You are on page 1of 194

ACETATE METABOLISM IN THE FUNGAL

PATHOGEN CRYPTOCOCCUS NEOFORMANS

A Dissertation
Presented to
the Graduate School of
Clemson University

In Partial Fulfillment
of the Requirements for the Degree
Doctor of Philosophy
Genetics

by
Oly Ahmed
May 2023

Accepted by
Dr. Kerry Smith, Committee Chair
Dr. Cheryl Ingram-Smith
Dr. James Morris
Dr. Meredith Morris
ABSTRACT

Cryptococcus neoformans is an environmental basidiomycetous fungus with

a worldwide distribution and a wide range of habitats. Inhalation of the desiccated

yeasts or spores of C. neoformans often leads to opportunistic pulmonary infections

in immunocompromised individuals, and in severe cases causes lethal meningitis

following hematogenous dissemination. During infection, depending on the tissue

and disease state, the invading fungi experience a range of nutrient

microenvironments within the host body. As a result, rapid metabolic adaptations

geared towards efficient utilization of carbon sources alternative to glucose become

one of the prime determinants of survival and growth for the pathogen. Incidentally,

cryptococcal infection has a long-standing association with acetate metabolism as

underlined by previous cryptococcal transcriptomic studies in infection models that

revealed an increased activity of alternative carbon metabolism during infection, and

highlighted by the observation that cryptococcomas (infectious granulomas) are

enriched with the two-carbon metabolite acetate, a physiologically important

alternative carbon source. In this work I investigated the direct transcriptomic

impacts of acetate utilization compared to glucose utilization via RNA-sequencing,

and noted that cryptococcal transcriptome bears signatures of nutrient starvation

and metabolic adaptations in acetate-grown conditions, as well as a remarkable

upregulation of virulence-associated genes. Moreover, to investigate the importance

of acetate production in C. neoformans we biochemically and kinetically

characterized acetate kinase (Ack), an enzyme involved in acetate metabolism, and

found that kinetically Ack has a preference in the acetate forming direction

compared acetyl phosphate forming direction. This observation is consistent with

ii
the previously discovered biochemical role of another cryptococcal enzyme xylulose-

5-phosphate/ fructose-6-phosphate phosphoketolase (Xfp2), which produces acetyl

phosphate from phosphoketose sugars. Taken together, here I propose that Ack and

Xfp2 forms a pathway of acetate production in C. neoformans, and hence are likely

to be involved in acetate homeostasis. Possibility of the Xfp-Ack pathway of acetate

production in the fungal species C. neoformans prompted us to search for Ack

sequences in other eukaryotic genomes. Here I employed a hidden Markov model

based strategy to predict Ack sequences in publicly available and curated eukaryotic

genomes, and found that an overwhelming majority of the predicted Ack sequences

can be observed in dikaryotic fungi. Maximum likelihood based phylogeny of the

predicted eukaryotic Ack sequences suggest an evolutionary trajectory involving

divergence from a common ancestor and loss in most eukaryotic lineages. Finally, to

interpret the observation that acetate utilization appears to be connected with the

upregulation of virulence-associated genes in C. neoformans, I have proposed a new

theoretical framework to explain the emergence of virulence in fungal pathogens.

The hypothesis posits that molecular response to various environmental stresses and

virulence phenotypes are modular in nature and hence are often not connected in

most fungal species. Occasionally, species may co-opt virulence modules, and

hierarchically nest it with stress-response modules, thereby acquiring the ability of

opportunistic pathogenicity in mammalian hosts. Overall, this work furthers our

current understanding of the impacts of acetate metabolism in the human fungal

pathogen C. neoformans.

iii
TABLE OF CONTENTS

TITLE PAGE...........................................................................................................i

ABSTRACT ........................................................................................................... ii

LIST OF TABLES................................................................................................... ix

LIST OF FIGURES .................................................................................................. x

CHAPTER 1 ..........................................................................................................1

Cryptococcal virulence in the context of acetate metabolism .................... 1

1.1. Cryptococcus and cryptococcosis ........................................................ 2

1.1.1. The genus Cryptococcus ................................................................2

1.1.2. Distribution and morphology........................................................ 3

1.1.3. Cryptococcosis ..............................................................................3

1.1.4. Pathogenesis and host immune response ..................................... 4

1.1.5. Cryptococcal virulence factors ...................................................... 6

1.2. Carbon metabolism and virulence in C. neoformans ........................... 7

1.2.1. Glucose metabolism: Glycolysis and gluconeogenesis................... 8

1.2.2. Peroxisome and the glyoxylate-shunt ......................................... 10

1.2.3. Mitochondria: β-oxidation, respiration, and the electron


transport chain (ETC) ..................................................................................... 11

iv
1.2.4. The pentose phosphate pathway (PPP), intracellular NADPH and
trehalose metabolism .................................................................................... 13

1.3. Cryptococcal virulence: acetate and acetyl-CoA ................................ 16

1.3.1. Routes to acetyl-CoA .................................................................. 17

1.3.2. Routes to acetate ....................................................................... 18

1.3.3. Acetate kinase ............................................................................ 20

1.4. Exploring acetate metabolism in C. neoformans ............................... 21

TABLE ...................................................................................................... 23

FIGURES .................................................................................................. 24

REFERENCES ............................................................................................ 27

CHAPTER 2 ........................................................................................................ 34

Analysis of transcriptomic response in fungal pathogen Cryptococcus


neoformans grown on acetate and glucose as carbon sources .......................... 34

ABSTRACT ............................................................................................... 35

2.1. INTRODUCTION ................................................................................ 36

2.2. RESULT AND DISCUSSION ................................................................. 40

2.2.1. Chromosomal distribution of differentially expressed genes ...... 41

2.2.2. Differential expression of putative transcription factors ............. 41

2.2.3. Downregulation of genes in acetate-grown conditions ............... 42

v
2.2.4. Genes for acetate-metabolism, the glyoxylate cycle, and
peroxisomal activities were upregulated. ...................................................... 44

2.2.5. Upregulated genes bear the signature of nutrient-starvation. .... 46

2.2.6. Gluconeogenesis and certain anabolic pathways are also


upregulated. .................................................................................................. 49

2.2.7. Some upregulated genes are implicated in virulence and


pathogenesis. ................................................................................................ 51

2.3. CONCLUSION .................................................................................... 51

2.4. METHODS AND MATERIALS .............................................................. 55

2.4.1. RNA isolation and sequencing .................................................... 55

2.4.2. Transcriptomic analysis .............................................................. 56

TABLES .................................................................................................... 57

FIGURES .................................................................................................. 71

REFERENCES ............................................................................................ 73

CHAPTER 3 ........................................................................................................ 80

Biochemical characterization of acetate kinase from the fungal pathogen


Cryptococcus neoformans and its role in acetate production ............................ 80

ABSTRACT ............................................................................................... 81

3.1. INTRODUCTION ................................................................................ 82

3.2. METHODS AND MATERIALS .............................................................. 84

vi
3.2.1. Production and purification of CnAck ......................................... 84

3.2.2. Kinetic assays of CnACK .............................................................. 85

3.2.3. Identification of putative eukaryotic Ack .................................... 85

3.2.4. Construction of the phylogenetic tree ........................................ 86

3.3. RESULTS ........................................................................................... 87

3.3.1. Ack in C. neoformans is possibly unidirectional in its biological


role ................................................................................................................ 87

3.3.2. Majority of the detected eukaryotic Ack are of fungal origin ...... 89

3.3.3. Non-fungal Ack may be functionally distinct from fungal-Ack ..... 90

3.4. DISCUSSION ...................................................................................... 91

TABLES .................................................................................................... 96

FIGURES .................................................................................................. 98

REFERENCES .......................................................................................... 101

CHAPTER 4 ...................................................................................................... 104

Modularity hypothesis of fungal virulence ............................................ 104

4.1. Why we need a regulatory network-based understanding of the origin


of virulence in fungi ........................................................................................ 105

4.2. Fungal infections in mammals are rare, and virulence factors can have
niche-specific protective roles for fungi ........................................................... 107

vii
4.3. Stress response and virulence can be conceived as modules of a larger
conceptual network ........................................................................................ 109

4.4. The hypothesis and its implications ................................................ 111

TABLE .................................................................................................... 113

FIGURES ................................................................................................ 114

REFERENCES .......................................................................................... 116

CHAPTER 5 ...................................................................................................... 119

Screening modularity in transcriptional hierarchy requires undertaking


larger exploratory studies ............................................................................... 119

5.1. Transcriptional circuitry of acetate utilization ................................ 120

5.2. Possibility of an acetate switch....................................................... 121

5.3. Xfp-Ack pathway as a route of acetate production ......................... 122

5.4. The trio: Xfp-Pta-Ack ...................................................................... 122

5.5. Gene co-expression network in C. neoformans ............................... 124

REFERENCE ............................................................................................ 125

APPENDIX A .......................................................................................... 126

RNA-sequencing data of Cryptococcus neoformans ............................ 126

APPENDIX B .......................................................................................... 148

Putative acetate kinases of eukaryotic origin ..................................... 148

viii
LIST OF TABLES

Table 1.1. Ten species of Cryptococcus............................................................... 23

Table 2.1. GO term enrichment of upregulated genes ........................................ 57

Table 2.2. GO term enrichment of downregulated genes ................................... 58

Table 2.3. KEGG pathway enrichment of upregulated genes............................... 59

Table 2.4. KEGG pathway enrichment of downregulated genes .......................... 60

Table 2.5. List of upregulated genes................................................................... 61

Table 2.6. List of downregulated genes .............................................................. 68

Table 2.7. Upregulated cryptococcal genes with reported association with


pathogenicity .................................................................................................... 69

Table 3.1. Kinetics of CnAck for common divalent metal cofactors ..................... 96

Table 3.2. Specific activities of CnAck for common NTPs..................................... 96

Table 3.3. Kinectic parameters of CnAck ............................................................ 97

Table 4.1. Common fungal infections in mammals ........................................... 113

ix
LIST OF FIGURES

Figure 1.1. Pathways of central carbon metabolism of C. neoformans ................ 24

Figure 1.2. Routes to acetate and acetyl-CoA synthesis in C. neoformans ........... 26

Figure 2.1. Effects of either acetate or glucose as sole carbon source on


transcriptome of C. neoformans......................................................................... 71

Figure 2.2. Chromosomal distribution of the proportion of differentially expressed


genes in C. neoformans...................................................................................... 72

Figure 3.1. Maximum likelihood tree of 526 eukaryotic Ack sequences ............... 98

Figure 3.2. Phylogenetic of eukaryotic Ack ......................................................... 99

Figure 3.3. Proposed XFP2/ACK pathway for acetate production in C. neoformans


....................................................................................................................... 100

Figure 4.1. Stress-response and virulence traits in fungi can be thought of as


separate modules of the same regulatory networks ......................................... 114

Figure 4.2. Modularity in biological networks ................................................... 115

x
CHAPTER 1

Cryptococcal virulence in the context of acetate

metabolism

Oly Ahmed

Department of Genetics and Biochemistry; Eukaryotic Pathogen Innovation Center,

Clemson University, SC, USA

1
1.1. Cryptococcus and cryptococcosis

1.1.1. The genus Cryptococcus

The organism that is the focus of this research is Cryptococcus neoformans,

an opportunistic fungal pathogen of humans. The genus Cryptococcus has the

potential to generate profuse confusion in the literature for historical reasons. One

example is the scale insect Cryptococcus fagisuga. Feeding behavior of this insect

leads to beech bark disease in American beech (Fagus grandifolia) (Ćalić et al., 2017).

Or, consider for example many other basidiomycetes fungi that are classified as

Cryptococcus based on morphological data, although modern molecular phylogeny

shows their paraphyletic relationship with C. neoformans. These paraphyletic

‘cryptococci’ still retain the generic epithet, now called basionym, in the literature,

but are no longer incorporated into the genus Cryptococcus (Liu et al., 2015). To date,

ten species of Cryptococcus have been reported, seven of which have pathogenic

potential for humans (Table 1.1).

Another problematic relic of the older literature is the differential

nomenclature of teleomorphic (sexually reproductive) and anamorphic forms

(asexual form) of the same fungus; Cryptococcus (anamorph) is also known as

Filobasidiella (teleomorph) in the literature. Fortunately, this practice was officially

discontinued in 2013 (Hawksworth, 2011), and the anamorphic name Cryptococcus

has been adopted as the generic name. Other confounding basionyms have been

renamed based on new molecular data. However, entomologists carry on to date

calling a genus of insect Cryptococcus.

2
1.1.2. Distribution and morphology

C. neoformans is not an obligate pathogen, rather an environmental species

with worldwide distribution. This species has been isolated from niches as diverse as

decaying materials in the soil and tree-hollows, and bird guano. It has also been

observed to proliferate inside soil amoebae and nematodes (May et al., 2016).

For the majority of its life cycle, C. neoformans is found in yeast forms;

however, as a true dimorphic species, switching between the yeast and the hyphal

morphotypes occurs regularly. This morphotypic switching has clinical importance

as the hyphal form is more immunogenic than the yeast form in mammalian hosts,

and spores are the infectious propagules that arise from sexual reproduction (both

bisexual and unisexual modes of reproduction are seen in C. neoformans). An

extensive review of the sexual reproduction and morphotypic switching in C.

neoformans is provided by (Zhao et al., 2019).

1.1.3. Cryptococcosis

Cryptococcosis is a generic term for mycosis caused by the two known

pathogenic species of the genus Cryptococcus: C. neoformans and C. gattii (Maziarz

and Perfect, 2016). Like almost every other fungal infection of mammals,

cryptococcosis can be localized on the skin (primary cutaneous cryptococcosis);

however, unlike most fungal infections, skin can be a secondary site of disseminated

cryptococcosis (Noguchi et al., 2019). Other rare secondary organs of disseminated

cryptococcosis include the eye and the prostate (Maziarz and Perfect, 2016).

3
Disseminated cryptococcosis usually follows successful initial infection of the

lungs. Infectious propagules such as desiccated yeasts or spores enter the lungs

through inhalation where successful evasion of the surveillance by the host

pulmonary components of both the innate and the adaptive immune systems may

lead to hematogenous disseminations to secondary organs (Setianingrum et al.,

2019). However, direct dissemination is exceedingly rare in immunocompetent

people, and hence cryptococcosis is typically classified as an opportunistic fungal

infection.

A relatively more usual route of infection is the reactivation of the latent form

of cryptococcosis. Childhood infection may be controlled by intact immunity and go

into dormancy, which may later resurface and disseminate if the person becomes

immunocompromised (e.g. HIV infection, etc.) or immunosuppressed (e.g. bone-

marrow transplant, etc.) later in life (Alanio, 2020). The most lethal form of

cryptococcosis is the infection of the central nervous system (CNS) that invariably

results in subacute meningoencephalitis, dubbed cryptococcal meningitis

(Williamson et al., 2017).

1.1.4. Pathogenesis and host immune response

Establishment of infection is thought to be initiated by inhalation of

desiccated yeast and/or spore of C. neoformans, where the small size of these

infectious propagules (e.g. spores, 1-3 µm; yeast form, ~3 µm; growing yeast cell, 4-10

µm) facilitates the deposition in deep alveoli (Rathore et al., 2022). The initial

response to the alveolar deposition of infectious propagules is phagocytosis of

cryptococcal cells by alveolar macrophages (AM) (Giles et al., 2009; Wang et al., 2022)

4
and opsonization following the detection of spores through binding with specific

pattern-recognition receptors (PRRs) (Romani, 2004; Wang et al., 2022).

Encapsulated yeast cells can deflect recognition and opsonization owing to low

immunogenicity of cryptococcal capsules, a virulence factor (see below). Moreover,

phagocytosis of cryptococcal cells is not guaranteed to lead to clearance as the

fungal cells are capable of proliferating inside the host macrophage as a facultative

intracellular pathogen (Wang et al., 2022). For effective clearance of the cryptococcal

cells both the innate and adaptive components of immunity are important, as

defective cell-mediated immune response leads to failure of complete clearance

(Wang et al., 2022).

Successful evasion of host immune response by C. neoformans in the lungs

can lead to hematogenous dissemination. In the bloodstream, the complement

system functions as the early humoral response against the migrating cryptococcal

cells; antibodies and B and T cell-mediated responses emerge afterwards (an

extensive review is provided by Wang et al., 2022). Secondary infection of the CNS

requires the migrating cryptococcal cells to cross the blood-brain barrier (BBB), a

layer of endothelial cells with tight junctions and selective permeability around the

brain. C. neoformans has been reported to employ the following processes to cross

the BBB: (1) paracytosis, or going around the cells through the tight junctions, where

the secreted cryptococcal proteins Mpr1 (metalloprotease) and urease have been

implicated in the process (Vu et al., 2014; Shi et al., 2010); (2) transcytosis, or traversing

the BBB through the cytoplasm of the endothelial cells (Chang et al., 2004); (3) a

Trojan horse strategy, where the engulfed, but intact, cryptococcal cells are carried

across the BBB via phagocytes and released non-lytically (Charlier et al., 2009; May et

5
al., 2016); and (4) direct invasion of the BBB via the enzymatic action of various

secreted cryptococcal proteins such as phospholipases and ureases (Rathore et al.,

2022).

1.1.5. Cryptococcal virulence factors

Traditionally the term “virulence factor” has been used loosely to refer to any

(1) morphological characteristics, (2) phenotypes, or (3) genes and gene products,

that are found to be implicated in pathogenesis. As such a comprehensive, and

uncontroversial, list of virulence factors is difficult to produce. A well-rounded review

of our current knowledge of cryptococcal virulence factors can be found here

(Zaragoza, 2019).

The ability of C. neoformans strains to grow at physiologically relevant

temperature is presumably the most important virulence factor as the mammalian

host’s core temperature creates an effective physical barrier to fungal growth in vivo

(Casadevall, 2012). Other physiologically relevant characteristics implicated in

virulence involve growth adaptations to different pH environments, tolerance to

oxidative and nitrosative stresses, and acquisition of important metals such as iron

(Zaragoza, 2019).

In addition to the above mentioned phenotypic virulence factors, some

virulence factors are gene products that are secreted by the fungus in the

surrounding host-tissue environment for nutrient acquisition (e.g. proteases and

lipases) and immunomodulation (e.g. urease and phospholipase B; see Wang et al.,

2022).

6
The most studied virulence factors are the morphological characteristics

unique to C. neoformans: polysaccharide capsule, melanin formation, and anti-

phagocytic protein 1 (App1). The unique cryptococcal capsule helps prevent

phagocytosis, complement and antibody opsonization, and thus contributes to

persistence and immune-evasion; tissue-specific factors such as temperature, pH,

CO2 and iron concentrations, all influence the capsule synthesis within mammalian

hosts (Momin and Webb, 2021).

The virulence factor melanin is relatively typical of fungal pathogens; in C.

neoformans melanization is thought to be involved in protection against oxidative

and nitrosative challenges from phagocytic activities of immune cells. Additionally,

the unique secreted protein App1 is known to inhibit phagocytosis via complement-

mediated mechanisms (Momin and Webb, 2021).

1.2. Carbon metabolism and virulence in C. neoformans

Our current understanding of carbon metabolism in C. neoformans is based

on extensive studies done in model organism Saccharomyces cerevisiae. However,

S. cerevisiae differs from C. neoformans in two important aspects. First, it is an

ascomycete fungus unlike Cryptococcus spp., which are basidiomycete fungi.

Second, S. cerevisiae shows Crabtree effect, which is the ability to carry on

fermentation under aerobic conditions provided a critical threshold of glucose

concentration is met. Despite such qualifications, searching for analogous metabolic

pathways in C. neoformans proved useful in discovering homologous genes and

describing nutritional conditions that are implicated in cryptococcal virulence

(Figure 1.1).

7
Historically, the importance of carbon metabolism in C. neoformans during

the disease progression was inferred from many transcriptional studies done under

various in vivo and in vitro conditions using a wide range of techniques including

differential display RT-PCR (ddRT-PCR), serial analysis of gene expression (SAGE), and

microarray (Kronstad et al, 2012; Hu et al, 2008; Steen et al, 2003; Rude et al, 2002).

From such studies one pattern became apparent: C. neoformans employs metabolic

reprogramming at different stages of the disease progression depending on the

host-tissue microenvironments. (For an earlier review see Kronstad et al., 2012.) In

recent years techniques such as RNA-seq improved the resolution of such

transcriptomic studies. For example, comparing the transcriptomes of clinical

isolates from the cerebrospinal fluid (CSF) of 31 cryptococcal meningitis patients with

that of the fungi grown on rabbit CSF, artificial CSF, and carbon-rich in vitro

conditions highlighted the importance of glycolysis, the TCA cycle, oxidative

phosphorylation, and sugar transport in cryptococcal meningitis and the dynamic

adaptability of this infectious fungus in the central nervous system of human hosts

(Yu et al., 2021).

1.2.1. Glucose metabolism: Glycolysis and gluconeogenesis

The mechanism of glucose sensing in C. neoformans is not understood in its

totality to date. Hexose transporters Hxs1 and Hxs2 in C. neoformans are described to

be important for glucose uptake, and expression of HXS1 is reduced in high glucose

concentration and important for growth in low glucose conditions (Liu et al., 2013).

However, the expression of HXS2 is not regulated by glucose, suggesting the role of

Hxs1 in glucose sensing and regulation of glucose uptake in C. neoformans.

8
Moreover, HXS1 is shown to be required for virulence in murine inhalation models

(knock-out experiment), indicating the importance of regulation of glucose

metabolism during pulmonary infection (Liu et al., 2013).

The impact of glucose utilization, as reflected in response to glucose

starvation in virulence, is also highlighted by the positive association observed

between incremental glucose starvation in media and increased expression of the

gene for antiphagocytic protein 1 (APP1). App1 is a unique cryptococcal virulence

protein involved in resistance to phagocytosis by macrophages (Williams and Del

Poeta, 2011).

The impact of glycolysis on cryptococcal virulence has been studied utilizing

deletion mutants. Two C. neoformans mutants defective in glycolytic pathway either

at the first step (hxk1Δ/hxk2Δ; hexokinases) or the last step (pyk1Δ; pyruvate kinase)

both show attenuated virulence in mouse models; however, there remains a

distinction between the ATP-generation capabilities of these two strains in ex vivo

human CSF. ATP production was found to be unaffected in the hxk1Δ/hxk2Δ mutant,

but was severely reduced in pyk1Δ mutant, suggesting that different mechanisms of

virulence attenuations are at play depending on which phase in glycolysis was

blocked (Price et al., 2011).

In parallel to glycolysis, the impact of gluconeogenesis on virulence has been

studied in mutant strains lacking phosphoenolpyruvate carboxykinase (PCK1), one of

four dedicated gluconeogenic enzymes. The C. neoformans pck1Δ strain can grow on

glucose rich media; however, as expected, it has reduced growth on acetate and the

3-carbon source lactate (Panepinto et al., 2005). Moreover, this mutant is severely

attenuated in a mouse model (Panepinto et al., 2005). In a later study, although the

9
PCK1 gene was observed to be highly expressed in the low-glucose environment of

rabbit CSF, deletion of PCK1 did not seem to have any impact on the persistence of

the fungus in the rabbit CSF model (Price et al., 2011).

1.2.2. Peroxisome and the glyoxylate-shunt

Peroxisomes, a single membraned, virtually universal eukaryotic organelle, is

involved in diverse cellular functions, including, but not limited to, β-oxidation, the

glyoxylate-shunt, and reactive oxygen species (ROS) metabolism. Generation of

acetyl-CoA and carbohydrates for cell wall synthesis and maintenance of turgor

pressure via peroxisomal metabolic pathways, as well as synthesis of melanin and

certain mycotoxins, implicated this organelle in virulence in the context of plant-

fungal pathogenicity (Falter and Reumann, 2022).

Peroxisome biogenesis requires the involvement of proteins known as

peroxins (or, PEX proteins). Knock-out mutants for PEX1 and PEX6 in C. neoformans

were shown to have reduced growth on fatty acid diets; however, these mutants

can grow on acetate as sole carbon source, implying a heavy reliance on

peroxisomes for fatty acid degradation in C. neoformans (Idnurm et al., 2007). One

unusual feature of C. neoformans, in contrast to other known fungi, is its apparent

need of intact peroxisomal functions for growth on simple sugars such as glucose,

fructose, and mannose (Idnurm et al., 2007). Surprisingly, PEX1 mutants were found

to be as virulent as the wild-type when tested in mouse models (Idnurm et al., 2007).

As noted above, C. neoformans defective in peroxisomal functions can still

grow on acetate as sole carbon source, suggesting plausible extra-peroxisomal

mechanisms of acetate utilization. However, disrupting the glyoxylate-shunt via

10
deletion of isocitrate lyase (ICL1), a gene encoding one of the two unique enzymes,

renders the fungus unable to grow on oleic acid or acetate (Rude et al., 2002).

Additionally, ICL1 is also shown to be upregulated by acetate or ethanol (2-carbon

metabolites), and suppressed at high concentration of glucose (Rude et al., 2002).

Deletion of the gene encoding malate synthase (MLS1), the other glyoxylate-shunt

enzyme, also leads to growth defects on acetate, highlighting the essentiality of the

glyoxylate-shunt in acetate utilization (Idnurm et al., 2007). However, despite the

importance of glyoxylate-shunt in acetate utilization, defects in this pathway do not

seem to impact virulence (Idnurm et al., 2007).

1.2.3. Mitochondria: β-oxidation, respiration, and the electron transport

chain (ETC)

In addition to peroxisomes, mitochondria also contribute to β-oxidation; hence

attempts were made early on to distinguish between the differential roles of these

two organelles in lipid metabolism and virulence. The Kronstad laboratory

(Kretschmer et al., 2012) created C. neoformans deletion strains for the enzymes

Mfe2 (multifunctional enzyme 2: peroxisomal β-oxidation), and Had1 and Had2

(Hydroxyacyl-CoA dehydrogenases: mitochondrial β-oxidation) and observed that

although had1Δ is as virulent as wild-type, mfe2Δ and mfe2Δ::had1Δ are hypovirulent

in mouse inhalation models, suggesting that peroxisomal β-oxidation might be

slightly more important for virulence than the mitochondrial version.

C. neoformans has been considered an obligate aerobe since the observation

that aeration of the culture media through rotation stimulates growth of the clinical

isolates tested, compared to static growth conditions, irrespective of the

11
concentration of glucose in the media (Odds et al., 1995). More than a decade later,

experimentation in CoCl2-based hypoxia-mimetic conditions with randomly

mutagenized (T-DNA insertions) C. neoformans strains showed growth-defect in

low-oxygen conditions demonstrated the essential connection between the

integrity of mitochondrial structure and function in respiration and ROS deactivation

and the survival of C. neoformans in low-oxygen environments (Ingavale et al., 2008).

This link is deemed important as oxygen levels available to C. neoformans for

respiration differs drastically among diverse environmental habitats and various

host-tissues. Partial pressure of O2 at sea level is 159 mmHg, while it is 104-108 mmHg

in alveoli, and 30-48 mmHg in the brain (Ortiz-Prado et al., 2019).

The importance of mitochondria in cryptococcal virulence has been

highlighted by some recent studies. Inhibition of either the classical (involving

proton efflux) or the alternative oxidase pathway (reduces oxygen without creating

proton gradient) of the electron transport chain (ETC) is found to be sufficient for

inhibition of capsule growth, one of the most salient cryptococcal virulence traits

(Trevijano-Contador et al., 2017). Additionally, this study also found that capsule

growth in C. neoformans seems positively associated with increased respiratory

activity in mitochondria, as reflected by upregulation of COX1 (cytochrome c oxidase;

electron transport), an increase in mitochondrial membrane potential, and a

consequent increase in ROS level (Trevijano-Contador et al., 2017). However, this

discovery is more perplexing considering that capsule synthesis is thought to be a

form of stress response stimulated by nutrient-starvation, and yet seems to be

heavily reliant on energy generation via the ETC.

12
Corroborating evidence of the requirement of the ETC in C. neoformans

survival came from study of the chaperone protein Mrj1 (mitochondrial respiration J-

domain protein 1). Mrj1 functions in respiration through its interaction with Complex

III of the ETC in the mitochondrion. The mrj1Δ and non-functional mrj1 (single amino

acid alteration) strains show growth defects at physiological temperature and on

alternative carbon sources, and are deficient in capsule elaboration. Moreover, mrj1Δ

strain is also avirulent in the mouse model (Horianopoulos et al., 2020).

In addition to virulence, mitochondrial activities may also moderate drug

resistance in C. neoformans. For example, it has recently been reported that low

glucose increases fluconazole tolerance in C. neoformans (Bhattacharya et al., 2021).

The proposed mechanism of fluconazole resistance involves transcriptional

modulation induced by low-glucose condition that leads to increased activities of

efflux proteins, enzymes of β-oxidation, and SOD (superoxide dismutase; ROS

inactivation pathway). Increased β-oxidation and SOD activities modify

mitochondrial functions causing increase in intracellular ATP concentration, which

in turn accelerates fluconazole efflux via actions of efflux proteins (Bhattacharya et

al., 2021).

1.2.4. The pentose phosphate pathway (PPP), intracellular NADPH and

trehalose metabolism

The disaccharide trehalose, known for its role in stress response in fungi, is

intricately linked to glucose metabolism. Genes of the trehalose biosynthetic

pathway have been studied for their involvement in the metabolism, cell wall

homeostasis, stress response and virulence in few pathogenic fungal species

13
(Thammahong et al., 2017). Not surprisingly, mutants in trehalose 6-phosphate

synthase (TPS1) of both the model fungus S. cerevisiae and the rice-blast fungus

Magnaporthe grisea show growth inhibition on glucose. However, M. grisea tps1Δ

showed a variable growth phenotype on glucose based on what other nitrogen

sources are available. No growth defect was seen on ammonium, nitrite, and amino

acids (except cysteine), but growth was impaired on nitrate, suggesting a link

between nitrate utilization and glucose metabolism (Wilson et al., 2007).

Additionally, in tps1Δ, the nitrogen metabolite repressor 1 (NMR1) was upregulated,

whereas nitrite and nitrate reductases were downregulated when grown on nitrate

(N-source) with glucose (C-source). In parallel, tps1Δ mutants showed an increased

accumulation of glucose-6-phosphate, but reduction of glucose-6-phosphate in

hxk1Δ (i.e. defective in glycolysis) and hxk1Δ::tps1Δ mutants when grown on glucose.

These observations led the authors to propose that TPS1 is presumably connecting

the carbon and nitrogen metabolism via influencing the intracellular level of NADPH

generated through the action of G6Pdh (glucose 6-phosphate dehydrogenase) in

the oxidative phase of PPP (Wilson et al., 2007; Masi et al., 2021).

Traditionally, the role of trehalose metabolism in fungal pathogenesis has

been studied from the perspective of trehalose’s involvement in germination/

sporulation and as stress response/ protectant, as such studies involving C.

neoformans were no different. Defect in trehalose synthesis via deletion of TPS1 in C.

neoformans has been shown to cause virulence attenuation and faster clearance

(i.e., reduced persistence) when tested both for persistence in an

immunosuppressed rabbit model and survival in a mouse inhalation model (Petzold

et al., 2006). On the other hand, the impact of trehalose utilization on cryptococcal

14
virulence was tested on a mice tail-vein model using null mutants of neutral

trehalase (NTH) — nth1Δ, nth2Δ, and nth1Δ::nth2Δ — and was observed that although

nth1Δ showed little defect in virulence, both nth2Δ and nth1Δ/nth2Δ were

hypervirulent. Moreover, deletion of NTH1 and NTH2 led to a defect in sporulation as

well (Botts et al., 2014).

As the example above of Magnaporthe illustrates, a plausible link between

carbon and nitrogen metabolism through cellular NADPH has the potential to

confound studies aimed at elucidating the effect of trehalose synthesis on fungal

virulence. This possibility becomes more likely considering the studied effect of

NADPH generating pathways on virulence in C. neoformans. There are at least three

different pathways for NADPH generation in C. neoformans: (1) the oxidative phase

of the PPP involving G6Pdh and 6Pgd (6-phosphogluconate dehydrogenase), (2)

alternative oxidative PPP involving Gnk (gluconate kinase), and (3) Idp (NADP+-

dependent isocitrate dehydrogenase). An earlier study reported that deletion of

G6PDH (zwf1Δ) does not seem to alter the sensitivity of C. neoformans to

temperature or oxidative stress, and intracellular concentration of NADPH appears

to stay constant, implying other active routes of NADPH generation (Brown et al.,

2010). Later, a different study reported that deletion of GNK (alternative oxidative

PPP) severely impairs sporulation (Jezewski et al., 2022). Finally, deletion of IDP in C.

neoformans leads to sensitivity to high temperature, a virulence factor, and

nitrosative stress (Brown et al., 2010). Taken together, the role of NADPH in cellular

and reproductive fitness, and in carbon and nitrogen metabolism might be

combined conceptually to better guide our future studies of cryptococcal virulence.

15
1.3. Cryptococcal virulence: acetate and acetyl-CoA

An NMR-based study of a rat pulmonary cryptococcoma model, an induced

infection of C. neoformans, revealed a high abundance of trehalose, mannitol,

glycerophosphorylcholine, ethanol, and acetate in cryptococcomas compared to

healthy lung tissue (Himmelreich et al., 2003). This study sparked the interest of the

community to explore the connection between acetate and cryptococcosis. Later a

serial analysis of gene expression (SAGE) study of a murine pulmonary infection

model suggested the potential involvement of transcriptional activation of acetyl-

CoA metabolism in C. neoformans during the early stages of pulmonary

cryptococcosis (Hu et al., 2008). The same study also found that deletion of ACS1, the

enzyme responsible for conversion of acetate to acetyl-CoA, leads to growth

impairment on acetate, and a slight delay in causing disease in the murine

inhalation model, presumably due to compensatory actions of other acetyl-CoA

generating pathways active in C. neoformans (Hu et al., 2008).

These earlier studies raised further questions about the utility of

acetate/acetyl-CoA metabolism in C. neoformans virulence; is acetate production

(and subsequent secretion) and/or is acetate utilization important for pathogenicity?

Both of these possibilities seem plausible in the context of cryptococcosis. For

example, acetate production and secretion may explain the observed acidic pH of

cryptococcoma in vivo (Wright et al., 2002); however, this acidification is unlikely to

be an effective defense mechanism against the cryptococcoma-clearance

(analogous to tumor-clearance) by natural killer (NK) cells of the innate immune

system (Islam et al., 2013). On the other hand, acetate utilization may be one of the

major pathways necessary for survival in the glucose-limited microenvironments of

16
host-tissues. For context, pulmonary airway surface liquid (ASL) contains roughly 10-

fold lower glucose than serum (Philips et al., 2003; Ries et al., 2018), whereas this

amount is 60-70% of the serum-level in CSF (Seehusen et al., 2003; Ries et al., 2018).

However, acetate utilization vs. acetate production/secretion does not have to

be a false dichotomy as appears to be here in the preceding discussion. The stages

of disease progression, tissue microenvironments, and interaction with the

components of the immune system are likely to wildly influence the state of acetate

metabolism in C. neoformans during infection. A regulated switching between

production and utilization of acetate could be co-opted for adaptation to the

constantly changing nutrient pool within the host body during infection; equivalent

systems of “acetate switch” have been documented first in prokaryotes (e.g. see the

classic Wolfe, 2005) and recently in an archaeon (Kuprat et al., 2021).

1.3.1. Routes to acetyl-CoA

As mentioned above, ACS1 deletion has mild impact on cryptococcal virulence

in a murine inhalation model (Hu et al., 2008), and the authors suggested that

defects in acetyl-CoA synthesis via ACS1 may have been compensated by other

routes of acetyl-CoA synthesis. Subsequently, another route of acetyl-CoA synthesis

was investigated by Griffiths et al. (Griffiths et al., 2012). Deletion of ACL1, encoding

ATP-citrate lyase which synthesizes acetyl-CoA from cytoplasmic pool of citrate,

leads to decreased survival of C. neoformans within murine macrophages, increased

susceptibility to fluconazole, and virulence defects in a murine inhalation model

(Griffiths et al., 2012).

17
Recently, another direct route of acetyl-CoA synthesis from acetoacetate, an

end product of various amino acid degradation pathways, has been discovered.

KBC1 encodes 3-keto-butanoyl-CoA synthetase which converts acetoacetate to

acetoacetyl-CoA, which can be further converted to acetyl-CoA via the action of

acetoacetyl-CoA lyases (Alden et al., 2022). Although the KBC1 knock-out does not

show any discernible defect in virulence, mice infected with acs1Δ::kbc1Δ double

mutant were found to have reduced CNS burden compared to those infected with

either single mutant or wild-type strain (Alden et al., 2022).

1.3.2. Routes to acetate

To date, there are at least two pathways identified in C. neoformans that can

potentially contribute to the intracellular acetate-pool: (1) conversion of pyruvate to

acetate (analogous to that in S. cerevisiae; see Orlandi et al., 2014 for an introduction);

and (2) conversion of phosphoketose sugar to acetate (analogous to that in bacteria,

see Henard et al., 2015 for an introduction); (Figure 1.2). The first pathway involves

decarboxylation of pyruvate into acetaldehyde by pyruvate decarboxylase (Pdc),

followed by oxidation of acetaldehyde to acetate by aldehyde dehydrogenase (Ald).

The second pathway is relatively rare in eukaryotic organisms, and involves first the

breakdown of xylulose 5-phosphate (or fructose 6-phosphate) to glyceraldehyde 3-

phosphate (or erythrose 4-phosphate) and acetyl phosphate using inorganic

phosphate catalyzed by xylulose 5-phosphate/ fructose 6-phosphate

phosphoketolase (Xfp), followed by conversion of acetyl phosphate to acetate via the

action of acetate kinase (Ack).

18
The Pdc-Ald pathway is known as Pyruvate-Acetaldehyde-Acetate (PAA)

pathway, and has been described in many fungi as a route of fermentation. In S.

cerevisiae, PAA pathway can lead to increased intracellular acetate, which in turn

may reduce ethanol utilization under respiration condition, as a form of catabolite

repression (Simpson-Lavy and Kupiec, 2019). Similarly, in pathogenic fungus

Gibberella zeae, the causative agent of head blight disease in cereals, the PAA

pathway is thought to play an important role in lipid synthesis in the aerial mycelia

(Son et al., 2012). These examples highlight the diverse biological roles that PAA

pathway plays in fungi.

It has been observed that many fungal genomes contain genes for Xfp and

Ack, raising the possibility of a potential pathway connecting xylulose metabolism

with acetate production (Ingram-Smith et al., 2006). Little is known about the

biological significance of the Xfp-Ack pathway, dubbed the phosphoketolase

pathway, in fungi. Metabolic flux analysis in a phosphoketolase (XFP) overexpression

mutant of the filamentous fungus Aspergillus nidulans suggests that this pathway

might have evolved to provide added flexibility to the central carbon metabolism

(Panagiotou et al., 2008).

The relative importance of the above-mentioned routes to acetate synthesis

in C. neoformans cellular metabolism is not fully understood. However, a case can be

argued for a metabolic scheme that links the Xfp-Ack pathway to the oxidative

phase of the PPP. In non-dividing cells, metabolic needs such as fatty acid synthesis

(growth and membrane integrity) and glutathione metabolism (oxidative damage)

create high demands for NADPH regeneration without the excess supply of ribose 5-

phosphate through the canonical PPP. In these situations the Xfp-Ack route can act

19
as a siphoning mechanism by utilizing the sugar phosphate intermediates and

thereby essentially decoupling the NADPH regeneration from excess ribose 5-

phosphate build-up. Additionally, such siphoning mechanism also implies the

existence of a possible regulatory system between the PPP and acetate synthesis via

Xfp-Ack route.

1.3.3. Acetate kinase

From our discussion above it follows that acetate kinase, Ack, is likely to play

important metabolic roles in addition to acetate metabolism in C. neoformans, and

presumably in other eukaryotic species as well. In fact, recently it has been proposed

that Ack in the amoebic parasite Entamoeba histolytica, a eukaryote, might not be

as important in acetate production as it is in maintaining intracellular NAD+/NADH

ratio during ethanol production via glycolysis (Dang et al., 2022).

Compared to the Ack of bacterial and archaeal origins, our knowledge of

eukaryotic Ack is limited. Eukaryotic Ack of E. histolytica differs from Acks of well-

studied bacterial genera (e.g. Escherichia, Bacillus, Salmonella etc.) and archaeal

genus Methanosarcina (see Ferry, 2011) in that it requires inorganic pyrophosphate

(PPi), rather than ATP, as phosphoryl donor (Tielens et al., 2010). Furthermore, in free-

living alga Chlamydomonas reinhardtii, ACK is expressed in aerobic condition and

Ack is localized in mitochondria (Atteia et al., 2006). However, transferring this alga to

an anaerobic, heterotrophic culture condition seems to induce transcription of

another ACK predicted to be localized in chloroplast (designated as ACK1), and the

accumulation kinetics of this transcript appears to correspond with acetate

accumulation in this species (Dubini et al., 2009; Tielens et al., 2010). Apart from such

20
kinetic or expression studies, eukaryotic Ack has largely remained understudied. The

structure of Ack enzymes from E. histolytica and C. neoformans have been resolved,

and it was observed that substrate specificity difference between these eukaryotic

Ack (i.e. ATP versus PPi) may be attributed to subtle amino acid changes in the active

site without drastically altering the backbone structure of these proteins (Thaker et

al., 2013). These structures of eukaryotic Ack are also observed to be consistent for

the reaction mechanism of direct in-line transfer of phosphoryl group proposed for

prokaryotic Acks (Thaker et al., 2013). Here it is worth mentioning that in C.

reinhardtii phosphotransacetylase (Pta) is assumed to form a pathway of acetate

production or assimilation with Ack analogous to that found in many bacterial

species, whereas in E. histolytica traditional partner of Ack such as Pta ro Xfp of

acetate pathway are absent (Ingram-Smith et al., 2006).

1.4. Exploring acetate metabolism in C. neoformans

In this work, I set out to explore the acetate metabolism in the fungal

pathogen C. neoformans. I began by asking how acetate utilization as the sole

carbon source brings about changes in the transcriptional landscape of this species

compared to that during glucose utilization. C. neoformans KN99 strains were

grown in identical condition except for the addition of two different carbon sources

in the media. RNA-seq experiments were performed on the fungus following growth

in these conditions to measure the changes in the expression levels of thousands of

cryptococcal genes (Chapter 2).

In Chapter 3, I describe the in vitro kinetic characteristics of C. neoformans

Ack, isolated following heterologous expression in Escherichia coli, to better

21
understand the role of Ack in C. neoformans. This work was also extended to

incorporate in silico identification of ACK in currently available genome sequences of

eukaryotic organisms and phylogenetic reconstruction of these putative eukaryotic

ACK genes to glean the evolutionary history of this enzyme within the domain

Eukarya.

Chapter 4 includes a synthesis of evolutionary network biology with fungal

pathogenesis. Here I have proposed the modularity hypothesis to explain the

emergence of virulence traits in present day fungi that enables them to cause

accidental pathogenicity in mammalian hosts. This expository piece introduces the

modularity hypothesis for the first time as an explanatory framework to describe

evolution of fungal pathogenicity, as well as a pragmatic framework to guide

research programs aimed at efficient discovery of molecular pathways for antifungal

drug targeting.

Finally, in Chapter 5, I conclude this dissertation by summarizing my findings

and suggesting possible future avenues in our exploration of acetate metabolism in

the fungal pathogen Cryptococcus neoformans.

22
TABLE

Table 1.1. Ten species of Cryptococcus

Species complex1 Species Pathogenic to humans?

C. neoformans (sensu stricto) Yes


C. neoformans
C. deneoformans Yes

C. gattii (sensu stricto) Yes

C. bacillisporus Yes

C. gattii C. deuterogattii Yes

C. tetragattii Yes

C. decagattii Yes

ND C. depauperatus No

ND C. luteus No

ND C. amylolentus No

Note:

1. Modern phylogeny supports a ‘seven species’ view over ‘two species’ view of C.
neoformans/gattii complex (Hagen et al., 2015)

ND: not defined

23
FIGURES

Figure 1.1. Pathways of central carbon metabolism in C. neoformans. Schematic

showing prominent intermediates and gene products (enzymes and transporters) of

cryptococcal central carbon metabolism; genes studied to date and mentioned in

the text are shown in blue, italics. (caption continues into the next page)

24
(Continued from the last page) Abbreviations (metabolites): G6P: glucose-6-

phosphate; F6P: fructose-6-phosphate; Gly 3P: glyceraldehyde-3-phosphate; DHAP:

dihydroxyacetone phosphate; E4P: erythrose-6-phosphate; R5P: ribose-5-phosphate;

X5P: xylulose-5-phosphate; Ru5P: ribulose-5-phosphate; 6PG: 6-phosphogluconate;

6PG lactone: 6-phosphogluconolactone; PEP: phosphoenolpyruvate; Pyr: pyruvate;

OxAc: oxaloacetate. Abbreviations (gene products): HXS: hexose transporter; HXK:

hexokinase; G6PDH: glucose-6-phosphate dehydrogenase; 6PGD: 6-

phosphogluconate dehydrogenase; GNK: gluconate kinase; PYK: pyruvate kinase;

PCK: phosphoenolpyruvate carboxykinase; ACS: acetyl-CoA synthetase; ACL: ATP-

citrate lyase; ICL: Iso-citrate lyase; MLS: malate synthase; MFE2: multifunctional

enzyme 2; HAD: 3-hydroxyacyl-CoA dehydrogenase. Colors and shapes: Thick

magenta vertical line denotes separation between the outside (left) and the inside

(right) of the cell; rectangular box on the magenta line denotes glucose transporter;

solid arrows denote single step in the pathway; dashed arrows indicate multiple

intermediate steps in the pathway omitted from display for improved clarity; red

arrows: gluconeogenic steps; purple arrows: the TCA cycle steps; orange arrows:

steps of the glyoxylate shunt; green arrows: oxidative and alternative oxidative

phases of the pentose phosphate pathway.

25
Figure 1.2. Routes to acetate and acetyl-CoA synthesis in Cryptococcus

neoformans. Abbreviations: Ack: acetate kinase; Acs: acetyl-CoA synthase; Acl: ATP-

citrate lyase; Ald: aldehyde dehydrogenase; Kbc: 3-keto butanoyl-CoA synthetase;

Pdc: pyruvate decarboxylase; Pdh: pyruvate dehydrogenase; Xfp: xylulose 5-

phosphate/ fructose 6-phosphate phosphoketolase.

26
REFERENCES

Alanio, A., 2020. Dormancy in Cryptococcus neoformans: 60 years of accumulating


evidence. J. Clin. Invest. 130, 3353–3360. https://doi.org/10.1172/JCI136223

Alden, K.M., Jezewski, A.J., Beattie, S.R., Fox, D., Krysan, D.J., 2022. Genetic interaction
analysis reveals that Cryptococcus neoformans utilizes multiple acetyl-CoA-
generating pathways during infection. mBio 13, e01279-22.
https://doi.org/10.1128/mbio.01279-22

Atteia, A., van Lis, R., Gelius-Dietrich, G., Adrait, A., Garin, J., Joyard, J., Rolland, N.,
Martin, W., 2006. Pyruvate formate-lyase and a novel route of eukaryotic ATP
synthesis in Chlamydomonas mitochondria. J. Biol. Chem. 281, 9909–9918.
https://doi.org/10.1074/jbc.M507862200

Bhattacharya, S., Oliveira, N.K., Savitt, A.G., Silva, V.K.A., Krausert, R.B., Ghebrehiwet, B.,
Fries, B.C., 2021. Low glucose mediated fluconazole tolerance in Cryptococcus
neoformans. J. Fungi 7, 489. https://doi.org/10.3390/jof7060489

Botts, M.R., Huang, M., Borchardt, R.K., Hull, C.M., 2014. Developmental cell fate and
virulence are linked to trehalose homeostasis in Cryptococcus neoformans.
Eukaryot. Cell 13, 1158–1168. https://doi.org/10.1128/EC.00152-14

Brown, S.M., Upadhya, R., Shoemaker, J.D., Lodge, J.K., 2010. Isocitrate dehydrogenase
is important for nitrosative stress resistance in Cryptococcus neoformans, but
oxidative stress resistance is not dependent on glucose-6-phosphate
dehydrogenase. Eukaryot. Cell 9, 971–980. https://doi.org/10.1128/EC.00271-09

Ćalić, I., Koch, J., Carey, D., Addo-Quaye, C., Carlson, J.E., Neale, D.B., 2017. Genome-
wide association study identifies a major gene for beech bark disease
resistance in American beech (Fagus grandifolia Ehrh.). BMC Genomics 18,
547. https://doi.org/10.1186/s12864-017-3931-z

Casadevall, A., 2012. Fungi and the rise of mammals. PLOS Pathog. 8, e1002808.
https://doi.org/10.1371/journal.ppat.1002808

Chang, Y.C., Stins, M.F., McCaffery, M.J., Miller, G.F., Pare, D.R., Dam, T., Paul-Satyaseela,
M., Kim, K.S., Kwon-Chung, K.J., Paul-Satyasee, M., 2004. Cryptococcal yeast
cells invade the central nervous system via transcellular penetration of the
blood-brain barrier. Infect. Immun. 72, 4985–4995.
https://doi.org/10.1128/IAI.72.9.4985-4995.2004

Charlier, C., Nielsen, K., Daou, S., Brigitte, M., Chretien, F., Dromer, F., 2009. Evidence of
a role for monocytes in dissemination and brain invasion by Cryptococcus
neoformans. Infect. Immun. 77, 120–127. https://doi.org/10.1128/IAI.01065-08

27
Dang, T., Angel, M., Cho, J., Nguyen, D., Ingram-Smith, C., 2022. The role of acetate
kinase in the human parasite Entamoeba histolytica. Parasitologia 2, 147–159.
https://doi.org/10.3390/parasitologia2020014

Dubini, A., Mus, F., Seibert, M., Grossman, A.R., Posewitz, M.C., 2009. Flexibility in
anaerobic metabolism as revealed in a mutant of Chlamydomonas reinhardtii
lacking hydrogenase activity. J. Biol. Chem. 284, 7201–7213.
https://doi.org/10.1074/jbc.M803917200

Falter, C., Reumann, S., 2022. The essential role of fungal peroxisomes in plant
infection. Mol. Plant Pathol. 23, 781–794. https://doi.org/10.1111/mpp.13180

Ferry, J.G., 2011. Chapter eleven - Acetate kinase and phosphotransacetylase, in:
Rosenzweig, A.C., Ragsdale, S.W. (Eds.), Methods in Enzymology, Methods in
Methane Metabolism, Part A. Academic Press, pp. 219–231.
https://doi.org/10.1016/B978-0-12-385112-3.00011-1

Giles, S.S., Dagenais, T.R.T., Botts, M.R., Keller, N.P., Hull, C.M., 2009. Elucidating the
pathogenesis of spores from the human fungal pathogen Cryptococcus
neoformans. Infect. Immun. 77, 3491–3500. https://doi.org/10.1128/IAI.00334-09

Griffiths, E.J., Hu, G., Fries, B., Caza, M., Wang, J., Gsponer, J., Gates-Hollingsworth, M.A.,
Kozel, T.R., De Repentigny, L., Kronstad, J.W., 2012. A defect in ATP-citrate lyase
links acetyl-CoA production, virulence factor elaboration and virulence in
Cryptococcus neoformans. Mol. Microbiol. 86, 1404–1423.
https://doi.org/10.1111/mmi.12065

Hagen, F., Khayhan, K., Theelen, B., Kolecka, A., Polacheck, I., Sionov, E., Falk, R.,
Parnmen, S., Lumbsch, H.T., Boekhout, T., 2015. Recognition of seven species in
the Cryptococcus gattii/Cryptococcus neoformans species complex. Fungal
Genet. Biol., Cryptococcus: model basidiomycetes and deadly pathogens 78,
16–48. https://doi.org/10.1016/j.fgb.2015.02.009

Hawksworth, D., 2011. A new dawn for the naming of fungi: impacts of decisions
made in Melbourne in July 2011 on the future publication and regulation of
fungal names. MycoKeys 1, 7–20. https://doi.org/10.3897/mycokeys.1.2062

Henard, C.A., Freed, E.F., Guarnieri, M.T., 2015. Phosphoketolase pathway engineering
for carbon-efficient biocatalysis. Curr. Opin. Biotechnol., Pathway engineering
36, 183–188. https://doi.org/10.1016/j.copbio.2015.08.018

Himmelreich, U., Allen, C., Dowd, S., Malik, R., Shehan, B.P., Mountford, C., Sorrell, T.C.,
2003. Identification of metabolites of importance in the pathogenesis of
pulmonary cryptococcoma using nuclear magnetic resonance spectroscopy.
Microbes Infect. 5, 285–290. https://doi.org/10.1016/S1286-4579(03)00028-5

28
Horianopoulos, L.C., Hu, G., Caza, M., Schmitt, K., Overby, P., Johnson, J.D., Valerius, O.,
Braus, G.H., Kronstad, J.W., 2020. The novel J-domain protein Mrj1 is required
for mitochondrial respiration and virulence in Cryptococcus neoformans.
mBio 11, e01127-20. https://doi.org/10.1128/mBio.01127-20

Hu, G., Cheng, P.-Y., Sham, A., Perfect, J.R., Kronstad, J.W., 2008. Metabolic adaptation
in Cryptococcus neoformans during early murine pulmonary infection. Mol.
Microbiol. 69, 1456–1475. https://doi.org/10.1111/j.1365-2958.2008.06374.x

Idnurm, A., Giles, S.S., Perfect, J.R., Heitman, J., 2007. Peroxisome function regulates
growth on glucose in the basidiomycete fungus Cryptococcus neoformans.
Eukaryot. Cell 6, 60–72. https://doi.org/10.1128/EC.00214-06

Ingavale, S.S., Chang, Y.C., Lee, H., McClelland, C.M., Leong, M.L., Kwon-Chung, K.J.,
2008. Importance of mitochondria in survival of Cryptococcus neoformans
under low oxygen conditions and tolerance to cobalt chloride. PLoS Pathog. 4,
e1000155. https://doi.org/10.1371/journal.ppat.1000155

Ingram-Smith, C., Martin, S.R., Smith, K.S., 2006. Acetate kinase: not just a bacterial
enzyme. Trends Microbiol. 14, 249–253. https://doi.org/10.1016/j.tim.2006.04.001

Islam, A., Li, S.S., Oykhman, P., Timm-McCann, M., Huston, S.M., Stack, D., Xiang, R.F.,
Kelly, M.M., Mody, C.H., 2013. An acidic microenvironment increases NK cell
killing of Cryptococcus neoformans and Cryptococcus gattii by enhancing
perforin degranulation. PLoS Pathog. 9, e1003439.
https://doi.org/10.1371/journal.ppat.1003439

Jezewski, A.J., Beattie, S.R., Alden, K.M., Krysan, D.J., 2022. Gluconate kinase is
required for gluconate assimilation and sporulation in Cryptococcus
neoformans. Microbiol. Spectr. 10, e00301-22.
https://doi.org/10.1128/spectrum.00301-22

Kretschmer, M., Wang, J., Kronstad, J.W., 2012. Peroxisomal and mitochondrial β-
oxidation pathways influence the virulence of the pathogenic fungus
Cryptococcus neoformans. Eukaryot. Cell 11, 1042–1054.
https://doi.org/10.1128/EC.00128-12

Kronstad, J., Saikia, S., Nielson, E.D., Kretschmer, M., Jung, W., Hu, G., Geddes, J.M.H.,
Griffiths, E.J., Choi, J., Cadieux, B., Caza, M., Attarian, R., 2012. Adaptation of
Cryptococcus neoformans to mammalian hosts: Integrated regulation of
metabolism and virulence. Eukaryot. Cell 11, 109–118.
https://doi.org/10.1128/EC.05273-11

Kuprat, T., Ortjohann, M., Johnsen, U., Schönheit, P., 2021. Glucose metabolism and
acetate switch in Archaea: the enzymes in Haloferax volcanii. J. Bacteriol. 203,
e00690-20. https://doi.org/10.1128/JB.00690-20

29
Liu, T.-B., Wang, Y., Baker, G.M., Fahmy, H., Jiang, L., Xue, C., 2013. The glucose sensor-
like protein Hxs1 is a high-affinity glucose transporter and required for
virulence in Cryptococcus neoformans. PLoS ONE 8, e64239.
https://doi.org/10.1371/journal.pone.0064239

Liu, X.-Z., Wang, Q.-M., Göker, M., Groenewald, M., Kachalkin, A.V., Lumbsch, H.T.,
Millanes, A.M., Wedin, M., Yurkov, A.M., Boekhout, T., Bai, F.-Y., 2015. Towards an
integrated phylogenetic classification of the Tremellomycetes. Stud. Mycol. 81,
85–147. https://doi.org/10.1016/j.simyco.2015.12.001

Masi, A., Mach, R.L., Mach-Aigner, A.R., 2021. The pentose phosphate pathway in
industrially relevant fungi: crucial insights for bioprocessing. Appl. Microbiol.
Biotechnol. 105, 4017–4031. https://doi.org/10.1007/s00253-021-11314-x

May, R.C., Stone, N.R.H., Wiesner, D.L., Bicanic, T., Nielsen, K., 2016. Cryptococcus: from
environmental saprophyte to global pathogen. Nat. Rev. Microbiol. 14, 106–117.
https://doi.org/10.1038/nrmicro.2015.6

Maziarz, E.K., Perfect, J.R., 2016. Cryptococcosis. Infect. Dis. Clin. North Am., Fungal
Infections 30, 179–206. https://doi.org/10.1016/j.idc.2015.10.006

Momin, M., Webb, G., 2021. The environmental effects on virulence factors and the
antifungal susceptibility of Cryptococcus neoformans. Int. J. Mol. Sci. 22, 6302.
https://doi.org/10.3390/ijms22126302

Noguchi, H., Matsumoto, T., Kimura, U., Hiruma, M., Kusuhara, M., Ihn, H., 2019.
Cutaneous cryptococcosis. Med. Mycol. J. 60, 101–107.
https://doi.org/10.3314/mmj.19.008

Odds, F.C., De Backer, T., Dams, G., Vranckx, L., Woestenborghs, F., 1995. Oxygen as
limiting nutrient for growth of Cryptococcus neoformans. J. Clin. Microbiol. 33,
995–997.

Orlandi, I., Coppola, D.P., Vai, M., 2014. Rewiring yeast acetate metabolism through
MPC1 loss of function leads to mitochondrial damage and decreases
chronological lifespan. Microb. Cell 1, 393–405.
https://doi.org/10.15698/mic2014.12.178

Ortiz-Prado, E., Dunn, J.F., Vasconez, J., Castillo, D., Viscor, G., 2019. Partial pressure of
oxygen in the human body: a general review. Am. J. Blood Res. 9, 1–14.

Panagiotou, G., Andersen, M.R., Grotkjær, T., Regueira, T.B., Hofmann, G., Nielsen, J.,
Olsson, L., 2008. Systems analysis unfolds the relationship between the
phosphoketolase pathway and growth in Aspergillus nidulans. PLOS ONE 3,
e3847. https://doi.org/10.1371/journal.pone.0003847

30
Panepinto, J., Liu, L., Ramos, J., Zhu, X., Valyi-Nagy, T., Eksi, S., Fu, J., Jaffe, H.A., Wickes,
B., Williamson, P.R., 2005. The DEAD-box RNA helicase Vad1 regulates multiple
virulence-associated genes in Cryptococcus neoformans. J. Clin. Invest. 115,
632–641. https://doi.org/10.1172/JCI23048

Petzold, E.W., Himmelreich, U., Mylonakis, E., Rude, T., Toffaletti, D., Cox, G.M., Miller,
J.L., Perfect, J.R., 2006. Characterization and regulation of the trehalose
synthesis pathway and its importance in the pathogenicity of Cryptococcus
neoformans. Infect. Immun. 74, 5877–5887. https://doi.org/10.1128/IAI.00624-06

Philips, B.J., Meguer, J.-X., Redman, J., Baker, E.H., 2003. Factors determining the
appearance of glucose in upper and lower respiratory tract secretions.
Intensive Care Med. 29, 2204–2210. https://doi.org/10.1007/s00134-003-1961-2

Price, M.S., Betancourt-Quiroz, M., Price, J.L., Toffaletti, D.L., Vora, H., Hu, G., Kronstad,
J.W., Perfect, J.R., 2011. Cryptococcus neoformans requires a functional
glycolytic pathway for disease but not persistence in the host. mBio 2, e00103-
11. https://doi.org/10.1128/mBio.00103-11

Rathore, S.S., Sathiyamoorthy, J., Lalitha, C., Ramakrishnan, J., 2022. A holistic review
on Cryptococcus neoformans. Microb. Pathog. 166, 105521.
https://doi.org/10.1016/j.micpath.2022.105521

Ries, L.N.A., Beattie, S., Cramer, R.A., Goldman, G.H., 2018. Overview of carbon and
nitrogen catabolite metabolism in the virulence of human pathogenic fungi.
Mol. Microbiol. 107, 277–297. https://doi.org/10.1111/mmi.13887

Romani, L., 2004. Immunity to fungal infections. Nat. Rev. Immunol. 4, 1–23.
https://doi.org/10.1038/nri1255

Rude, T.H., Toffaletti, D.L., Cox, G.M., Perfect, J.R., 2002. Relationship of the glyoxylate
pathway to the pathogenesis of Cryptococcus neoformans. Infect. Immun. 70,
5684–5694. https://doi.org/10.1128/IAI.70.10.5684-5694.2002

Seehusen, D.A., Reeves, M.M., Fomin, D.A., 2003. Cerebrospinal fluid analysis. Am.
Fam. Physician 68, 1103–1109.

Setianingrum, F., Rautemaa-Richardson, R., Denning, D.W., 2019. Pulmonary


cryptococcosis: A review of pathobiology and clinical aspects. Med. Mycol. 57,
133–150. https://doi.org/10.1093/mmy/myy086

Shi, M., Li, S.S., Zheng, C., Jones, G.J., Kim, K.S., Zhou, H., Kubes, P., Mody, C.H., 2010.
Real-time imaging of trapping and urease-dependent transmigration of
Cryptococcus neoformans in mouse brain. J. Clin. Invest. 120, 1683–1693.
https://doi.org/10.1172/JCI41963

31
Simpson-Lavy, K., Kupiec, M., 2019. Carbon catabolite repression: not only for glucose.
Curr. Genet. 65, 1321–1323. https://doi.org/10.1007/s00294-019-00996-6

Son, H., Min, K., Lee, J., Choi, G.J., Kim, J.-C., Lee, Y.-W., 2012. Differential roles of
pyruvate decarboxylase in aerial and embedded mycelia of the ascomycete
Gibberella zeae. FEMS Microbiol. Lett. 329, 123–130. https://doi.org/10.1111/j.1574-
6968.2012.02511.x

Thaker, T.M., Tanabe, M., Fowler, M.L., Preininger, A.M., Ingram-Smith, C., Smith, K.S.,
Iverson, T.M., 2013. Crystal structures of acetate kinases from the eukaryotic
pathogens Entamoeba histolytica and Cryptococcus neoformans. J. Struct.
Biol. 181, 185–189. https://doi.org/10.1016/j.jsb.2012.11.001

Thammahong, A., Puttikamonkul, S., Perfect, J.R., Brennan, R.G., Cramer, R.A., 2017.
Central role of the trehalose biosynthesis pathway in the pathogenesis of
human fungal infections: Opportunities and challenges for therapeutic
development. Microbiol. Mol. Biol. Rev. MMBR 81, e00053-16.
https://doi.org/10.1128/MMBR.00053-16

Tielens, A.G.M., van Grinsven, K.W.A., Henze, K., van Hellemond, J.J., Martin, W., 2010.
Acetate formation in the energy metabolism of parasitic helminths and
protists. Int. J. Parasitol. 40, 387–397. https://doi.org/10.1016/j.ijpara.2009.12.006

Trevijano-Contador, N., Rossi, S.A., Alves, E., Landín-Ferreiroa, S., Zaragoza, O., 2017.
Capsule enlargement in Cryptococcus neoformans is dependent on
mitochondrial activity. Front. Microbiol. 8.

Vu, K., Tham, R., Uhrig, J.P., Thompson, G.R., Na Pombejra, S., Jamklang, M., Bautos,
J.M., Gelli, A., 2014. Invasion of the central nervous system by Cryptococcus
neoformans requires a secreted fungal metalloprotease. mBio 5, e01101-01114.
https://doi.org/10.1128/mBio.01101-14

Wang, Y., Pawar, S., Dutta, O., Wang, K., Rivera, A., Xue, C., 2022. Macrophage
mediated immunomodulation during Cryptococcus pulmonary infection.
Front. Cell. Infect. Microbiol. 12.

Williams, V., Del Poeta, M., 2011. Role of glucose in the expression of Cryptococcus
neoformans Antiphagocytic Protein 1, App1. Eukaryot. Cell 10, 293–301.
https://doi.org/10.1128/EC.00252-10

Williamson, P.R., Jarvis, J.N., Panackal, A.A., Fisher, M.C., Molloy, S.F., Loyse, A.,
Harrison, T.S., 2017. Cryptococcal meningitis: epidemiology, immunology,
diagnosis and therapy. Nat. Rev. Neurol. 13, 13–24.
https://doi.org/10.1038/nrneurol.2016.167

32
Wilson, R.A., Jenkinson, J.M., Gibson, R.P., Littlechild, J.A., Wang, Z.-Y., Talbot, N.J.,
2007. Tps1 regulates the pentose phosphate pathway, nitrogen metabolism
and fungal virulence. EMBO J. 26, 3673–3685.
https://doi.org/10.1038/sj.emboj.7601795

Wolfe, A.J., 2005. The Acetate Switch. Microbiol. Mol. Biol. Rev. 69, 12–50.
https://doi.org/10.1128/MMBR.69.1.12-50.2005

Wright, L., Bubb, W., Davidson, J., Santangelo, R., Krockenberger, M., Himmelreich, U.,
Sorrell, T., 2002. Metabolites released by Cryptococcus neoformans var.
neoformans and var. gattii differentially affect human neutrophil function.
Microbes Infect. 4, 1427–1438. https://doi.org/10.1016/S1286-4579(02)00024-2

Yu, C.-H., Sephton-Clark, P., Tenor, J.L., Toffaletti, D.L., Giamberardino, C., Haverkamp,
M., Cuomo, C.A., Perfect, J.R., 2021. Gene expression of diverse Cryptococcus
isolates during infection of the human central nervous system. mBio 12,
e02313-21. https://doi.org/10.1128/mBio.02313-21

Zaragoza, O., 2019. Basic principles of the virulence of Cryptococcus. Virulence 10,
490–501. https://doi.org/10.1080/21505594.2019.1614383

Zhao, Y., Lin, J., Fan, Y., Lin, X., 2019. Life cycle of Cryptococcus neoformans. Annu. Rev.
Microbiol. 73, 17–42. https://doi.org/10.1146/annurev-micro-020518-120210

33
CHAPTER 2

Analysis of transcriptomic response in fungal

pathogen Cryptococcus neoformans grown on

acetate and glucose as carbon sources

Oly Ahmed, Bridget Luckie, and Kerry Smith

Department of Genetics and Biochemistry; Eukaryotic Pathogen Innovation Center,

Clemson University, SC, USA

Authors contribution:

KS: conceived and designed the experiment; BL: cultured the cells and harvested

RNA for sequencing; OA: analyzed the RNA-seq data and prepared the manuscript

34
ABSTRACT

The two-carbon short-chain fatty acid acetate is an important alternative

carbon source in fungal pathogens, as its abundance and the abundance of the

preferred carbon source glucose varies significantly depending on the host-tissue

microenvironments. This supports the importance of rapid metabolic adaptation in

the invading fungal pathogens for efficient utilization of alternative carbon sources

for survival and disease progression. Previous transcriptomic and proteomic studies

of pathogenic ascomycete fungi such as Aspergillus fumigatus, Paracoccidioides

spp., and Candida glabrata highlight the species-specific metabolic reprogramming

resulting in these fungi following acetate utilization compared to glucose utilization.

Here we report the impact of acetate utilization on the transcriptome of the

pathogenic basidiomycete Cryptococcus neoformans for the first time. Of the 6,967

cryptococcal genes quantified in this study, 282 and 74 genes were upregulated and

downregulated, respectively, in acetate-grown cells compared to glucose-grown

cells. Acetate utilization resulted in an induction of genes in the metabolic pathways

such as the TCA, the glyoxylate cycle, acetyl-CoA production, fatty acid and polyol

metabolism, gluconeogenesis, glycogen synthesis and glucose mobilization, amino

acid degradation, lysine biosynthesis, and many putative transporters involved in

nutrient acquisition. Acetate utilization also resulted in an induction of fifteen

transcription factors (TF); however, no TFs were found to be downregulated.

Additionally, many genes previously reported to be associated with cryptococcal

virulence were also found to be upregulated in acetate-grown cells. Taken together,

35
these observations suggest that acetate utilization in C. neoformans leads to

metabolic reprogramming for efficient growth and survival along with elaboration of

virulence traits as a response to nutritional stress.

2.1. INTRODUCTION

Acetate (C2H3O-), a short chain fatty acid, is known to be an important carbon

source for fueling the cellular metabolism in Bacteria, Archaea, and Eukarya (Pandey

et al., 2018). For fungal pathogens of humans, the ability to exploit carbon sources in

various host tissue-environments is more important for survival as the concentration

of the preferred carbon source (typically glucose; C6H12O6) varies from tissue to tissue

within the host body. For example, the glucose level in human bloodstream is

around 0.06% (3.3 mM)- 0.1% (5.5 mM) (Gow and Brown, 2018), in vaginal lumen the

level is around 0.5% (27.8 mM) (Owen and Katz, 1999), in airway surface liquid the

level is about 0.008% (0.4 mM) (Baker and Baines, 2018), and in cerebrospinal fluid

the level is approximately 0.06-0.07% (3.3 mM - 3.8 mM) (Seehusen et al., 2003).

Until quite recently, the impact of acetate utilization on the fitness and

pathogenicity of fungal pathogens has rarely been studied. Ries et al. (Ries et al.,

2021) have investigated the role of acetate utilization in shaping the virulence

phenotypes of the human fungal pathogen Aspergillus fumigatus, an ascomycete,

and observed that the transcription factor FacB is a master regulator of acetate

utilization, and acetate utilization is also under carbon catabolite repression in the

presence of glucose. The authors also examined secreted secondary metabolites in

A. fumigatus under low glucose (0.1% w/v) and acetate (0.1% w/v and 1% w/v)

conditions and the level of secretion of secondary metabolites, which might be

36
essential for growth and survival of the fungus, is presumably influenced by the

available carbon sources and carbon starvation in host environments (Ries et al.,

2021). Additionally, growth on acetate appears to modify the cell wall composition in

A. fumigatus by increasing the concentration of chitin and the structural

polysaccharide β-1-3-glucan, and decreasing the concentration of the cementing

polysaccharide α-1-3-glucan, which is supposed to have notable consequences for

the fungal survival owing to the importance of the cell wall composition in

conferring resistance to oxidative, antifungal, and neutrophil-induced stresses (Ries

et al., 2021).

In ascomycete genus Paracoccidioides, proteomic changes with respect to

acetate utilization have also been studied. The study included one isolate of

Paracoccidioides lutzii and three isolates of Paracoccidioides brasiliensis, and

observed that utilization of acetate as the sole carbon source results in metabolic

reprogramming through induction of proteins in the pathways such as

gluconeogenesis, the glyoxylate cycle, and degradation of amino acids (Baeza et al.,

2017). Moreover, depending on the isolates, induction of the tricarboxylic acid (TCA)

cycle, the electron transport chain (ETC), ethanol production, reassimilation of cell

wall components for gluconeogenesis, β-oxidation, and the methylcitrate cycle were

also observed (Baeza et al., 2017).

Similarly, gluconeogenesis and the glyoxylate cycle a were upregulated in

transcriptomic study of the acetate-grown ascomycete pathogen Candida glabrata.

Other pathways were downregulated, including the biosynthesis of certain amino

acids and the pentose phosphate pathway (PPP) (Chew et al., 2021). The same study

also performed proteomic experiments and reported upregulation of oxidative

37
phosphorylation related proteins in acetate-grown cells compared to glucose-grown

ones, suggesting the involvement of an active electron transport chain (ETC) for ATP

generation in the acetate-grown cells under glucose-deprived conditions (Chew et

al., 2021).

By contrast, the impact of acetate utilization on the expression and

elaboration of virulence phenotypes has not yet been studied in the fungal

pathogen C. neoformans. C. neoformans is an aerobic, basidiomycete yeast with a

worldwide distribution (Bahn et al., 2020). Cryptococcus spp. are found in diverse

ecological niches, with a particular abundance in bird guano (Springer et al., 2017).

The medical relevance of C. neoformans involves its opportunistic pathogenicity in

immunocompromised people; inhaled fungal propagules from the environment

might not be cleared by the immune system, and may lead to a systemic infection

following hematogenous dissemination (May et al., 2016). In addition to organs such

as skin, eyes, prostrate, and bones, major sites of infection by C. neoformans are

lungs and the central nervous system (CNS) (Maziarz and Perfect, 2016), where an

infection of the CNS leads to a lethal meningoencephalitis in humans.

At various stages of cryptococcal infection, yeast cells can act as both an

intracellular and an extracellular parasite (Feldmesser et al., 2000), which implies

that C. neoformans is highly capable of rapid metabolic adaptations within the host

body to cope with the changing nutrient landscape depending on the types of

tissue or cell under invasion. For example, previous gene expression studies of

cryptococcal cells that are phagocytosed revealed an indication of nutrient-limited

and stressful environments within the phagocytic cells (Kronstad et al., 2012).

Additionally, transcriptomic profiles of C. neoformans isolated from a murine

38
pulmonary infection model compared to those grown under standard in vitro

condition showed an elevated expression of genes involved in the glyoxylate

pathway, gluconeogenesis, β-oxidation, amino acid biosynthesis and acetyl-CoA

utilization (Hu et al., 2008), implying that utilization of the alternative carbon sources

such as lactate and acetate are likely to be important early in the establishment of

pulmonary infections (Price et al., 2011). Similarly, transcriptional profiles of C.

neoformans isolated from cerebrospinal fluid (CSF) of human patients showed

elevated expression of genes involved in cell wall metabolism, extracellular transport,

fatty acids and glucose metabolism, iron regulation, pH response, pentose

phosphate pathway, and translational machineries (Yu et al., 2021). These

transcriptional studies have drawn our attention to the possible connection between

a reliance on acetate utilization for survival in the course of disease progression by C.

neoformans and the subsequent and/or concurrent expression and elaboration of

virulence phenotypes in this pathogen.

Similar to the glucose concentrations mentioned above, concentration of

acetate in the human body varies from tissue to tissue. For example, the plasma

concentration of acetate in humans is estimated to be around 0.0003% - 0.0012%

(0.05 mM - 0.2 mM), which may rise above 0.003% (0.5 mM) following alcohol

ingestion (Bose et al., 2019); and in vaginal lumen the estimation is 0.02% - 0.05% (3.4

mM - 8.5 mM) (Owen and Katz, 1999). The aim of this study is to investigate the

effects of acetate utilization on the global transcription of C. neoformans genome.

Here, we report for the first time the changes in the transcriptome of C. neoformans

attributable to the differential utilization of either acetate or glucose as the sole

carbon source.

39
2.2. RESULT AND DISCUSSION

We grew C. neoformans in the presence of either glucose or acetate as the

sole carbon source to observe the effect of these carbon sources on the global gene

expression in this pathogenic fungus. We identified and quantified the expression of

a total of 6,967 genes, of which 282 were upregulated and 274 were downregulated

in acetate-grown cells compared to glucose-grown cells (Summary of the

differentially expressed genes, DEGs can be found in Figure 2.1 (A) and (B);

expression pattern of DEGs across all conditions and replicates is represented in

Figure 2.1 (C) as a heatmap). Owing to the incompleteness of the functional

annotation of C. neoformans genome (nearly 42% of the total protein coding genes

lacks functional annotation), we used a stringent cutoff value for GO enrichment

analysis (using the built-in tools in the FungiDB database for the “Biological Process”

terms): p-value < 0.01 and FDR < 0.05. Table 2.1 lists GO term enrichment of

upregulated genes, and Table 2.2 lists GO terms enrichment of downregulated

genes. At any rate, we are in favor of a conservative interpretation of GO enrichment

analysis for C. neoformans genome owing to its poor functional annotation status, as

statistical tests such as Fisher’s exact test employed in enrichment analysis loses

explanatory power for incompletely annotated genomes.

Moreover, we analyzed the differentially expressed gene (DEG) set for

enrichment in KEGG pathways (using built-in tools in the FungiDB database: p-value

< 0.01 and FDR < 0.05). Table 2.3 lists KEGG pathway enrichment of upregulated

genes, and Table 2.4 lists KEGG pathway enrichment of the downregulated genes.

40
2.2.1. Chromosomal distribution of differentially expressed genes

Histone acetylation is known to be sensitive to the intracellular availability of

acetyl-CoA (Pietrocola et al., 2015). It is plausible that acetate-grown cells likely

experience a global reshuffling of nuclear organization through histone acetylation,

thereby adding nuances to the transcriptional control of acetate utilization. A

cursory analysis of chromosomal distribution of differentially regulated genes

(normalized as proportion) suggests a possible association between a gene’s

regulation status (up-, down-, or non-regulated) and the chromosome it is located in

(Figure 2.2). However, a Chi-square test of independence (14 × 3 contingency, df = 26)

does not point to any significant association (p = 0.071, Cramer’s V = 0.052) between

regulation status and chromosome position. This might be due to the violation of a

tacit assumption of the contingency analysis (i.e., observations are independent of

one another) by the commonplace trans-activation of non-cis-chromosomal genes

during global transcription.

However, in this study, we did not attempt to confirm such epigenetic

regulation of differential gene expression. Moreover, one of the upregulated genes,

cytochrome c oxidase subunit III (CNAG_09004), is on the mitochondrial genome.

Interestingly, expression of no other mitochondrial gene was impacted by the choice

of nutrient source.

2.2.2. Differential expression of putative transcription factors

To investigate the role of transcription factors in differential regulation of

nutrient-utilization genes, we searched the C. neoformans genome for potential

transcription factors (TF). For this study, we used the deep neural network-based tool

41
DeepTFactor (Kim et al., 2021) which resulted in the prediction of 200 TFs (Table A2).

None of these predicted and known TFs were found to be present in the set of

downregulated genes; however, a total of 15 predicted and known TFs were

upregulated in acetate-grown cells, suggesting their potential involvement in

acetate-utilization, stress-response, and/or glucose-deprivation. The list of

upregulated TFs comprises 4 previously unknown TFs predicted in this study

(CNAG_02525, CNAG_04618, CNAG_01847, CNAG_05990), as well as 11 TFs previously

described elsewhere ( (Jung et al., 2015); (Lee et al., 2020)): CCD4 (CNAG_03279),

FZC15 (CNAG_06188), FZC31 (CNAG_03741), FZC34 (CNAG_00896), YRM103

(CNAG_04093), STB4 (CNAG_05785), LIV3 (CNAG_05835), BZP4 (CNAG_03346),

ARO8001 (CNAG_04345), MLN1 (CNAG_04837), GAT201 (CNAG_01551).

2.2.3. Downregulation of genes in acetate-grown conditions

Many downregulated genes are enriched in KEGG pathways such as

glycolysis/ gluconeogenesis; pentose phosphate pathway; glycine, serine and

threonine metabolism; and steroid biosynthesis; pentose and glucuronate

interconversions; metabolism of xenobiotics; tyrosine metabolism among many

others (Table 2.4). About 40% (30/74) downregulated genes have unknown

functions (‘hypothetical proteins’). Two prominent genes of central carbon

metabolism, fructose bis-phosphate aldolase 1 (FBA1; CNAG_06770) and pyruvate

decarboxylase (PDC1; CNAG_04659), were downregulated, whereas expression of

other glycolytic/ gluconeogenic genes were not significantly impacted by glucose

uptake and metabolism. Notably, FBA1 is an essential gene for survival in C.

42
neoformans; however, by contrast, targeted mutagenesis experiment shows that

PDC1 is not essential for its viability (Ianiri and Idnurm, 2015).

Pentose phosphate pathway genes such as 6-phosphogluconate

dehydrogenase (CNAG_07561) and a transketolase (CNAG_07445), were also

downregulated in acetate-grown cells. Other genes with putative roles in carbon

metabolism found in our set of downregulated genes include xylulose-5-

phsophate/fructose-6-phosphate phosphoketolase (XFP2; CNAG_06923),

phosphoglycerate dehydrogenase (CNAG_00774), mannitol dehydrogenase

(CNAG_00515), enoyl reductase (CNAG_04652), endo-1,3(4)-beta-glucanase

(CNAG_05458), cellulase (CNAG_00799), aldo-keto reductase (CNAG_01954), and an

NADP-alcohol dehydrogenase (CNAG_01896). However, no TCA cycle gene was

downregulated.

Putative transporters such as CNAG_01960 (efflux protein), CNAG_03438

(HXT1), CNAG_06377 (Solute carrier 25, mitochondrial phosphate transporter),

CNAG_00749 (alternative sulfate transporter) were upregulated in glucose-grown

cells (i.e. downregulated in acetate-grown cells), suggesting their involvement in

glucose uptake and/or secretion of metabolites.

Additionally, cells grown on acetate showed decreased expression of

ribosomal protein genes such as CNAG_06447 (L17), CNAG_02928 (L5e), CNAG_03577

(LP0), CNAG_04114 (S0), CNAG_02754 (S12e). Also downregulated is a chaperone

regulator (CNAG_05252) and conidiation-specific protein 6 (CNAG_03759), a homolog

of a conserved fungal gene involved in sporulation in Neurospora crassa (Wang et

al., 2012). One cyclophilin A homolog, CPA2 (CNAG_03621), which is known to be

43
important in normal growth but plays an ancillary role in virulence was

downregulated in acetate-grown cells (Wang et al., 2001).

Other noteworthy downregulated genes with known functions include a thiol

peroxidase (TSA1; CNAG_03482), known to be important for growth at physiological

temperature (Missall et al., 2004); aspartate kinase (CNAG_04347), an essential kinase

involved in threonine biosynthesis (Kingsbury and McCusker, 2008); and a putative

calcineurin substrate, HAD-like hydrolase (CNAG_01744), required for cell wall

integrity, virulence, and growth at physiological temperatures (Jung et al., 2018).

2.2.4. Genes for acetate-metabolism, the glyoxylate cycle, and

peroxisomal activities were upregulated.

Almost 35% (98/282) of the upregulated genes in acetate-grown cells are

functionally uncharacterized (‘hypothetical proteins’), while many other genes are

enriched in KEGG pathways such as valine, leucine and isoleucine degradation; fatty

acid degradation; propanoate and butanoate metabolism; β-alanine metabolism;

lysine degradation; fatty acid elongation; pyruvate metabolism; glyoxylate and

dicarboxylate metabolism among many others (Table 2.3).

Increased acetyl-CoA production and utilization in acetate-grown cells was

suggested by the higher expression of acetyl CoA synthetase (ACS; CNAG_00797),

and two acetyl CoA acyltransferases (CNAG_00490 and CNAG_00524). Three

carnitine O-acetyltransferases (CNAG_00537, CNAG_06551, and CNAG_05042) also

had increased expression in acetate-grown cells compared to glucose-grown ones;

however, information regarding their cellular localization has not been reported in

the literature. In connection to acetate assimilation, the TCA and the glyoxylate

44
cycles play an important role: two glyoxylate cycle enzymes, isocitrate lyase (ICL1,

CNAG_05303) and malate synthase A (MLS1, CNAG_05653), and a putative TCA cycle

enzyme malate dehydrogenase (MDH; CNAG_06374), were found to be upregulated

in acetate-grown cells. Interestingly, although C. neoformans is known to be a

Crabtree negative yeast (Ries et al., 2018), a homolog of fumarate reductase

(CNAG_07862), an enzyme essential for maintaining redox balance during anaerobic

growth in S. cerevisiae (Camarasa et al., 2007), was found in our upregulated gene-

set.

Peroxisome-related genes such as PEX1 (CNAG_07403), PEX2 (CNAG_00171),

PEX3 (CNAG_02939), PEX5 (CNAG_03729), PEX12 (CNAG_04937), PEX16

(CNAG_02096), one peroxisome targeting signal receptor (CNAG_01926), and two

peroxisomal long-chain fatty acid importers CNAG_00651 and CNAG_02764 (PXA2),

were found in the upregulated gene-set, indicating increased peroxisomal

biogenesis in acetate-grown cells. Interestingly, a gene for a putative NIPSNAP

family protein (CNAG_05310) is upregulated in acetate-grown cells. NIPSNAP

proteins are evolutionarily conserved and known to act as a mitophagy-inducing

signal following their accumulation on the mitochondrial surface (Kumar and

Reichert, 2021). Other interesting genes upregulated in the acetate-grown cells

include CNAG_03051 (polyamine transporter) and CNAG_04794 (spermine

transporter), which are involved in polyamine transport. This observation is

particularly interesting as polyamine metabolism has been linked to cell survival

during stress (Valdés-Santiago and Ruiz-Herrera, 2013).

45
2.2.5. Upregulated genes bear the signature of nutrient-starvation.

The transcriptome of acetate-grown cells in our study bore the signature of

nutrient-starvation as signaled by the upregulation of various catabolic genes. Table

2.5 lists a subset of the upregulated genes grouped by their involvement in various

biological pathways. A similar subset of downregulated genes is compiled in Table

2.6. Table A3 contains the comprehensive list of DEGs found in this study.

Upregulated genes involved in fatty acid catabolism included long chain acyl-

CoA synthetase (CNAG_03019), peroxisomal 2,4-dienoyl-CoA reductase

(CNAG_04238), three acyl-CoA dehydrogenases (CNAG_04688, CNAG_03666,

CNAG_02562), enoyl-CoA hydratase (CNAG_04531), multifunctional beta-oxidation

protein (MFE2; CNAG_05721,), enoyl-CoA hydratase/isomerase (CNAG_03010), 3-

hydroxyacyl-CoA dehydrogenase (HAD1; CNAG_04308), and 3-hydroxybutyryl-CoA

dehydrogenase (HAD2; CNAG_03134).

Genes involved in glucose mobilization and metabolism such as glucan 1,3-

beta-glucosidase (EXG104; CNAG_02225,), glycosyl hydrolase (CNAG_05991), beta-

glucosidase (CNAG_07600), glyceraldehyde-3-phosphate dehydrogenase type I

(CNAG_04523), glycosyl-hydrolase (CNAG_04291), UDP-glucose, sterol transferase

(ATG2602; CNAG_03133), alpha-glucosidase (CNAG_05913), and exo-beta-1,3-

glucanase (CNAG_05803) were also upregulated in acetate-grown cells, suggesting a

possible mechanism of reassimilation of cell wall components and other ancillary

storage molecules in response to glucose-deprived conditions.

A few genes involved in protein and amino acid degradation were also

upregulated in acetate-grown fungi: Gly-Xaa carboxypeptidase (CNAG_03960),

glycine cleavage system T protein (CNAG_02818), L-serine ammonia-lyase

46
(CNAG_06913), sarcosine oxidase (CNAG_05115), aryl-alcohol dehydrogenase

(CNAG_01952), and proline oxidase (CNAG_02049, PUT1). An unusually high number

of genes involved in branched-chain amino acids (BCAA) degradation was found to

be upregulated in acetate-grown cells; these genes include 2-oxoisovalerate

dehydrogenase E1 components (alpha subunit; CNAG_02284, and beta subunit;

CNAG_00397), 2-oxoisovalerate dehydrogenase E2 component (dihydrolipoyl

transacylase: CNAG_00484), isovaleryl-CoA dehydrogenase (CNAG_00452), 3-oxoacid

CoA-transferase (CNAG_05031), 3-methylcrotonyl-CoA carboxylase alpha subunit

(CNAG_01680), and two forms of methylmalonate-semialdehyde dehydrogenase

(acylating) (CNAG_01075 and CNAG_04351). Other upregulated genes involved in

nitrogen and sulfur metabolism were urea carboxylase (CNAG_07944), glutamate

dehydrogenase (CNAG_00879), cystathionine beta-synthase (CNAG_00637),

cystathionine gamma-lyase (CNAG_06448), and two isoforms of 2-nitropropane

dioxygenase (CNAG_05644 and CNAG_03243).

Polyols play an important role as carbohydrate reserves in fungi (Jennings,

1985). As such, a number of genes for transporters and enzymes involved in polyol

metabolism were found to be upregulated in acetate-grown cells. Enzymes of polyol

metabolism that were upregulated include sorbitol dehydrogenase (CNAG_00269),

and two isoforms of L-iditol 2-dehydrogenase (CNAG_05658 and CNAG_06977).

Upregulated transporters include the polyol transporter protein 1 (PTP1;

CNAG_05662), and three myo-inositol transporters: ITR2 (CNAG_00864), ITR3c

(CNAG_05381), and ITR6 (CNAG_03910).

Other enzymes of carbon metabolism that were found to be upregulated

include dihydroxyacetone kinase (DAK101, CNAG_00826), formate dehydrogenase

47
(CNAG_06672), aldehyde dehydrogenase (NAD) (CNAG_06628), acetoacetate-CoA

ligase (CNAG_02045), alcohol dehydrogenase (CNAG_02489), and D-glycerate 3-

kinase (CNAG_07779). Other notable catabolic genes that were also upregulated

include uracil phosphoribosyltransferase (CNAG_02344) from the pyrimidine salvage

pathway, and cytochrome C peroxidase (CNAG_02147), a cytosolic enzyme involved

in peroxide neutralization. The xenobiotic degradation enzyme alpha-ketoglutarate-

dependent 2,4- dichlorophenoxyacetate dioxygenase (CNAG_04417), a taurine

metabolic enzyme taurine catabolism dioxygenase (TAUD, CNAG_01542), and a

hydroxymethylglutaryl-CoA lyase (CNAG_03067) involved in fat and protein

catabolism were also upregulated.

Two enzymes of the pentose phosphate pathways (PPP), transketolase

(CNAG_06172) and ribose 5-phosphate isomerase (CNAG_00827), were also

upregulated in acetate-grown cells. Although the PPP has a major role in

biosynthetic processes and is less likely to be upregulated during nutrient-starvation

(glucose-deprived conditions in our case), it is not surprising to observe upregulation

of certain PPP genes in acetate-grown cells considering the special role PPP plays in

generation of reducing equivalents (Deacon and Deacon, 2006).

Moreover, many putative transporters involved in the uptake of various

nutrient sources were also upregulated in acetate-grown cells. Such transporters are

involved in sugar transport: CNAG_00048, CNAG_03772 (HXS1), CNAG_04931 (HXS2),

CNAG_04092, CNAG_02586, CNAG_05324, CNAG_05867 (L-fucose transporter),

CNAG_01936, CNAG_07874, CNAG_05387 (galactose transporter), CNAG_02733,

CNAG_07641; amino acid transport: CNAG_07367, CNAG_00597 (DIP5), CNAG_05685,

CNAG_07902; nicotinic acid transport: CNAG_00028, CNAG_00598, CNAG_03242,

48
CNAG_06942, CNAG_04536, CNAG_04795, CNAG_06204; organic acids and alcohols

transport: CNAG_04038 (quinate), CNAG_06817 (uric acid xanthine permease, UAP1),

CNAG_02254 (quinate, LPI12), CNAG_06561 (allantoate), CNAG_04142 (tartrate),

CNAG_04704 (JEN4; lactate), CNAG_05119 (gamma-aminobutyric acid); and

phosphate transport: CNAG_02777 (PHO84). Other upregulated genes with putative

transporter function include CNAG_00796 (MDR1), CNAG_05718, CNAG_00284,

CNAG_00499, CNAG_05300, CNAG_00904, CNAG_02288, CNAG_03242; as well as

alpha-glucoside: H+ symporters CNAG_05330, CNAG_06527, CNAG_05914,

CNAG_05929, CNAG_06259.

2.2.6. Gluconeogenesis and certain anabolic pathways are also

upregulated.

In addition to two glyoxylate cycle enzymes (see above), the gluconeogenesis-

defining enzyme, phosphoenolpyruvate carboxykinase (PCK1, CNAG_04217), was

upregulated in acetate-grown cells. Fungal cells grown on acetate as the sole carbon

source depend on gluconeogenesis for production of sugars that are eventually

utilized in synthesis of cell wall, nucleotides, storage molecules and many other

biosynthetic processes (Deacon and Deacon, 2006). Additionally, certain genes

involved in glycogen and fatty acid synthesis were also upregulated in acetate-

grown cells. Upregulated glycogen synthesis genes include 1,4-alpha-glucan-

branching enzyme (CNAG_00393), glycogen synthase (CNAG_04621), and glycogenin

glucosyltransferase (CNAG_05293). Other upregulated fatty acid synthesis genes

include acetyl/propionyl CoA carboxylase (CNAG_01671) and beta-ketoacyl reductase

(CNAG_01116).

49
Other noteworthy anabolic enzymes include the proline synthesis pathway

gene ornithine-oxo-acid transaminase (CNAG_05134), a putative tetrahydrobiopterin

synthesis pathway gene cyclohydrolase (CNAG_06009), and a ferroxidase (CFO1,

CNAG_06241) involved in iron uptake.

One particularly interesting observation was that two lysine biosynthesis

genes, saccharopine reductase (LYS9, CNAG_00247) and dihydrodipicolinate

synthase (CNAG_04346), were upregulated. Saccharopine reductase [saccharopine

dehydrogenase (NADP+, L-glutamate forming)] catalyzes the formation of the

penultimate intermediate of lysine biosynthesis through the α-aminoadipate (AAA)

pathway and was observed to be a part of chimera with spermidine synthase (SPE3)

in C. neoformans (Kingsbury et al., 2004). On the other hand, dihydrodipicolinate

synthase catalyzes the first committed step of lysine biosynthesis via the

diaminopimelate (DAP) pathway, which is found in bacteria, lower fungi, and plants

and is believed to be absent in ascomycetes and basidiomycetes fungi (Zabriskie

and Jackson, 2000). Presence of a dihydrodipicolinate synthase (DHDPS) gene in

many fungal genomes, including C. neoformans, appears to contradict the claim

that AAA, and not DAP, is the only route of lysine production in higher fungi.

However, recently, Desbois and colleagues (Desbois et al., 2018) have shown that

DHDPS from another pathogenic fungus Coccidioides immitis, does not possess the

said function, rather has the 2-keto-3-deoxygluconate aldolase (KDGA) and 4-

hydroxy-2-oxoglutarate aldolase (HOGA) activities, suggesting an instance of

functional misannotation.

50
2.2.7. Some upregulated genes are implicated in virulence and

pathogenesis.

Recent studies in cryptococcal virulence mechanisms and pathogenesis have

shed light on three primary processes: adaptation to host environment, evasion of

the immune system, and production of virulence factors (Zaragoza, 2019). While our

study was aimed at learning adaptive mechanisms in Cryptococcus at transcription

level in response to low-glucose and high-acetate environments, we have found that

a handful of genes with connection to virulence and pathogenesis were

upregulated. Table 2.7 lists the virulence-associated genes that were found to be

significantly upregulated in this study.

2.3. CONCLUSION

For a fungus like Cryptococcus neoformans, invasion of and mobilization to

different organ systems within a multicellular host offers a range of challenges in

addition to the ones offered by the immune system of the host. Nutrient challenge is

one of those challenges as different organs differ in nutrient profiles in terms of

availability of various nutritional resources. Prior to this study we assumed that such

a flexibility in the metabolic response to alternative carbon source is existent in

Cryptococcus spp. and should be reflected in the transcriptional profiles. Hence, we

set out to seek for such transcription-level modifications through capturing a

snapshot of global transcriptomic landscape of C. neoformans during the growth in

quasi-starvation (YNB w/o amino acids) environments with either glucose or acetate

as sole carbon source. To magnify any subtle transcription-level modifications, we

have deliberately chosen the media concentration of the carbon sources to be high

51
[compare the concentrations used in this study for glucose (2% w/v or 111 mM) and

acetate (2% w/v or 339 mM) with those of glucose (5.5 mM) and acetate (0.2 mM) in

human blood]. Additionally, under the assumption that, in glucose-limited

environments C. neoformans will prioritize the uptake and utilization of available

alternative carbon sources, acetate appears to be a straightforward candidate for

alternative carbon source to study owing to its abundance in various Cryptococcus-

infected tissues, as well as its direct assimilation in energy metabolism pathway

through conversion to acetyl-CoA that sits at the crossroad of anabolism and

catabolism: i.e. its overall physiological relevance. Our data suggests, as

hypothesized, C. neoformans does show variation in transcriptome in response to

variation of carbon source, and such differential regulation accounts for about 5% of

the genome.

Not surprisingly, the transcriptome of acetate-grown cells shows

characteristic signature of starvation via upregulation of genes involved in certain

catabolic pathways to readily supply the fuel, and certain other anabolic pathways

for synthesis of sugar and energy storage molecules. In addition to that, pathways

involved in the maintenance and construction of cellular structures appear to have

decelerated in acetate-only conditions. None of these seems unusual within the

context of growth in a glucose-limited environment. However, many genes, hitherto

reported to have implications in cryptococcal virulence and pathogenicity in various

host models, are also upregulated in acetate-grown cells. This suggests a possible

connection between acetate utilization and induction of pathogenicity-associated

phenotypes. Modulation of fungal phenotypes to shape the disease outcome of the

52
host, mediated by acetate utilization, has already been observed, quite recently, in

another human pathogenic fungus, Aspergillus fumigatus (Ries et al., 2021).

The mechanism through which acetate utilization can accentuate pathogenic

phenotypes of C. neoformans is not clear; however, we would like to propose

modularity as a plausible mechanism in this context. It is possible that the regulatory

circuitry evolved in C. neoformans to deal with acetate as a carbon source might

have also incorporated virulence-associated genes, and such regulatory circuitries

are modular in structure (e.g., modules as observed in various transcriptional

regulatory networks or TRNs).

This speculation is consistent with the idea that C. neoformans’ capacity for

virulence is accidental (Fu et al., 2021)— an exaptation (Naranjo-Ortiz and Gabaldón,

2019)— and has evolved under the selection pressure from amoebic predation in soil

(Casadevall et al., 2019). Since adaptation through natural selection can occur only in

response to ever-present or ever-recurring features of the surroundings (Garvey,

2007), this hypothesis can be extended to incorporate other naturally occurring

stresses imposed on the soil-dwelling fungi by their physical surroundings.

It is more likely that these physical stresses— including, but not limited to,

oxidative, osmotic, thermal, radiational, and nutritional stresses— come in a bundle

as a recurring geographic and seasonal feature of the soil-dwelling fungi’s

surroundings. Most of the pathogenicity-associated phenotypes are, probably not

surprisingly, found to be correlated. For example, growth at physiological

temperature (thermal stress), otherwise considered too high for environmental

fungi, is also associated with antioxidant response (oxidative stress), trehalose

accumulation (osmotic stress), and others (Zaragoza, 2019). Similarly, some virulence

53
factors, such as capsule and melanin production, can even provide the fungi with

actual physical protection in the wild (Naranjo-Ortiz and Gabaldón, 2019). This

implies that a more efficient way for adaptation to occur would be for C. neoformans

to evolve a blanket-defense against the recurring bundle of environmental stresses.

This may explain why most of those defense- (and, in the context of immunology,

‘virulence’)-related phenotypes are expressed simultaneously in C. neoformans.

In our study, we found transcriptomic hints of simultaneous activation of

various stress responses (e.g., starvation avoidance, maintenance of redox balance,

synthesis and uptake of compounds with osmotic potential), as well as changes in

expression of a few pathogenicity-associated genes with seemingly no involvement

in acetate utilization. Here, we propose to explain this observation by the partial

activation of blanket defense against environmental stressor-bundle (that we posit

to be extant in C. neoformans owing to natural selection) through pleiotropic effect

of the regulatory network involved in acetate utilization. While this hypothesis will

require further experimental validation, this framework holds a potential for

understanding cryptococcal pathogenicity in a different way. For example, if the

blanket defense mechanism is at work, then developing drugs that target only one

virulence factor at a time may soon lead to drug resistance.

54
2.4. METHODS AND MATERIALS

2.4.1. RNA isolation and sequencing

Cryptococcus neoformans var grubii KN99 cells were grown on Yeast

Nitrogen Base without amino acids (YNB w/o aa) with 2% (w/v) glucose, overnight, to

a density of ~1.5 x 108 cells/mL. These starting cultures were washed with the pre-

warmed media accordingly and were transferred to either YNB without amino acids

with 2% (w/v) acetate (acetate-grown) or YNB without amino acids with 2% (w/v)

glucose (glucose-grown) media for incubation on a rotor shaker for 5 hours.

Following incubation on experimental growth conditions, cells were harvested and

flash-frozen in liquid nitrogen. Total RNA was isolated (three biological replicates per

condition) using RiboPureTM RNA Purification Kit, yeast (InvitrogenTM, Thermo Fisher

Scientific, MA, USA) following manufacturer’s instruction. RNA-seq was performed on

an Illumina Hiseq 2500 sequencing platform using NEBNext® Ultra™ RNA Library

Prep Kit for Illumina (NEB, USA). Prior to sequencing, the index-coded samples were

clustered in a cBot Cluster Generation System using PE Cluster Kit cBot-HS

(Illumina). Library construction, quality control and sequencing were outsourced to

Novogene Co., Ltd, China. Sequencing generated an average of 57,360,137 raw and

54,766,198 clean reads (Table A1).

55
2.4.2. Transcriptomic analysis

Quality of the raw reads were checked using FastQC (v-0.11.7). Raw reads were

aligned against the C. neoformans KN99 genomes (fasta and GFF3) from FungiDB

48th release. STAR (v-2.7.5), Bowtie2 (v-2.3.5.1), and RSEM (v-1.3.3) were used for

alignment, mapping, and counting of the reads. Differential expression analysis was

conducted on the RSEM-generated transcript counts using R package DESeq2.

Figures for data visualization were prepared in RStudio (v-1.4). For ease and

consistency, final functional annotation of the KN99 strain’s transcripts were based

on the corresponding CNAG-IDs of the genome of C. neoformans H99 strain.

56
TABLES

Table 2.1. GO term enrichment of upregulated genes

Total
gene Upregulated Fold
GO term ID Biological process count gene count enrichment FDR

GO:0055085 transmembrane transport 350 57 3.88 9.7E-18

GO:0006810 transport 649 59 2.16 4.4E-07

GO:0051234 establishment of localization 654 59 2.15 4.4E-07

GO:0051179 localization 674 59 2.08 1.1E-06

GO:0007031 peroxisome organization 9 6 15.87 3.4E-05

GO:0008643 carbohydrate transport 23 8 8.28 1.7E-04

GO:0055114 obsolete oxidation-reduction process 335 33 2.35 1.7E-04

GO:0016054 organic acid catabolic process 24 8 7.94 1.7E-04

GO:0046395 carboxylic acid catabolic process 24 8 7.94 1.7E-04

GO:0044282 small molecule catabolic process 43 10 5.54 3.5E-04

GO:0009063 cellular amino acid catabolic process 19 5 6.27 3.6E-02

GO:0032787 monocarboxylic acid metabolic process 51 8 3.73 4.2E-02

GO:0072329 monocarboxylic acid catabolic process 6 3 11.9 4.3E-02

GO:0034440 lipid oxidation 2 2 23.81 4.3E-02

GO:0019395 fatty acid oxidation 2 2 23.81 4.3E-02

GO:0009062 fatty acid catabolic process 2 2 23.81 4.3E-02

GO:0006635 fatty acid beta-oxidation 2 2 23.81 4.3E-02

GO:0005978 glycogen biosynthetic process 2 2 23.81 4.3E-02

GO:0043436 oxoacid metabolic process 204 18 2.1 4.6E-02

GO:0019752 carboxylic acid metabolic process 204 18 2.1 4.6E-02

GO:0006631 fatty acid metabolic process 14 4 6.8 4.6E-02

GO:0006082 organic acid metabolic process 207 18 2.07 4.9E-02

57
Table 2.2. GO term enrichment of downregulated genes

Total
gene Downregulated Fold
GO term ID Biological process count gene count enrichment FDR

obsolete oxidation-reduction
GO:0055114 process 335 14 4.55 0.00026

GO:0019752 carboxylic acid metabolic process 204 8 4.27 0.03472

GO:0043436 oxoacid metabolic process 204 8 4.27 0.03472

GO:0006082 organic acid metabolic process 207 8 4.21 0.03472

GO:0008152 metabolic process 2044 30 1.6 0.03969

GO:0046394 carboxylic acid biosynthetic process 85 5 6.41 0.03969

GO:0044283 small molecule biosynthetic process 128 6 5.11 0.03969

GO:0016053 organic acid biosynthetic process 90 5 6.05 0.04442

GO:0005975 carbohydrate metabolic process 140 6 4.67 0.04916

58
Table 2.3. KEGG pathway enrichment of upregulated genes

Total gene Upregulated Fold


KEGG ID KEGG pathway count gene count enrichment FDR
Valine, leucine and isoleucine
ec00280 degradation 61 23 7.51 8.1E-14

ec00071 Fatty acid degradation 78 22 5.62 2.0E-10

ec00281 Geraniol degradation 30 14 9.3 7.8E-10

ec00640 Propanoate metabolism 109 24 4.39 2.9E-09

ec00650 Butanoate metabolism 110 24 4.35 2.9E-09

ec00410 beta-Alanine metabolism 49 15 6.1 8.7E-08

ec00930 Caprolactam degradation 52 12 4.6 9.2E-05


Carbon fixation pathways in
ec00720 prokaryotes 91 15 3.28 4.0E-04

ec00310 Lysine degradation 94 15 3.18 4.8E-04


alpha-Linolenic acid
ec00592 metabolism 94 15 3.18 4.8E-04

ec00062 Fatty acid elongation 23 7 6.06 9.7E-04

ec00620 Pyruvate metabolism 113 16 2.82 9.9E-04

ec00362 Benzoate degradation 81 13 3.2 1.2E-03


Glyoxylate and dicarboxylate
ec00630 metabolism 72 12 3.32 1.5E-03

59
Table 2.4. KEGG pathway enrichment of downregulated genes

Total gene Downregulated Fold


KEGG ID KEGG pathway count gene count enrichment FDR

ec00592 alpha-Linolenic acid metabolism 94 8 5.8 0.0032


Glycine, serine and threonine
ec00260 metabolism 109 8 5 0.0047

ec00930 Caprolactam degradation 52 5 6.56 0.0238

ec00010 Glycolysis / Gluconeogenesis 126 7 3.79 0.0327


Metabolism of xenobiotics by
ec00980 cytochrome P450 65 5 5.25 0.0327
Carbon fixation in photosynthetic
ec00710 organisms 40 4 6.82 0.0327

ec00830 Retinol metabolism 104 6 3.93 0.0327

ec00626 Naphthalene degradation 106 6 3.86 0.0327

ec00561 Glycerolipid metabolism 73 5 4.67 0.0327

ec00350 Tyrosine metabolism 187 8 2.92 0.0327

ec00902 Monoterpenoid biosynthesis 7 2 19.48 0.0327


Pentose and glucuronate
ec00040 interconversions 77 5 4.43 0.0327

ec00071 Fatty acid degradation 78 5 4.37 0.0327

ec00620 Pyruvate metabolism 113 6 3.62 0.0327

ec00030 Pentose phosphate pathway 80 5 4.26 0.0332


Drug metabolism - cytochrome
ec00982 P450 87 5 3.92 0.0446

ec00100 Steroid biosynthesis 30 3 6.82 0.0499

60
Table 2.5. List of upregulated genes in acetate-grown C. neoformans compared to
glucose-grown cells

Gene Name log2 Fold-


Gene ID Gene annotation or Symbol Chromosome Change FDR

Acetyl-CoA metabolism

CNAG_00490 acetyl-CoA acyltransferase N/A 1 4.150523898 0

CNAG_00524 acetyl-CoA acyltransferase 2 N/A 1 2.230319617 9.46E-135

CNAG_00537 Carnitine O-acetyltransferase N/A 1 5.18957041 0

CNAG_00797 acetyl-CoA synthetase N/A 1 2.593962989 1.06E-135

CNAG_05042 carnitine acetyltransferase N/A 4 2.339243859 9.32E-136

CNAG_06551 Carnitine O-acetyltransferase N/A 7 4.252895603 0

Branched-chain amino acid degradation pathway


2-oxoisovalerate dehydrogenase E1
CNAG_00397 component, beta subunit N/A 1 3.327788507 4.05E-204

CNAG_00452 Isovaleryl-CoA dehydrogenase N/A 1 4.381863876 3.22E-133


2-oxoisovalerate dehydrogenase E2
component (dihydrolipoyl
CNAG_00484 transacylase) N/A 1 2.659748528 8.88E-180
methylmalonate-semialdehyde
CNAG_01075 dehydrogenase (acylating) N/A 5 2.139409989 9.16E-54
3-methylcrotonyl-CoA carboxylase
CNAG_01680 alpha subunit N/A 11 4.653411548 0
2-oxoisovalerate dehydrogenase E1
CNAG_02284 component, alpha subunit N/A 6 2.059866905 8.43E-78
methylmalonate-semialdehyde
CNAG_04351 dehydrogenase (acylating) N/A 9 2.044370708 3.26E-57

CNAG_05031 3-oxoacid CoA-transferase N/A 4 3.444662639 0

Cell cycle

CNAG_00442 cyclin N/A 1 2.465150614 1.41E-123

Fatty acid catabolism

CNAG_02562 acyl-CoA dehydrogenase N/A 3 3.63243594 1.89E-266

61
CNAG_03010 enoyl-CoA hydratase/isomerase N/A 3 4.251823055 0

CNAG_03019 long-chain acyl-CoA synthetase N/A 3 3.790271235 0

CNAG_03134 3-hydroxybutyryl-CoA dehydrogenase HAD2 8 2.006338715 1.23E-53

CNAG_03666 acyl-CoA dehydrogenase N/A 2 4.942071419 0


peroxisomal 2,4-dienoyl-CoA
CNAG_04238 reductase N/A 9 3.041104459 6.05E-230

CNAG_04308 3-hydroxyacyl-CoA dehydrogenase HAD1 9 2.103182554 9.28E-98

CNAG_04531 Enoyl-CoA hydratase N/A 9 2.432157388 2.46E-129

CNAG_04688 acyl-CoA dehydrogenase N/A 10 2.273564711 4.60E-121


multifunctional beta-oxidation
CNAG_05721 protein MFE2 7 4.022808323 0

Fatty acid synthesis

CNAG_01116 beta-ketoacyl reductase N/A 5 2.767779821 4.50E-79

CNAG_01671 acetyl/propionyl CoA carboxylase N/A 11 3.563038208 0

Protein catabolism

CNAG_03067 hydroxymethylglutaryl-CoA lyase N/A 3 3.176153974 2.07E-274

Gluconeogenesis
phosphoenolpyruvate carboxykinase
CNAG_04217 (ATP) PCK1 9 3.26714125 6.60E-163

Glucose mobilization

CNAG_02225 glucan 1,3-beta-glucosidase EXG104 6 3.371935116 2.52E-249

CNAG_03133 UDP-glucose,sterol transferase ATG2602 8 2.111181869 4.92E-101

CNAG_04291 glycosyl-hydrolase N/A 9 2.44488372 2.31E-117


glyceraldehyde-3-phosphate
CNAG_04523 dehydrogenase, type I N/A 9 2.281267594 9.36E-153

CNAG_05803 exo-beta-1,3-glucanase N/A 7 2.501116431 3.85E-120

CNAG_05913 Alpha-glucosidase N/A 7 4.759365762 0

CNAG_05991 glycosyl hydrolase family 88 N/A 12 2.774307079 2.36E-176

CNAG_07600 Beta-glucosidase N/A 11 2.565018508 5.55E-137

Glycogen synthesis

CNAG_00393 1,4-alpha-glucan-branching enzyme N/A 1 2.535777799 2.86E-184

CNAG_04621 glycogen(starch) synthase, variant N/A 10 2.940035403 5.84E-184

CNAG_05293 glycogenin glucosyltransferase N/A 4 2.353347363 2.03E-55

62
Glyoxylate cycle

CNAG_05303 isocitrate lyase ICL1 4 6.900794065 0

CNAG_05653 malate synthase A MLS1 14 4.950928228 0

Iron uptake

CNAG_06241 acidic laccase CFO1 12 2.22727825 7.13E-83

Lysine synthesis
alpha-aminoadipic semialdehyde
CNAG_00247 synthase LYS9 1 4.647012781 0

CNAG_04346 dihydrodipicolinate synthase N/A 9 2.238276473 7.54E-60

Carbon metabolism

CNAG_00826 dihydroxyacetone kinase DAK101 1 3.981063756 0

CNAG_02045 acetoacetate-CoA ligase N/A 6 4.784163223 0


alcohol dehydrogenase, propanol-
CNAG_02489 preferring N/A 6 6.655716374 0

CNAG_06628 aldehyde dehydrogenase (NAD) N/A 7 3.554003489 2.07E-269

CNAG_06672 formate dehydrogenase N/A 7 2.019028288 5.82E-38

CNAG_07779 D-glycerate 3-kinase TDA10 9 2.205650117 5.10E-43

Mitophagy

CNAG_05310 nipsnap family protein N/A 4 5.003188347 1.07E-195

Nitrogen and Sulfur metabolism

CNAG_00637 Cystathionine beta-synthase N/A 1 3.059496886 5.75E-290

CNAG_00879 Glutamate dehydrogenase N/A 5 5.152721407 0

CNAG_03243 2-Nitropropane dioxygenase N/A 8 4.194067019 2.51E-143

CNAG_05644 2-Nitropropane dioxygenase N/A 14 3.729505188 0

CNAG_06448 Cystathionine gamma-lyase N/A 13 2.437034447 7.63E-64

CNAG_07944 Urea carboxylase N/A 13 3.394992027 0

Peroxide neutralization

CNAG_02147 cytochrome c peroxidase N/A 6 4.190300827 4.20E-208

Peroxisomal organization and transport

CNAG_00171 peroxin-2 N/A 1 2.002408397 2.54E-89


ATP-binding cassette, subfamily D
(ALD), peroxisomal long-chain fatty
CNAG_00651 acid import protein N/A 1 3.10426851 0

63
CNAG_01926 peroxisome targeting signal receptor PEX5 11 2.424142535 1.99E-146

CNAG_02096 peroxin-16 N/A 6 2.367732307 1.17E-77


ATP-binding cassette, subfamily D
(ALD), peroxisomal long-chain fatty
CNAG_02764 acid import protein PXA2 3 2.582242246 5.34E-143

CNAG_02939 peroxin-3 N/A 3 2.206201302 1.33E-121

CNAG_03729 peroxin-5 N/A 2 2.91316122 3.71E-263

CNAG_04937 peroxin-12 N/A 4 2.199480608 8.13E-88

CNAG_07403 Peroxisomal ATPase PEX1 PEX1 5 2.50683658 9.57E-158

Phagocytosis resistance

CNAG_06574 antiphagocytic protein 1 APP1 7 3.088293978 3.40E-122

Polyamine metabolism

CNAG_03051 polyamine transporter N/A 3 2.098521697 2.80E-74

CNAG_04794 spermine transporter N/A 10 2.584536287 1.19E-135

Polyol metabolism

CNAG_00269 Sorbitol dehydrogenase N/A 1 2.472085629 7.34E-54

CNAG_00864 myo-inositol transporter, putative ITR2 5 3.035563112 9.98E-182

CNAG_03910 myo-inositol transporter, putative ITR6 2 6.517998365 0

CNAG_05381 myo-inositol transporter, putative ITR3C 14 2.065739418 9.34E-37

CNAG_05658 L-iditol 2-dehydrogenase N/A 14 4.051104215 3.05E-223

CNAG_05662 Polyol transporter protein 1 PTP1 14 7.859393735 0

CNAG_06977 L-iditol 2-dehydrogenase N/A 8 2.836455175 1.82E-12

Pentose phosphate pathway

CNAG_06172 transketolase N/A 12 2.848885737 9.10E-75

CNAG_00827 ribose 5-phosphate isomerase N/A 1 4.99834915 1.33E-143

Proline synthesis

CNAG_05134 ornithine-oxo-acid transaminase N/A 4 2.241942301 7.64E-110

Protein and amino acid degradation

CNAG_01952 aryl-alcohol dehydrogenase N/A 11 2.509646461 5.96E-67

CNAG_02049 Proline dehydrogenase PUT1 6 3.09586931 7.34E-92

CNAG_02818 Glycine cleavage system T protein N/A 3 2.037473869 5.18E-181

CNAG_03960 Gly-Xaa carboxypeptidase N/A 2 2.174955689 1.79E-141

64
CNAG_05115 sarcosine oxidase N/A 4 2.406894282 9.41E-91

CNAG_06913 L-serine ammonia-lyase N/A 3 2.744533923 2.68E-146

Pyrimidine salvage pathway

CNAG_02344 Uracil phosphoribosyltransferase N/A 6 2.909303709 4.74E-132

Redox balance

CNAG_07862 fumarate reductase N/A 10 5.402443646 0

Taurine metabolism

CNAG_01542 taurine catabolism dioxygenase TauD N/A 11 4.087034641 1.74E-305

Tricarboxylic acid cycle


malate dehydrogenase (oxaloacetate-
CNAG_06374 decarboxylating)(NADP) N/A 13 2.564851405 8.46E-69

Tetrahydrobiopterin synthesis

CNAG_06009 cyclohydrolase N/A 12 2.962930691 1.30E-190

Transcription factors

CNAG_00896 transcription factor FZC34 5 2.628890692 9.44E-74

CNAG_01551 GATA family transcription factor GAT201 11 3.675353875 5.24E-221

CNAG_01847 hypothetical protein N/A 11 2.678662233 1.49E-183

CNAG_02525 hypothetical protein N/A 6 2.201674493 2.36E-83

CNAG_03279 hypothetical protein CCD4 8 2.071978705 6.91E-79

CNAG_03346 bZip transcription factor, putative BZP4 8 2.060783343 7.19E-57


0.0062721257
CNAG_03741 hypothetical protein FZC31 2 2.265537065 68

CNAG_04093 putative transcription factor YRM103 2 2.821424664 9.87E-138


specific RNA polymerase II
CNAG_04345 transcription factor ARO8001 9 4.057102724 0

CNAG_04618 hypothetical protein N/A 10 2.109468659 3.34E-66

CNAG_04837 bHLH family transcription factor MLN1 10 5.658104561 0

CNAG_05785 putative transcription factor STB4 7 4.142636148 2.09E-218

CNAG_05835 wor1/pac2 family transcription factor LIV3 7 2.020214829 1.01E-114

CNAG_05990 hypothetical protein N/A 12 3.520892501 1.66E-217

CNAG_06188 hypothetical protein FZC15 12 2.051555525 1.23E-43

Transport of metabolites

65
high-affinity nicotinic acid
CNAG_00028 transporter, variant N/A 1 2.833517394 1.41E-167

CNAG_00048 Sugar transporter N/A 1 2.165703801 2.68E-70

CNAG_00284 efflux protein EncT N/A 1 2.051495086 1.00E-40


solute carrier family 25 (mitochondrial
carnitine/acylcarnitine transporter),
CNAG_00499 member 20/29 N/A 1 3.69160675 0

CNAG_00597 amino acid transporter DIP5 1 4.518136614 0


nicotinamide mononucleotide
CNAG_00598 permease N/A 1 5.436810579 0
ATP-binding cassette, subfamily B
CNAG_00796 (MDR/TAP), member 1 MDR1 1 2.664776154 4.21E-53

CNAG_00904 aflatoxin efflux pump AFLT N/A 5 2.533320701 2.99E-10

CNAG_01936 Sugar transporter N/A 11 5.398218881 0

CNAG_02254 quinate permease LPI12 6 3.294651913 9.77E-295


solute carrier family 25 (mitochondrial
CNAG_02288 citrate transporter), member 1 N/A 6 5.725525264 0

CNAG_02586 Sugar transporter N/A 3 3.213961171 2.20E-108

CNAG_02733 Monosaccharide transporter N/A 3 4.838391609 7.50E-302

CNAG_02777 phosphate:H symporter, variant PHO84 3 2.235545414 2.05E-09


solute carrier family 25 (peroxisomal
adenine nucleotide transporter),
CNAG_03242 member 17 N/A 8 2.249113409 2.69E-81

CNAG_03772 high-affinity glucose transporter HXS1 2 8.11764169 0

CNAG_04038 MFS quinate transporter QutD N/A 2 2.289231867 3.47E-79

CNAG_04092 Sugar transporter N/A 2 2.350600318 3.95E-79

CNAG_04142 tartrate transporter N/A 9 5.329645146 0


nicotinamide mononucleotide
CNAG_04536 permease N/A 9 4.261264549 5.91E-218
MFS transporter, SHS family, lactate
CNAG_04704 transporter N/A 10 4.877479239 4.20E-260

CNAG_04795 adenine nucleotide transporter N/A 10 2.528076177 3.98E-97


0.0077011887
CNAG_04931 Putative hexose transporter HXS2 10 4.37816534 05

CNAG_05119 GABA permease N/A 4 2.758643737 1.60E-87

CNAG_05300 mitochondrial carrier protein N/A 4 2.730932228 1.93E-140

CNAG_05324 Sugar transporter N/A 4 4.434844391 0

66
MFS transporter, SP family, general
CNAG_05330 alpha glucoside:H symporter N/A 4 5.501723558 1.13E-94

CNAG_05387 galactose transporter N/A 14 5.431201544 0

CNAG_05685 neutral amino acid transporter N/A 7 3.83528674 0

CNAG_05718 multidrug resistance protein fnx1 N/A 7 2.073586105 9.63E-64

CNAG_05867 L-fucose transporter N/A 7 7.624377564 0


MFS transporter, SP family, general
CNAG_05914 alpha glucoside:H symporter N/A 7 7.804574725 0
MFS transporter, SP family, general
CNAG_05929 alpha glucoside:H symporter N/A 7 2.893258898 5.48E-134

CNAG_06204 high-affinity nicotinic acid transporter N/A 12 2.529064146 1.85E-150


MFS transporter, SP family, general
CNAG_06259 alpha glucoside:H symporter N/A 13 2.221697213 4.28E-52
MFS transporter, SP family, general
alpha glucoside:H symporter, variant
CNAG_06527 2 N/A 7 3.840821525 4.20E-149

CNAG_06561 allantoate transporter N/A 7 2.745084933 6.78E-57

CNAG_06817 uric acid xanthine permease UAP1 5 2.413959934 1.46E-110


nicotinamide mononucleotide
CNAG_06942 permease N/A 8 2.42053468 1.61E-137

CNAG_07367 amino acid transporter, variant N/A 1 2.147164551 1.15E-33

CNAG_07641 Monosaccharide transporter N/A 6 3.418297251 1.06E-252

CNAG_07874 Sugar transporter N/A 14 4.271990142 2.21E-150

CNAG_07902 AAT family amino acid transporter N/A 12 2.079429146 1.41E-74

Xenobiotic metabolism
alpha-ketoglutarate-dependent 2,4-
CNAG_04417 dichlorophenoxyacetate dioxygenase N/A 9 3.539262119 3.83E-88

67
Table 2.6. List of downregulated genes in acetate-grown C. neoformans compared to
glucose-grown one
Gene Name log2 Fold-
Gene ID Gene annotation or Symbol Chromosome Change FDR

Carbon metabolism

CNAG_00515 mannitol dehydrogenase N/A 1 -2.056718088 4.69E-46

CNAG_00774 phosphoglycerate dehydrogenase N/A 1 -2.320407504 3.88E-102

CNAG_00799 cellulase N/A 1 -2.169536781 3.35E-61

CNAG_01896 alcohol dehydrogenase (NADP) N/A 11 -2.146751373 1.19E-74

CNAG_01954 aldo-keto reductase N/A 11 -2.155508762 7.61E-60

CNAG_04652 enoyl reductase N/A 10 -2.326270665 1.76E-80

CNAG_04659 Pyruvate decarboxylase PDC1 10 -2.640566552 1.33E-184

CNAG_05458 endo-1,3(4)-beta-glucanase N/A 14 -2.647163092 1.60E-125

CNAG_06770 fructose-bisphosphate aldolase 1 FBA1 2 -2.061906962 7.67E-97


xylulose 5-phosphate/fructose 6-
CNAG_06923 phosphate phosphoketolase XFP2 3 -4.900496595 0

CNAG_02542 fructosamine kinase IRK2 6 -2.425279714 8.50E-56


ketol-acid reductoisomerase,
CNAG_05725 mitochondrial N/A 7 -2.89523887 1.73E-149
alcohol dehydrogenase, propanol-
CNAG_07745 preferring MPD1 8 -3.292011686 1.18E-193

Pentose phosphate pathway

CNAG_07445 transketolase N/A 5 -2.170823672 1.27E-54


6-phosphogluconate dehydrogenase,
CNAG_07561 decarboxylating 1 N/A 3 -2.431142436 1.95E-88

Ribosomal organization

CNAG_02754 small subunit ribosomal protein S12e N/A 3 -2.111585423 1.84E-118

CNAG_02928 large subunit ribosomal protein L5e N/A 3 -2.002523096 9.01E-82

CNAG_03577 large subunit ribosomal protein LP0 N/A 8 -2.155719974 5.55E-69

CNAG_04114 small subunit ribosomal protein S0 N/A 9 -2.199848081 2.32E-50


Large subunit ribosomal protein L17,
CNAG_06447 putative RPL17 13 -2.005085008 1.11E-127

Transport

CNAG_00749 alternative sulfate transporter N/A 1 -2.733300095 1.15E-66

68
CNAG_01960 efflux protein EncT N/A 11 -2.313456416 6.47E-57

CNAG_03438 hexose transporter HXT1 8 -2.501359805 9.23E-144

solute carrier family 25 (mitochondrial


CNAG_06377 phosphate transporter), member 3 N/A 13 -2.152642318 3.17E-72

Table 2.7. Upregulated cryptococcal genes with reported association with


pathogenicity

Gene CNAG ID Function and Reference

APP1 CNAG_06574 Inhibition of phagocytosis (Luberto et al., 2003)

CFO1 CNAG_06241 Loss of CFO1 attenuates virulence (Jung et al., 2009);


Required for adapting to oxidative, osmotic, genotoxic, and
heavy metal stress (Lee et al., 2014);
Mutant shows susceptibility to fluconazole treatment (Kim et al.,
2012);

PCK1 CNAG_04217 Deletion mutant shows attenuated virulence in murine


inhalation model, but shows persistence in rabbit CSF model
(Price et al., 2011)

LIV3 CNAG_05835 Mutant is defective in lung infectivity (Liu et al., 2008);


TF implicated in quorum sensing (Homer et al., 2016)

MLN1 CNAG_04837 Increased expression in in vivo and ex vivo CSF conditions


compared to YPD media (Chen et al., 2014)

DAK101 CNAG_00826 Core-virulence kinase, and is required for lung and brain
infectivity (Lee et al., 2020)

DPP101 CNAG_06499 One of the phosphatases that were upregulated during host
infection; further characterization required (Jin et al., 2020)

PHO84 CNAG_02777 One of the deleted genes in the triple mutant phoΔΔΔ (pho84,
pho840, pho89) which fails to replicate within macrophage
(Kretschmer et al., 2014)

BZP4 CNAG_03346 TF involved in melanin synthesis; and is expressed at lower


levels in clinical isolates compared to environmental ones (Yu et
al., 2020); loss-of-function leads to reduced melanization
(Desjardins et al., 2017)

69
FZC31 CNAG_03741 Core-virulence TF required for lung and brain infectivity (Jung
et al., 2015)

FZC34 CNAG_00896 Upregulated (moderately, since FDR > 0.05) in ex vivo and in
vivo CSF conditions compared to YPD media (Chen et al., 2014)

GAT201 CNAG_01551 Regulation of anti-phagocytic mechanism (Santiago-Tirado et


al., 2015);
deletion leads to hypocapsular phenotype (Maier et al., 2015);
core -virulence TF for lung and brain infectivity (Lee et al., 2020)

RCV1 CNAG_01983 Deletion leads to impaired GXM production and secretion,


defective cytokinesis, reduced growth at 37°C, and loss of
virulence in murine model (Paes et al., 2018)

STB4 CNAG_05785 TF required for brain infectivity (Lee et al., 2020)

ICL1 CNAG_05303 Highly expressed during infection, however not required for
infection progression (Rude et al., 2002)

MFE2 CNAG_05721 loss -of-function leads to reduction in capsule production and


diminished ability to colonize brain (Kretschmer et al., 2012)

ISP6 CNAG_00456 Upregulated in in vivo and ex vivo CSF conditions compared to


YPD media, a putative target for Gat201 TF (Chen et al., 2014);
one of the proteins enriched in cryptococcal spore (Huang et al.,
2015)

MLS1 CNAG_05653 Mutant is unable to grow on acetate, however does not show
any reduction in virulence (Idnurm et al., 2007)

EXG104 CNAG_02225 Extracellular protein; downregulated in kcs1Δ mutants [kcs1Δ do


not elicit strong immune response and are not readily
phagocytosed] (Lev et al., 2015)

PXA2 CNAG_02764 Mutants showed survival defects in CSF model (Lee et al., 2010)

MEP1 CNAG_04735 Target of Gat201 [master virulence TF] (Brown and Madhani,
2012);
reduced expression in kcs1Δ mutants [kcs1Δ do not elicit strong
immune response and are not readily phagocytosed] (Lev et al.,
2015)

70
FIGURES

Figure 2.1. Effect of either acetate or glucose as sole carbon source on transcriptome

of Cryptococcus neoformans. (A) Venn diagram showing number of DEGs for

acetate-grown and glucose-grown cells. (B) Volcano plot depicting DEGs as

turquoise dots and non-DEGs as orange dots. (C) Heatmap showing expression

patterns of all 356 DEGs across six replicates under two different carbon conditions.

71
Figure 2.2. Chromosomal distribution of the proportion of Differentially expressed

genes (DEGs) in C. neoformans. Proportion of total genes that are either upregulated

(lime green bar) or downregulated (purple bar) in each of the 14 chromosomes

calculated as the percentage of the total genes on the corresponding chromosome.

Vertical lines represent global expected proportions (i.e. overall percentage of the

upregulated [green line] or downregulated [purple line] gene for all 14

chromosomes).

72
REFERENCES

Baeza, L.C., da Mata, F.R., Pigosso, L.L., Pereira, M., de Souza, G.H.M.F., Coelho, A.S.G.,
de Almeida Soares, C.M., 2017. Differential metabolism of a two-carbon
substrate by members of the Paracoccidioides genus. Front. Microbiol. 8.

Bahn, Y.-S., Sun, S., Heitman, J., Lin, X., 2020. Microbe profile: Cryptococcus
neoformans species complex. Microbiology 166, 797.
https://doi.org/10.1099/mic.0.000973

Baker, E.H., Baines, D.L., 2018. Airway glucose homeostasis: A new target in the
prevention and treatment of pulmonary infection. CHEST 153, 507–514.
https://doi.org/10.1016/j.chest.2017.05.031

Bose, S., Ramesh, V., Locasale, J.W., 2019. Acetate metabolism in physiology, cancer,
and beyond. Trends Cell Biol. 29, 695–703.
https://doi.org/10.1016/j.tcb.2019.05.005

Brown, J.C.S., Madhani, H.D., 2012. Approaching the functional annotation of fungal
virulence factors using cross-species genetic interaction profiling. PLOS
Genet. 8, e1003168. https://doi.org/10.1371/journal.pgen.1003168

Camarasa, C., Faucet, V., Dequin, S., 2007. Role in anaerobiosis of the isoenzymes for
Saccharomyces cerevisiae fumarate reductase encoded by OSM1 and FRDS1.
Yeast 24, 391–401. https://doi.org/10.1002/yea.1467

Casadevall, A., Fu, M.S., Guimaraes, A.J., Albuquerque, P., 2019. The ‘Amoeboid
Predator-Fungal Animal Virulence’ hypothesis. J. Fungi 5, 10.
https://doi.org/10.3390/jof5010010

Chen, Y., Toffaletti, D.L., Tenor, J.L., Litvintseva, A.P., Fang, C., Mitchell, T.G., McDonald,
T.R., Nielsen, K., Boulware, D.R., Bicanic, T., Perfect, J.R., 2014. The Cryptococcus
neoformans transcriptome at the site of human meningitis. mBio 5, e01087-13.
https://doi.org/10.1128/mBio.01087-13

Chew, S.Y., Brown, A.J.P., Lau, B.Y.C., Cheah, Y.K., Ho, K.L., Sandai, D., Yahaya, H., Than,
L.T.L., 2021. Transcriptomic and proteomic profiling revealed reprogramming
of carbon metabolism in acetate-grown human pathogen Candida glabrata.
J. Biomed. Sci. 28, 1. https://doi.org/10.1186/s12929-020-00700-8

Deacon, J.W., Deacon, J.W., 2006. Fungal biology, 4th ed. ed. Blackwell Pub, Malden,
MA.

Desbois, S., John, U.P., Perugini, M.A., 2018. Dihydrodipicolinate synthase is absent in
fungi. Biochimie 152, 73–84. https://doi.org/10.1016/j.biochi.2018.06.017

73
Desjardins, C.A., Giamberardino, C., Sykes, S.M., Yu, C.-H., Tenor, J.L., Chen, Y., Yang, T.,
Jones, A.M., Sun, S., Haverkamp, M.R., Heitman, J., Litvintseva, A.P., Perfect, J.R.,
Cuomo, C.A., 2017. Population genomics and the evolution of virulence in the
fungal pathogen Cryptococcus neoformans. Genome Res. 27, 1207–1219.
https://doi.org/10.1101/gr.218727.116

Feldmesser, M., Kress, Y., Novikoff, P., Casadevall, A., 2000. Cryptococcus neoformans
is a facultative intracellular pathogen in murine pulmonary infection. Infect.
Immun. 68, 4225–4237. https://doi.org/10.1128/IAI.68.7.4225-4237.2000

Fu, M.S., Liporagi-Lopes, L.C., dos Santos, S.R., Tenor, J.L., Perfect, J.R., Cuomo, C.A.,
Casadevall, A., 2021. Amoeba predation of Cryptococcus neoformans results in
pleiotropic changes to traits associated with virulence. mBio 12, e00567-21.
https://doi.org/10.1128/mBio.00567-21

Garvey, B., 2007. Philosophy of biology, Philosophy and science. Acumen, Stocksfield.

Gow, N.A.R., Brown, A.J.P., 2018. Physiology and metabolism of fungal pathogens, in:
Kibbler, C.C., Barton, R., Gow, N.A.R., Howell, S., MacCallum, D.M., Manuel, R.J.,
Kibbler, C.C., Barton, R., Gow, N.A.R., Howell, S., MacCallum, D.M., Manuel, R.J.
(Eds.), Oxford Textbook of Medical Mycology. Oxford University Press, p. 0.
https://doi.org/10.1093/med/9780198755388.003.0003

Homer, C.M., Summers, D.K., Goranov, A.I., Clarke, S.C., Wiesner, D.L., Diedrich, J.K.,
Moresco, J.J., Toffaletti, D., Upadhya, R., Caradonna, I., Petnic, S., Pessino, V.,
Cuomo, C.A., Lodge, J.K., Perfect, J., Yates, J.R., Nielsen, K., Craik, C.S., Madhani,
H.D., 2016. Intracellular action of a secreted peptide required for fungal
virulence. Cell Host Microbe 19, 849–864.
https://doi.org/10.1016/j.chom.2016.05.001

Hu, G., Cheng, P.-Y., Sham, A., Perfect, J.R., Kronstad, J.W., 2008. Metabolic adaptation
in Cryptococcus neoformans during early murine pulmonary infection. Mol.
Microbiol. 69, 1456–1475. https://doi.org/10.1111/j.1365-2958.2008.06374.x

Huang, M., Hebert, A.S., Coon, J.J., Hull, C.M., 2015. Protein composition of infectious
spores reveals novel sexual development and germination factors in
Cryptococcus. PLOS Genet. 11, e1005490.
https://doi.org/10.1371/journal.pgen.1005490

Ianiri, G., Idnurm, A., 2015. Essential gene discovery in the basidiomycete
Cryptococcus neoformans for antifungal drug target prioritization. mBio 6,
e02334-14. https://doi.org/10.1128/mBio.02334-14

Idnurm, A., Giles, S.S., Perfect, J.R., Heitman, J., 2007. Peroxisome function regulates
growth on glucose in the basidiomycete fungus Cryptococcus neoformans.
Eukaryot. Cell 6, 60–72. https://doi.org/10.1128/EC.00214-06

74
Jennings, D.H., 1985. Polyol metabolism in fungi, in: Rose, A.H., Tempest, D.W. (Eds.),
Advances in Microbial Physiology. Academic Press, pp. 149–193.
https://doi.org/10.1016/S0065-2911(08)60292-1

Jin, J.-H., Lee, K.-T., Hong, J., Lee, D., Jang, E.-H., Kim, J.-Y., Lee, Yeonseon, Lee, S.-H., So,
Y.-S., Jung, K.-W., Lee, D.-G., Jeong, E., Lee, M., Jang, Y.-B., Choi, Y., Lee, M.H.,
Kim, J.-S., Yu, S.-R., Choi, J.-T., La, J.-W., Choi, H., Kim, S.-W., Seo, K.J., Lee, Yelin,
Thak, E.J., Choi, J., Averette, A.F., Lee, Y.-H., Heitman, J., Kang, H.A., Cheong, E.,
Bahn, Y.-S., 2020. Genome-wide functional analysis of phosphatases in the
pathogenic fungus Cryptococcus neoformans. Nat. Commun. 11, 4212.
https://doi.org/10.1038/s41467-020-18028-0

Jung, K.-W., Yang, D.-H., Maeng, S., Lee, K.-T., So, Y.-S., Hong, J., Choi, J., Byun, H.-J.,
Kim, H., Bang, S., Song, M.-H., Lee, J.-W., Kim, M.S., Kim, S.-Y., Ji, J.-H., Park, G.,
Kwon, H., Cha, S., Meyers, G.L., Wang, L.L., Jang, J., Janbon, G., Adedoyin, G.,
Kim, T., Averette, A.K., Heitman, J., Cheong, E., Lee, Y.-H., Lee, Y.-W., Bahn, Y.-S.,
2015. Systematic functional profiling of transcription factor networks in
Cryptococcus neoformans. Nat. Commun. 6, 6757.
https://doi.org/10.1038/ncomms7757

Jung, W.H., Hu, G., Kuo, W., Kronstad, J.W., 2009. Role of Ferroxidases in iron uptake
and virulence of Cryptococcus neoformans. Eukaryot. Cell 8, 1511–1520.
https://doi.org/10.1128/EC.00166-09

Jung, W.-H., Son, Y.-E., Oh, S.-H., Fu, C., Kim, H.S., Kwak, J.-H., Cardenas, M.E., Heitman,
J., Park, H.-S., 2018. Had1 is required for cell wall integrity and fungal virulence
in Cryptococcus neoformans. G3 GenesGenomesGenetics 8, 643–652.
https://doi.org/10.1534/g3.117.300444

Kim, G.B., Gao, Y., Palsson, B.O., Lee, S.Y., 2021. DeepTFactor: A deep learning-based
tool for the prediction of transcription factors. Proc. Natl. Acad. Sci. 118,
e2021171118. https://doi.org/10.1073/pnas.2021171118

Kim, J., Cho, Y.-J., Do, E., Choi, J., Hu, G., Cadieux, B., Chun, J., Lee, Y., Kronstad, J.W.,
Jung, W.H., 2012. A defect in iron uptake enhances the susceptibility of
Cryptococcus neoformans to azole antifungal drugs. Fungal Genet. Biol. 49,
955–966. https://doi.org/10.1016/j.fgb.2012.08.006

Kingsbury, J.M., McCusker, J.H.Y. 2008, 2008. Threonine biosynthetic genes are
essential in Cryptococcus neoformans. Microbiology 154, 2767–2775.
https://doi.org/10.1099/mic.0.2008/019729-0

Kingsbury, J.M., Yang, Z., Ganous, T.M., Cox, G.M., McCusker, J.H., 2004. Novel chimeric
spermidine synthase-saccharopine dehydrogenase gene (SPE3-LYS9) in the
human pathogen Cryptococcus neoformans. Eukaryot. Cell 3, 752–763.
https://doi.org/10.1128/EC.3.3.752-763.2004

75
Kretschmer, M., Reiner, E., Hu, G., Tam, N., Oliveira, D.L., Caza, M., Yeon, J.H., Kim, J.,
Kastrup, C.J., Jung, W.H., Kronstad, J.W., 2014. Defects in phosphate acquisition
and storage influence virulence of Cryptococcus neoformans. Infect. Immun.
82, 2697–2712. https://doi.org/10.1128/IAI.01607-14

Kretschmer, M., Wang, J., Kronstad, J.W., 2012. Peroxisomal and mitochondrial β-
oxidation pathways influence the virulence of the pathogenic fungus
Cryptococcus neoformans. Eukaryot. Cell 11, 1042–1054.
https://doi.org/10.1128/ec.00128-12

Kronstad, J., Saikia, S., Nielson, E.D., Kretschmer, M., Jung, W., Hu, G., Geddes, J.M.H.,
Griffiths, E.J., Choi, J., Cadieux, B., Caza, M., Attarian, R., 2012. Adaptation of
Cryptococcus neoformans to mammalian hosts: integrated regulation of
metabolism and virulence. Eukaryot. Cell 11, 109–118.
https://doi.org/10.1128/EC.05273-11

Kumar, R., Reichert, A.S., 2021. Common principles and specific mechanisms of
mitophagy from yeast to humans. Int. J. Mol. Sci. 22, 4363.
https://doi.org/10.3390/ijms22094363

Lee, A., Toffaletti, D.L., Tenor, J., Soderblom, E.J., Thompson, J.W., Moseley, M.A., Price,
M., Perfect, J.R., 2010. Survival defects of Cryptococcus neoformans mutants
exposed to human cerebrospinal fluid result in attenuated virulence in an
experimental model of meningitis. Infect. Immun. 78, 4213–4225.
https://doi.org/10.1128/IAI.00551-10

Lee, K.-T., Hong, J., Lee, D.-G., Lee, M., Cha, S., Lim, Y.-G., Jung, K.-W., Hwangbo, A., Lee,
Y., Yu, S.-J., Chen, Y.-L., Lee, J.-S., Cheong, E., Bahn, Y.-S., 2020. Fungal kinases
and transcription factors regulating brain infection in Cryptococcus
neoformans. Nat. Commun. 11, 1521. https://doi.org/10.1038/s41467-020-15329-2

Lee, K.-T., Lee, J.-W., Lee, D., Jung, W.-H., Bahn, Y.-S., 2014. A ferroxidase, Cfo1,
regulates diverse environmental stress responses of Cryptococcus
neoformans through the HOG pathway. Mycobiology 42, 152–157.
https://doi.org/10.5941/MYCO.2014.42.2.152

Lev, S., Li, C., Desmarini, D., Saiardi, A., Fewings, N.L., Schibeci, S.D., Sharma, R., Sorrell,
T.C., Djordjevic, J.T., 2015. Fungal inositol pyrophosphate IP7 Is crucial for
metabolic adaptation to the host environment and pathogenicity. mBio 6,
e00531-00515. https://doi.org/10.1128/mBio.00531-15

Liu, O.W., Chun, C.D., Chow, E.D., Chen, C., Madhani, H.D., Noble, S.M., 2008.
Systematic genetic analysis of virulence in the human fungal pathogen
Cryptococcus neoformans. Cell 135, 174–188.
https://doi.org/10.1016/j.cell.2008.07.046

76
Luberto, C., Martinez-Mariño, B., Taraskiewicz, D., Bolaños, B., Chitano, P., Toffaletti,
D.L., Cox, G.M., Perfect, J.R., Hannun, Y.A., Balish, E., Del Poeta, M., 2003.
Identification of App1 as a regulator of phagocytosis and virulence of
Cryptococcus neoformans. J. Clin. Invest. 112, 1080–1094.
https://doi.org/10.1172/JCI18309

Maier, E.J., Haynes, B.C., Gish, S.R., Wang, Z.A., Skowyra, M.L., Marulli, A.L., Doering, T.L.,
Brent, M.R., 2015. Model-driven mapping of transcriptional networks reveals
the circuitry and dynamics of virulence regulation. Genome Res. 25, 690–700.
https://doi.org/10.1101/gr.184101.114

May, R.C., Stone, N.R.H., Wiesner, D.L., Bicanic, T., Nielsen, K., 2016. Cryptococcus: from
environmental saprophyte to global pathogen. Nat. Rev. Microbiol. 14, 106–117.
https://doi.org/10.1038/nrmicro.2015.6

Maziarz, E.K., Perfect, J.R., 2016. Cryptococcosis. Infect. Dis. Clin. North Am., Fungal
Infections 30, 179–206. https://doi.org/10.1016/j.idc.2015.10.006

Missall, T.A., Pusateri, M.E., Lodge, J.K., 2004. Thiol peroxidase is critical for virulence
and resistance to nitric oxide and peroxide in the fungal pathogen,
Cryptococcus neoformans. Mol. Microbiol. 51, 1447–1458.
https://doi.org/10.1111/j.1365-2958.2004.03921.x

Naranjo-Ortiz, M.A., Gabaldón, T., 2019. Fungal evolution: major ecological


adaptations and evolutionary transitions. Biol. Rev. Camb. Philos. Soc. 94,
1443–1476. https://doi.org/10.1111/brv.12510

Owen, D.H., Katz, D.F., 1999. A vaginal fluid simulant. Contraception 59, 91–95.
https://doi.org/10.1016/S0010-7824(99)00010-4

Paes, H.C., Derengowski, L. da S., Peconick, L.D.F., Albuquerque, P., Pappas, G.J.,
Nicola, A.M., Silva, F.B.A., Vallim, M.A., Alspaugh, J.A., Felipe, M.S.S., Fernandes,
L., 2018. A Wor1-like transcription factor is essential for virulence of
Cryptococcus neoformans. Front. Cell. Infect. Microbiol. 8, 369.
https://doi.org/10.3389/fcimb.2018.00369

Pandey, S.K., Yadav, S., Temre, M.K., Singh, S.M., 2018. Tracking acetate through a
journey of living world: Evolution as alternative cellular fuel with potential for
application in cancer therapeutics. Life Sci. 215, 86–95.
https://doi.org/10.1016/j.lfs.2018.11.004

Pietrocola, F., Galluzzi, L., Bravo-San Pedro, J.M., Madeo, F., Kroemer, G., 2015. Acetyl
coenzyme A: a central metabolite and second messenger. Cell Metab. 21, 805–
821. https://doi.org/10.1016/j.cmet.2015.05.014

Price, M.S., Betancourt-Quiroz, M., Price, J.L., Toffaletti, D.L., Vora, H., Hu, G., Kronstad,
J.W., Perfect, J.R., 2011. Cryptococcus neoformans requires a functional
glycolytic pathway for disease but not persistence in the host. mBio 2, e00103-
00111. https://doi.org/10.1128/mBio.00103-11

77
Ries, L.N.A., Alves de Castro, P., Pereira Silva, L., Valero, C., Dos Reis, T.F., Saborano, R.,
Duarte, I.F., Persinoti, G.F., Steenwyk, J.L., Rokas, A., Almeida, F., Costa, J.H., Fill,
T., Sze Wah Wong, S., Aimanianda, V., Rodrigues, F.J.S., Gonçales, R.A., Duarte-
Oliveira, C., Carvalho, A., Goldman, G.H., 2021. Aspergillus fumigatus acetate
utilization impacts virulence traits and pathogenicity. mBio 12, e0168221.
https://doi.org/10.1128/mBio.01682-21

Ries, L.N.A., Beattie, S., Cramer, R.A., Goldman, G.H., 2018. Overview of carbon and
nitrogen catabolite metabolism in the virulence of human pathogenic fungi.
Mol. Microbiol. 107, 277–297. https://doi.org/10.1111/mmi.13887

Rude, T.H., Toffaletti, D.L., Cox, G.M., Perfect, J.R., 2002. Relationship of the glyoxylate
pathway to the pathogenesis of Cryptococcus neoformans. Infect. Immun. 70,
5684–5694. https://doi.org/10.1128/IAI.70.10.5684-5694.2002

Santiago-Tirado, F.H., Peng, T., Yang, M., Hang, H.C., Doering, T.L., 2015. A single
protein S-acyl transferase acts through diverse substrates to determine
cryptococcal morphology, stress tolerance, and pathogenic outcome. PLoS
Pathog. 11, e1004908. https://doi.org/10.1371/journal.ppat.1004908

Seehusen, D.A., Reeves, M.M., Fomin, D.A., 2003. Cerebrospinal fluid analysis. Am.
Fam. Physician 68, 1103–1109.

Springer, D.J., Mohan, R., Heitman, J., 2017. Plants promote mating and dispersal of
the human pathogenic fungus Cryptococcus. PloS One 12, e0171695.
https://doi.org/10.1371/journal.pone.0171695

Valdés-Santiago, L., Ruiz-Herrera, J., 2013. Stress and polyamine metabolism in fungi.
Front. Chem. 1, 42. https://doi.org/10.3389/fchem.2013.00042

Wang, P., Cardenas, M.E., Cox, G.M., Perfect, J.R., Heitman, J., 2001. Two cyclophilin A
homologs with shared and distinct functions important for growth and
virulence of Cryptococcus neoformans. EMBO Rep. 2, 511–518.
https://doi.org/10.1093/embo-reports/kve109

Wang, Z., Kin, K., López-Giráldez, F., Johannesson, H., Townsend, J.P., 2012. Sex-
specific gene expression during asexual development of Neurospora crassa.
Fungal Genet. Biol. FG B 49, 533–543. https://doi.org/10.1016/j.fgb.2012.05.004

Yu, C.-H., Chen, Y., Desjardins, C.A., Tenor, J.L., Toffaletti, D.L., Giamberardino, C.,
Litvintseva, A., Perfect, J.R., Cuomo, C.A., 2020. Landscape of gene expression
variation of natural isolates of Cryptococcus neoformans in response to
biologically relevant stresses. Microb. Genomics 6.
https://doi.org/10.1099/mgen.0.000319

Yu, C.-H., Sephton-Clark, P., Tenor, J.L., Toffaletti, D.L., Giamberardino, C., Haverkamp,
M., Cuomo, C.A., Perfect, J.R., 2021. Gene expression of diverse Cryptococcus
isolates during infection of the human central nervous system. mBio 12,
e02313-21. https://doi.org/10.1128/mBio.02313-21

78
Zabriskie, T.M., Jackson, M.D., 2000. Lysine biosynthesis and metabolism in fungi. Nat.
Prod. Rep. 17, 85–97. https://doi.org/10.1039/a801345d

Zaragoza, O., 2019. Basic principles of the virulence of Cryptococcus. Virulence 10,
490–501. https://doi.org/10.1080/21505594.2019.1614383

79
CHAPTER 3

Biochemical characterization of acetate kinase

from the fungal pathogen Cryptococcus

neoformans and its role in acetate production

Oly Ahmed, Cheryl Ingram-Smith, Ann Guggisberg, Sheena


Henry, Ashley Mattison, Aigerim Bizhanova, and Kerry S. Smith

Department of Genetics and Biochemistry; Eukaryotic Pathogen Innovation Center,

Clemson University, SC, USA

Authors contribution:

KS and CIS conceived and designed the experiments; CIS, AG, SH, AM, AB cloned the
cDNA, designed the protocol for production and isolation of Ack, and performed
kinetic assays in acetyl phosphate forming direction; OA performed the kinetic
assays in acetate forming direction, designed and performed the in silico
identification and phylogenetic analysis of the eukaryotic Acks, and prepared the
manuscript.

80
ABSTRACT

Acetate kinase (EC 2.7.2.1; Ack) catalyzes the interconversion of acetyl

phosphate (AcP) and acetate where the phosphoryl group donor/acceptor for the

reaction is either ATP/ADP in genera of Bacteria and the genus Methanosarcina of

Archaea or inorganic pyrophosphate (PPi)/inorganic phosphate (Pi) in eukaryotic

amoebic pathogen Entamoeba histolytica. Excepting the example of E. histolytica

Ack, which is a PPi/Pi-dependant Ack, there is a dearth of information regarding the

kinetic properties of eukaryotic Ack (eAck). Ack from the human fungal pathogen

Cryptococcus neoformans (CnAck) is thought to play an important role in

acetate/acetyl-CoA metabolism, which has been implicated in cryptococcal

infections. Here we studied the kinetic behavior of CnAck and observed that

kinetically CnAck has a preference in the direction of acetate formation over that of

AcP formation. Based on the directionality of Ack catalyzed reaction we propose that

Ack forms a pathway with Xfp2 (xylulose-5-phosphate /fructose-6-phosphate

phosphoketolase; xylulose-5-phosphate / fructose-6-phosphate + Pi →

glyceraldehyde-3-phosphate / erythrose-4-phosphate + AcP) to form a pathway of

acetate production in C. neoformans. In addition, we have predicted 526 putative

eAck from a database constructed from 2,119 published proteomes/genomes, and

built a maximum likelihood tree. A majority of the predicted eAcks were of fungal

origin and monophyletic, with a comparable rate of evolution within the fungal

superclade suggesting a history of common ancestry. This is the first report of

81
biochemical characterization of an ATP/ADP-dependent Ack of eukaryotic origin.

Moreover, this Ack’s possible involvement in an acetate production pathway in C.

neoformans may have interesting implications for cryptococcal virulence and hence

warrants further study.

3.1. INTRODUCTION

Acetate kinase (Ack; EC 2.7.2.1) catalyzes the reaction acetyl phosphate (AcP) +

ADP ↔ acetate + ATP, and is abundant in Bacteria and one methanogenic genus

from Archaea (Ferry, 2011). In Bacteria, Ack partners with phosphotransacetylase (Pta,

EC 2.3.1.8) to form a pathway of acetate formation during fermentative growth

(Zhang et al., 2021); whereas in the Archaea, such as Methanosarcina spp., Ack-Pta

pathway is utilized for methane formation (Ferry, 2011). Additionally, Ack has been

predicted in many eukaryotic genomes, notably in fungi (Ingram-Smith et al., 2006).

Outside the kingdom Fungi, the Ack from the eukaryotic amoebic pathogen

Entamoeba histolytica has been studied. The E. histolytica Ack differs from well-

studied prokaryotic Acks in its preference for inorganic pyrophosphate

(PPi)/inorganic phosphate (Pi) over ATP/ADP as phosphoryl group donor/acceptor

(Fowler et al., 2012). Surprisingly, Pta, the typical Ack partner, is not present in E.

histolytica (Ingram-Smith et al., 2006; Fowler et al., 2012). Instead, Ack is suggested to

play a role in maintaining intracellular NAD+/NADH ratio during growth on glucose

through an extended glycolytic pathway (Dang et al., 2022). Thus, variation in the

metabolic role of Ack appears to be niche and species-dependent, which opens up

the possibility of exploiting Ack’s functional relevance in other pathogens for

targeted intervention.

82
The structure of Ack from the archaeon Methanosarcina thermophila was the

first one to be solved (Buss et al., 2001; Fowler et al., 2012). Later employment of

transition-state analog (Miles et al., 2002; Gorrell et al., 2005) and substrate analog

(Gorrell et al., 2005) to study the structural and kinetic properties of M. thermophila

Ack provided evidence for a direct in-line mechanism of phosphoryl group transfer.

Subsequently, the structures of the eukaryotic Ack from the amoebic species E.

histolytica and fungal species Cryptococcus neoformans were solved. The difference

between the fungal and amoebic Ack with respect to their specificity for ATP and

PPi, respectively, was explained by subtle amino acid substitutions within the

catalytic cleft (Thaker et al., 2013). The structures of the eukaryotic Ack strongly

resemble those of their prokaryotic counterparts and thus are considered to be

consistent with the proposed direct in-line mechanism of phosphoryl group transfer

(Thaker et al., 2013).

Human fungal pathogen Cryptococcus neoformans belongs to the phylum

Basidiomycota, and has a wider distribution both geographically and ecologically.

Spores and desiccated yeast of C. neoformans can become airborne and are inhaled

by its mammalian hosts, including humans. This mostly leads to initial pulmonary

infection in immunocompromised people, which can subsequently lead to fungal

meningitis through systemic dissemination. During pulmonary infection, C.

neoformans is exposed to a low glucose environment, and appears to metabolically

adapt to the pulmonary microenvironment through upregulation of pathways of

alternative carbon source metabolism (for an updated and extensive review see

(Berguson et al., 2022)). In the absence of simple sugars, C. neoformans relies on 2-C

compounds (e.g., acetate, ethanol) via upregulating the glyoxylate cycle; however

83
this pathway seems to be dispensable for virulence (Berguson et al., 2022). Thus, it is

plausible that C. neoformans utilizes other non-canonical pathway(s) of acetate

metabolism both in the context of alternative carbon utilization and maintenance of

intracellular pool of acetate for growth (e.g. lipid synthesis) and epigenetic

regulations (e.g. histone acetylation). Presence of Ack in C. neoformans (FungiDB

gene ID: CNAG_06432) raises the possibility of a hitherto unknown acetate metabolic

pathway.

In this study, we investigated the kinetic characteristics of Ack from C.

neoformans (CnAck), which is the first report of biochemical characterization of a

fungal Ack. Moreover, we surveyed a curated, reference proteome database to

identify putative Ack across the domain Eukarya, and reconstructed a phylogeny to

better understand the possible evolutionary relationships among this large set of

identified eukaryotic acetate kinases (eAck).

3.2. METHODS AND MATERIALS

3.2.1. Production and purification of CnAck

Production and purification of CnAck were achieved following methods

described in (Thaker et al., 2013) with slight modifications. Briefly, the cDNA encoding

CnAck was cloned into the pET21b expression plasmid and transformed into E. coli

Rosetta2 (DE3) cells. Transformants were cultured in LB broth with appropriate

antibiotics at 37°C with shaking to a final OD600 of 0.75, followed by induction of the

ack gene expression through addition of IPTG to a final concentration of 1 mM. Cells

were harvested by centrifugation and resuspended in the purification buffer as

84
described: 25 mM Tris, 150 mM NaCl, 20 mM imidazole, and 10% glycerol, pH 7.4

(Thaker et al., 2013). Harvested cells were lysed using French press (20,000 psi),

followed by clarification of the lysate through ultracentrifugation. Finally, CnAck was

purified to electrophoretic homogeneity on a 5 mL HisTrap Ni2+-affinity column using

the exact elution protocol described previously (Thaker et al., 2013).

3.2.2. Kinetic assays of CnACK

The reaction in the direction of AcP/ADP formation was measured using

hydroxamate assay exactly as previously described (Ingram-Smith et al., 2015). The

reaction in the direction of acetate/ATP formation was measured by coupling CnAck

reaction with hexokinase and glucose-6-phosphate dehydrogenase at 37°C in a

reaction buffer consisting of 100 mM Tris-HCl (pH 7.6), 10 mM MgCl2, 5.5 mM

dextrose, 0.2 mM DTT, 1 mM NADP+, with one substrate held constant at saturation

level and the other varied, and measuring the formation of NADPH at 340 nm.

Measurements for various metal cofactors and NTPs were performed using

hydroxamate assays in the AcP/ADP formation direction. Reactions were done in

triplicate; error was calculated in standard deviation; Prism 5 (GraphPad Software,

Dotmatics) was used for curve-fitting.

3.2.3. Identification of putative eukaryotic Ack

To reconstruct the phylogenetic relations among the eukaryotic Ack we

selected a total of 526 putative Ack sequences that were identified using a profile

hidden Markov model (HMM) and searched against a database of 34,059,850 protein

85
sequences from 2,119 genomes (proteomes) (see below). The HMM profile and the

query database used in this study were based on the available latest releases from

PRIAM (Claudel‐Renard et al., 2003) and Uniprot (The UniProt Consortium, 2021),

respectively (PRIAM ID: PRI000886, 2018 release; Uniprot Reference Proteomes,

release 2022_03).

The search was conducted using HMMER version 3.3.2 (Eddy, 2011), and the

returned positive hits were further filtered by the reported E-values and bit scores.

Sequences with dissimilar E-values/ bit scores for ‘overall’ hit versus ‘best-domain’ hit

were considered spurious. Remaining sequences were aligned using KALIGN

(default parameters; Unipro UGENE v 43.0 (Okonechnikov et al., 2012)). The resulting

multiple sequence alignment (MSA) was manually edited to remove excess gaps

and mismatches. Shorter sequences (i.e. < 95% of the length of the longest sequence

in the MSA; modal length is 331 amino acids) were further removed from the final

alignment to avoid suspicious branching patterns due to the long-branch attraction

effect in the subsequent tree-construction. Finally, redundant genomes (i.e. multiple

sequenced genomes and/or strains of the same species) were discarded from the

analysis (only one Ack per species was used in the final tree). A complete list of eACK

used for construction of the phylogenetic tree, along with their corresponding

UniProt accession ID can be found in Table B1.

3.2.4. Construction of the phylogenetic tree

A maximum likelihood (ML) tree was constructed using IQ-TREE 2 (Minh et al.,

2020). Following evolutionary models were tested: Blosum 62, Dayhoff, JTT, LG,

Poisson, VT, and WAG. Based on the Bayesian Information Criterion (BIC), LG+I+I+R9

86
(i.e, LG substitution model and, both invariant site and FreeRate with 9 categories)

were chosen for the subsequence ML tree construction. To reduce the

computational burden, 2,000 bootstrap replicates were constructed using UFBoot

(an ultrafast bootstrap approximation (Hoang et al., 2018)) (with option appended for

further optimization in case of severe model violation). Readers should be cautious

about the interpretation of the bootstrap values obtained using UFBoot as they

differ from those of standard nonparametric bootstrapping: the authors of IQ-TREE 2

recommend using 95% bootstrap as a reasonable cutoff. The tree was rooted using

Ack from the bacteria Salmonella typhimurium (PDB: 3SLC). The tree was visualized

using the web interface of the Interactive Tree of Life (https://itol.embl.de/).

3.3. RESULTS

3.3.1. Ack in C. neoformans is possibly unidirectional in its biological role

Acetate kinase (Ack) catalyzes the interconversion between ATP/acetate and

ADP/acetyl phosphate, and is reliant on divalent metal cofactors for the reaction

(Chittori et al., 2012). Previously, crystallographic structures of two eAck enzymes

have been reported: E. histolytica and C. neoformans (Thaker et al., 2013).

Catalytically, the major difference between these two eAcks is the preference for

either ATP/ADP or PPi/Pi as the phosphoryl group donor/acceptor in CnAck and

EhAck catalyzed reactions respectively (Thaker et al., 2013). Although kinetic

experiments were performed on EhAck previously (Fowler et al., 2011; Pineda et al.,

2016; Dang and Ingram-Smith, 2017), a biochemical characterization of an ATP-

87
dependent eAck has not been reported yet. Our current study of CnAck kinetics is

thus the first report of kinetic properties of any ATP-dependent eAck.

In the current study we observed that the catalytic efficiency of CnAck is

higher for Mg2+ compared to Mn2+, Co2+, and Zn2+ (in the AcP forming direction; Table

3.1); hence in the subsequent kinetic assays, we used Mg2+ as the preferred cofactor

of CnAck. It should be noted that although the catalytic efficiencies (kcat/KM) of CnAck

with Zn2+ and Mg2+ are somewhat comparable, the KM is higher for Zn2+ than for Mg2+.

CnACK specific activity was determined with different NTPs to identify the

best phosphoryl donor. The specific activity of CnAck was found to be considerably

higher with ATP than CTP, GTP, TTP, UTP, or ITP (assays were performed in the AcP

forming direction; Table 3.2), indicating ATP is the preferred phosphoryl donor.

As Ack is classified as a member of the acetate/propionate kinase superfamily

(InterPro ID: IPR004372; Chittori et al., 2012), we examined CnAck’s ability to use

propionate as a substrate. and found the apparent KM for propionate to be 80.7 ± 8.3

mM, 2.8-fold higher than that for acetate, but the kcat was only 0.05 ± 0.01 sec-1, a

0.03-fold decrease compared to that for acetate. In the acetyl phosphate/ADP

forming direction, CnAck has a high apparent KM for the substrate acetate (KM = 28.8

± 1.1 mM) compared to ATP (KM = 1.5 ± 0.1 mM) and the substrates in the acetate

forming direction (AcP KM = 0.5 ± 0.1 mM; ADP KM = 2.3 ± 0.4 mM). The catalytic

efficiency (kcat/KM) of CnAck is also low for the substrates in the AcP forming direction

compared to acetate forming direction, and has a 19-fold higher Kcat for AcP versus

acetate forming direction (Table 3.3). Taken together, our results suggest a strong

preference of CnAck for the acetate/ATP forming direction in vitro.

88
3.3.2. Majority of the detected eukaryotic Ack are of fungal origin

Historically, Ack was thought to be found only in genera of the Bacteria and

the single archaeon M. thermophila (Ingram-Smith et al., 2006). However, with the

increase in the number of sequenced genomes, it was soon discovered that many

eukaryotic organisms belonging to Ascomycota, Basidiomycota, Chytridiomycota,

Oomycota, Amoebozoa, and Chlorophyta have Ack genes (Ingram-Smith et al.,

2006). In this work, we also wanted to update that study as the modern genome

databases experienced an exponential increase in the number of genome

sequences deposited, and computation techniques and resources have improved

significantly ever since the previous study. Figure 3.1 summarizes the number of

species in various taxonomic groups with putative Ack genes identified in this study

along with the constructed maximum likelihood tree showing the phylogenetic

relation among putative eAcks.

Here we found that the majority of the Ack detected in this study form a

single monophyletic clade with comparable branch lengths (Figure 3.2 A); not

surprisingly these Ack are of fungal origin, belonging to phylum Ascomycota (63.5%)

and Basidiomycota (22.4%). Other fungal species (dubbed as ‘Enigmatic Fungi’ in

reference to the uncertainty around their placement in the tree of life) also form

sister clades to the abovementioned dikaryotic fungi groups, representing a small

fraction of the dataset (i.e. 2.9%; comprising of Mucoromycota, Chytridiomycota,

Olpidiomycota, and Zoopagomycota). The only non-fungal Ack that forms a tight

cluster with all the fungal Ack is detected in the dicot (also the only dicot found in

the dataset) plant Carpinus fangiana (Monkeytail hornbeam). It is not certain if this

dicot plant has acquired the gene for Ack through some parallel evolutionary

89
mechanism (possibly via endophytic fungus); however, C. fangiana Ack shares 40.8%

identity with Ack of C. neoformans. With the exception of C. fangiana, all the

remaining non-fungal Ack (i.e. 11%) are paraphyletic to the said fungal super-clade

(Figure 3.2 B).

3.3.3. Non-fungal Ack may be functionally distinct from fungal-Ack

The second largest taxonomic group in our dataset is the SAR (Stramenopiles,

Alveolata, Rhizaria) supergroup, followed by Viridiplantae (green plants). As

mentioned above, except for the dicot C. fangiana, other members of the

Viridiplantae are paraphyletic to the fungal superclade. From the SAR group, only

stramenopiles and alveolates are represented in our dataset; however presence of

Ack in Rhizaria cannot be ruled out at this point since the taxonomic group of

Rhizaria is not well represented in the proteome database built on UniProt release

(proteomes of only three species of Rhizaria are present out of 2,119 reference

proteomes). This also raises the possibility of underrepresentation of certain

taxonomic groups, and hence our uncertainty regarding the presence of Ack in

many other minor eukaryotic groups yet to be sequenced.

Putative Ack sequences are also detected in metazoan species: two

arthropods (Tigriopus californicus and Hyalella azteca), a cnidarian (Clytia

hemisphaerica), and a nematode (Caenorhabditis remanei). Additionally, three

amoebic species also have Ack: Planoprotostelium fungivorum, Entamoeba

histolytica, and Entamoeba invadens. Unlike the CnAck reported here, Ack in E.

90
histolytica utilizes inorganic pyrophosphate as sole phosphate donor (Fowler et al.,

2012); this functional distinction between fungal Ack and amoebic Ack leads us to

speculate about the potential existence of similar pyrophosphate-dependent Ack in

non-fungal eukaryotes.

3.4. DISCUSSION

Here we studied the kinetic parameters of CnAck heterologously produced in

E. coli and found that CnAck preferentially utilizes Mg2+ as the divalent metal

cofactor for interconversion between AcP and acetate with ATP/ADP as preferred

phosphoryl group donor/acceptor. Comparison of KM values and catalytic efficiency

for the substrates acetate and AcP suggests that Ack in C. neoformans may be

metabolically unidirectional with a strong preference in the acetate-forming

direction.

In many fermentative bacteria and the archaea Methanosarcina, Ack and Pta

form a pathway which is, in principle, bidirectional. Moreover, the relevance of the

Ack-Pta pathway is niche-specific. For example, Ack-Pta pathway is abundant in

fermentative bacteria, where it functions in the conversion of acetyl-CoA to acetate

and ATP production, or in archaea Methanosarcina spp. as the initial steps of

methanogenesis (Gorrell and Ferry, 2007). This pathway is also an important player

in the acetate switch, i.e., the cellular metabolic process involved in regulating either

acetate secretion or assimilation based on nutrient availability, in many bacteria

(Barnhart et al., 2015). Depending on the species and niche this pathway may

actually be very essential for survival as well. For example, in methanogenic

archaeon Methanosarcina acetivorans, the Ack-Pta pathway is required for growth

91
on acetate as a carbon and energy source for methane formation and on carbon

monoxide as the sole carbon source as this pathway is involved in acetogenesis from

carbon monoxide (Rother and Metcalf, 2004). Outside the bacteria and

Methanosarcina, the Ack-Pta pathway has only been investigated in the unicellular,

soil-dwelling, green alga Chlamydomonas reinhardtii, where both the enzymes have

mitochondrial and chloroplastic homologs. Ack1-Pat2 localizes in chloroplast and

plays a more prominent role in acetate production under dark, anoxic conditions

compared to Ack2-Pat1, the mitochondrial counterpart (note the abbreviation for Pta

in this species is Pat) (Yang et al., 2014). The other example of eukaryotic Ack studied

so far is from amoeba E. histolytica. However, in E. histolytica, Ack lacks the typical

Ack partners Pta and Xfp (Fowler et al., 2012), and it does not seem to play a major

role in acetate production, but rather helps maintain the NAD+/NADH balance

during ethanol synthesis in the extended glycolytic pathway (Dang et al., 2022).

The gene for Pta is missing from C. neoformans genome, hence Ack-Pta

pathway is not tenable in this organism. Incidentally, AcP, the substrate for Ack, can

also be synthesized from xylulose 5-phosphate (X5P) or fructose 6-phosphate (F6P)

and inorganic phosphate (Pi) via the action of the enzyme xylulose 5-phosphate/

fructose 6-phosphate phosphoketolase (Xfp) (Ingram-Smith et al., 2006); previously

an isozyme of Xfp from C. neoformans has been reported and characterized (Glenn

et al., 2014). Xfp can form a pathway with Ack for acetate and ATP production (e.g.

bifid shunt in bacteria Bifidobacterium spp. which is responsible for conversion of

hexose sugar into acetate), or Xfp can also form a pathway with Pta for acetyl-CoA

production from hexoses (e.g. as in lactic acid bacteria and bifidobacteria) (Henard et

al., 2015).

92
The cryptococcal isozyme Xfp2 is allosterically inhibited by ATP (energy

pathway), phosphoenol pyruvate (glycolysis), and oxaloacetic acid (the TCA cycle);

and is activated by AMP (low energy condition) (Glenn et al., 2014). Thus Xfp2, in

combination with Ack, is likely to form a pathway that leads to the formation of ATP

and acetate from phosphoketose sugars and inorganic phosphate (Figure 3.3).

There is a dearth of information regarding the metabolism of AcP in eukaryotic cells;

however in recent times application of modern analytical techniques such as 2D-

NMR spectroscopy helped detect the transient nature of AcP formation in human

mitochondria, suggestive of its role as an intermediate of acetate formation in

mammalian cells, although this may not be the primary mechanism (Xu et al., 2018).

A plausible analogous mechanism of fast conversion of AcP to acetate in C.

neoformans becomes explicable through our observation that CnAck shows high

catalytic efficiency in acetate formation direction.

To understand the evolutionary history of Ack in Eukarya, we have used a

hidden Markov model (HMM)-based searching strategy against a curated database

of 2,119 sequenced eukaryotic genomes. Our conservative estimation is that 526

eukaryotic species appear to have putative Ack sequences in their genomes (note

that this estimate excludes extraneous multiple versions or strains of the same

species). The majority (~88%) of this list belongs to the group Fungi. Phylogenetic

analysis shows a tight clustering of the fungal Ack in a single super-clade with

comparable branch lengths. Such monophyly and nearly uniform rate of evolution

across the fungal Ack is strongly suggestive of a common ancestry. Similarly, most, if

not all, non-fungal Ack also show reasonable monophyletic relation within a given

taxa; however the branch lengths differ notably from one taxon to the next,

93
suggesting taxon-specific rates of divergence which may be reflective of the

organism’s niche and life-history.

At the moment nothing conclusive can be said from our dataset about the

distributions of Ack among fungi and non-fungal eukaryotes, since the

disproportionate distribution (i.e. 88.8% fungi vs. 11.2% non-fungi) is likely to have

arisen from a corresponding disproportionate representation of these two groups in

the total genomes sequenced to date. Despite such methodological bias it became

apparent from our phylogenetic tree that modern day protozoan, fungal, and algal

Acks have likely been derived from a common ancestral sequence. The diversity of

the ecological niches occupied and the sheer variations in the life-histories of the

species identified here make convergent evolution of this enzyme improbable;

however, overlap between species-taxon and Ack-clade, and similar branch-length

within the clades are consistent with both the scenarios of parallel and divergent

evolution. Widespread distribution of Ack in Bacteria, and in few species of Archaea,

points to the existence of a more ancient ancestral sequence and subsequent

divergence; in fact, from the consideration of structural conservation, a possibility of

an ancestral sequence has previously been suggested which might have led to the

evolution of both Ack and Acetyl-CoA synthetase (Acs), another crucial enzyme of

acetate metabolism, via duplication and divergence in all three domains of life

(Barnhart et al., 2015). Moreover, if AcP is indeed a primordial energy currency of the

ancient world that potentially coupled the energy with carbon flux as argued by

(Whicher et al., 2018), it will not be surprising to assume that any ancestral enzyme

that enhanced the rate of phosphorylation by AcP and thereby catalyzed the crucial

process of coupling carbon flux and energy generation, might have belonged to the

94
Last Universal Common Ancestor (LUCA)-protein set (Elias and Tawfik, 2012) and is at

the heart of the ancient cellular evolution.

However, we do not believe that all the eAck sequences in our dataset (and

many more that might be evident with future sequencing projects) have a neat

history of duplication and divergence. Some eukaryotes might have undergone

reticulate evolution and have likely come to possess Ack through horizontal gene

transfer (HGT). Consider, for example, that from the land-plant group (Streptophyta)

Ack is found in charophytic alga Chara braunii, liverwort Marchantia polymorpha,

moss Physcomitrium patens, dicot plant Carpinus fangiana. Although we believe

that our strategy for detecting eAck was limited by the size and the distribution of

various taxa in the genome database (e.g. Rhizaria: 0.1% representation vs. Metazoa:

41.0% representation), the situation cannot be argued to be the same for land plants

(Streptophyta: 9.0%; Chlorophyta: 1.2%; Amoebozoa: 0.4%, vs. Fungi: 41.2%

representation). The near-absence of Ack in Streptophyta is best explained by loss of

Ack in their last common ancestor, followed by reacquisition through HGT in a

handful of lineages. Similar reasoning seems more applicable for metazoan Ack

detected in our study.

95
TABLES

Table 3.1. Kinetics of CnAck for common divalent metal cofactors in the acetyl
phosphate forming direction

kcat KM kcat/KM
Cofactor
(s-1) (mM) (s-1 mM-1)
Mg2+ 1.42 ± 0.02 1.5 ± 0.1 0.95 ± 0.06
Zn2+ 2.39 ± 0.07 2.8 ± 0.2 0.85 ± 0.07
Mn2+ 2.36 ± 0.02 4.1 ± 0.1 0.58 ± 0.01
Co2+ 2.57 ± 0.07 4.5 ± 0.3 0.57 ± 0.04

Table 3.2. Specific activities of CnAck for common nucleotide triphosphates


(NTPs) in the acetyl phosphate forming direction

Specific activity Percentage activity


NTP
(µmol min-1 mg-1) (%)

ATP 0.93 ± 0.021 100.0

ITP 0.28 ± 0.001 30.2

CTP 0.26 ± 0.024 28.4

GTP 0.24 ± 0.010 25.9

UTP 0.22 ± 0.003 24.2

TTP 0.03 ± 0.004 3.1

96
Table 3.3. Kinetic parameters of CnAck

kcat KM kcat/KM
Substrate
(s-1) (mM) (s-1 mM-1)

Direction of acetyl phosphate formation


Acetate 1.54 ± 0.03 28.8 ± 1.1 0.05 ± 0.01
ATP 1.42 ± 0.02 1.5 ± 0.1 0.95 ± 0.06
Direction of acetate formation
Acetyl phosphate 30.22 ± 1.04 0.5 ± 0.1 60.44 ± 12.27
ADP 61.53 ± 4.08 2.3 ± 0.4 26.75 ± 4.98

97
FIGURES

Figure 3.1. Maximum Likelihood tree of 526 eukaryotic Ack sequences, rooted using

Ack of bacteria Salmonella typhimurium (red branch). Different taxonomic groups

are color coded (legend). For visual clarity of the tree topology, bootstrap values

(2000 replicates) and branch lengths are not shown. Legend also includes summary

(chart on the right) of the counts (column 3) and corresponding distribution of Ack in

various taxonomic supergroups (bold, left column) and subgroups (middle column).

98
Figure 3.2. (A) Ack of fungal origin (blue, orange, and yellow) and one dicot plant

(green) are highlighted as they form a monophyletic superclade (see text);

remaining paraphyletic non-fungal Ack are not color-coded. (B) the superclade in (A)

is collapsed (gray triangle) and non-fungal Ack are highlighted. In both (A) and (B)

branch lengths are shown. Outgroup branch is colored red.

99
Figure 3.3. Proposed XFP2/ACK pathway for acetate production in C. neoformans.

XFP2 catalyzes the conversion of xylulose-5-phosphate (X5P) or fructose-6-

phosphate (F6P) into glyceraldehyde-3-phosphate (G3P) or erythrose-4-phosphate

(E4P) and acetyl phosphate (AcP) using inorganic phosphate (Pi). Acetate kinase

(ACK) can then convert AcP to acetate by transferring the phosphate group to

adenosine diphosphate (ADP). Phosphoenol pyruvate (PEP), oxaloacetic acid (OAA),

and ATP (adenosine triphosphate), generated via metabolic processes such as

glycolysis, and the tricarboxylic acid (TCA) cycle negatively regulate activity of XFP2

(Xylulose-5-phosphate/ fructose-6-phosphate phosphoketolase 2), whereas

adenosine monophosphate (AMP) positively upregulates XPF2 (Glenn et al, 2014).

Abbreviation: PPP - pentose phosphate pathway.

100
REFERENCES

Barnhart, E.P., McClure, M.A., Johnson, K., Cleveland, S., Hunt, K.A., Fields, M.W., 2015.
Potential role of acetyl-CoA synthetase (acs) and malate dehydrogenase (mae)
in the evolution of the acetate switch in Bacteria and Archaea. Sci. Rep. 5,
12498. https://doi.org/10.1038/srep12498

Berguson, H.P., Caulfield, L.W., Price, M.S., 2022. Influence of pathogen carbon
metabolism on interactions with host immunity. Front. Cell. Infect. Microbiol.
12.

Buss, K.A., Cooper, D.R., Ingram-Smith, C., Ferry, J.G., Sanders, D.A., Hasson, M.S., 2001.
Urkinase: Structure of acetate kinase, a member of the ASKHA superfamily of
phosphotransferases. J. Bacteriol. 183, 680–686.
https://doi.org/10.1128/JB.183.2.680-686.2001

Chittori, S., Savithri, H.S., Murthy, M.R., 2012. Structural and mechanistic investigations
on Salmonella typhimurium acetate kinase (AckA): identification of a putative
ligand binding pocket at the dimeric interface. BMC Struct. Biol. 12, 24.
https://doi.org/10.1186/1472-6807-12-24

Claudel‐Renard, C., Chevalet, C., Faraut, T., Kahn, D., 2003. Enzyme‐specific profiles for
genome annotation: PRIAM. Nucleic Acids Res. 31, 6633–6639.
https://doi.org/10.1093/nar/gkg847

Dang, T., Angel, M., Cho, J., Nguyen, D., Ingram-Smith, C., 2022. The Role of acetate
kinase in the human parasite Entamoeba histolytica. Parasitologia 2, 147–159.
https://doi.org/10.3390/parasitologia2020014

Dang, T., Ingram-Smith, C., 2017. Investigation of pyrophosphate versus ATP


substrate selection in the Entamoeba histolytica acetate kinase. Sci. Rep. 7,
5912. https://doi.org/10.1038/s41598-017-06156-5

Eddy, S.R., 2011. Accelerated profile HMM searches. PLOS Comput. Biol. 7, e1002195.
https://doi.org/10.1371/journal.pcbi.1002195

Elias, M., Tawfik, D.S., 2012. Divergence and convergence in enzyme evolution: Parallel
Evolution of Paraoxonases from Quorum-quenching Lactonases*. J. Biol.
Chem. 287, 11–20. https://doi.org/10.1074/jbc.R111.257329

Ferry, J.G., 2011. Chapter eleven - Acetate kinase and phosphotransacetylase, in:
Rosenzweig, A.C., Ragsdale, S.W. (Eds.), Methods in Enzymology, Methods in
Methane Metabolism, Part A. Academic Press, pp. 219–231.
https://doi.org/10.1016/B978-0-12-385112-3.00011-1

Fowler, M.L., Ingram-Smith, C., Smith, K.S., 2012. Novel pyrophosphate-forming


acetate kinase from the protist Entamoeba histolytica. Eukaryot. Cell 11, 1249–
1256. https://doi.org/10.1128/EC.00169-12

101
Fowler, M.L., Ingram-Smith, C.J., Smith, K.S., 2011. Direct detection of the acetate-
forming activity of the enzyme acetate kinase. J. Vis. Exp. JoVE 3474.
https://doi.org/10.3791/3474

Glenn, K., Ingram-Smith, C., Smith, K.S., 2014. Biochemical and kinetic
characterization of xylulose 5-phosphate/fructose 6-phosphate
phosphoketolase 2 (Xfp2) from Cryptococcus neoformans. Eukaryot. Cell 13,
657–663. https://doi.org/10.1128/EC.00055-14

Gorrell, A., Ferry, J.G., 2007. Investigation of the Methanosarcina thermophila acetate
kinase mechanism by fluorescence quenching. Biochemistry 46, 14170–14176.
https://doi.org/10.1021/bi701292a

Gorrell, A., Lawrence, S.H., Ferry, J.G., 2005. Structural and kinetic analyses of arginine
residues in the active site of the acetate kinase from Methanosarcina
thermophila*. J. Biol. Chem. 280, 10731–10742.
https://doi.org/10.1074/jbc.M412118200

Henard, C.A., Freed, E.F., Guarnieri, M.T., 2015. Phosphoketolase pathway engineering
for carbon-efficient biocatalysis. Curr. Opin. Biotechnol., Pathway engineering
36, 183–188. https://doi.org/10.1016/j.copbio.2015.08.018

Hoang, D.T., Chernomor, O., von Haeseler, A., Minh, B.Q., Vinh, L.S., 2018. UFBoot2:
Improving the Ultrafast bootstrap approximation. Mol. Biol. Evol. 35, 518–522.
https://doi.org/10.1093/molbev/msx281

Ingram-Smith, C., Martin, S.R., Smith, K.S., 2006. Acetate kinase: not just a bacterial
enzyme. Trends Microbiol. 14, 249–253. https://doi.org/10.1016/j.tim.2006.04.001

Ingram-Smith, C., Wharton, J., Reinholz, C., Doucet, T., Hesler, R., Smith, K., 2015. The
Role of active site residues in ATP binding and catalysis in the
Methanosarcina thermophila acetate kinase. Life Basel Switz. 5, 861–871.
https://doi.org/10.3390/life5010861

Miles, R.D., Gorrell, A., Ferry, J.G., 2002. Evidence for a transition state analog, MgADP-
aluminum fluoride-acetate, in acetate kinase from Methanosarcina
thermophila *. J. Biol. Chem. 277, 22547–22552.
https://doi.org/10.1074/jbc.M105921200

Minh, B.Q., Schmidt, H.A., Chernomor, O., Schrempf, D., Woodhams, M.D., von
Haeseler, A., Lanfear, R., 2020. IQ-TREE 2: New models and efficient methods
for phylogenetic inference in the genomic era. Mol. Biol. Evol. 37, 1530–1534.
https://doi.org/10.1093/molbev/msaa015

Okonechnikov, K., Golosova, O., Fursov, M., the UGENE team, 2012. Unipro UGENE: a
unified bioinformatics toolkit. Bioinformatics 28, 1166–1167.
https://doi.org/10.1093/bioinformatics/bts091

102
Pineda, E., Vázquez, C., Encalada, R., Nozaki, T., Sato, E., Hanadate, Y., Néquiz, M.,
Olivos-García, A., Moreno-Sánchez, R., Saavedra, E., 2016. Roles of acetyl-CoA
synthetase (ADP-forming) and acetate kinase (PPi-forming) in ATP and PPi
supply in Entamoeba histolytica. Biochim. Biophys. Acta BBA - Gen. Subj.
1860, 1163–1172. https://doi.org/10.1016/j.bbagen.2016.02.010

Rother, M., Metcalf, W.W., 2004. Anaerobic growth of Methanosarcina acetivorans


C2A on carbon monoxide: An unusual way of life for a methanogenic
archaeon. Proc. Natl. Acad. Sci. 101, 16929–16934.
https://doi.org/10.1073/pnas.0407486101

Thaker, T.M., Tanabe, M., Fowler, M.L., Preininger, A.M., Ingram-Smith, C., Smith, K.S.,
Iverson, T.M., 2013. Crystal structures of acetate kinases from the eukaryotic
pathogens Entamoeba histolytica and Cryptococcus neoformans. J. Struct.
Biol. 181, 185–189. https://doi.org/10.1016/j.jsb.2012.11.001

The UniProt Consortium, 2021. UniProt: the universal protein knowledgebase in 2021.
Nucleic Acids Res. 49, D480–D489. https://doi.org/10.1093/nar/gkaa1100

Whicher, A., Camprubi, E., Pinna, S., Herschy, B., Lane, N., 2018. Acetyl phosphate as a
primordial energy currency at the origin of life. Orig. Life Evol. Biospheres 48,
159–179. https://doi.org/10.1007/s11084-018-9555-8

Xu, W.J., Wen, H., Kim, H.S., Ko, Y.-J., Dong, S.-M., Park, I.-S., Yook, J.I., Park, S., 2018.
Observation of acetyl phosphate formation in mammalian mitochondria
using real-time in-organelle NMR metabolomics. Proc. Natl. Acad. Sci. 115,
4152–4157. https://doi.org/10.1073/pnas.1720908115

Yang, W., Catalanotti, C., D’Adamo, S., Wittkopp, T.M., Ingram-Smith, C.J., Mackinder,
L., Miller, T.E., Heuberger, A.L., Peers, G., Smith, K.S., Jonikas, M.C., Grossman,
A.R., Posewitz, M.C., 2014. Alternative acetate production pathways in
Chlamydomonas reinhardtii during dark anoxia and the dominant role of
chloroplasts in fermentative acetate production. Plant Cell 26, 4499–4518.
https://doi.org/10.1105/tpc.114.129965

Zhang, B., Lingga, C., Bowman, C., Hackmann, T.J., 2021. A new pathway for forming
acetate and synthesizing ATP during fermentation in bacteria. Appl. Environ.
Microbiol. 87, e02959-20. https://doi.org/10.1128/AEM.02959-20

103
CHAPTER 4

Modularity hypothesis of fungal virulence

Oly Ahmed, and Kerry Smith

Department of Genetics and Biochemistry; Eukaryotic Pathogen Innovation Center,


Clemson University, SC, USA

104
4.1. Why we need a regulatory network-based understanding
of the origin of virulence in fungi

To date fungal infection persists as a substantial problem in terms of human

health and healthcare cost (Warnock, 2017). An estimated one billion people

worldwide are affected by fungi, especially through superficial infections, resulting in

1.6 million infections annually (Siscar-Lewin et al., 2022). It is generally assumed that

throughout human history fungal infection was not a substantial threat; however, in

recent times we have noticed a significant rise in fungal infections (Dangarembizi et

al., 2022). This rise can at least be attributed to two anthropogenic sources: first,

climate change, which appears to be driving the emergence of new traits such as

thermotolerance in relatively harmless fungi, and thus evolution of pathogenic traits

in these species (Nnadi and Carter, 2021). Second, improved medical interventions,

for example immunosuppressant therapy, which renders the recipient of the

therapy ill-suited for fighting against opportunistic fungal infections (Dangarembizi

et al., 2022).

And yet, the unfortunate reality is that we are limited in our arsenal to fight

against fungal pathogens. The antifungal agents at our disposal suffer from

limitations with regards to activity, pharmacological properties, potential toxicity,

and cost of administration and management (Lewis, 2017). Additionally, secondary

resistance to antifungal drugs has also been reported (Johnson, 2017). Under such

circumstances, researchers are working incessantly to discover new antifungal

therapy via drug screening and drug repurposing (Donlin and Meyers, 2022).

Our understanding of the fungal virulence factors and mechanism (Brunke et

al., 2016) and the role of immunity in defense and disease progression (Casadevall,

105
2022) has advanced significantly. A noteworthy achievement of the community is

the proposal of a hypothesis that enables us to glean the evolutionary story of the

origin of virulence in fungi. The thesis posits that interactions in nature among

environmental fungi and their amoebic predators worked as a backdrop whereby

virulence factors imparted selective advantages to the fungi, which might have been

co-opted (as an exaptation) for pathogenicity in mammalian hosts (Casadevall et al.,

2019; Radosa and Hillmann, 2021). Although this hypothesis provides a theoretical

framework that is of tremendous use for understanding the evolutionary origin of

virulence traits in fungi, it is limited in expounding a framework useful for

understanding the emergence of such virulence traits from a cellular and molecular

biology perspective. Understanding the cell and molecular biology of fungal

virulence is paramount for developing research strategies that alleviates the

problem of blind screening for novel drugs.

Here we propose a hypothesis to explain the emergence of virulence in fungi

based on an underlying cellular regulatory network architecture that connects stress

perception and response traits modules with virulence traits modules in fungi. This

hypothesis, in our view, has a potential to predict connectedness between such

modules, and thus directs us to such links which, in principle, can be targeted to be

severed to remove or diminish pathogenicity in fungi.

106
4.2. Fungal infections in mammals are rare, and virulence
factors can have niche-specific protective roles for fungi

Fungal pathogenicity in mammals appears to be accidental as they do not

seem to suffer many fungal infections, and most cases are found in

immunocompromised individuals (for a recent review, see Gnat et al., 2021). Table 4.1

compiles a list of common fungal diseases reported in mammalian hosts.

It is worthwhile to ponder about the rarity of fungal pathogens in mammals.

First, out of an estimated 2.2 ~ 3.8 million fungal species, only a handful of the species

(presumably 0.01%; Köhler et al., 2017) cause any serious infection in plants and

animals (a more elaborate list of infectious fungi can be found in Sun et al., 2020).

This implies that not all fungi have pathogenic potential.

Second, not all fungal infections lead to disease progression. Disease, the

penultimate outcome of virulence, requires successful completion of multiple steps

involving different virulence factors and physiological adaptations on the invading

pathogen’s part within the invaded host body (Diard and Hardt, 2017). Thus, success

or failure at each of the pre-disease steps of infection is important in driving the

evolution and/or maintenance of virulence traits/factors in any pathogen (Diard and

Hardt, 2017). Mammalian body temperature and possession of a complex immune

system are thus two major challenges for fungal pathogens inside mammalian hosts

(Casadevall, 2022).

Attempts have been made to understand how a handful of fungal pathogens

might have developed (or perhaps, retained) the ability for pathogenicity in

mammalian hosts (dubbed as “The amoeboid predator - fungal animal virulence

hypothesis”, see the previous section and Casadevall et al., 2019). What all these

107
potential pathogenic fungi have in common — perhaps, in contrast to their non-

pathogenic counterparts — is a set of morphological and biochemical traits (dubbed

as virulence factors) that enables them to establish initial infection, evade the

patrolling immune system, and effectively disseminate systemically. However, what

is considered as a virulence factor is unique to the pathogenic species (an excellent

review of the examples of species-specific virulence factors can be found in Brunke

et al., 2016).

Our current understanding is that these virulence factors have tangible

utilities in terms of providing protection to the fungi from naturally occurring

predators and various environmental stresses. For example, morphological virulence

factors, such as dimorphic transition (Candida spp.), capsule formation

(Cryptococcus spp.), or growth at high temperatures (seen in all pathogenic fungi),

as well as biochemical factors, such as secretion of hydrolytic enzymes

(dermatophytes, Malassezia, Cryptococcus spp.), synthesis of reactive oxygen

species (ROS) and radiation protective compound melanin (Aspergillus,

Cryptococcus), etc. are likely to help fungi counter environmental changes such as

nutrient exhaustion, amoebic predation, rise in temperature, micro-nutrient

deprivation, change in pH and UV intensity, etc (Brunke et al., 2016). Thus, it is highly

likely that such ecologically relevant, niche-specific fungal phenotypes (known to us

as virulence factors) are also useful toolkits that can be co-opted for accidental

pathogenicity in an unfortunate mammalian host (Rokas, 2022).

108
4.3. Stress response and virulence can be conceived as
modules of a larger conceptual network

From gene knock-out experiments to -omics studies, investigations of

infectious agents (bacterial, fungal, protozoan) have pointed to numerous examples

illustrating the relatedness of virulence and stress response. Simply searching the

PubMed database on November 13, 2022 for “stress response and virulence” resulted

in 7,804 publications. The underlying assumption is that any infectious agent,

obligate or opportunistic, is constantly being exposed to various physical and

chemical stresses within the host; hence, a rapid adaptation to this host-specific

stresses are imperative for growth and survival of the infecting organism (Brown and

Goldman, 2016). Thus, it appears that a microbe’s ability of quick response and

adjustment to host-imposed stresses are necessary, if not sufficient, for

pathogenicity. On the flip side, non-pathogenic environmental species appear to

lack virulence traits, but they cannot possibly do without any niche-specific stress-

response mechanisms. These considerations point us to thinking of stress response

and virulence as separate modules (Figure 4.1).

Modularity can be roughly defined as a community of nodes in a given

network that shares a significantly high number of within-connections compared to

their connections with other such modules elsewhere in the network (Alcala-Corona

et al, 2021; Serban 2020) (Figure 4.2). Imagining stress response and virulence as

separate modules urges us to look for connections (or links, or edges) between

them. This is not inconceivable. For example, a single gene product such as YjbH

appears to tie oxidative stress response and virulence in opportunistic bacterial

pathogen Staphylococcus aureus (Paudel et al, 2021), whereas cAMP/PKA pathway is

109
speculated to be involved in nutrient sensing and adaptation, and virulence in

opportunistic fungal pathogen Cryptococcus neoformans (Caza and Kronstad, 2019).

In recent years, advancement in transcriptomic and proteomic technologies

and computational techniques has opened up an opportunity to take a systems-

level examination at our pathogens of interest. Quite recently an attempt was made

to reconstruct the Transcriptional Regulatory Network (TRN) of plant pathogenic

fungus Fusarium graminearum which revealed a modular architecture of important

traits such as virulence, sexual reproduction, and mycotoxin production (Guo et al,

2020). Similarly, Gene Co-expression Network (GCN) analysis has also been exploited

to discover gene regulatory modules that hint at previously unknown virulence

mechanisms in plant pathogenic bacteria Pseudomonas syringae (Nobori et al,

2020).

Although the examples above suggest a possibility of existence of virulence

and stress response as modules at gene regulatory levels, this need not be the case,

as modularity can be sought for across various scales in a given biological system

(Serban, 2020). Keeping aside any debate about the ontological status of such

modules, thinking in terms of a network and its (subnetwork) modules to

understand the regulatory connectivity of virulence traits and stress sensing and

response traits, we believe, does have an immense pragmatic importance.

110
4.4. The hypothesis and its implications

To summarize, stress sensing and response are connected to virulence in

many pathogenic organisms, whereas non-pathogenic organisms have functioning

mechanisms for stress sensing/response without any virulence traits, implying a

plausible existence of two separate modules. Moreover, previous TRN and GCN

analyses have been used to explore the modular nature of the underlying regulatory

network in connection to virulence in plant pathogens.

From such considerations, we hypothesized that various stress responses (and

associated sensing and signaling) and virulence traits in fungal pathogens of

mammals can be thought of as interconnected modules, where the connections

between them are likely to be regulatory in nature.

This picture of fungal virulence can also be used to explain the emergence of

pathogenicity, since modularity leads to distinctive functions with a possibility of

subsequent co-option, thereby facilitating evolutionary innovations (Sambamoorthy

et al, 2019). For example, in mycobacteria, the evolution of host-specific adaptations,

as well as the ability to emerge in new niches might have come about through

acquisition of new genetic materials, which needed functional repurposing following

integration. Such successful functional repurposing suggests an underlying modular

structure for evolution to work on (see Pepperell, 2022 for an updated review of the

evolution of tuberculosis pathogenesis).

Another major implication of the proposed hypothesis is that it predicts the

existence of interconnections between stress response and virulence modules. In

terms of gene regulatory networks, such interconnectedness may be exemplified by

the existence of one or few transcription factors or other regulatory elements. A

111
conception of fungal virulence in terms of network topology — a plausible

evolutionary design principle — is sure to lead to an understanding of connections

between stress response and virulence factors in fungal pathogens. Such discoveries

have the potential to usher in rational strategies for developing molecular and

pharmacological interventions for fungal infections in mammalian hosts.

112
TABLE

Table 4.1. Common fungal infections in mammals (Compiled and adapted from
Sayedmousavi et al, 2018)

Opportunistic infection (non-transmissible) Aspergillosis


Mucormycosis
Candidiasis
Cryptococcosis
Various black-fungi related
infections

Environmental sources (indirect Coccidioidomycosis


transmission) Histoplasmosis
Paracoccidioidomycosis
Blastomycosis

Other infections (direct transmission and White-nose syndrome


zoonosis) Microsporum infection
Sporothrix infection
Trichophyton infections
Encephalitozoon infections
Enterocytozoon infections

113
FIGURES

Figure 4.1. Stress-response and virulence-associated traits in fungi can be


thought of as separate modules of the same regulatory network. Fungi both with
and without pathogenic potential have all the necessary signaling and response
mechanisms for recurrent stresses experienced in their natural habitat, such as
nutrient starvation, pH, osmotic and temperature fluctuations, radiation, predators
etc. Moreover, pathogenic species possess virulence-associated phenotypes and
genetic factors which are absent in their non-pathogenic counterparts. Virulence
traits/factors are thought to play roles in stress-response in these pathogenic species
outside their host body, and are presumably co-opted for causing accidental
damages to the hosts. However, absence of virulence traits/factors in non-
pathogenic species does not diminish their ability to cope with the environmental
stresses. In pathogenic species, virulence traits/factors appear to be linked to the
stress response within the host body, and hence are presumably connected via
regulatory mechanism. Such consideration points to the possibility of stress-
response and virulence as two separate, but regulatorily interconnected modules in
pathogenic fungi.

114
Figure 4.2. Modularity in biological networks. Modularity is a widespread
phenomenon in any complex system such as ecological interactions, gene co-
expression and transcriptional regulatory networks in biology. Within a given
network (here boxes represent nodes of a hypothetical network, e.g. genes, and solid
connecting lines represent relationships among the nodes, e.g. co-expression or
transcriptional regulation), modules are clusters of nodes that are heavily
interconnected compared to other nodes elsewhere in the same network (here
nodes within a module share the same color). Two most common types of
modularity are depicted here: (A) simple modules, and (B) hierarchical modules. In
contrast to the simple modules, a module can be nested within a larger module
(depicted by the large red rectangle in B) giving rise to a hierarchical topology.
Virulence traits/ factors can be thought of as a module nested under stress-response
module in most pathogenic fungi.

115
REFERENCES

Alcalá-Corona, S.A., Sandoval-Motta, S., Espinal-Enríquez, J., Hernández-Lemus, E.,


2021. Modularity in biological networks. Front. Genet. 12.

Brown, N.A., Goldman, G.H., 2016. The contribution of Aspergillus fumigatus stress
responses to virulence and antifungal resistance. J. Microbiol. 54, 243–253.
https://doi.org/10.1007/s12275-016-5510-4

Brunke, S., Mogavero, S., Kasper, L., Hube, B., 2016. Virulence factors in fungal
pathogens of man. Curr. Opin. Microbiol., Host-microbe interactions:
parasites/fungi/viruses 32, 89–95. https://doi.org/10.1016/j.mib.2016.05.010

Casadevall, A., 2022. Immunity to invasive fungal diseases. Annu. Rev. Immunol. 40,
121–141. https://doi.org/10.1146/annurev-immunol-101220-034306

Casadevall, A., Fu, M.S., Guimaraes, A.J., Albuquerque, P., 2019. The “Amoeboid
Predator-Fungal Animal Virulence” hypothesis. J. Fungi Basel Switz. 5, E10.
https://doi.org/10.3390/jof5010010

Caza, M., Kronstad, J.W., 2019. The cAMP/Protein Kinase A pathway regulates
virulence and adaptation to host conditions in Cryptococcus neoformans.
Front. Cell. Infect. Microbiol. 9.

Dangarembizi, R., Wasserman, S., Hoving, J.C., 2022. Emerging and re-emerging
fungal threats in Africa. Parasite Immunol. e12953.
https://doi.org/10.1111/pim.12953

Diard, M., Hardt, W.-D., 2017. Evolution of bacterial virulence. FEMS Microbiol. Rev. 41,
679–697. https://doi.org/10.1093/femsre/fux023

Donlin, M.J., Meyers, M.J., 2022. Repurposing and optimization of drugs for discovery
of novel antifungals. Drug Discov. Today 27, 2008–2014.
https://doi.org/10.1016/j.drudis.2022.04.021

Gnat, S., Łagowski, D., Nowakiewicz, A., Dyląg, M., 2021. A global view on fungal
infections in humans and animals: opportunistic infections and
microsporidioses. J. Appl. Microbiol. 131, 2095–2113.
https://doi.org/10.1111/jam.15032

Guo, L., Ji, M., Ye, K., 2020. Dynamic network inference and association computation
discover gene modules regulating virulence, mycotoxin and sexual
reproduction in Fusarium graminearum. BMC Genomics 21, 179.
https://doi.org/10.1186/s12864-020-6596-y

Johnson, E.M., 2017. Antifungal susceptibility testing and resistance.


https://doi.org/10.1093/med/9780198755388.003.0047

116
Köhler, J.R., Hube, B., Puccia, R., Casadevall, A., Perfect, J.R., 2017. Fungi that infect
humans. Microbiol. Spectr. 5, 5.3.08.
https://doi.org/10.1128/microbiolspec.FUNK-0014-2016

Lewis, R.E., 2017. Principles of antifungal therapy.


https://doi.org/10.1093/med/9780198755388.003.0045

Nnadi, N.E., Carter, D.A., 2021. Climate change and the emergence of fungal
pathogens. PLOS Pathog. 17, e1009503.
https://doi.org/10.1371/journal.ppat.1009503

Nobori, T., Wang, Y., Wu, J., Stolze, S.C., Tsuda, Y., Finkemeier, I., Nakagami, H., Tsuda,
K., 2020. Multidimensional gene regulatory landscape of a bacterial pathogen
in plants. Nat. Plants 6, 883–896. https://doi.org/10.1038/s41477-020-0690-7

Paudel, A., Panthee, S., Hamamoto, H., Grunert, T., Sekimizu, K., 2021. YjbH regulates
virulence genes expression and oxidative stress resistance in Staphylococcus
aureus. Virulence 12, 470–480. https://doi.org/10.1080/21505594.2021.1875683

Pepperell, 2022. Evolution of tuberculosis pathogenesis [WWW Document].


https://doi.org/10.1146/annurev-micro-121321-093031

Radosa, S., Hillmann, F., 2021. Host-pathogen interactions: lessons from phagocytic
predation on fungi. Curr. Opin. Microbiol. 62, 38–44.
https://doi.org/10.1016/j.mib.2021.04.010

Rokas, A., 2022. Evolution of the human pathogenic lifestyle in fungi. Nat. Microbiol. 7,
607–619. https://doi.org/10.1038/s41564-022-01112-0

Sambamoorthy, G., Sinha, H., Raman, K., 2019. Evolutionary design principles in
metabolism. Proc. R. Soc. B Biol. Sci. 286, 20190098.
https://doi.org/10.1098/rspb.2019.0098

Serban, M., 2020. Exploring modularity in biological networks. Philos. Trans. R. Soc. B
Biol. Sci. 375, 20190316. https://doi.org/10.1098/rstb.2019.0316

Seyedmousavi, S., Bosco, S. de M.G., de Hoog, S., Ebel, F., Elad, D., Gomes, R.R.,
Jacobsen, I.D., Jensen, H.E., Martel, A., Mignon, B., Pasmans, F., Piecková, E.,
Rodrigues, A.M., Singh, K., Vicente, V.A., Wibbelt, G., Wiederhold, N.P., Guillot, J.,
2018. Fungal infections in animals: a patchwork of different situations. Med.
Mycol. 56, S165–S187. https://doi.org/10.1093/mmy/myx104

Siscar-Lewin, S., Hube, B., Brunke, S., 2022. Emergence and evolution of virulence in
human pathogenic fungi. Trends Microbiol. 30, 693–704.
https://doi.org/10.1016/j.tim.2021.12.013

117
Sun, S., Hoy, M.J., Heitman, J., 2020. Fungal pathogens. Curr. Biol. 30, R1163–R1169.
https://doi.org/10.1016/j.cub.2020.07.032

Warnock, D.W., 2017. Introduction to medical mycology.


https://doi.org/10.1093/med/9780198755388.003.0001

118
CHAPTER 5

Screening modularity in transcriptional hierarchy

requires undertaking larger exploratory studies

Oly Ahmed

Department of Genetics and Biochemistry; Eukaryotic Pathogen Innovation Center,

Clemson University, SC, USA

119
In this work I set out to explore the mechanism(s) and effect(s) of acetate

metabolism in the human pathogenic fungus Cryptococcus neoformans owing to

the suggested relevance of acetate utilization in the survival and virulence

elaboration of this organism during the course of infection (Chapter 1). I have also

investigated biochemical properties of acetate kinase (Ack) from C. neoformans as

Ack forms a potential pathway for acetate production in this species (Chapter 3). In

this chapter I aim to suggest potential studies designed to answer important

questions raised in my work.

5.1. Transcriptional circuitry of acetate utilization

I initiated our exploration through the investigation of direct effects of acetate

utilization, in lieu of glucose as the sole carbon source, on the transcriptomic

landscape, and noticed that a subset of genes is regulated as to bear the signature

of nutrient stress, glucose deprivation, and global adaptation for efficient nutrient

acquisition (Chapter 2). Additionally, I have noted significant upregulation of a

number of genes previously described in the literature for their association with

virulence and fitness in C. neoformans (Chapter 2).

In this study (Chapter 2) I have computationally predicted a total of 200

putative transcription factors (TFs), of which a subset of 15 TFs were found to be

upregulated during acetate utilization, suggesting a possible role of these TFs in the

regulation of acetate metabolism. To date 11 out of 15 of these TFs have been

characterized previously, and the remaining 4 TFs are predicted for the first time in

the current study (Chapter 2). This opens up the possibility for studying the

120
individual and interactive roles of these putative TFs in acetate utilization. I suggest

that the first step towards that aim may be achieved through in silico detection of

the binding sites of these putative TFs. Under the assumption that these DNA-

binding sites (DBS) are likely to act as the initial cis-regulatory elements (CRE) for

proximal genes (for simplicity, the trans-acting regulatory mechanism should be

ignored to begin with), comparison among the proximal genes of the predicted DBS

with the differentially regulated gene-set from the transcriptomic study in Chapter 2

can be used, in principle, to predict the primary regulatory circuitry of acetate

utilization in C. neoformans.

Furthermore, the involvement of these 15 upregulated TFs need not be

restricted to acetate utilization. Expression of this subset of TFs can be monitored

using quantitative RT-PCR techniques under various growth conditions with carbon

sources such as lactate and trehalose, two other important physiologically relevant

alternative carbon sources (Chapter 1). These experiments will shed light on TFs

uniquely connected to acetate utilization by comparison.

5.2. Possibility of an acetate switch

In chapter 1 I discussed the possibility of the acetate switch (i.e., the ability of

either secreting out or assimilating in acetate from the surroundings depending on

the availability of the nutrients) in C. neoformans depending on the tissue of

infection and the disease state of the host as these parameters influence the

availability of acetate in various host microenvironments. An extension of this

hypothesis posits that C. neoformans thereby may possess the ability of acetate

switch. To investigate the possibility of an acetate-switch phenotype in cryptococcal

121
species, acetate concentrations in the spent media can be measured at equally

spaced time intervals (colorimetric assays, LC-MS/MS etc.), followed by time-series

and autocorrelation analyses. This may help us probe the presence of such

phenotype in unicellular fungal species, presumably, for the first time in this group

of eukaryotes.

5.3. Xfp-Ack pathway as a route of acetate production

The kinetic characterization of acetate kinase (Ack) reported here and

previous reports of kinetic properties of xylulose-5-phosphate/fructose-6-phosphate

phosphoketolase 2 (Xfp2) from C. neoformans consistently point to a pathway of

acetate synthesis from phosphoketose sugars (Chapter 3). Thus, a coupled reaction

involving both Xfp2 and Ack can be devised to characterize the properties of the

entire pathway, as a single metabolic entity, as the proof of principle.

Moreover, C. neoformans genome encodes for a paralog of Xfp, Xfp1, which is

yet to be characterized biochemically. The Xfp1 gene from C. neoformans thus can

be cloned, recombinantly produced, purified and kinetically characterized in

addition to examining its post-translational regulation. This will broaden our

understanding of the role of the Xfp paralogs in the Xfp-Ack pathway in C.

neoformans.

5.4. The trio: Xfp-Pta-Ack

In Chapter 3, I predicted Ack sequences from the published and curated

genomes of over two thousand species and constructed a phylogeny to investigate

the evolutionary history of this enzyme in Eukarya. The pathway of acetate

122
production (or assimilation) in prokaryotes and some eukaryotes involves Xfp and/or

phosphotransacetylase (Pta) (Chapter 3). Thus, a similar in silico study for predicting

and reconstructing the evolutionary history of Xfp and Pta, the two prominent

partners of Ack, can be compared across the domain Eukarya to further our

understanding of the evolution of the entire pathway in eukaryotes.

Moreover, special attention can be focused on the kingdom Fungi for their

biotechnological and medical relevance and potential. Strong functional ties

between Ack and its partner (say Xfp) are likely to lead to chromosomal proximity of

the genes and hence linkage disequilibrium (LD). Although it will be logistically

challenging to perform an LD experiment of this magnitude (i.e. across the kingdom

Fungi), available published genomes are easy to scan for in silico distance

measurement between Ack and partner genes.

In case of LD between Ack and partner genes, this new information will be

quite useful down the line. For example, Ack and the partner gene can be used to

check for synteny in closely related fungal species. Syntenic relationship combined

with LD may be suggestive of the selective sweep as a mechanism of evolution for

the entire pathway in eukaryotes. This implies that these genes have the

contributory potential along with other well-known near-universal single-copy

orthologs (e.g., OrthoDB, Zdobnov et al., 2021) for multi-loci phylogenomic studies in

Fungi for better phylogenomic resolution.

123
5.5. Gene co-expression network in C. neoformans

In chapter 4 I have proposed a new theoretical framework to explain the

possible mechanism of molecular evolution of various virulence traits in pathogenic

fungi. In chapter 2 we first saw a glimpse of the modular nature of virulence

phenotype and stress response phenotype in C. neoformans by noting the

association between the transcriptional regulation of the known stress response

genes with the known virulence genes in C. neoformans when grown with acetate

as the sole carbon source.

An extension of my work depicted in chapter 2, and a proof of concept of my

hypothesis introduced in chapter 4, can be systematically tested via gene co-

expression network (GCN) studies. Note that modularity at the transcriptional level is

not an absolute requirement for the emergence of fungal virulence, however,

logistically GCN is a good starting point. Hence, cryptococcal cells can be exposed to

various stress conditions, including but not limited to, heat stress, oxidative and/or

nitrosative stress, carbon and/or nitrogen starvation, radiation stress, and the

resultant transcriptomic profiles under each stress condition can be combined to

construct a GCN. Such analyses easily reveal the underlying modular structure of the

network; however, the challenge would be to tease out the recurrent patterns of

connectivity under various stress conditions between the genes of the

corresponding stress response and those of virulence and pathogenicity. An offshoot

result of GCN can be the classic guilt-by-association analysis where constant

expression of one or few virulence-associated genes under all or most stress

conditions tested may help us narrow down the most important virulence genes as

well.

124
REFERENCE

Zdobnov, E.M., Kuznetsov, D., Tegenfeldt, F., Manni, M., Berkeley, M., Kriventseva, E.V.,
2021. OrthoDB in 2020: evolutionary and functional annotations of orthologs.
Nucleic Acids Res. 49, D389–D393. https://doi.org/10.1093/nar/gkaa1009

125
APPENDIX A

RNA-sequencing data of Cryptococcus neoformans

Table A1. Preliminary stats of RNA-sequencing data

Clean read
Sample ID Raw read counts counts Clean bases GC content

Glucose-1 61,638,480 60,023,740 9.0 Gb 52.35 %

Glucose-2 55,954,158 53,621,164 8.0 Gb 52.49 %

Glucose-3 58,790,866 56,394,978 8.5 Gb 52.74 %

Acetate-1 53,419,220 51,696,148 7.8 Gb 51.57 %

Acetate-2 55,160,712 51,022,146 7.7 Gb 51.62 %

Acetate-3 59,197,388 55,839,016 8.4 Gb 51.6 %

Average 57,360,137.33 54,766,198.67 N/A 52.06 %

Median 57,372,512 54,730,090 N/A 51.99 %

Table A2. List of putative transcription factors identified in the study

CNAG_ID Chromosome Gene Name or Product Description


Symbol
CNAG_00014 1 N/A hypothetical protein

CNAG_00017 1 N/A hypothetical protein

CNAG_00018 1 FZC6 hypothetical protein

CNAG_00027 1 N/A transcriptional activator

CNAG_00031 1 MLR1 hypothetical protein

CNAG_00035 1 N/A hypothetical protein

CNAG_00039 1 ZFC6 hypothetical protein

CNAG_00043 1 N/A hypothetical protein, variant

CNAG_00068 1 N/A specific RNA polymerase II transcription factor

126
CNAG_00132 1 N/A hypothetical protein, variant

CNAG_00239 1 YAP1 hypothetical protein

CNAG_00460 1 LIV1 virulence related protein of unknown function

CNAG_00505 1 FZC28 hypothetical protein

CNAG_00514 1 N/A hypothetical protein

CNAG_00559 1 BZP3 bZip transcription factor, putative

CNAG_00653 1 N/A hypothetical protein

CNAG_00670 1 FZC12 hypothetical protein

CNAG_00732 1 CCD3 hypothetical protein

CNAG_00828 1 SIP401 pathway-specific nitrogen regulator

CNAG_00830 1 FZC38 hypothetical protein

CNAG_00841 1 N/A hypothetical protein

CNAG_00871 5 CLR3 hypothetical protein

CNAG_00883 5 ECM2201 Putative Zn2-Cys6 zinc-finger transcription factor

CNAG_01069 5 N/A hypothetical protein

CNAG_01136 5 N/A hypothetical protein

CNAG_01256 5 N/A hypothetical protein

CNAG_01315 5 N/A protein transporter

CNAG_01448 5 N/A hypothetical protein

CNAG_01454 5 STE12 alpha transcription factor STE12

CNAG_01487 11 N/A hypothetical protein

CNAG_01551 11 GAT201 GATA family transcription factor

CNAG_01620 11 N/A hypothetical protein

CNAG_01645 11 N/A hypothetical protein

CNAG_01781 11 N/A hypothetical protein, variant 3

CNAG_01847 11 N/A hypothetical protein

CNAG_01858 11 HOB2 hypothetical protein

CNAG_01948 11 FZC36 nuclear protein

CNAG_01965 11 N/A hypothetical protein

CNAG_01977 11 N/A hypothetical protein

CNAG_01999 11 N/A hypothetical protein

127
CNAG_02022 6 N/A DNA-directed RNA polymerase II subunit RPB3

CNAG_02046 6 PBP1 poly(A)-binding protein binding protein

CNAG_02066 6 FZC13 hypothetical protein

CNAG_02080 6 N/A hypothetical protein

CNAG_02134 6 RSC8 rsc chromatin remodeling complex subunit

CNAG_02215 6 HAP3 transcriptional activator, variant

CNAG_02305 6 FZC45 hypothetical protein

CNAG_02322 6 FZC17 hypothetical protein

CNAG_02359 6 N/A small subunit ribosomal protein S25e

CNAG_02364 6 FZC19 hypothetical protein

CNAG_02408 6 N/A hypothetical protein

CNAG_02413 6 N/A hypothetical protein, variant

CNAG_02435 6 BWC2 white collar 2 protein

CNAG_02466 6 N/A hypothetical protein

CNAG_02476 6 YRM101 putative transcription factor

CNAG_02516 6 HLH5 hypothetical protein

CNAG_02525 6 N/A hypothetical protein

CNAG_02530 6 N/A hypothetical protein

CNAG_02555 6 SIP402 Putative Zn2-Cys6 zinc-finger transcription factor

CNAG_02566 3 FKH2 hepatocyte nuclear factor

CNAG_02603 3 N/A early growth response protein 1

CNAG_02691 3 N/A hypothetical protein

CNAG_02723 3 FZC23 hypothetical protein

CNAG_02739 3 N/A hypothetical protein

CNAG_02767 3 N/A transcription initiation factor TFIIE subunit alpha

CNAG_02774 3 MAL13 Fungal Zn(2)-Cys(6) binuclear cluster domain protein

CNAG_02822 3 N/A hypothetical protein

CNAG_02877 3 FZC51 hypothetical protein

CNAG_02951 3 N/A hypothetical protein

CNAG_02986 3 YSA1 ADP-ribose pyrophosphatase

CNAG_03018 3 ASG101 Putative zinc finger transcription factor

128
CNAG_03086 8 FZC20 hypothetical protein

CNAG_03092 8 N/A hypothetical protein

CNAG_03115 8 FZC46 hypothetical protein

CNAG_03116 8 HCM1 forkhead transcription factor 3

CNAG_03129 8 N/A transcription factor IIIA

CNAG_03132 8 N/A hypothetical protein

CNAG_03155 8 N/A ENTH domain-containing protein

CNAG_03177 8 N/A hypothetical protein, variant

CNAG_03212 8 HCM101 forkhead domain-containing protein

CNAG_03229 8 YOX101 specific transcriptional repressor

CNAG_03279 8 CCD4 hypothetical protein

CNAG_03346 8 BZP4 bZip transcription factor, putative

CNAG_03366 8 ZNF2 C2H2 type zinc finger transcription factor

CNAG_03527 8 HEL2 cytoplasmic protein

CNAG_03607 2 N/A hypothetical protein

CNAG_03608 2 N/A hypothetical protein

CNAG_03636 2 N/A DNA-directed RNA polymerase I and III subunit RPAC2

CNAG_03710 2 ECM22 hypothetical protein

CNAG_03741 2 FZC31 hypothetical protein

CNAG_03745 2 N/A hypothetical protein

CNAG_03768 2 FZC32 hypothetical protein

CNAG_03894 2 PDR802 Putative Zn2-Cys6 zinc-finger transcription factor

CNAG_03902 2 RDS2 Regulator of drug sensitivity 2, putative

CNAG_03914 2 FZC14 hypothetical protein

CNAG_03976 2 N/A hypothetical protein

CNAG_03998 2 RLM1 MADS-box transcription factor

CNAG_04012 2 FZC18 hypothetical protein

CNAG_04023 2 N/A hypothetical protein

CNAG_04090 2 ATF1 activating transcription factor

CNAG_04130 9 N/A hypothetical protein

CNAG_04176 9 HSF2 hypothetical protein

129
CNAG_04178 9 N/A hypothetical protein

CNAG_04263 9 BZP2 hypothetical protein

CNAG_04303 9 N/A E3 ubiquitin-protein ligase bre1

CNAG_04316 9 UTR1 NAD kinase

CNAG_04345 9 ARO8001 specific RNA polymerase II transcription factor

CNAG_04353 9 CLR1 hypothetical protein, variant

CNAG_04372 9 N/A DNA-directed RNA polymerase I, II, and III subunit RPABC2

CNAG_04382 9 N/A hypothetical protein

CNAG_04398 9 ARO80 specific RNA polymerase II transcription factor

CNAG_04457 9 FZC30 hypothetical protein

CNAG_04473 9 N/A hypothetical protein

CNAG_04492 9 N/A hypothetical protein

CNAG_04534 9 N/A hypothetical protein, variant

CNAG_04586 10 HOB7 LIM-homeobox protein

CNAG_04588 10 ERT1 hypothetical protein

CNAG_04594 10 FZC27 hypothetical protein

CNAG_04618 10 N/A hypothetical protein

CNAG_04630 10 YAP2 hypothetical protein

CNAG_04708 10 N/A hypothetical protein

CNAG_04768 10 N/A hypothetical protein

CNAG_04798 10 N/A regulatory protein Cys-3

CNAG_04807 10 FZC8 hypothetical protein

CNAG_04836 10 FZC10 nuclear protein

CNAG_04837 10 MLN1 bHLH family transcription factor

CNAG_04841 10 FZC43 transcriptional regulatory protein, variant

CNAG_04864 10 CIR1 iron regulator 1

CNAG_04878 10 FZC1 hypothetical protein, variant

CNAG_04916 10 FZC16 hypothetical protein

CNAG_04925 10 N/A nuclear protein

CNAG_04933 4 N/A hypothetical protein, variant 2

CNAG_05010 4 ZFC7 hypothetical protein

130
CNAG_05019 4 FZC21 hypothetical protein

CNAG_05049 4 PIP201 hypothetical protein

CNAG_05067 4 CLR5 hypothetical protein

CNAG_05093 4 HOB6 hypothetical protein

CNAG_05112 4 FZC42 hypothetical protein

CNAG_05153 4 GAT5 hypothetical protein

CNAG_05170 4 PIP2 hypothetical protein

CNAG_05221 4 H2A-4 histone H2A.Z

CNAG_05222 4 NRG1 transcriptional regulator Nrg1

CNAG_05255 4 FZC2 hypothetical protein

CNAG_05333 14 N/A hypothetical protein

CNAG_05371 14 N/A hypothetical protein

CNAG_05375 14 N/A hypothetical protein

CNAG_05380 14 FZC44 hypothetical protein

CNAG_05392 14 ZAP104 specific RNA polymerase II transcription factor

CNAG_05420 14 USV101 Nutrient and stress factor 1, putative

CNAG_05431 14 RIM101 pH-response transcription factor pacC/RIM101

CNAG_05442 14 N/A TATA-box-binding protein

CNAG_05479 14 N/A Fungal specific transcription factor, putative

CNAG_05603 14 N/A hypothetical protein

CNAG_05749 7 N/A hypothetical protein

CNAG_05761 7 N/A cyclin-dependent kinase regulatory subunit CKS1

CNAG_05785 7 STB4 putative transcription factor

CNAG_05786 7 N/A hypothetical protein, variant

CNAG_05990 12 N/A hypothetical protein

CNAG_06005 12 N/A U3 small nucleolar RNA-associated protein 4

CNAG_06097 12 N/A hypothetical protein

CNAG_06134 12 BZP1 hypothetical protein

CNAG_06156 12 FZC7 hypothetical protein

CNAG_06163 12 N/A hypothetical protein

CNAG_06203 12 MAT2 HMG-box transcription factor, putative

131
CNAG_06246 12 N/A hypothetical protein

CNAG_06276 13 CEP3 Centromere DNA-binding protein complex CBF3 subunit B,


putative
CNAG_06327 13 MIG1 DNA-binding protein creA

CNAG_06331 13 N/A hypothetical protein

CNAG_06339 13 FZC35 hypothetical protein

CNAG_06425 13 PPR1 fungal specific transcription factor

CNAG_06475 13 N/A hypothetical protein

CNAG_06531 7 N/A hypothetical protein

CNAG_06641 7 N/A DNA-directed RNA polymerase III subunit RPC6

CNAG_06719 2 FZC49 hypothetical protein

CNAG_06734 2 N/A DNA-directed RNA polymerase I, II, and III subunit rpabc5

CNAG_06751 2 HLH3 hypothetical protein

CNAG_06818 5 HAP1 hypothetical protein

CNAG_06871 5 FZC41 hypothetical protein

CNAG_06921 3 HOB4 hypothetical protein

CNAG_07011 12 N/A hypothetical protein

CNAG_07312 1 N/A hypothetical protein

CNAG_07332 1 N/A hypothetical protein

CNAG_07345 1 N/A hypothetical protein

CNAG_07435 5 HAP2 Transcriptional activator HAP2, putative

CNAG_07443 5 HLH4 hypothetical protein

CNAG_07460 2 N/A heat shock transcription factor

CNAG_07494 2 N/A U3 small nucleolar RNA-associated protein MPP10

CNAG_07532 3 PHO23 hypothetical protein

CNAG_07545 3 N/A nuclear protein

CNAG_07590 11 N/A V-type H -transporting ATPase subunit C

CNAG_07597 11 N/A hypothetical protein, variant

CNAG_07680 7 HAP5 transcriptional activator hap5

CNAG_07867 10 N/A hypothetical protein

CNAG_07901 12 FZC29 hypothetical protein

CNAG_07921 13 N/A hypothetical protein

132
CNAG_07922 13 FZC4 transcription factor

CNAG_07924 13 MCM1 Pheromone receptor transcription factor, putative

CNAG_07940 13 BZP5 hypothetical protein

CNAG_07977 4 N/A hypothetical protein

CNAG_08008 11 N/A hypothetical protein

Table A3. List of differentially expressed genes (DEGs) in C. neoformans on


acetate-grown versus glucose-grown conditions

Gene log2 Fold-


Gene Annotation Symbol Chromosome Change FDR

xylulose 5-phosphate/fructose
CNAG_06923 6-phosphate phosphoketolase XFP2 3 -4.900496595 0

CNAG_05079 hypothetical protein N/A 4 -3.91100802 6.10E-15


Myo-inositol-1-phosphate
CNAG_01539 synthase N/A 11 -3.884280467 1.69E-261

CNAG_06298 hypothetical protein N/A 13 -3.783990989 1.15E-74

CNAG_06388 hypothetical protein N/A 13 -3.759133589 5.59E-21

CNAG_03759 conidiation-specific protein 6 N/A 2 -3.494376442 6.76E-16

CNAG_04857 hypothetical protein N/A 10 -3.424973515 1.39E-89

CNAG_03539 hypothetical protein N/A 8 -3.330761743 1.55E-73

alcohol dehydrogenase,
CNAG_07745 propanol-preferring MPD1 8 -3.292011686 1.18E-193

CNAG_04963 hypothetical protein N/A 4 -3.180091892 9.53E-98

CNAG_07981 hypothetical protein N/A 5 -3.141751498 6.78E-08

CNAG_01272 hypothetical protein N/A 5 -3.098706727 1.75E-96

CNAG_04313 NADPH2 dehydrogenase N/A 9 -3.0553792 8.97E-163

CNAG_02692 nitroreductase N/A 3 -3.00643794 8.88E-162

CNAG_02418 asparagine-tRNA ligase N/A 6 -2.99493202 1.78E-299

CNAG_06194 hypothetical protein N/A 12 -2.965511175 2.10E-37

133
ketol-acid reductoisomerase,
CNAG_05725 mitochondrial N/A 7 -2.89523887 1.73E-149

CNAG_06297 hypothetical protein N/A 13 -2.886470142 3.13E-10

CNAG_01562 pr4/barwin domain protein BLP4 11 -2.849443331 7.93E-145

CNAG_01277 hypothetical protein N/A 5 -2.836037827 6.67E-124

CNAG_06823 hypothetical protein N/A 5 -2.764776556 2.62E-20

CNAG_03983 oxidoreductase N/A 2 -2.764136537 4.95E-219

glucose-methanol-choline
CNAG_05258 oxidoreductase SMG1 4 -2.753581817 1.05E-66

CNAG_00749 alternative sulfate transporter N/A 1 -2.733300095 1.15E-66

CNAG_05278 hypothetical protein N/A 4 -2.721464246 1.80E-20

CNAG_05458 endo-1,3(4)-beta-glucanase N/A 14 -2.647163092 1.60E-125

CNAG_04659 Pyruvate decarboxylase PDC1 10 -2.640566552 1.33E-184

CNAG_00023 hypothetical protein N/A 1 -2.626687698 8.62E-70

CNAG_07638 hypothetical protein N/A 6 -2.547505476 4.35E-57

CNAG_03482 Thiol peroxidase TSA1 8 -2.539433406 5.94E-55

CNAG_02235 hypothetical protein N/A 6 -2.528804357 2.95E-44

CNAG_03438 hexose transporter HXT1 8 -2.501359805 9.23E-144

CNAG_05638 hypothetical protein N/A 14 -2.473390634 9.85E-102

transglycosylase SLT domain-


CNAG_04861 containing protein N/A 10 -2.449181211 2.91E-48
6-phosphogluconate
dehydrogenase,
CNAG_07561 decarboxylating 1 N/A 3 -2.431142436 1.95E-88

CNAG_02542 fructosamine kinase IRK2 6 -2.425279714 8.50E-56

CNAG_03477 hypothetical protein N/A 8 -2.422611089 1.34E-43

CNAG_04652 enoyl reductase N/A 10 -2.326270665 1.76E-80


phosphoglycerate
CNAG_00774 dehydrogenase N/A 1 -2.320407504 3.88E-102

CNAG_01464 Flavohemoglobin FHB1 11 -2.317039085 1.70E-49

CNAG_01960 efflux protein EncT N/A 11 -2.313456416 6.47E-57

CNAG_05874 hypothetical protein N/A 7 -2.25633409 5.95E-41

CNAG_04901 hypothetical protein N/A 10 -2.252047249 2.23E-30

134
CNAG_01907 PLK/PLK1 protein kinase N/A 11 -2.244250572 1.69E-40

CNAG_04347 Aspartate kinase N/A 9 -2.241789382 1.03E-158

CNAG_00261 hypothetical protein N/A 1 -2.233626148 5.20E-66

CNAG_05379 regucalcin N/A 14 -2.211205825 6.16E-75


small subunit ribosomal protein
CNAG_04114 S0 N/A 9 -2.199848081 2.32E-50

CNAG_07826 hypothetical protein N/A 4 -2.193322851 7.05E-21

CNAG_07445 transketolase N/A 5 -2.170823672 1.27E-54

CNAG_00799 cellulase N/A 1 -2.169536781 3.35E-61

large subunit ribosomal protein


CNAG_03577 LP0 N/A 8 -2.155719974 5.55E-69

CNAG_01954 aldo-keto reductase N/A 11 -2.155508762 7.61E-60

solute carrier family 25


(mitochondrial phosphate
CNAG_06377 transporter), member 3 N/A 13 -2.152642318 3.17E-72

L-aminoadipate-semialdehyde
CNAG_03588 dehydrogenase LYS2 8 -2.152612104 5.88E-188
0.019946052
CNAG_04551 hypothetical protein N/A 10 -2.152320791 42

CNAG_01896 alcohol dehydrogenase (NADP) N/A 11 -2.146751373 1.19E-74


C-4 methyl sterol oxidase,
CNAG_01737 putative ERG25 11 -2.14256466 3.93E-103

CNAG_03621 Cyclophilin A CPA2 2 -2.127098841 1.83E-103

CNAG_01744 phosphatase HAD1 11 -2.126660815 4.36E-77

CNAG_04183 hypothetical protein N/A 9 -2.111870737 1.24E-36

small subunit ribosomal protein


CNAG_02754 S12e N/A 3 -2.111585423 1.84E-118

CNAG_05623 chorismate synthase N/A 14 -2.062610548 1.17E-155

fructose-bisphosphate aldolase
CNAG_06770 1 FBA1 2 -2.061906962 7.67E-97

CNAG_04018 hypothetical protein N/A 2 -2.05957224 4.17E-49

CNAG_00515 mannitol dehydrogenase N/A 1 -2.056718088 4.69E-46

CNAG_02857 hypothetical protein N/A 3 -2.030723616 4.09E-51

CNAG_01803 hypothetical protein N/A 11 -2.030538073 6.25E-27

135
CNAG_00519 lathosterol oxidase ERG3 1 -2.030366682 2.58E-76

CNAG_07842 hypothetical protein N/A 10 -2.020548665 4.84E-06

Large subunit ribosomal


CNAG_06447 protein L17, putative RPL17 13 -2.005085008 1.11E-127

CNAG_05252 chaperone regulator N/A 4 -2.005078939 1.78E-79

large subunit ribosomal protein


CNAG_02928 L5e N/A 3 -2.002523096 9.01E-82

CNAG_05641 hypothetical protein, variant N/A 14 -2.000948487 2.24E-18

CNAG_00171 peroxin-2 N/A 1 2.002408397 2.54E-89


3-hydroxybutyryl-CoA
CNAG_03134 dehydrogenase HAD2 8 2.006338715 1.23E-53

CNAG_04890 hypothetical protein N/A 10 2.011876656 2.24E-06

CNAG_06672 formate dehydrogenase N/A 7 2.019028288 5.82E-38


wor1/pac2 family transcription
CNAG_05835 factor LIV3 7 2.020214829 1.01E-114
BNR/Asp-box repeat family
CNAG_02255 protein N/A 6 2.027277009 2.77E-61

CNAG_06000 glycoprotein N/A 12 2.029189017 9.97E-42

CNAG_03552 hypothetical protein N/A 8 2.033627971 8.25E-31


Glycine cleavage system T
CNAG_02818 protein N/A 3 2.037473869 5.18E-181

CNAG_06499 hypothetical protein DPP101 13 2.040968099 2.72E-111

methylmalonate-semialdehyde
CNAG_04351 dehydrogenase (acylating) N/A 9 2.044370708 3.26E-57

CNAG_06136 hypothetical protein N/A 12 2.049180999 3.91E-85

CNAG_00284 efflux protein EncT N/A 1 2.051495086 1.00E-40

CNAG_06188 hypothetical protein FZC15 12 2.051555525 1.23E-43

CNAG_06484 hypothetical protein N/A 13 2.055249818 6.64E-46

CNAG_04892 hypothetical protein, variant 4 N/A 10 2.057446252 6.20E-59

2-oxoisovalerate
dehydrogenase E1 component,
CNAG_02284 alpha subunit N/A 6 2.059866905 8.43E-78
bZip transcription factor,
CNAG_03346 putative BZP4 8 2.060783343 7.19E-57
myo-inositol transporter,
CNAG_05381 putative ITR3C 14 2.065739418 9.34E-37

CNAG_02593 hypothetical protein N/A 3 2.07028751 5.33E-85

136
CNAG_03279 hypothetical protein CCD4 8 2.071978705 6.91E-79
multidrug resistance protein
CNAG_05718 fnx1 N/A 7 2.073586105 9.63E-64
AAT family amino acid
CNAG_07902 transporter N/A 12 2.079429146 1.41E-74

CNAG_03692 hypothetical protein N/A 2 2.08347551 1.98E-101

CNAG_04100 hypothetical protein N/A 9 2.08616525 6.82E-16

CNAG_01835 hypothetical protein N/A 11 2.09318771 2.45E-128

CNAG_03051 polyamine transporter N/A 3 2.098521697 2.80E-74


3-hydroxyacyl-CoA
CNAG_04308 dehydrogenase HAD1 9 2.103182554 9.28E-98

CNAG_04618 hypothetical protein N/A 10 2.109468659 3.34E-66

CNAG_03133 UDP-glucose,sterol transferase ATG2602 8 2.111181869 4.92E-101

CNAG_01795 hypothetical protein N/A 11 2.118026923 2.18E-33

CNAG_01601 lipase N/A 11 2.125510186 1.15E-94

CNAG_00488 hypothetical protein, variant N/A 1 2.126659399 2.94E-57

CNAG_02164 hypothetical protein N/A 6 2.128792396 1.44E-113

CNAG_06186 hypothetical protein, variant 2 N/A 12 2.139409259 3.82E-84

methylmalonate-semialdehyde
CNAG_01075 dehydrogenase (acylating) N/A 5 2.139409989 9.16E-54

CNAG_07555 hypothetical protein N/A 3 2.145979311 4.35E-134

CNAG_07367 amino acid transporter, variant N/A 1 2.147164551 1.15E-33

CNAG_00521 hypothetical protein N/A 1 2.155386842 2.57E-51

CNAG_03696 hypothetical protein N/A 2 2.160906339 3.47E-75

CNAG_07993 hypothetical protein N/A 8 2.162635866 1.94E-06

CNAG_00048 Sugar transporter N/A 1 2.165703801 2.68E-70

CNAG_05266 membrane protein N/A 4 2.169742691 1.19E-36

CNAG_03960 Gly-Xaa carboxypeptidase N/A 2 2.174955689 1.79E-141

CNAG_04616 hypothetical protein N/A 10 2.179999719 2.79E-92

CNAG_04937 peroxin-12 N/A 4 2.199480608 8.13E-88

CNAG_00192 hypothetical protein N/A 1 2.201574687 5.30E-114

CNAG_02525 hypothetical protein N/A 6 2.201674493 2.36E-83

137
CNAG_07779 D-glycerate 3-kinase TDA10 9 2.205650117 5.10E-43

CNAG_02939 peroxin-3 N/A 3 2.206201302 1.33E-121

CNAG_02463 hypothetical protein N/A 6 2.208675884 1.07E-53


0.0211717349
CNAG_00446 hypothetical protein N/A 1 2.216615026 1

MFS transporter, SP family,


general alpha glucoside:H
CNAG_06259 symporter N/A 13 2.221697213 4.28E-52

CNAG_04585 hypothetical protein N/A 10 2.222530528 3.74E-27

CNAG_06241 acidic laccase CFO1 12 2.22727825 7.13E-83

CNAG_05313 hypothetical protein N/A 4 2.228133777 1.91E-107

CNAG_00524 acetyl-CoA acyltransferase 2 N/A 1 2.230319617 9.46E-135

CNAG_02777 phosphate:H symporter, variant PHO84 3 2.235545414 2.05E-09

CNAG_04346 dihydrodipicolinate synthase N/A 9 2.238276473 7.54E-60


ornithine-oxo-acid
CNAG_05134 transaminase N/A 4 2.241942301 7.64E-110

CNAG_08014 hypothetical protein N/A 12 2.243331657 4.50E-29

CNAG_06651 amidohydrolase N/A 7 2.244280371 5.90E-30

CNAG_02132 hypothetical protein N/A 6 2.248299335 4.56E-50

solute carrier family 25


(peroxisomal adenine
nucleotide transporter),
CNAG_03242 member 17 N/A 8 2.249113409 2.69E-81

CNAG_05547 hypothetical protein N/A 14 2.250676243 1.22E-54

protein-L-isoaspartate O-
CNAG_03837 methyltransferase N/A 2 2.264403293 7.91E-60
0.006272125
CNAG_03741 hypothetical protein FZC31 2 2.265537065 768

CNAG_01207 hypothetical protein N/A 5 2.271271362 8.24E-157

CNAG_04688 acyl-CoA dehydrogenase N/A 10 2.273564711 4.60E-121

glyceraldehyde-3-phosphate
CNAG_04523 dehydrogenase, type I N/A 9 2.281267594 9.36E-153

CNAG_04038 MFS quinate transporter QutD N/A 2 2.289231867 3.47E-79

CNAG_03046 hypothetical protein N/A 3 2.291459593 4.42E-83

CNAG_03511 hypothetical protein N/A 8 2.294631289 4.92E-21

138
CNAG_03050 hypothetical protein N/A 3 2.299742276 2.08E-100

CNAG_00456 Identified spore protein 6 ISP6 1 2.332279337 3.51E-110

CNAG_05042 carnitine acetyltransferase N/A 4 2.339243859 9.32E-136

CNAG_00190 hypothetical protein N/A 1 2.339409745 2.14E-37

CNAG_03782 hypothetical protein N/A 2 2.347378239 2.86E-113

CNAG_04092 Sugar transporter N/A 2 2.350600318 3.95E-79

CNAG_01865 hypothetical protein N/A 11 2.353302513 2.62E-128

CNAG_05293 glycogenin glucosyltransferase N/A 4 2.353347363 2.03E-55

CNAG_02096 peroxin-16 N/A 6 2.367732307 1.17E-77

CNAG_04185 hypothetical protein N/A 9 2.371692905 1.62E-66

CNAG_04096 racemase N/A 2 2.40182565 1.68E-126

CNAG_05115 sarcosine oxidase N/A 4 2.406894282 9.41E-91

CNAG_07341 hypothetical protein N/A 1 2.407757284 1.00E-187

CNAG_06817 uric acid xanthine permease UAP1 5 2.413959934 1.46E-110

nicotinamide mononucleotide
CNAG_06942 permease N/A 8 2.42053468 1.61E-137

CNAG_02978 hypothetical protein N/A 3 2.421534287 5.42E-81


peroxisome targeting signal
CNAG_01926 receptor PEX5 11 2.424142535 1.99E-146

CNAG_04531 Enoyl-CoA hydratase N/A 9 2.432157388 2.46E-129

CNAG_06568 RAN protein kinase SKS1 7 2.433420944 3.38E-88

CNAG_07512 hypothetical protein N/A 3 2.434398806 9.42E-118

CNAG_06448 Cystathionine gamma-lyase N/A 13 2.437034447 7.63E-64

CNAG_02524 hypothetical protein N/A 6 2.441590009 5.47E-88

CNAG_04291 glycosyl-hydrolase N/A 9 2.44488372 2.31E-117

CNAG_03697 hormone-sensitive lipase N/A 2 2.45694956 1.94E-194

CNAG_00442 cyclin N/A 1 2.465150614 1.41E-123

CNAG_04869 para-nitrobenzyl esterase PNB1 10 2.471842926 2.99E-117

CNAG_00269 Sorbitol dehydrogenase N/A 1 2.472085629 7.34E-54

CNAG_00425 hypothetical protein N/A 1 2.479120357 9.99E-34

139
CNAG_05180 hypothetical protein N/A 4 2.482393514 1.05E-242

CNAG_07458 hypothetical protein, variant N/A 5 2.482496041 5.05E-88

CNAG_06776 membrane protein N/A 2 2.485269115 2.00E-59

CNAG_07767 hypothetical protein N/A 9 2.494049409 6.76E-31

CNAG_03199 FAD dependent oxidoreductase N/A 8 2.500247586 7.11E-96

CNAG_05803 exo-beta-1,3-glucanase N/A 7 2.501116431 3.85E-120

CNAG_07403 Peroxisomal ATPase PEX1 PEX1 5 2.50683658 9.57E-158

CNAG_01952 aryl-alcohol dehydrogenase N/A 11 2.509646461 5.96E-67

CNAG_07913 hypothetical protein N/A 12 2.51058481 6.01E-09

Protein required for cytokinesis


CNAG_01983 and virulence 1 RCV1 11 2.524657503 1.67E-115
cardiolipin-specific
CNAG_02505 phospholipase N/A 6 2.52513257 1.47E-50

CNAG_04795 adenine nucleotide transporter N/A 10 2.528076177 3.98E-97

high-affinity nicotinic acid


CNAG_06204 transporter N/A 12 2.529064146 1.85E-150

CNAG_00904 aflatoxin efflux pump AFLT N/A 5 2.533320701 2.99E-10


1,4-alpha-glucan-branching
CNAG_00393 enzyme N/A 1 2.535777799 2.86E-184

malate dehydrogenase
(oxaloacetate-
CNAG_06374 decarboxylating)(NADP) N/A 13 2.564851405 8.46E-69

CNAG_07600 Beta-glucosidase N/A 11 2.565018508 5.55E-137

CNAG_06433 AMP-binding protein N/A 13 2.571617498 8.00E-197

ATP-binding cassette, subfamily


D (ALD), peroxisomal long-chain
CNAG_02764 fatty acid import protein PXA2 3 2.582242246 5.34E-143

CNAG_04794 spermine transporter N/A 10 2.584536287 1.19E-135

CNAG_00797 acetyl-CoA synthetase N/A 1 2.593962989 1.06E-135

CNAG_03394 hypothetical protein N/A 8 2.600980048 8.25E-157

CNAG_00079 hypothetical protein N/A 1 2.624742636 1.96E-59

CNAG_00896 transcription factor FZC34 5 2.628890692 9.44E-74

CNAG_07026 hypothetical protein N/A 12 2.632246795 1.47E-20

140
CNAG_07497 hypothetical protein N/A 3 2.641162697 8.49E-05

2-oxoisovalerate
dehydrogenase E2 component
CNAG_00484 (dihydrolipoyl transacylase) N/A 1 2.659748528 8.88E-180

CNAG_01493 hypothetical protein, variant N/A 11 2.66108414 1.28E-08

ATP-binding cassette, subfamily


CNAG_00796 B (MDR/TAP), member 1 MDR1 1 2.664776154 4.21E-53

CNAG_01847 hypothetical protein N/A 11 2.678662233 1.49E-183

CNAG_04960 hypothetical protein N/A 4 2.689977732 1.22E-82

CNAG_03551 hypothetical protein, variant N/A 8 2.701902734 2.81E-13


0.000177492
CNAG_06957 hypothetical protein N/A 11 2.709053016 4349

CNAG_01851 hypothetical protein N/A 11 2.709963284 3.86E-32

CNAG_05300 mitochondrial carrier protein N/A 4 2.730932228 1.93E-140

CNAG_06913 L-serine ammonia-lyase N/A 3 2.744533923 2.68E-146

CNAG_06561 allantoate transporter N/A 7 2.745084933 6.78E-57

CNAG_04759 hypothetical protein N/A 10 2.7466253 3.50E-76

CNAG_04459 hypothetical protein N/A 9 2.752032087 4.33E-68

CNAG_05119 GABA permease N/A 4 2.758643737 1.60E-87

extracellular elastinolytic
CNAG_04735 metalloproteinase MEP1 10 2.759144137 5.75E-151

CNAG_02118 hypothetical protein N/A 6 2.76036818 2.65E-243

CNAG_01481 hypothetical protein N/A 11 2.763816155 3.85E-189

CNAG_06493 hypothetical protein N/A 13 2.764588623 7.96E-79

CNAG_01116 beta-ketoacyl reductase N/A 5 2.767779821 4.50E-79

CNAG_01982 hypothetical protein N/A 11 2.771505722 9.80E-165

CNAG_05991 glycosyl hydrolase family 88 N/A 12 2.774307079 2.36E-176

CNAG_02473 endoplasmic reticulum protein N/A 6 2.780784919 4.57E-184

CNAG_00047 pyruvate dehydrogenase kinase PKP1 1 2.818316495 1.03E-112

CNAG_04093 putative transcription factor YRM103 2 2.821424664 9.87E-138

141
CNAG_05326 hypothetical protein N/A 4 2.827660976 5.21E-95

CNAG_06028 hypothetical protein N/A 12 2.8313924 7.72E-165

CNAG_04582 hypothetical protein N/A 10 2.83349268 2.16E-139

high-affinity nicotinic acid


CNAG_00028 transporter, variant N/A 1 2.833517394 1.41E-167

CNAG_06977 L-iditol 2-dehydrogenase N/A 8 2.836455175 1.82E-12

CNAG_06172 transketolase N/A 12 2.848885737 9.10E-75

CNAG_01925 hypothetical protein N/A 11 2.878018723 3.47E-226

MFS transporter, SP family,


general alpha glucoside:H
CNAG_05929 symporter N/A 7 2.893258898 5.48E-134

CNAG_00001 hypothetical protein N/A 1 2.898307253 1.00E-15


Uracil
CNAG_02344 phosphoribosyltransferase N/A 6 2.909303709 4.74E-132

CNAG_03729 peroxin-5 N/A 2 2.91316122 3.71E-263

CNAG_05299 oxidoreductase N/A 4 2.914464089 8.17E-75

CNAG_02439 hypothetical protein, variant N/A 6 2.932982057 1.55E-57


glycogen(starch) synthase,
CNAG_04621 variant N/A 10 2.940035403 5.84E-184

CNAG_02934 hypothetical protein, variant N/A 3 2.945057505 2.59E-250

CNAG_01384 hypothetical protein N/A 5 2.952408893 1.08E-116

CNAG_07704 hypothetical protein N/A 7 2.960740455 8.17E-13

CNAG_06009 cyclohydrolase N/A 12 2.962930691 1.30E-190

CNAG_00015 hypothetical protein N/A 1 3.019537088 1.08E-129


myo-inositol transporter,
CNAG_00864 putative ITR2 5 3.035563112 9.98E-182

CNAG_05639 hypothetical protein PPS1 14 3.038252281 0

peroxisomal 2,4-dienoyl-CoA
CNAG_04238 reductase N/A 9 3.041104459 6.05E-230

CNAG_00637 Cystathionine beta-synthase N/A 1 3.059496886 5.75E-290

CNAG_00141 monooxygenase N/A 1 3.082581858 4.71E-45

CNAG_06574 antiphagocytic protein 1 APP1 7 3.088293978 3.40E-122

CNAG_02049 Proline dehydrogenase PUT1 6 3.09586931 7.34E-92

142
ATP-binding cassette, subfamily
D (ALD), peroxisomal long-chain
CNAG_00651 fatty acid import protein N/A 1 3.10426851 0

CNAG_03295 hypothetical protein N/A 8 3.122326156 2.02E-82

CNAG_01180 hypothetical protein N/A 5 3.153286452 4.14E-223

CNAG_07305 hypothetical protein N/A 1 3.173833764 7.97E-20


hydroxymethylglutaryl-CoA
CNAG_03067 lyase N/A 3 3.176153974 2.07E-274

CNAG_07549 hypothetical protein N/A 3 3.197187595 3.38E-178

CNAG_02586 Sugar transporter N/A 3 3.213961171 2.20E-108


NAD-dependent
CNAG_00094 epimerase/dehydratase N/A 1 3.227931305 1.14E-261

CNAG_05599 hypothetical protein N/A 14 3.233960578 5.37E-260

CNAG_06522 hypothetical protein N/A 7 3.260241994 7.87E-23


phosphoenolpyruvate
CNAG_04217 carboxykinase (ATP) PCK1 9 3.26714125 6.60E-163

CNAG_07747 acyl-CoA oxidase N/A 8 3.284455701 2.29E-250

CNAG_02254 quinate permease LPI12 6 3.294651913 9.77E-295

CNAG_07303 hypothetical protein, variant N/A 1 3.302928715 2.98E-10

2-oxoisovalerate
dehydrogenase E1 component,
CNAG_00397 beta subunit N/A 1 3.327788507 4.05E-204

CNAG_05251 hypothetical protein N/A 4 3.3666192 1.38E-64

CNAG_02225 glucan 1,3-beta-glucosidase EXG104 6 3.371935116 2.52E-249

CNAG_07944 Urea carboxylase N/A 13 3.394992027 0

CNAG_07984 hypothetical protein N/A 6 3.406703097 1.29E-67

CNAG_07641 Monosaccharide transporter N/A 6 3.418297251 1.06E-252

CNAG_05031 3-oxoacid CoA-transferase N/A 4 3.444662639 0

CNAG_02165 cytoplasmic protein N/A 6 3.468757517 1.26E-167

CNAG_06926 hypothetical protein N/A 3 3.51291347 0

CNAG_05990 hypothetical protein N/A 12 3.520892501 1.66E-217

CNAG_00844 hypothetical protein N/A 1 3.52507934 1.31E-162

alpha-ketoglutarate-dependent
2,4- dichlorophenoxyacetate
CNAG_04417 dioxygenase N/A 9 3.539262119 3.83E-88

143
CNAG_06628 aldehyde dehydrogenase (NAD) N/A 7 3.554003489 2.07E-269
acetyl/propionyl CoA
CNAG_01671 carboxylase N/A 11 3.563038208 0

CNAG_02562 acyl-CoA dehydrogenase N/A 3 3.63243594 1.89E-266

CNAG_04392 sterol-binding protein MFE2 9 3.64229792 9.91E-116

CNAG_01551 GATA family transcription factor GAT201 11 3.675353875 5.24E-221

solute carrier family 25


(mitochondrial
carnitine/acylcarnitine
CNAG_00499 transporter), member 20/29 N/A 1 3.69160675 0

aldehyde dehydrogenase family


CNAG_00735 7 member A1 N/A 1 3.713063302 1.61E-259

CNAG_02188 hypothetical protein N/A 6 3.72088673 0

CNAG_05644 2-Nitropropane dioxygenase N/A 14 3.729505188 0

phosphotransferase enzyme
CNAG_02295 family protein N/A 6 3.785156862 1.02E-301

CNAG_03019 long-chain acyl-CoA synthetase N/A 3 3.790271235 0

CNAG_05685 neutral amino acid transporter N/A 7 3.83528674 0

thiosulfate/3-mercaptopyruvate
CNAG_01252 sulfurtransferase N/A 5 3.839575537 1.09E-151

MFS transporter, SP family,


general alpha glucoside:H
CNAG_06527 symporter, variant 2 N/A 7 3.840821525 4.20E-149

CNAG_02044 hypothetical protein N/A 6 3.860207276 5.16E-124

CNAG_03516 hypothetical protein N/A 8 3.861739181 2.40E-243

CNAG_04891 hypothetical protein N/A 10 3.90197729 4.30E-247

CNAG_00826 dihydroxyacetone kinase DAK101 1 3.981063756 0


multifunctional beta-oxidation
CNAG_05721 protein MFE2 7 4.022808323 0

CNAG_05658 L-iditol 2-dehydrogenase N/A 14 4.051104215 3.05E-223

specific RNA polymerase II


CNAG_04345 transcription factor ARO8001 9 4.057102724 0
taurine catabolism dioxygenase
CNAG_01542 TauD N/A 11 4.087034641 1.74E-305

CNAG_04548 hypothetical protein N/A 1 4.09792973 5.07E-44

144
0.04074920
CNAG_06938 hypothetical protein N/A 12 4.107921681 641

CNAG_05229 Stomatin family protein N/A 4 4.133932227 1.17E-232

CNAG_05785 putative transcription factor STB4 7 4.142636148 2.09E-218

CNAG_00490 acetyl-CoA acyltransferase N/A 1 4.150523898 0

CNAG_00726 hypothetical protein N/A 1 4.187090684 9.17E-91

CNAG_02147 cytochrome c peroxidase N/A 6 4.190300827 4.20E-208

CNAG_03243 2-Nitropropane dioxygenase N/A 8 4.194067019 2.51E-143

CNAG_03010 enoyl-CoA hydratase/isomerase N/A 3 4.251823055 0

CNAG_06551 Carnitine O-acetyltransferase N/A 7 4.252895603 0

nicotinamide mononucleotide
CNAG_04536 permease N/A 9 4.261264549 5.91E-218

CNAG_07874 Sugar transporter N/A 14 4.271990142 2.21E-150

CNAG_06583 hypothetical protein N/A 7 4.344699995 0


0.007701188
CNAG_04931 Putative hexose transporter HXS2 10 4.37816534 705

CNAG_00452 Isovaleryl-CoA dehydrogenase N/A 1 4.381863876 3.22E-133

CNAG_06312 hypothetical protein N/A 13 4.394641355 7.53E-84


0.045335569
CNAG_07791 hypothetical protein N/A 9 4.431599325 56

CNAG_05324 Sugar transporter N/A 4 4.434844391 0

CNAG_00474 hypothetical protein N/A 1 4.509539051 1.57E-24

CNAG_00597 amino acid transporter DIP5 1 4.518136614 0

CNAG_00588 hypothetical protein N/A 1 4.636458197 7.03E-202

alpha-aminoadipic
CNAG_00247 semialdehyde synthase LYS9 1 4.647012781 0

3-methylcrotonyl-CoA
CNAG_01680 carboxylase alpha subunit N/A 11 4.653411548 0

CNAG_06294 hypothetical protein N/A 13 4.728962956 0

CNAG_05913 Alpha-glucosidase N/A 7 4.759365762 0


0.007461162
CNAG_07919 hypothetical protein N/A 13 4.763226734 725

CNAG_02045 acetoacetate-CoA ligase N/A 6 4.784163223 0

145
CNAG_07832 hypothetical protein N/A 4 4.801798277 3.44E-40

CNAG_02733 Monosaccharide transporter N/A 3 4.838391609 7.50E-302

MFS transporter, SHS family,


CNAG_04704 lactate transporter N/A 10 4.877479239 4.20E-260

CNAG_07912 hypothetical protein N/A 12 4.893173881 2.57E-193

CNAG_06431 acyl-CoA oxidase N/A 13 4.916807567 0

CNAG_03666 acyl-CoA dehydrogenase N/A 2 4.942071419 0

CNAG_05653 malate synthase A MLS1 14 4.950928228 0

CNAG_00827 ribose 5-phosphate isomerase N/A 1 4.99834915 1.33E-143

CNAG_05310 nipsnap family protein N/A 4 5.003188347 1.07E-195

CNAG_00879 Glutamate dehydrogenase N/A 5 5.152721407 0

CNAG_00537 Carnitine O-acetyltransferase N/A 1 5.18957041 0

CNAG_04142 tartrate transporter N/A 9 5.329645146 0


0.001460120
CNAG_07496 hypothetical protein N/A 3 5.338398519 406

CNAG_01936 Sugar transporter N/A 11 5.398218881 0

CNAG_07862 fumarate reductase N/A 10 5.402443646 0

CNAG_05387 galactose transporter N/A 14 5.431201544 0

nicotinamide mononucleotide
CNAG_00598 permease N/A 1 5.436810579 0

MFS transporter, SP family,


general alpha glucoside:H
CNAG_05330 symporter N/A 4 5.501723558 1.13E-94

CNAG_04837 bHLH family transcription factor MLN1 10 5.658104561 0

solute carrier family 25


(mitochondrial citrate
CNAG_02288 transporter), member 1 N/A 6 5.725525264 0

CNAG_00679 hypothetical protein N/A 1 6.375293407 3.38E-266


myo-inositol transporter,
CNAG_03910 putative ITR6 2 6.517998365 0

alcohol dehydrogenase,
CNAG_02489 propanol-preferring N/A 6 6.655716374 0

CNAG_05303 isocitrate lyase ICL1 4 6.900794065 0

CNAG_07869 hypothetical protein N/A 14 7.488596165 0

146
CNAG_05867 L-fucose transporter N/A 7 7.624377564 0

MFS transporter, SP family,


general alpha glucoside:H
CNAG_05914 symporter N/A 7 7.804574725 0

CNAG_05662 Polyol transporter protein 1 PTP1 14 7.859393735 0


high-affinity glucose
CNAG_03772 transporter HXS1 2 8.11764169 0

Notes:
1. List contains DEGs of the nuclear genome
2. The 5th column is color coded: yellow for downregulated genes and red for
upregulate genes

147
APPENDIX B

Putative acetate kinases of eukaryotic origin

Table B1. List of putative eukaryotic acetate kinases identified in the study

UniProt Length Species Taxon


accession (amino
acid)
A0A2P6P0X2 428 Planoprotostelium fungivorum Amoebozoa

A0A0A1U119 394 Entamoeba invadens IP1 Amoebozoa

C4M1C3 392 Entamoeba histolytica Amoebozoa

A0A0F4GWL1 420 Zymoseptoria brevis Ascomycota

A0A165G327 414 Xylona heveae (strain CBS 132557 / TC161) Ascomycota

A0A3M9Y9Y9 415 Verticillium nonalfalfae Ascomycota

A0A0D2A747 421 Verruconis gallopava Ascomycota

A0A370U2T7 457 Venustampulla echinocandica Ascomycota

A0A4Z1NHJ3 426 Venturia nashicola Ascomycota

A0A8H3U544 426 Venturia inaequalis (Apple scab fungus) Ascomycota

A0A517LHE8 426 Venturia effusa Ascomycota

A0A423W2V6 472 Valsa sordida Ascomycota

A0A423X684 460 Valsa malicola Ascomycota

A0A194UQZ8 447 Valsa mali var. pyri Ascomycota

A0A8E5MF26 420 Ustilaginoidea virens (Rice false smut fungus) Ascomycota


(Villosiclava virens)

148
C4JZC3 417 Uncinocarpus reesii (strain UAMH 1704) Ascomycota

A0A178EUJ9 410 Trichophyton rubrum (Athlete's foot fungus) Ascomycota


(Epidermophyton rubrum)

A0A059J4T7 401 Trichophyton interdigitale (strain MR816) Ascomycota

A0A2T4CH84 415 Trichoderma longibrachiatum ATCC 18648 Ascomycota

A0A0F9WTJ8 804 Trichoderma harzianum (Hypocrea lixii) Ascomycota

A0A2P4Z935 417 Trichoderma gamsii Ascomycota

A0A2T3ZQT0 417 Trichoderma asperellum CBS 433.97 Ascomycota

A0A395NJV2 417 Trichoderma arundinaceum Ascomycota

A0A0A1TI16 417 Torrubiella hemipterigena Ascomycota

A0A2S4L229 414 Tolypocladium paradoxum Ascomycota

A0A0L0N8L1 1157 Tolypocladium ophioglossoides (strain CBS 100239) Ascomycota


(Elaphocordyceps ophioglossoides)

A0A507BFA5 1602 Thyridium curvatum Ascomycota

A0A0F4Z849 414 Thielaviopsis punctulata Ascomycota

G2R1K7 436 Thermothielavioides terrestris (strain ATCC 38088 / Ascomycota


NRRL 8126) (Thielavia terrestris)

149
R4XA18 415 Taphrina deformans (strain PYCC 5710 / ATCC 11124 / Ascomycota
CBS 356.35 / IMI 108563 / JCM 9778 / NBRC 8474)
(Peach leaf curl fungus) (Lalaria deformans)

B8M7Z6 415 Talaromyces stipitatus (strain ATCC 10500 / CBS 375.48 Ascomycota
/ QM 6759 / NRRL 1006) (Penicillium stipitatum)

A0A7H8QRV8 420 Talaromyces rugulosus Ascomycota

B6Q5T9 452 Talaromyces marneffei (strain ATCC 18224 / CBS 334.59 Ascomycota
/ QM 7333) (Penicillium marneffei)

A0A0U1LTG9 421 Talaromyces islandicus (Penicillium islandicum) Ascomycota

A0A1Q5Q8B9 413 Talaromyces atroroseus Ascomycota

A0A364KLK3 415 Talaromyces amestolkiae Ascomycota

A0A364N0R6 419 Stemphylium lycopersici Ascomycota

A0A178AIA2 419 Stagonospora sp. SRC1lsM3a Ascomycota

A0A084QIB7 416 Stachybotrys chlorohalonata (strain IBT 40285) Ascomycota

U7Q038 443 Sporothrix schenckii (strain ATCC 58251 / de Perez Ascomycota


2211183) (Rose-picker's disease fungus)

150
A0A162I8M3 423 Sporothrix insectorum RCEF 264 Ascomycota

M3B0H8 421 Sphaerulina musiva (strain SO2202) (Poplar stem Ascomycota


canker fungus) (Septoria musiva)

F7W1G1 461 Sordaria macrospora (strain ATCC MYA-333 / DSM 997 / Ascomycota
K(L3346) / K-hell)

A0A3N2PRN0 415 Sodiomyces alkalinus F11 Ascomycota

R0ICZ8 419 Setosphaeria turcica (strain 28A) (Northern leaf blight Ascomycota
fungus) (Exserohilum turcicum)

A0A425BU85 396 Scytalidium sp. 3C Ascomycota

A0A8H2VYZ8 419 Sclerotinia trifoliorum Ascomycota

W9CR99 423 Sclerotinia borealis (strain F-4128) Ascomycota

A0A4U0U5S2 421 Salinomyces thailandica Ascomycota

A0A0E9N8B6 510 Saitoella complicata (strain BCRC 22490 / CBS 7301 / Ascomycota
JCM 7358 / NBRC 10748 / NRRL Y-17804)

A0A2S7PA73 428 Rutstroemia sp. NJR-2017a BVV2 Ascomycota

A0A1E1K2L5 417 Rhynchosporium commune Ascomycota

A0A0D2ITC7 412 Rhinocladiella mackenziei CBS 650.93 Ascomycota

A0A2D3VN54 424 Ramularia collo-cygni Ascomycota

151
B2WH19 412 Pyrenophora tritici-repentis (strain Pt-1C-BFP) (Wheat Ascomycota
tan spot fungus) (Drechslera tritici-repentis)

E3RRQ8 419 Pyrenophora teres f. teres (strain 0-1) (Barley net blotch Ascomycota
fungus) (Drechslera teres f. teres)

A0A178DWV0 419 Pyrenochaeta sp. DS3sAY3a Ascomycota

A0A179HNH1 414 Purpureocillium lilacinum (Paecilomyces lilacinus) Ascomycota

A0A1B8GHL8 419 Pseudogymnoascus verrucosus Ascomycota

L8FPJ8 419 Pseudogymnoascus destructans (strain ATCC MYA- Ascomycota


4855 / 20631-21) (Bat white-nose syndrome fungus)
(Geomyces destructans)

A0A8H6VPD4 419 Pseudocercospora fuligena Ascomycota

M3AJ32 419 Pseudocercospora fijiensis (strain CIRAD86) (Black leaf Ascomycota


streak disease fungus) (Mycosphaerella fijiensis)

A0A139GYJ5 419 Pseudocercospora eumusae Ascomycota

A0A1Y2FSE6 408 Protomyces lactucae-debilis Ascomycota

A0A2B7YE97 415 Polytolypa hystricis UAMH7299 Ascomycota

152
B2ALS8 437 Podospora anserina (strain S / ATCC MYA-4624 / DSM Ascomycota
980 / FGSC 10383) (Pleurage anserina)

A0A179FGW7 415 Pochonia chlamydosporia 170 Ascomycota

A0A0N0NP39 407 Phialophora attinorum Ascomycota

A0A0D2E2K0 418 Phialophora americana Ascomycota

A0A1L7X722 416 Phialocephala subalpina Ascomycota

Q0U2Y7 419 Phaeosphaeria nodorum (strain SN15 / ATCC MYA- Ascomycota


4574 / FGSC 10173) (Glume blotch fungus)
(Parastagonospora nodorum)

A0A0G2GIQ0 413 Phaeomoniella chlamydospora Ascomycota

R8BCA5 433 Phaeoacremonium minimum (strain UCR-PA7) (Esca Ascomycota


disease fungus) (Togninia minima)

A0A5N6FI44 409 Petromyces alliaceus (Aspergillus alliaceus) Ascomycota

W3XNS0 422 Pestalotiopsis fici (strain W106-1 / CGMCC3.15140) Ascomycota

A0A2V1D5Z2 419 Periconia macrospinosa Ascomycota

A0A1V6SBD3 403 Penicillium vulpinum Ascomycota

A0A1Q5TGI6 418 Penicillium subrubescens Ascomycota

A0A1V6T0M6 414 Penicillium steckii Ascomycota

153
A0A1V6RQP5 389 Penicillium solitum Ascomycota

B6HBY2 437 Penicillium rubens (strain ATCC 28089 / DSM 1075 / Ascomycota
NRRL 1951 / Wisconsin 54-1255) (Penicillium
chrysogenum)

A0A1V6P4I6 414 Penicillium polonicum Ascomycota

A0A135LIQ9 421 Penicillium patulum (Penicillium griseofulvum) Ascomycota

S7ZIS8 414 Penicillium oxalicum (strain 114-2 / CGMCC 5302) Ascomycota


(Penicillium decumbens)

A0A0M8NYD5 414 Penicillium nordicum Ascomycota

A0A1V6Z6R9 463 Penicillium nalgiovense Ascomycota

A0A0A2KAM9 395 Penicillium italicum (Blue mold) Ascomycota

A0A101MGX6 414 Penicillium freii Ascomycota

A0A1V6SPN5 431 Penicillium flavigenum Ascomycota

A0A0A2KEA4 459 Penicillium expansum (Blue mold rot fungus) Ascomycota

A0A1V6PMA8 414 Penicillium decumbens Ascomycota

A0A1V6V7D8 403 Penicillium coprophilum Ascomycota

A0A0F7TI52 414 Penicillium brasilianum Ascomycota

A0A1F5LCL1 423 Penicillium arizonense Ascomycota

A0A1V6QLJ1 407 Penicillium antarcticum Ascomycota

154
A0A1L9SX51 416 Penicilliopsis zonata CBS 506.65 Ascomycota

A0A6H0Y0H4 419 Peltaster fructicola Ascomycota

A0A177BWQ2 418 Paraphaeosphaeria sporulosa Ascomycota

C1H8T9 439 Paracoccidioides lutzii (strain ATCC MYA-826 / Pb01) Ascomycota


(Paracoccidioides brasiliensis)

S3CRH5 435 Ophiostoma piceae (strain UAMH 11346) (Sap stain Ascomycota
fungus)

A0A2C5XTH9 413 Ophiocordyceps australis Ascomycota

A0A0C3CSB4 417 Oidiodendron maius Zn Ascomycota

G4U9W7 469 Neurospora tetrasperma (strain FGSC 2509 / P0656) Ascomycota

Q7SH17 469 Neurospora crassa (strain ATCC 24698 / 74-OR23-1A / Ascomycota


CBS 708.71 / DSM 1257 / FGSC 987)

A1D6Z2 425 Neosartorya fischeri (strain ATCC 1020 / DSM 3700 / Ascomycota
CBS 544.65 / FGSC A1164 / JCM 1740 / NRRL 181 / WB 181)
(Aspergillus fischerianus)

A0A0P7BCD8 842 Neonectria ditissima Ascomycota

A0A6A6Y5A7 417 Mytilinidion resinicola Ascomycota

155
G2Q2D7 447 Myceliophthora thermophila (strain ATCC 42464 / Ascomycota
BCRC 31852 / DSM 1799) (Sporotrichum thermophile)

A0A3N4L0W6 435 Morchella conica CCBAS932 Ascomycota

A0A4Q4T869 415 Monosporascus sp. MC13-8B Ascomycota

A0A4Q4TS32 415 Monosporascus ibericus Ascomycota

A0A5N6JU89 470 Monilinia laxa (Brown rot fungus) (Sclerotinia laxa) Ascomycota

A0A5M9JCQ9 501 Monilinia fructicola (Brown rot fungus) (Ciboria Ascomycota


fructicola)

A0A507QMX6 421 Monascus purpureus (Red mold) (Monascus anka) Ascomycota

A0A194XKF5 416 Mollisia scopiformis Ascomycota

A0A168B4M1 416 Moelleriella libera RCEF 2490 Ascomycota

A0A136IWQ0 434 Microdochium bolleyi Ascomycota

A0A162J5H1 415 Metarhizium rileyi (strain RCEF 4871) (Nomuraea rileyi) Ascomycota

A0A0D9PCA3 416 Metarhizium anisopliae BRIP 53293 Ascomycota

A0A0B2WYL4 416 Metarhizium album (strain ARSEF 1941) Ascomycota

E9EB84 416 Metarhizium acridum (strain CQMa 102) Ascomycota

A0A7C8HY93 440 Massariosphaeria phaeospora Ascomycota

156
A0A218Z4D0 553 Marssonina coronariae Ascomycota

K1X5B9 404 Marssonina brunnea f. sp. multigermtubi (strain Ascomycota


MB_m1) (Marssonina leaf spot fungus)

G4MPH1 432 Magnaporthe oryzae (strain 70-15 / ATCC MYA-4617 / Ascomycota


FGSC 8958) (Rice blast fungus) (Pyricularia oryzae)

A0A6P8B9W0 432 Magnaporthe grisea (Crabgrass-specific blast fungus) Ascomycota


(Pyricularia grisea)

A0A175W805 423 Madurella mycetomatis Ascomycota

K2RPL2 443 Macrophomina phaseolina (strain MS6) (Charcoal rot Ascomycota


fungus)

A0A1E3QE05 419 Lipomyces starkeyi NRRL Y-11557 Ascomycota

A0A8H6CKG4 412 Letharia lupina Ascomycota

E5A190 425 Leptosphaeria maculans (strain JN3 / isolate v23.1.3 / Ascomycota


race Av1-4-5-6-7-8) (Blackleg fungus) (Phoma lingam)

A0A8E2E485 421 Lepidopterella palustris CBS 459.81 Ascomycota

A0A5N5DLW6 422 Lasiodiplodia theobromae Ascomycota

A0A559M5D1 406 Lachnellula willkommii Ascomycota

A0A8H8RTH2 417 Lachnellula subtilissima Ascomycota

157
A0A8H8UK53 417 Lachnellula occidentalis Ascomycota

A0A8H8QT93 417 Lachnellula hyalina Ascomycota

A0A7D8UN39 427 Lachnellula cervina Ascomycota

A0A1Y2UG58 417 Hypoxylon sp. CO27-5 Ascomycota

G9N905 417 Hypocrea virens (strain Gv29-8 / FGSC 10586) Ascomycota


(Gliocladium virens) (Trichoderma virens)

G0RDG0 415 Hypocrea jecorina (strain QM6a) (Trichoderma reesei) Ascomycota

G9P656 417 Hypocrea atroviridis (strain ATCC 20476 / IMI 206040) Ascomycota
(Trichoderma atroviride)

A0A2J6SD10 416 Hyaloscypha variabilis F Ascomycota

A0A2J6PU90 417 Hyaloscypha hepaticicola Ascomycota

A0A2J6SYZ9 417 Hyaloscypha bicolor E Ascomycota

A0A1Z5SL83 505 Hortaea werneckii EXF-2000 Ascomycota

A0A8H7Z444 421 Histoplasma ohiense (nom. inval.) Ascomycota

A0A0F7ZSK2 415 Hirsutella minnesotensis 3608 Ascomycota

A0A2B7XQ83 417 Helicocarpus griseus UAMH5409 Ascomycota

158
F0XC80 655 Grosmannia clavigera (strain kw1407 / UAMH 11150) Ascomycota
(Blue stain fungus) (Graphiocladiella clavigera)

A0A8E2ETB1 417 Glonium stellatum Ascomycota

S3DEK6 416 Glarea lozoyensis (strain ATCC 20868 / MF5171) Ascomycota

A0A0E0RQ82 415 Gibberella zeae (strain ATCC MYA-4620 / CBS 123657 / Ascomycota
FGSC 9075 / NRRL 31084 / PH-1) (Wheat head blight
fungus) (Fusarium graminearum)

A0A8H5P1M9 835 Gibberella subglutinans (Fusarium subglutinans) Ascomycota

A0A2K0WDE8 415 Gibberella nygamai (Bean root rot disease fungus) Ascomycota
(Fusarium nygamai)

W7LRF8 415 Gibberella moniliformis (strain M3125 / FGSC 7600) Ascomycota


(Maize ear and stalk rot fungus) (Fusarium
verticillioides)

S0DI41 415 Gibberella fujikuroi (strain CBS 195.34 / IMI 58289 / Ascomycota
NRRL A-6831) (Bakanae and foot rot disease fungus)
(Fusarium fujikuroi)

159
J3NJ82 468 Gaeumannomyces tritici (strain R3-111a-1) (Wheat and Ascomycota
barley take-all root rot fungus) (Gaeumannomyces
graminis var. tritici)

A0A8H4U7Z2 414 Fusarium zealandicum Ascomycota

A0A2L2TNY0 415 Fusarium venenatum Ascomycota

C7YK38 414 Fusarium vanettenii (strain ATCC MYA-4622 / CBS Ascomycota


123669 / FGSC 9596 / NRRL 45880 / 77-13-4) (Fusarium
solani subsp. pisi)

A0A395SX31 419 Fusarium sporotrichioides Ascomycota

A0A8H4SWU7 414 Fusarium sarcochroum Ascomycota

A0A8H5L2W2 415 Fusarium pseudocircinatum Ascomycota

A0A1B8B3W3 415 Fusarium poae Ascomycota

A0A8H5IME6 908 Fusarium phyllophilum Ascomycota

A0A0D2X879 541 Fusarium oxysporum f. sp. lycopersici (strain 4287 / Ascomycota


CBS 123668 / FGSC 9935 / NRRL 34936) (Fusarium
vascular wilt of tomato)

A0A8H5IJJ4 403 Fusarium napiforme Ascomycota

A0A8H6D3E6 415 Fusarium mundagurra Ascomycota

A0A395RYV7 419 Fusarium longipes Ascomycota

A0A3M2S4Z2 414 Fusarium kuroshium Ascomycota

160
A0A8H5XXI0 415 Fusarium globosum Ascomycota

A0A8H4WR54 838 Fusarium gaditjirri Ascomycota

A0A395MIC5 415 Fusarium flagelliforme Ascomycota

A0A2T4HAR1 710 Fusarium culmorum Ascomycota

A0A8H4P5B2 415 Fusarium austroafricanum Ascomycota

A0A8H4L2G8 414 Fusarium albosuccineum Ascomycota

A0A8H4NFG0 415 Fusarium acutatum Ascomycota

A0A4V6WL79 1761 Friedmanniomyces simplex Ascomycota

A0A4U0VI43 430 Friedmanniomyces endolithicus Ascomycota

A0A0D2GMI4 436 Fonsecaea pedrosoi CBS 271.37 Ascomycota

A0A0D2JUH1 451 Fonsecaea multimorphosa CBS 102226 Ascomycota

A0A178ZNV0 417 Fonsecaea erecta Ascomycota

A0A0D2EQT4 421 Exophiala xenobiotica Ascomycota

A0A0D2C375 421 Exophiala spinifera Ascomycota

A0A0D1XF34 424 Exophiala sideris Ascomycota

A0A0D2B121 421 Exophiala oligosperma Ascomycota

A0A0D1ZH21 414 Exophiala mesophila (Black yeast) Ascomycota

H6BV64 423 Exophiala dermatitidis (strain ATCC 34100 / CBS 525.76 Ascomycota
/ NIH/UT8656) (Black yeast) (Wangiella dermatitidis)

161
A0A072PV08 423 Exophiala aquamarina CBS 119918 Ascomycota

M7T3L5 418 Eutypa lata (strain UCR-EL1) (Grapevine dieback Ascomycota


disease fungus) (Eutypa armeniacae)

A0A6G1G3F4 412 Eremomyces bilateralis CBS 781.70 Ascomycota

A0A1Y2LXG4 419 Epicoccum nigrum (Soil fungus) (Epicoccum Ascomycota


purpurascens)

A0A7S9PV96 421 Epichloe festucae (strain Fl1) Ascomycota

A0A2B7ZID5 433 Emmonsia crescens Ascomycota

Q5B3G6 420 Emericella nidulans (strain FGSC A4 / ATCC 38163 / CBS Ascomycota
112.46 / NRRL 194 / M139) (Aspergillus nidulans)

A0A1J9QSB5 426 Emergomyces pasteurianus Ep9510 Ascomycota

A0A1B7P5T7 527 Emergomyces africanus Ascomycota

A0A232LPM9 414 Elaphomyces granulatus Ascomycota

N1PRG8 423 Dothistroma septosporum (strain NZE10 / CBS 128990) Ascomycota


(Red band needle blight fungus) (Mycosphaerella pini)

A0A1S8BEI3 422 Diplodia seriata Ascomycota

A0A1J9RQK1 422 Diplodia corticola Ascomycota

162
A0A162WCX8 419 Didymella rabiei (Chickpea ascochyta blight fungus) Ascomycota
(Mycosphaerella rabiei)

A0A1Y2XC97 415 Daldinia sp. EC12 Ascomycota

A0A423XIR8 446 Cytospora leucostoma Ascomycota

W2RIW9 417 Cyphellophora europaea CBS 101466 Ascomycota

A0A4U0XC36 424 Cryomyces minteri Ascomycota

A0A179I9K0 412 Cordyceps confragosa (Lecanicillium lecanii) Ascomycota

R7Z350 432 Coniosporium apollinis (strain CBS 100218) (Rock- Ascomycota


inhabiting black yeast)

A0A420XXW1 416 Coniochaeta pulveracea Ascomycota

A0A1J7J5G2 781 Coniochaeta ligniaria NRRL 30616 Ascomycota

A0A2T3ABD5 462 Coniella lustricola Ascomycota

A0A4R8RIR5 414 Colletotrichum trifolii Ascomycota

A0A161YKA7 494 Colletotrichum tofieldiae Ascomycota

A0A4U6XVE9 416 Colletotrichum tanaceti Ascomycota

A0A066Y2I4 438 Colletotrichum sublineola (Sorghum anthracnose Ascomycota


fungus)

A0A5Q4C5U8 415 Colletotrichum shisoi Ascomycota

A0A135UQL4 417 Colletotrichum salicis Ascomycota

163
A0A8H6JX88 414 Colletotrichum plurivorum Ascomycota

A0A1G4APG6 417 Colletotrichum orchidophilum Ascomycota

N4V0K9 414 Colletotrichum orbiculare (strain 104-T / ATCC 96160 / Ascomycota


CBS 514.97 / LARS 414 / MAFF 240422) (Cucumber
anthracnose fungus) (Colletotrichum lagenarium)

A0A8H6U8R8 414 Colletotrichum musicola Ascomycota

A0A167B7J0 449 Colletotrichum incanum Ascomycota

A0A4T0VN07 414 Colletotrichum higginsianum Ascomycota

E3QCU7 415 Colletotrichum graminicola (strain M1.001 / M2 / FGSC Ascomycota


10212) (Maize anthracnose fungus) (Glomerella
graminicola)

A0A8H4FI56 692 Colletotrichum gloeosporioides (Anthracnose fungus) Ascomycota


(Glomerella cingulata)

L2FYU5 415 Colletotrichum fructicola (strain Nara gc5) Ascomycota


(Anthracnose fungus) (Colletotrichum gloeosporioides
(strain Nara gc5))

A0A010RVP2 805 Colletotrichum fioriniae PJ7 Ascomycota

A0A1Q8RZ87 414 Colletotrichum chlorophyti Ascomycota

164
A0A8H3W5T5 415 Colletotrichum asianum Ascomycota

A0A3D8SRF7 419 Coleophoma cylindrospora Ascomycota

A0A8H5ZG49 419 Cochliobolus sativus (Common root rot and spot Ascomycota
blotch fungus) (Bipolaris sorokiniana)

M2UZ24 419 Cochliobolus heterostrophus (strain C5 / ATCC 48332 / Ascomycota


race O) (Southern corn leaf blight fungus) (Bipolaris
maydis)

E9D099 415 Coccidioides posadasii (strain RMSCC 757 / Silveira) Ascomycota


(Valley fever fungus)

J3KC18 415 Coccidioides immitis (strain RS) (Valley fever fungus) Ascomycota

A0A1Y1ZKL3 419 Clohesyomyces aquaticus Ascomycota

M1VYV3 415 Claviceps purpurea (strain 20.1) (Ergot fungus) Ascomycota


(Sphacelia segetum)

W9W7X7 418 Cladophialophora yegresii CBS 114405 Ascomycota

W9XCJ0 416 Cladophialophora psammophila CBS 110553 Ascomycota

A0A0D2AW14 437 Cladophialophora immunda Ascomycota

A0A1C1CVD9 476 Cladophialophora carrionii Ascomycota

A0A3M7N4L1 400 Chaetothyriales sp. CBS 132003 Ascomycota

165
G0S0H4 431 Chaetomium thermophilum (strain DSM 1495 / CBS Ascomycota
144.50 / IMI 039719)

Q2HAJ4 459 Chaetomium globosum (strain ATCC 6205 / CBS 148.51 Ascomycota
/ DSM 1962 / NBRC 6347 / NRRL 1970) (Soil fungus)

A0A2I0RV10 420 Cercospora zeina Ascomycota

A0A2S6BZ10 420 Cercospora berteroae Ascomycota

A0A0F8B793 416 Ceratocystis fimbriata f. sp. platani Ascomycota

W9YRV4 423 Capronia epimyces CBS 606.96 Ascomycota

W9ZHA7 423 Capronia coronata CBS 617.96 Ascomycota

A0A8H7WGP5 416 Cadophora malorum Ascomycota

A0A443HQM9 415 Byssochlamys spectabilis (Paecilomyces variotii) Ascomycota

A0A4Z1KSA5 420 Botrytis porri Ascomycota

A0A8H6ECL4 415 Botrytis fragariae Ascomycota

A0A4Z1HMR2 419 Botryotinia narcissicola Ascomycota

A0A384JW64 419 Botryotinia fuckeliana (strain B05.10) (Noble rot Ascomycota


fungus) (Botrytis cinerea)

A0A4Y8CXK7 407 Botryotinia calthae Ascomycota

R1GMM0 424 Botryosphaeria parva (strain UCR-NP2) (Grapevine Ascomycota


canker fungus) (Neofusicoccum parvum)

166
A0A8H4IQ47 425 Botryosphaeria dothidea Ascomycota

A0A0H1BBD6 429 Blastomyces silverae Ascomycota

A0A1J9R0J2 433 Blastomyces percursus Ascomycota

A0A2B7WQB0 422 Blastomyces parvus Ascomycota

W6Y338 419 Bipolaris zeicola 26-R-13 Ascomycota

J4VTG6 427 Beauveria bassiana (strain ARSEF 2860) (White Ascomycota


muscardine disease fungus) (Tritirachium shiotae)

M2NIJ1 422 Baudoinia panamericana (strain UAMH 10762) (Angels' Ascomycota


share fungus) (Baudoinia compniacensis (strain UAMH
10762))

A0A074YGF3 421 Aureobasidium subglaciale (strain EXF-2481) Ascomycota


(Aureobasidium pullulans var. subglaciale)

A0A074Y071 421 Aureobasidium pullulans EXF-150 Ascomycota

A0A074WH41 421 Aureobasidium namibiae CBS 147.97 Ascomycota

A0A074W1K0 421 Aureobasidium melanogenum CBS 110374 Ascomycota

A0A1L9RW86 411 Aspergillus wentii DTO 134E9 Ascomycota

A0A3F3PUF1 417 Aspergillus welwitschiae Ascomycota

167
A0A2V5HGC9 409 Aspergillus violaceofuscus (strain CBS 115571) Ascomycota

A0A1L9PG63 419 Aspergillus versicolor CBS 583.65 Ascomycota

A0A319BZ17 421 Aspergillus uvarum CBS 121591 Ascomycota

A0A8E0QMC4 416 Aspergillus udagawae Ascomycota

A0A229Z1T7 416 Aspergillus turcosus Ascomycota

A0A397GCA0 415 Aspergillus thermomutatus Ascomycota

A0A5M3YT85 415 Aspergillus terreus Ascomycota

A0A5N6UBN6 433 Aspergillus tamarii Ascomycota

A0A1L9TNW8 422 Aspergillus sydowii CBS 593.65 Ascomycota

A0A2I2GJ85 421 Aspergillus steynii IBT 23096 Ascomycota

A0A317WE84 448 Aspergillus sclerotioniger CBS 115572 Ascomycota

A0A319FM33 421 Aspergillus sclerotiicarbonarius (strain CBS 121057 / IBT Ascomycota


28362)

A0A3A2ZKT8 415 Aspergillus sclerotialis Ascomycota

A0A318ZCW3 421 Aspergillus saccharolyticus JOP 1030-1 Ascomycota

A0A017SCU8 415 Aspergillus ruber CBS 135680 Ascomycota

A0A0F8XMD7 425 Aspergillus rambellii Ascomycota

A0A7R8AKY2 419 Aspergillus puulaauensis Ascomycota

168
A0A5N6SME9 430 Aspergillus pseudotamarii Ascomycota

A0A5N6HTP9 432 Aspergillus pseudonomiae Ascomycota

A0A0F0IGJ5 417 Aspergillus parasiticus (strain ATCC 56775 / NRRL 5862 Ascomycota
/ SRRC 143 / SU-1)

A0A1S9DMN7 420 Aspergillus oryzae (Yellow koji mold) Ascomycota

A0A5N6E964 433 Aspergillus novoparasiticus Ascomycota

A0A2I1C393 423 Aspergillus novofumigatus (strain IBT 16806) Ascomycota

A0A0L1J511 441 Aspergillus nomiae NRRL 13137 Ascomycota

A0A100I4C4 417 Aspergillus niger Ascomycota

A0A3D8SJ69 420 Aspergillus mulundensis Ascomycota

A0A5N6IJR6 433 Aspergillus minisclerotigenes Ascomycota

A0A5N5WUW1 433 Aspergillus leporis Ascomycota

A0A7R7WT65 417 Aspergillus kawachii (White koji mold) (Aspergillus Ascomycota


awamori var. kawachi)

A0A395H5T0 420 Aspergillus ibericus CBS 121593 Ascomycota

A0A395I5J3 421 Aspergillus homomorphus (strain CBS 101889) Ascomycota

A0A8H6Q8F1 416 Aspergillus hiratsukae Ascomycota

A0A317X1F7 424 Aspergillus heteromorphus CBS 117.55 Ascomycota

169
A0A1L9VNY8 415 Aspergillus glaucus CBS 516.65 Ascomycota

A0A8H4H1M9 416 Aspergillus fumigatiaffinis Ascomycota

A0A8H6QN79 416 Aspergillus felis Ascomycota

A0A319DPW6 416 Aspergillus ellipticus CBS 707.79 Ascomycota

A0A1E3BFQ1 1383 Aspergillus cristatus Ascomycota

A0A5N6YXL9 418 Aspergillus coremiiformis Ascomycota

A1CKH9 416 Aspergillus clavatus (strain ATCC 1007 / CBS 513.65 / Ascomycota
DSM 816 / NCTC 3887 / NRRL 1 / QM 1276 / 107)

A0A7R7ZLY7 414 Aspergillus chevalieri (Eurotium chevalieri) Ascomycota

A0A1R3RZ98 421 Aspergillus carbonarius (strain ITEM 5010) Ascomycota

A0A2I2F688 421 Aspergillus candidus Ascomycota

A0A0U5CDX5 415 Aspergillus calidoustus Ascomycota

A0A5N6ZP99 433 Aspergillus caelatus Ascomycota

A0A8H6ED23 420 Aspergillus burnettii Ascomycota

A0A1L9UX41 417 Aspergillus brasiliensis (strain CBS 101740 / IMI 381727 / Ascomycota
IBT 21946)

A0A1F7ZS77 420 Aspergillus bombycis Ascomycota

A0A5N7BLA0 417 Aspergillus bertholletiae Ascomycota

170
A0A401KGW0 417 Aspergillus awamori (Black koji mold) Ascomycota

A0A5N6TUU5 420 Aspergillus avenaceus Ascomycota

A0A2G7G6G2 433 Aspergillus arachidicola Ascomycota

A0A1L9X9V3 421 Aspergillus aculeatus (strain ATCC 16872 / CBS 172.66 / Ascomycota
WB 5094)

A0A168CDI7 1026 Ascosphaera apis ARSEF 7405 Ascomycota

A0A4S2N8C9 441 Ascodesmis nigricans Ascomycota

A0A8H7IVU4 419 Ascochyta lentis Ascomycota

A0A3N4I9P2 458 Ascobolus immersus RN42 Ascomycota

C5FE50 425 Arthroderma otae (strain ATCC MYA-4605 / CBS Ascomycota


113480) (Microsporum canis)

E5R3R6 421 Arthroderma gypseum (strain ATCC MYA-4604 / CBS Ascomycota


118893) (Microsporum gypseum)

D4AN76 428 Arthroderma benhamiae (strain ATCC MYA-4681 / CBS Ascomycota


112371) (Trichophyton mentagrophytes)

A0A436ZWC8 435 Arthrobotrys flagrans Ascomycota

A0A2T3B3F1 416 Amorphotheca resinae ATCC 22711 Ascomycota

A0A8H7B9U5 419 Alternaria burnsii Ascomycota

171
A0A177E2D3 419 Alternaria alternata (Alternaria rot fungus) (Torula Ascomycota
alternata)

A0A8H3II19 409 Alectoria fallacina Ascomycota

A0A168AV52 416 Akanthomyces lecanii RCEF 1005 Ascomycota

A0A086T8T6 428 Acremonium chrysogenum (strain ATCC 11550 / CBS Ascomycota


779.69 / DSM 880 / IAM 14645 / JCM 23072 / IMI 49137)

A0A2H3JKU9 438 Wolfiporia cocos (strain MD-104) (Brown rot fungus) Basidiomycota

A0A4T0SWP1 1142 Wallemia mellicola Basidiomycota

A0A4T0JHR1 2038 Wallemia ichthyophaga Basidiomycota

A0A4T0FS99 2528 Wallemia hederae Basidiomycota

A0A0D1DVK7 488 Ustilago maydis (strain 521 / FGSC 9021) (Corn smut Basidiomycota
fungus)

I2FWP7 493 Ustilago hordei (strain Uh4875-4) (Barley covered smut Basidiomycota
fungus)

A0A8H6HDF1 457 Tulosesus angulatus Basidiomycota

A0A0C3Q271 434 Tulasnella calospora MUT 4182 Basidiomycota

K1W6E2 446 Trichosporon asahii var. asahii (strain CBS 8904) Basidiomycota
(Yeast)

A0A4Q1BIA8 428 Tremella mesenterica (Jelly fungus) Basidiomycota

A0A1Y2J0G5 449 Trametes coccinea BRFM310 Basidiomycota

172
A0A066VW28 475 Tilletiaria anomala (strain ATCC 24038 / CBS 436.72 / Basidiomycota
UBC 951)

A0A177ULE8 608 Tilletia walkeri Basidiomycota

A0A0B7G051 446 Thanatephorus cucumeris (strain AG1-IB / isolate Basidiomycota


7/3/14) (Lettuce bottom rot fungus) (Rhizoctonia solani)

A0A0D0BEH4 431 Suillus luteus UH-Slu-Lm8-n1 Basidiomycota

A0A0C9UJU9 434 Sphaerobolus stellatus (strain SS14) Basidiomycota

A0A401H0V8 455 Sparassis crispa Basidiomycota

A0A166EG66 428 Sistotremastrum suecicum HHB10207 ss-3 Basidiomycota

F8PZ61 424 Serpula lacrymans var. lacrymans (strain S7.3) (Dry rot Basidiomycota
fungus)

A0A0C3AK39 403 Serendipita vermifera MAFF 305830 Basidiomycota

G4TD17 411 Serendipita indica (strain DSM 11827) (Root endophyte Basidiomycota
fungus) (Piriformospora indica)

A0A0C2ZWQ1 460 Scleroderma citrinum Foug A Basidiomycota

A0A0H2RWH9 442 Schizopora paradoxa Basidiomycota

D8Q9B6 411 Schizophyllum commune (strain H4-8 / FGSC 9210) Basidiomycota


(Split gill fungus)

A0A427YIM2 430 Saitozyma podzolica Basidiomycota

A0A4Y7PPQ7 435 Rickenella mellea Basidiomycota

173
A0A2S5BFW5 445 Rhodotorula taiwanensis Basidiomycota

A0A5C5FVE3 464 Rhodotorula diobovata Basidiomycota

A0A0K3C947 445 Rhodosporidium toruloides (Yeast) (Rhodotorula Basidiomycota


gracilis)

A0A1B7NE26 432 Rhizopogon vinicolor AM-OR11-026 Basidiomycota

A0A1J8R3R8 431 Rhizopogon vesiculosus Basidiomycota

A0A074S3V5 445 Rhizoctonia solani 123E Basidiomycota

A0A060SBW5 449 Pycnoporus cinnabarinus (Cinnabar-red polypore) Basidiomycota


(Trametes cinnabarina)

A0A5C3QLT6 511 Pterula gracilis Basidiomycota

A0A409WF51 469 Psilocybe cyanescens Basidiomycota

A0A8H7Y0P0 451 Psilocybe cubensis (Psychedelic mushroom) Basidiomycota


(Stropharia cubensis)

A0A5C3F5Y1 508 Pseudozyma flocculosa Basidiomycota

A0A1X6MN41 428 Postia placenta MAD-698-R-SB12 Basidiomycota

A0A371CPC3 451 Polyporus brumalis Basidiomycota

A0A5C3PIC0 448 Polyporus arcularius HHB13444 Basidiomycota

A0A067NK60 456 Pleurotus ostreatus PC15 Basidiomycota

A0A0C3PYC9 456 Pisolithus tinctorius Marx 270 Basidiomycota

174
A0A0D0A9Q1 448 Pisolithus microcarpus 441 Basidiomycota

A0A0C3BD59 441 Piloderma croceum (strain F 1598) Basidiomycota

A0A0C3RZX5 449 Phlebiopsis gigantea 11061_1 CR5-6 Basidiomycota

A0A4S4LAC0 427 Phellinidium pouzarii Basidiomycota

K5WFK2 453 Phanerochaete carnosa (strain HHB-10118-sp) (White- Basidiomycota


rot fungus) (Peniophora carnosa)

A0A165ZJB9 428 Peniophora sp. CONT Basidiomycota

A0A0D0EC35 436 Paxillus rubicundulus Ve08.2h10 Basidiomycota

A0A0C9U4C7 430 Paxillus involutus ATCC 200175 Basidiomycota

A0A409VE85 446 Panaeolus cyanescens Basidiomycota

A0A8E2J5T2 465 Obba rivulosa Basidiomycota

A0A8H3TY10 438 Naganishia liquefaciens Basidiomycota

A0A1Y2B794 430 Naematelia encephala Basidiomycota

A0A8H7CAS0 423 Mycena venus Basidiomycota

A0A8H6ZFB5 401 Mycena sanguinolenta Basidiomycota

A0A8H6U2C7 856 Mycena kentingensis (nom. inval.) Basidiomycota

A0A8H6WH39 877 Mycena indigotica Basidiomycota

A0A8H6WGQ6 434 Mycena chlorophos (Agaric fungus) (Agaricus Basidiomycota


chlorophos)

A0A0W0EXN9 454 Moniliophthora roreri Basidiomycota

175
A0A2X0NU47 454 Microbotryum silenes-dioicae Basidiomycota

A0A2X0LUS4 457 Microbotryum saponariae Basidiomycota

U5HFH2 454 Microbotryum lychnidis-dioicae (strain p1A1 Lamole / Basidiomycota


MvSl-1064) (Anther smut fungus)

A0A238FFF7 454 Microbotryum intermedium Basidiomycota

A0A1Y2G1A2 441 Leucosporidium creatinivorum Basidiomycota

A0A8H5LJK8 423 Leucoagaricus leucothites Basidiomycota

A0A5C2SCR8 449 Lentinus tigrinus ALCF2SS1-6 Basidiomycota

A0A5M6C8X4 426 Kwoniella shandongensis Basidiomycota

A0A1B9I896 432 Kwoniella pini CBS 10737 Basidiomycota

A0A1B9INM5 434 Kwoniella mangroviensis CBS 10435 Basidiomycota

A0A1B9GVS7 430 Kwoniella heveanensis BCC8398 Basidiomycota

A0A1A6A2N3 432 Kwoniella dejecticola CBS 10117 Basidiomycota

A0A1B9G008 434 Kwoniella bestiolae CBS 10118 Basidiomycota

A0A1Y1UJC5 430 Kockovaella imperatae Basidiomycota

A0A369JYC4 455 Hypsizygus marmoreus (White beech mushroom) Basidiomycota


(Agaricus marmoreus)

A0A0C9W6X7 432 Hydnomerulius pinastri MD-312 Basidiomycota

176
A0A2R6RQ14 456 Hermanssonia centrifuga Basidiomycota

A0A4Y9ZXK1 445 Hericium alpestre Basidiomycota

A0A5C3MMS3 439 Heliocybe sulcata Basidiomycota

A0A409XZM8 440 Gymnopilus dilepis Basidiomycota

S7RN70 438 Gloeophyllum trabeum (strain ATCC 11539 / FP-39264 / Basidiomycota


Madison 617) (Brown rot fungus)

A0A2G8S8E1 452 Ganoderma sinense ZZ0214-1 Basidiomycota

A0A067TGK8 446 Galerina marginata (strain CBS 339.88) Basidiomycota

A0A4Y9YDG2 452 Fomitopsis rosea Basidiomycota

S8E9U1 448 Fomitopsis pinicola (strain FP-58527) (Brown rot Basidiomycota


fungus)

J4I8V7 455 Fibroporia radiculosa Basidiomycota

A0A165JH73 433 Exidia glandulosa HHB12029 Basidiomycota

A0A4Q9NLQ0 437 Dichomitus squalens Basidiomycota

A0A4S8MBV0 448 Dendrothele bispora (strain CBS 962.96) Basidiomycota

M5G1H6 450 Dacryopinax primogenitus (strain DJM 731) (Brown rot Basidiomycota
fungus)

A0A0D7BPQ1 429 Cylindrobasidium torrendii FP15055 ss-10 Basidiomycota

A0A0J0XQ60 432 Cutaneotrichosporon oleaginosum Basidiomycota

177
P0CL98 430 Cryptococcus neoformans var. neoformans serotype D Basidiomycota
(strain JEC21 / ATCC MYA-565) (Filobasidiella
neoformans)

A0A0D0VDK0 430 Cryptococcus gattii CA1280 Basidiomycota

A0A1E3I9K5 430 Cryptococcus depauperatus CBS 7841 Basidiomycota

A0A1E3HDW9 431 Cryptococcus amylolentus CBS 6039 Basidiomycota

A0A5C3M3Y8 452 Crucibulum laeve Basidiomycota

A0A5C3L588 452 Coprinopsis marcescibilis Basidiomycota

A8NT51 470 Coprinopsis cinerea (strain Okayama-7 / 130 / ATCC Basidiomycota


MYA-4618 / FGSC 9003) (Inky cap fungus)
(Hormographiella aspergillata)

A0A4Y7STS6 455 Coprinellus micaceus Basidiomycota

M2R586 466 Ceriporiopsis subvermispora (strain B) (White-rot Basidiomycota


fungus) (Gelatoporia subvermispora)

A0A5N5QWH6 443 Ceratobasidium theobromae Basidiomycota

A0A8H8IUK0 439 Ceratobasidium sp. AG-Ba Basidiomycota

A0A4Q2DE89 442 Candolleomyces aberdarensis Basidiomycota

A0A165DAE4 450 Calocera cornea HHB12733 Basidiomycota

A0A067N0D0 449 Botryobasidium botryosum FD-172 SS1 Basidiomycota

178
A0A4S4LYZ9 440 Bondarzewia mesenterica Basidiomycota

A0A8I2YLL0 432 Boletus reticuloceps Basidiomycota

A0A550CYV4 472 Auriculariopsis ampla Basidiomycota

A0A8H7FRU8 425 Athelia sp. TMB Basidiomycota

A0A2H3BXR6 442 Armillaria solidipes Basidiomycota

A0A284R5K4 426 Armillaria ostoyae (Armillaria root rot fungus) Basidiomycota

A0A2H3E7I4 427 Armillaria gallica (Bulbous honey fungus) (Armillaria Basidiomycota


bulbosa)

A0A427YBS6 437 Apiotrichum porosum Basidiomycota

A0A4S4MQJ6 505 Antrodiella citrinella Basidiomycota

A0A2A9NSX0 423 Amanita thiersii Skay4041 Basidiomycota

A0A0C2T4V5 425 Amanita muscaria (strain Koide BX008) Basidiomycota

A0A8H4QNH9 449 Agrocybe pediades Basidiomycota

A0A8H7PVX7 399 Umbelopsis vinacea Mucor

A0A507BZE9 906 Synchytrium microbalum Chytrid

A0A507D9N5 454 Synchytrium endobioticum Chytrid

A0A0L0HQX9 430 Spizellomyces punctatus (strain DAOM BR117) Chytrid

A0A015K194 441 Rhizophagus irregularis (strain DAOM 197198w) Mucor


(Glomus intraradices)

A0A2Z6QKJ3 437 Rhizophagus clarus Mucor

179
A0A507DYC3 406 Powellomyces hirtus Chytrid

A0A8H7ZQU2 507 Olpidium bornovanus Olpidiomycota

A0A8H7U6M7 399 Mortierella isabellina (Filamentous fungus) Mucor


(Umbelopsis isabellina)

A0A068RX46 399 Lichtheimia corymbifera JMRC:FSU:9682 Mucor

A0A433D2I3 419 Jimgerdemannia flammicorona Mucor

A0A397TCR1 440 Glomus cerebriforme Mucor

A0A433PL30 420 Endogone sp. FLAS-F59071 Mucor

A0A261XTH6 415 Bifiguratus adelaidae Mucor

A0A1Y1XRM4 417 Basidiobolus meristosporus CBS 931.73 Zoopagomycota

A0A553PHG5 403 Tigriopus californicus (Marine copepod) Arthropod

A0A8B7NMF7 294 Hyalella azteca (Amphipod) Arthropod

A0A7M5WS78 431 Clytia hemisphaerica Cnidaria

E3NU68 417 Caenorhabditis remanei (Caenorhabditis vulgaris) Nematode

A0A0L0DCK8 1513 Thecamonas trahens ATCC 50062 Zooflagellate

A0A5J4Z201 483 Porphyridium purpureum (Red alga) (Porphyridium Rhodophyte


cruentum)

180
L1JQA4 435 Guillardia theta (strain CCMP2712) (Cryptophyte) Cryptophyte

R1DK22 404 Emiliania huxleyi (Coccolithophore) (Pontosphaera Coccolithophore


huxleyi)

A0A1V9ZFZ7 412 Thraustotheca clavata Stramenopiles

A0A812R312 959 Symbiodinium pilosum (Dinoflagellate) Alveolata

A0A812PTZ6 1053 Symbiodinium necroappetens Alveolata

A0A1Q9EJJ4 1092 Symbiodinium microadriaticum (Dinoflagellate) Alveolata


(Zooxanthella microadriatica)

A0A067CR29 414 Saprolegnia parasitica (strain CBS 223.65) Stramenopiles

T0PVD3 414 Saprolegnia diclina (strain VS20) Stramenopiles

A0A5D6Y223 433 Pythium brassicum Stramenopiles

A0A0P1AMX3 427 Plasmopara halstedii (Downy mildew of sunflower) Stramenopiles

G4ZDQ9 429 Phytophthora sojae (strain P6497) (Soybean stem and Stramenopiles
root rot agent) (Phytophthora megasperma f. sp.
glycines)

A0A6A3N222 429 Phytophthora rubi Stramenopiles

H3G8B6 431 Phytophthora ramorum (Sudden oak death agent) Stramenopiles

V9EEB5 428 Phytophthora parasitica P1569 Stramenopiles

A0A2P4WYF8 428 Phytophthora palmivora var. palmivora Stramenopiles

181
A0A0W8E017 428 Phytophthora nicotianae (Buckeye rot agent) Stramenopiles

A0A225V054 425 Phytophthora megakarya Stramenopiles

A0A3R7KB39 776 Phytophthora kernoviae Stramenopiles

A0A833TB11 428 Phytophthora infestans (Potato late blight agent) Stramenopiles


(Botrytis infestans)

A0A6A4E9J1 429 Phytophthora fragariae Stramenopiles

A0A329S542 428 Phytophthora cactorum Stramenopiles

B7G0U3 401 Phaeodactylum tricornutum (strain CCAP 1055/1) Stramenopiles

A0A3M6VFC5 428 Peronospora effusa Stramenopiles

A0A7J6NHW6 449 Perkinsus olseni (Perkinsus atlanticus) Alveolata

C5KJY4 436 Perkinsus marinus (strain ATCC 50983 / TXsc) Alveolata

A0A7J6MWY0 434 Perkinsus chesapeaki (Clam parasite) (Perkinsus Alveolata


andrewsi)

A0A662Y1L8 432 Nothophytophthora sp. Chile5 Stramenopiles

M4BZ04 434 Hyaloperonospora arabidopsidis (strain Emoy2) Stramenopiles


(Downy mildew agent) (Peronospora arabidopsidis)

K3WFJ6 435 Globisporangium ultimum (strain ATCC 200006 / CBS Stramenopiles


805.95 / DAOM BR144) (Pythium ultimum)

182
A0A484DVB3 434 Bremia lactucae (Lettuce downy mildew) Stramenopiles

A0A485LS90 500 Aphanomyces stellatus Stramenopiles

A0A024URZ4 406 Aphanomyces invadans Stramenopiles

A0A6G0XF49 407 Aphanomyces euteiches Stramenopiles

W4HBB2 406 Aphanomyces astaci Stramenopiles

A0A024GN63 2399 Albugo candida Stramenopiles

A0A1V9ZUK1 411 Achlya hypogyna Stramenopiles

D8TQ81 452 Volvox carteri f. nagariensis Chlorophyte

A0A383VQY2 414 Tetradesmus obliquus (Green alga) (Acutodesmus Chlorophyte


obliquus)

A0A7J7Q2V3 415 Scenedesmus sp. NREL 46B-D3 Chlorophyte

A0A2V0NTE8 416 Raphidocelis subcapitata Chlorophyte

A9TTS6 407 Physcomitrium patens (Spreading-leaved earth moss) Streptophyte


(Physcomitrella patens)

A0A2P6VED2 885 Micractinium conductrix Chlorophyte

A0A176VRW3 407 Marchantia polymorpha subsp. ruderalis Streptophyte

I0Z3H5 408 Coccomyxa subellipsoidea (strain C-169) (Green Chlorophyte


microalga)

A0A2P6TRH5 2189 Chlorella sorokiniana (Freshwater green alga) Chlorophyte

183
A0A2K3CPG8 414 Chlamydomonas reinhardtii (Chlamydomonas smithii) Chlorophyte

A0A835SSY5 431 Chlamydomonas incerta Chlorophyte

A0A388JX09 406 Chara braunii (Braun's stonewort) Streptophyte

A0A5N6KXJ9 421 Carpinus fangiana Dicot

A0A087SL08 406 Auxenochlorella protothecoides (Green microalga) Chlorophyte


(Chlorella protothecoides)

184

You might also like