DIAGNOSTICS, GENOMICS AND POPULATION STUDIES ON oxysporum FORMAE SPECIALES ASSOCIATED WITH ORNAMENTAL PALMS

By

SUSHMA V. PONUKUMATI

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2017

© 2017 Sushma V. Ponukumati

To my family for their unconditional love and support

ACKNOWLEDGMENTS

I would like to express my gratitude to Dr. Monica Elliott, my advisor, for supporting me in my research in all possible ways. She always encouraged me to be a passionate, independent researcher and gave me freedom to explore new research fields. She helped me to develop the habits of logical thinking and proper planning. Dr.Elliott selfless time and care were sometimes all that kept me going during the writing of this dissertation.

I would like to thank Dr. Jeffrey Rollins for inducing passion towards bioinfomatics and helping me understand the importance of critical thinking. I am also appreciative to Dr. Rollins for his constant guidance and dedication in his role as co-advisor.

Profound gratitiude goes to my other Ph.D committee members, Dr. Mathew Smith and

Dr. Robin Giblin Davis, for serving on my committee and for providing their time and expertise.

Their advice and recommendations were valuable resources for my dissertation. I will always cherish Dr. Smith for his inspirational chats. Dr.Smith has been a good role model in teaching, who encouraged and expected students to think independently with a purpose regarding any experiment or project. I am also thankful to Dr. Robin Giblin Davis for his constant support and motivation.

My profound gratitude goes to Dr. Jose C. Huguet Tapia who took extreme care and time to teach me bioinformatics. He was always available to trouble shoot and advise me regarding any bioinformatics related questions during the dissertation.

A special mention goes to Ms. Beth Des Jardin for providing me a good support system in Dr. Elliott’s lab. She encouraged me and gave fantastic plant pathology skills training.

4

I would like to extend thanks to many people (friends, Dr. Elliott’s interns and co- graduate students) both at Fort Lauderdale Research and Extension Center and Plant Pathology

Department in Gainesville who have supported me by their friendship and their advice.

I would like to thank my husband Rajasekhar Ponukumati for his unconditional love and support; without it, this dissertation would not be possible. Finally, I express my gratitude to my parents and in-laws, my brother and sister for their belief in me and for their unending love and encouragement.

5

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS ...... 4

LIST OF TABLES ...... 9

LIST OF FIGURES ...... 11

ABSTRACT ...... 13

CHAPTER

1 LITERATURE REVIEW ...... 15

Introduction on Palms ...... 15 Fusarium oxysporum ...... 16 Fusarium Wilt Pathogens in Palms ...... 18 Fusarium oxysporum Pathogenesis ...... 23 Cell wall degrading enzymes ...... 24 Signaling proteins ...... 24 Transcriptional factors ...... 25 Blocking host defenses ...... 25 Fusarium oxysporum Genomic Studies ...... 26 Secreted in Xylem Gene Family ...... 28 Objectives ...... 30

2 DIAGNOSTIC PRIMER PAIR TO IDENTIFY FUSARIUM WILT PATHOGENS ON ORNAMENTAL PALMS IN FLORIDA ...... 32

Introduction ...... 32 Material and Methods ...... 35 Isolates ...... 35 Genomic DNA and EF-1α Sequencing ...... 36 Primer Design ...... 36 Genomic DNA Amplification with PFW Primer Pair ...... 37

6

DNA Amplification with Agar Plug and PFW Primer Pair ...... 38 DNA Amplification with the FOA Primer Pair ...... 39 Phylogenetic Analysis ...... 39 Results...... 40 Genomic DNA and EF-1α Sequencing ...... 40 Primer Design ...... 40 Genomic DNA Amplification with PFW Primer Pair ...... 40 DNA Amplification with Agar Plug and PFW Primer Pair ...... 41 DNA Amplification with FOA Primer Pair ...... 41 Phylogenetic Analysis ...... 41 Discussion ...... 43

3 GENOME CHARACTERIZATION OF Fusarium oxysporum f. sp. palmarum ...... 61

Introduction ...... 61 Material and Methods ...... 66 Isolates ...... 66 DNA Extraction and Whole Genome Sequencing ...... 66 Genome Sequencing ...... 67 Benchmarking Universal Single Copy Orthologs ...... 68 Comparative Genomics to Identify Orthologous Gene Families ...... 68 Identification of Secretory Proteins ...... 69 Detection of Carbohydrate Active Enzymes ...... 69 Detection of Secondary Metabolite Clusters ...... 70 Secreted in Xylem (SIX) Genes in FOC and FOP ...... 70 Detection of Virulence Genes in FOP and FOC ...... 71 Results...... 71 Genome Sequencing ...... 71 Benchmarking Universal Single Copy Orthologs ...... 72 Proteome Analysis of FOC and FOP ...... 72 Identifying Orthologous Families among Fusarium Genomes ...... 73

7

Secretome Analysis ...... 74 Carbohydrate Active Enzymes of FOC and FOP ...... 75 Detection of Secondary Metabolite Clusters ...... 76 Secreted in Xylem (SIX) Gene Family ...... 78 Virulence Genes ...... 78 Discussion ...... 79

4 SCREENING FOR SECRETED IN XYLEM (SIX) GENES IN FUSARIUM WILT PATHOGENS OF ORNAMENTAL PALMS ...... 108

Introduction ...... 108 Material and Methods ...... 112 Isolates ...... 112 DNA Extraction and Quality Assessment ...... 112 Molecular Characterization by EF-1α Gene ...... 113 Screening of SIX Genes from Assembled Genomes ...... 113 Detection of SIX Genes by PCR ...... 113 Phylogenetic Analysis ...... 114 Results...... 115 Molecular Characterization by EF-1α Gene ...... 115 Screening of SIX Genes from Assembled Genomes ...... 116 Detection of SIX Genes by PCR ...... 116 SIX1 ...... 116 SIX7 ...... 117 SIX10 ...... 117 SIX12 ...... 118 SIX8 and SIX9 ...... 119 Discussion ...... 119

5 RESEARCH SUMMARY ...... 145

LIST OF REFERENCES ...... 148

BIOGRAPHICAL SKETCH ...... 163

8

LIST OF TABLES

Table page

2-1 Fusarium isolates used in the study, indicating isolate, palm host, geographic location and tissue sampled. Genomic DNA or a fungal colonized agar plug was used as template for PCR using PFW primers and three different polymerases...... 48

3-1 List of the Fusarium genomes used in the study...... 90

3-2 Genomic statistics of F. oxysporum f. sp. canariensis and F. oxysporum f. sp. palmarum with selected Fusarium oxysporum formae speciales...... 91

3-3 Comparison of functional annotations among Fusarium oxysporum f. sp. canariensis and Fusarium oxysporum f. sp. palmarum...... 92

3-4 Important proteins grouped by Pfam domain/families based on Fusarium oxysporum f.sp. canariensis (FOC) and Fusarium oxysporum f. sp. palmarum (FOP) predicted proteomes...... 93

3-5 Comparison of important PFAM domains/families between Fusarium oxysporum f. sp. canariensis (FOC) and Fusarium oxysporum f. sp.palmarum (FOP)...... 94

3-6 The common Pfam domains that are present in the secretome of Fusarium oxysporum f. sp. palmarum...... 95

3-7 Fusarium oxysporum f. sp. palmarum expanded gene clusters based on orthoMCL comparison with 15 Fusarium genomes...... 96

3-8 Expanded gene families of Fusarium oxysporum f. sp. palmarum with genes and their Pfam domain description...... 97

3-9 Carbohydrate active enzymes (CAZymes) families involved in cellulose, hemicellulose and pectin degradation in Fusarium genomes...... 99

3-10 Description of Fusarium oxysporum f. sp. palmarum secondary metabolite gene clusters and its orthologous genes present Fusarium oxysporum genomes that were studied by ORTHOMCL...... 100

3-11 Secreted in Xylem (SIX) genes recovered from Fusarium. oxysporum f. sp. canariensis and F. oxysporum f. sp. palmarum genome assemblies using multiple bioinformatics approaches. . 102

3-12 Putative virulence gene distribution in Fusarium oxysporum f. sp. canariensis and F. oxysporum f. sp palmarum predicted by PHI-Base...... 102

3-13 Putative virulence associated genes orthologs identified in predicted proteome of Fusarium oxysporum f. sp. palmarum by PHI-Base, classified based on role in infection process...... 103

9

4-1 Previous studies on prominent Fusarium oxysporum formae speciales plus F. foetens with their SIX gene profiles elucidated by various approaches...... 124

4-2 Isolates used in the study and PCR based presence or absence of Secreted in Xylem (SIX) genes...... 125

4-3 PCR primer pairs that were used in this study, with their annealing temperature and amplicon size...... 128

4-4 Secreted in Xylem (SIX) genes recovered from Fusarium oxysporum f. sp. canariensis and Fusarium oxysporum f. sp. palmarum draft genomes using multiple bioinformatics approaches...... 129

10

LIST OF FIGURES

Figure page

2-1 Comparison of genomic DNA amplification using three different polymerases with the PFW primer pair as viewed on a 1.5% agarose gel...... 54

2-2 Comparison of DNA amplification using mycelial agar plug as template for PCR with three different polymerases and PFW primer pair as viewed on a 1.5% agarose gel...... 55

2-3 Unrooted consensus tree inferred from 10 most parsimonious trees using the Maximum Parsimony method based on a portion of the translation elongation factor (EF-1α) gene...... 56

2-4 Unrooted consensus tree inferred from 10 most parsimonious trees using the Maximum Parsimony method based on the PFW primer...... 58

2-5 Flowchart of the proposed diagnostic method for identification of Fusarium wilt pathogens on ornamental palms...... 60

3-1 Distribution of Fusarium oxysporum f. sp.canariensis (FOC) and F.oxysporum f. sp. palmarum (FOP) proteomes into sub categories of Eukaryotic Orthologous Groups (KOG) gene families...... 105

3-2 The carbohydrate active enzymes (CAZymes) class distribution in Fusarium genomes...... 106

3-3 SMURF secondary metabolite (SM) backbone gene prediction of Fusarium proteomes used in this study...... 107

4-1 EF-1α based phylogenetic tree of Fusarium oxysporum f. sp. canariensis (FOC) and F. oxysporum f. sp. palmarum (FOP) including various formae speciales inferred by Maximum Likelihood...... 130

4-2 SIX1 based phylogenetic tree of Fusarium oxysporum f. sp. canariensis (FOC) by Maximum Likelihood based on the Kimura 2-parameter model...... 132

4-3 SIX7 based phylogenetic tree of Fusarium oxysporum f. sp. canariensis (FOC) by Maximum Likelihood method based on the Kimura 2-parameter model...... 134

4-4 SIX10 based phylogenetic tree of Fusarium oxysporum f. sp. canariensis (FOC) along with F.oxysporum f. sp.palmarum (FOP) using the Maximum Likelihood method based on the Jukes-Cantor model...... 135

4-5 SIX12 based phylogenetic tree of Fusarium oxysporum f. sp. canariensis (FOC) inferred by Maximum Likelihood based on the Jukes-Cantor model...... 136

4-6 SIX8 based phylogenetic tree of Fusarium oxysporum f. sp. palmarum (FOP) by Maximum Likelihood method based on the Kimura 2-parameter model...... 137

11

4-7 SIX9 based phylogenetic tree of Fusarium oxysporum f. sp. palmarum (FOP) inferred by Maximum Likelihood method based on the Kimura 2-parameter model. The tree with the highest log likelihood (-968.1485) is shown...... 138

4-8 Genomic DNA amplification of Fusarium oxysporum f. sp. canariensis (FOC) isolates using SIX1 primer pair as viewed on 1.5 % agarose gel stained with GelRed...... 139

4-9 Genomic DNA amplification of Fusarium oxysporum f. sp. canariensis (FOC) and F. oxysporum f.sp.palmarum (FOP) isolates using SIX7 primer pair as viewed on 1.5 % agarose gel stained with GelRed...... 140

4-10 Genomic DNA amplification of Fusarium oxysporum f. sp. palmarum (FOP) isolates using SIX8 primer pair as viewed on 1.5 % agarose gel stained with GelRed...... 141

4-11 Genomic DNA amplification of Fusarium oxysporum f. sp. palmarum (FOP) and F. oxysporum f.sp. canariensis (FOC) isolates using SIX9 primer pair as viewed on 1.5 % agarose gel stained with GelRed...... 142

4-12 Genomic DNA amplification of Fusarium oxysporum f. sp. canariensis (FOC) and F. oxysporum f. sp. palmarum (FOP) isolates using SIX10 primer pair as viewed on 1.5 % agarose gel stained with GelRed...... 143

4-13 Genomic DNA amplification of Fusarium oxysporum f. sp. canariensis (FOC) isolates using SIX12 primer pair as viewed on 1.5 % agarose gel stained with GelRed...... 144

12

Abstract of Dissertation Presented to the Graduate School of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

DIAGNOSTICS, GENOMICS AND POPULATION STUDIES ON Fusarium oxysporum FORMAE SPECIALES ASSOCIATED WITH ORNAMENTAL PALMS

By

Sushma V.Ponukumati

August 2017

Chair: Monica L. Elliott Cochair: Jeffrey A. Rollins Major: Plant Pathology

Fusarium wilt on ornamental palms in Florida is a lethal disease caused by Fusarium oxysporum f. sp. canariensis (FOC) and F. oxysporum f. sp. palmarum (FOP). FOP has a wider palm host range, causes an aggressive disease and is primarily localized to Florida, whereas FOC is found worldwide and restricted primarily to one palm species. The first objective was to design an efficient diagnostic method to differentiate palm Fusarium wilt pathogens from other

Fusarium species often associated with palms. A translation elongation factor based primer pair

(PFW) was designed. All FOP and FOC isolates, which were obtained from symptomatic palm petiole or rachis tissue, were amplified. No other Fusarium species were amplified with PFW primers. Results indicate that symptomatic petiole or rachis is the only tissue to use for isolation of palm Fusarium wilt pathogens. For the second objective, the FOP and FOC genomes were sequenced, assembled and annotated using multiple bioinformatic approaches to compare FOC and FOP proteomes to identify common and unique gene families. Results indicate FOP and

FOC share similar copies of conserved pathogenicity factors. FOP has more secondary

13

metabolite potential than FOC. At least three nonribosomal peptide synthetase (NRPS) and one polyketide synthase (PKS) clusters appear novel to FOP compared to FOC and other Fusarium genomes, suggesting the possible role of secondary metabolites in FOP pathogenicity. The third objective was screening for Secreted in Xylem (SIX) genes that produce small secretory proteins, which may play a role in virulence and in the infection process of Fusarium oxysporum formae speciales. Whole genome sequencing of FOC and FOP indicates the presence of SIX7, SIX10 and

SIX12 in FOC and SIX8 and SIX9 in FOP. A PCR based approach was used to screen 47 isolates of FOC and FOP, obtained from different palm hosts and geographical locations, to detect the presence of SIX genes with published primers. PCR results confirmed the genomic results but

SIX1 was also detected in FOC isolates. Insights about these formae speciales of F. oxysporum will enhance our understanding of host specific interactions and may lead to effective management strategies.

14

CHAPTER 1 LITERATURE REVIEW

Introduction on Palms

Carl Linnaeus described palms as “Principes–the princes of the plant kingdom” (Plotkin and Balick 1984). The family Arecaceae consists of approximately 2,600 palm species placed in more than 200 genera (Christenhusz and Byng 2016). Palms are grown in tropical, subtropical and Mediterranean climates with an extended life span averaging greater than 60 years (Meerow

1994; Tomlinson and Huggett 2012). Across the world, palms are grown for food, oil, fruits, wine and economic products such as rope, mats, tannin and roofing material (Elliott et al. 2004).

Within the continental USA, palms are primarily used for aesthetic purposes in commercial and residential landscapes.

Palms are monocots with adventitious roots and scattered vascular bundles in the stem that survive the complete life span of the palm. Palms have no vascular cambium, which means there is no secondary growth. Moreover, with a single apical meristem per stem, palms are vulnerable to any damage to the meristem, which can result in death. Palm leaves are some of the largest in the plant kingdom and can be either fan shape (palmate, costa-palmate), feather

(pinnate) or entire. All palm leaves are divided into blade, petiole and leaf base. There are four stages of palm development: seedling, juvenile (leaves but no above-ground stem), adult vegetative (above-ground stem but not mature enough to flower) and adult mature phase

(reproductive). Even though the palm flowers are tiny, they produce some of the largest fruits and seeds in the plant kingdom (Elliott et al. 2004).

In the USA, palms are extensively planted in Hawaii and southern states as ornamentals.

As of 2010, palms rank fourth in the list of nursery sales in Florida accounting for a $404 million

15

dollar industry (Khachatryan and Hodges 2012). Common palms that are grown in Florida include Bismarckia nobilis (Bismarckia palm), Butia odorato (pindo palm), x Butyagrus nabonnandi (mule palm), Cocos nucifera (coconut), Dypsis lutescens (areca palm), Phoenix canariensis (Canary Island date palm), Phoenix dactylifera (date palm), Phoenix reclinata

(Senegal date palm), Phoenix sylvestris (wild date palm), Ravenea rivularus (majesty palm),

Roystonea regia (royal palm), Sabal palmetto (cabbage palm), Syagrus romanzoffiana (queen palm), and Washingtonia robusta (Mexican fan palm) (Broschat and Black 2016). The average cost of palms in Florida ranges from $100 (areca palm) - $3800 (date palm)

(http://realpalmtrees.com). With few exceptions, the majority of these palms are not native to the

USA.

Similar to other plants, palms are subjected to various diseases, insects, nutritional, and physiological disorders. Palms are susceptible to diseases caused by a wide variety of organisms that include algae, bacteria, fungi, oomycetes, nematodes, phytoplasmas and viruses. Among these, phytoplasmas and fungi are more common in Florida. Common fungal diseases of palms in Florida include bud rot, Fusarium wilt of Canary Island date palm, Fusarium wilt of queen palm and Mexican fan palm, Ganoderma butt rot, Graphiola leaf spot, petiole blight and

Thielaviopsis trunk rot (Elliott et al. 2004).

Fusarium oxysporum

The genus Fusarium belongs to Domain Eukaryota, Kingdom Fungi, Phylum

Ascomycota, Subphylum Pezizimycotina, Class , Order , and

Family Nectriceae. H. F. Link first described the Fusarium genus in 1809 as Fusisporium based on banana shaped conidia (Gordon and Martyn 1997; Leslie and Summerell 2006). In 1935,

Wollenweber and Reinking grouped at least 1000 proposed Fusarium like species into 16 16

sections, comprising 65 species, 55 varieties and 22 forms based on culture characteristics in media and spore morphology (macroconidia, microconidia and chlamydospores) (Nelson et al.

1983). They published the first atlas, Die Fusarien, describing these 142 Fusarium-like species.

In 1940, Synder and Hansen consolidated 16 sections into nine species using single spore isolates. They further proposed 25 formae speciales based on the hosts they infect (Gordon and

Martyn 1997). Fusarium oxysporum, one of the nine species, was first described by D. F. L. von

Schlechtendahl in 1824, and later amended by Synder and Hansen in 1940 (Booth 1971; Nelson et al. 1983).

Fusarium oxysporum Schlechtend: Fr. is an asexual species complex comprising a wide variety of biological forms including biocontrol, endophytic, saprophytic and pathogenic isolates. The pathogenic forms affect animals, humans, insects and plants. The plant pathogenic strains are necrotrophs and affect a wide variety of economically important plants, causing either vascular wilt, bulb rot, crown rot or root rot (Michielse and Rep 2009). Fusarium oxysporum is ranked fifth in the list of most notorious plant pathogens (Dean et al. 2012). Plant pathogenic F. oxysporum strains are divided into formae speciales (f. sp singular; ff. spp. plural) based on host plants that they infect. Currently more than 120 formae speciales have been classified, and a few are further classified into races based on host lines or cultivars that they infect (Armstrong and

Armstrong 1981; Baayen et al. 2000). For example, the forma specialis that infects Canary Island date palm is named F. oxysporum f. sp. canariensis, the forma specialis that infects banana is named F. oxysporum f. sp. cubense and so forth.

Fusarium oxysporum has no known teleomorph and is considered to reproduce clonally.

It produces three types of asexual spores: microconidia, macroconidia and chlamydospores

(Nelson et al. 1983). Microconidia are usually one-celled and oval to elliptical in shape. 17

Macroconidia are produced abundantly on infected plants and consist of three to five-celled spores. Chlamydospores are thick-walled, survival structures that can survive in soil for at least

25 years (Downer et al. 2013). They can be produced terminally or intercalary. Morphological characterization of F. oxysporum include shape of macroconidia, microconidia and chlamydospores (Leslie and Summerell 2006).

Fusarium oxysporum vascular wilts are initated primarily as root infections, in which the chlamydospores present on plant debris or in soil germinate and mycelia penetrate into plant roots through natural openings or wounds. The mycelium advances into root cortex reaching xylem vessels through the pits. The mycelium is confined to xylem vessels, but continues to grow and produce microconidia. These microconidia are carried upward by xylem sap, where they germinate and invade more xylem vessels. By this time, plants are in severe water stress, resulting in wilt symptoms as the xylem vessels are clogged by mycelium, conidia, gels, gum and tyloses that are produced by plant defenses. Finally, the invades neighboring tissues of the plant and becomes exposed to the external environment. Newly produce micro and macrocondia may be dispersed to new plants by wind, water or insects (Agrios 2005).

Fusarium Wilt Pathogens in Palms

Currently, there are four Fusarium wilt diseases on palms, each with lethal consequences.

F. oxysporum f. sp. albedinis (FOA) causes date palm wilt on Phoenix dactylifera and is observed in Algeria, Morocco and Mauritania (EPPO 2003). FOA appears to be monophyletic with all isolates belonging to one vegetative compatibility group (Fernandez et al. 1998). Disease progression is slow, and it is usually spread by transport of infected plant material (EPPO 2003).

18

F. oxysporum f. sp. eleidis (FOE) causes oil palm wilt on Elaeis guineensis (African oil palm) and is reported from countries in Africa, and localized areas in Brazil and Ecuador (Flood

2006; Rusli 2012). FOE may be spread by contaminated seeds. FOE attacks oil palms at various growth stages and has two disease types: acute wilt, which is a fast-paced disease where palms die within 2-3 months, and chronic wilt, which is a slow, progressive decline of trees over years

(Flood 2006).

Canary Island date palm wilt, caused by F. oxysporum f. sp. canariensis (FOC), is present worldwide on Phoenix canariensis. It was first reported in France in 1973 (Mercier and

Louvet 1973), followed by the USA in 1976 (Feather et al. 1989), Japan in 1977 (Arai and

Yamamoto 1977), Spain in 1987 (Fernandez et al. 1997), Australia in 1990 (Priest and Letham

1996), Greece in 2002 (Elena 2005), Italy (Sardinia) in 2004 (Migheli et al. 2005) and Argentina in 2005 (Palmucci 2005). In the USA, FOC was first reported on Phoenix canariensis in

California in 1976 (Feather et al. 1989) and then spread to Florida in 1994 (Simone 2004). It has been reported in Nevada (Wang and McKie 2007), Louisiana (Singh et al. 2011) and South

Carolina and Texas (Elliott et al. 2011). FOC has also been documented on Phoenix reclinata in

Florida (Elliott 2015) and California (personal communication, California Department of Food and Agriculture).

Fusarium wilt of queen (Syagrus romanzoffiana) and Mexican fan (Washingtonia robusta) palms is caused by F. oxysporum f. sp. palmarum (FOP) and was first observed in

Florida (Elliott et al. 2010). Except for one report from a single location in Texas (Giesbrecht et al. 2013), this pathogen is not known anywhere else in the world. FOP has a wider host range than most F. oxysporum formae speciales, as it not only causes wilt diseases of queen and

Mexican fan palms, but also on Phoenix canariensis, and x Butyagrus nabonnandii (Elliott 2011; 19

Elliott et al. 2017). It has also been detected in wilt symptomatic specimens of Bismarckia nobilis and Phoenix reclinata in Florida (Elliott, personal communication).

Fusarium wilts caused by FOC and FOP share similarities, including symptoms and disease cycle, with an overlapping host range, and yet they are distinct diseases and pathogens.

With FOC and FOP, initial symptoms are usually observed in lower fronds and include typical one-sided wilt and necrosis of the leaflets. A brownish streak on the petiole and rachis is always associated with necrotic dead leaflets. Cross sections of the petiole show an internal discoloration that corresponds to the necrotic dead leaflets. Eventually the whole leaf dies and the process is repeated on the next oldest leaf. The primary difference between these two Fusarium wilt pathogens is disease progression. From the onset of symptoms, the disease progression caused by

FOC is slow, taking up to a year to two for palms to die. In the disease caused by FOP, progression is fast, killing the palms within weeks to months (Elliott et al. 2010). This suggests

FOP is a more aggressive pathogen than FOC. Compared to other formae speciales, FOC and

FOP are special cases in a few regards. Typically, a F. oxysporum forma specialis is host specific with one or a few hosts. However, FOC can infect at least two species within the genus Phoenix, and FOP can infect multiple genera within the Arecaceae family including Phoenix, Syagrus and

Washingtonia. FOA and FOE primarily infect palms via root infections, whereas, FOC and FOP primarily infect leaves, but can also infect via the roots.

Fusarium wilt of Canary Island date palm is spread by contaminated pruning tools, infested soil and dead infected plant material. FOP is assumed to be spread primarily by wind- blown spores (Elliott et al. 2010). There are no cures or treatment for either of the Fusarium wilts caused by FOP or FOC. Prevention is the only way to limit the spread of the disease, and early detection is the key to implement management strategies. Management strategies for Fusarium 20

wilt caused by both pathogens include use of clean and disinfected tools when pruning healthy or diseased palm trees, and removal of the diseased tree as quickly as possible. It is strongly suggested not to replant a susceptible palm species in the infested soils (Elliott et al. 2004).

Phylogenetic studies have revealed that isolates of different formae speciales are often more closely related than isolates within the same forma specialis (Kistler 1997; Lievens et al.

2008). Many formae speciales are polyphyletic in nature where isolates may have adapted to the same host by convergent evolution as seen in F. oxysporum f. sp. cubense and F. oxysporum f. sp. lycopersici (O’Donnell et al. 1998; Ma et al. 2013). A few formae specialis, such as F. oxysporum f. sp. loti and F. oxysporum f. sp. ciceris, are monophyletic with isolates of the same forma specialis all resolved within a clonal lineage (Jimenez-Gasco et al. 2002; Wunsch et al.

2009).

Vegetative compatibility grouping (VCG), polymerase chain reaction assays and DNA fingerprinting are routinely used to understand genetic diversity in the Fusarium oxysporum species complex (Baayen et al. 2000; Gunn and Summerell 2002). A few studies have explored genetic diversity among FOC, but only one study examined FOP isolates. Most FOC isolates

(71%) group into a single VCG group (0240), and mitochondrial DNA markers revealed FOC isolates belonged to a few haplotypes, implying the presence of low to moderate diversity and that FOC is monophyletic (Plyler et al. 2000). Translation elongation factor gene (EF-1α) phylogenetic analysis based on Australian and non-Australian isolates, including USA FOC isolates, indicated that Australian FOC isolates are more diverse than non-Australian isolates with at least three different lineages present, suggesting the Australian FOC are polyphyletic

(Gunn and Summerell 2002; Laurence et al. 2015). A study conducted by O’Donnell et al.

(2009), using two-locus DNA sequence typing, revealed the presence of 256 sequence types (ST) 21

among 850 F. oxysporum species complex isolates. The two loci were EF-1α and nuclear ribosomal intergenic spacer region (IGS). This robust study, which included five isolates of FOP obtained in Florida, revealed the presence of three sequence types in FOP (ST250, ST251 and

ST284) and two sequence types in FOC (ST41 and ST200). None of the FOC isolates were from

Florida. No correlation among sequence types, palm host or disease severity has been observed among FOP isolates.

The most reliable method to identify F. oxysporum formae speciales are through pathogenicity assay in which host plants are inoculated with the pathogens to observe disease development. Additionally, morphological and cultural characteristics and symptoms on infected host tissues are used to identify F. oxysporum (Baayen et al. 2000). However, F. oxysporum formae speciales are morphologically similar and pathogenicity assays are time consuming and laborious. Molecular methods provide an alternative to identify F.oxysporum formae speciales

(Lievens et al. 2008). Designing a diagnostic tool to identify F. oxysporum formae speciales is often challenging and time consuming due to their polyphyletic lineages (Plyler et al. 2000; Suga et al. 2013).

Diagnosis of Fusarium wilt caused by FOC and FOP include isolation of the pathogen from the infected palm tissue, observation of colony and spore morphology (microconidia and macroconidia) and molecular assays. For the latter, DNA is extracted from single-spored isolates. Molecular identification of FOC includes a positive PCR amplification using the

HK66/HK67 primer pair (Plyler et al. 1999). However, this primer pair routinely detects F. proliferatum, which is sometimes isolated from symptomatic palms (Viteroli 2008). Presently,

FOC and FOP are molecularly identified by PCR amplification using the primer pair ef1 and ef2, which amplifies a portion of the EF-1α (Geiser et al. 2004). The amplicon is sequenced and a 22

similarity match is identified using the NCBI database and BLASTn (O’Donnell et al. 1998;

Elliott et al. 2010; Geiser et al. 2004).

Fusarium oxysporum Pathogenesis

To establish a successful infection, fungal plant pathogens have to invade the plant host and overcome host defenses, either by suppressing or counteracting them. Pathogenesis describes

“the process of disease development in the host from initial infection to onset of symptom”

(Lucas 1998). The fungal genes that are required for pathogenesis are defined in multiple ways.

Idnurm and Howlett (2001) broadly defined pathogenicity genes as “genes necessary for disease development, but not essential for its life cycle”. A subclass of pathogenic genes called effectors are defined as small-secreted proteins that function within the plant host to alter host cell structure, promote infections and suppress plant immunity responses. These proteins are generally cysteine rich and lack homologs in closely-related species (Van de Wouw and Howlett

2011). Some of these effectors, “defeated effectors”, can act as avirulence genes that can be recognized by the plant host and contribute to plant resistance. Casadevall and Pirofski (1999) defined virulence as “relative capacity to cause damage in a host”. Virulence factors are determinants of pathogenicity.

Several studies using forward and reverse genetic approaches along with comparative genomic studies have explored pathogenicity genes of F. oxysporum. In this study, pathogenic genes are defined as those that are essential for disease development. A mini-review of some well-known pathogenicity related genes of Fusarium oxysporum are discussed below.

23

Cell wall degrading enzymes

Plants cell walls are complex, made up of cellulose, hemicellulose, lignin, and pectin. Along with the cuticle, they act as the first barriers against pathogens. Fusarium oxysporum, being a nectrotroph produces an array of cell wall degrading enzymes, such as cutinases (CUT), cellulases, polygalacturnases (PG1, PG5, PGX), pectate lyases (PL1), xylanases (XYL3, XYL4,

XYL5) and proteases like subtilase (PRT1), which assist in host penetration and colonization

(Beckman 1987). However, target mutagenesis studies failed to implicitate them in virulence (Di

Pietro et al. 2003). In F. oxysporum f. sp. lycopersici, xylanases were studied for their involvement in pathogenicity by inactivating transcription factor XlnR that regulates xylanases and celluloses. Although it resulted in low expression of xylanase, no effect on pathogenicity was observed (Calero-Nieto et al. 2007). CTF1, a transcriptional activator for cutinase (CUT1) and lipase (LIP1), was studied in axenic conditions (Rocha et al. 2008), and it was thought to be involved in esterase and lipase activities. Yet, in planta studies suggested that CTF1 does not regulate LIP1 and CUT1, as it was expressed in low levels during root infections. Cell wall degrading enzymes also influence carbohydrate metabolism by carbon catabolite repression.

Disruption of protein kinases like SNF1 and FRP1 (F-box protein) resulted in poor root colonization due to impaired carbohydrate metabolism (Duyvesteijin et al. 2005; Jonkers et al.

2009; Ospina-Giraldo et al. 2003). It is difficult to prove the role of cell wall degrading enzymes in pathogenicity due to their functional redundancy.

Signaling proteins

For successful infection, a pathogen has to recognize the host and adapt to their environment by triggering signaling pathways. In F. oxysporum cAMP-PKA (cyclic adenosine monophosphate-protein kinase) and MAPK (mitogen-activated protein kinase) cascades, activate

24

transcriptional factors which regulate pathogenicity genes. Studies have shown that MAPK

(FMK1), G protein subunits α (FGA1) and β (FGB1) are required for pathogenicity (review by

Di Pietro et al., 2003, Michielse and Rep 2009a). FGA2 disrupted mutants of F. oxysporum f. sp. cucumerinum resulted in altered colony morphology, reduced conidiation and reduced pathogenicity (Jain et al. 2005). This suggests that kinases proteins have pleiotropic effects on housekeeping and pathogenicity related genes.

Transcriptional factors

Recent studies have indicated that transcriptional factors play a role in virulence as they might regulate pathogenicity genes and are gaining importance in understanding pathogenicity processes. Mutations of FOW2 in F. oxysporum f. sp. melonis and SGE1 in F. oxysporum f. sp. lycopersici have resulted in loss of pathogenicity (Imazaki et al. 2007; Michielse et al. 2009b). In

F. oxysporum f. sp. phaseoli, FTF, a multi-copy transcriptional factor gene, is present only in virulent strains and located on a lineage specific chromosome. The genes that are regulated by

FTF1 are currently unknown (Ramos et al. 2007).

Blocking host defenses

In order to survive in hosts, F. oxysporum has to overcome physical barriers and antifungal compounds that are secreted by plant hosts (Beckman 1987; VanEtten et al. 1994). In a well studied example, α-tomatine is produced by tomato plants and attacks fungal membranes resulting in their lysis (Roddick 1977). F. oxysporum f. sp. lycopersici secretes tomatinases which neutralizes α-tomatine (Larini et al. 1996). Disruption of one of the tomatinases, TOM1, in

F. oxysporum f. sp. lycopersici resulted in mutants with decreased progression of symptoms, suggesting their involvement in virulence (Pareja-Jaime et al. 2008). In response to pathogens, host plants deposit cell wall bound phenolic compounds, which are sensed by F. oxysporum and 25

trigger the production of mycotoxins and enzymes like pectinase, cellulose and amylase. Studies involving insertional mutagenesis of carboxy-cis, cis–muconate cylase (CMLE1) resulted in a decreased degradation of phenolic compounds suggesting their role in pathogenesis (Michielse and Rep 2009; Michielse et al. 2009a).

Fusarium oxysporum Genomic Studies

Reverse genetic approaches have focused on selective genes and model organisms, but comparative genomic studies offer the potential for greater insights into profiles of pathogenic genes in non-model organisms and helps to understand the origins and clues regarding the evolution of these genes. An advantage of comparing genomes comes with the ability to identify the presence of orthologous genes (Okagaki et al. 2016). Orthologs are homologs that are derived from a common ancestor and often retain similar functions. Moreover, comparative genomics can also reveal unique genes that are specific or novel to each genome. Comparative genomic studies are widely being used to address different aspects of fungal plant pathogen biology.

Krijger et al. (2014) conducted a comparative study involving 33 fungal pathogen secretomes, revealing that pathogens with similar lifestyles share certain gene families, but the expansion and contraction of gene families are strongly lineage specific and not confined by fungal lifestyle. In another study Guyon et al. (2014) identified 486 candidate virulence genes in Sclerotinia sclerotiorum using bioinformatic tools and evolution analysis, of which 78 were effector candidates. This research will facilitate future studies to understand host-pathgoen interactions by necrotrophic pathogens. For clinical purposes, a comparative genomic study was used to design a single step PCR based diagnostic assay for F. oxysporum f. sp. conglutians (Ling et al.

2016).

26

There are multiple platforms with annotated genomes that are publicly available and can be used as reference, and numerous bioinformatics tools are available to study phylogenic, structural and functional aspects of identified gene families. The two main platforms include

Mycocosm, maintained by the Department of Energy Joint Genome Institute, which provides information on 1000 fungal genomes (Grigoriev et al. 2012), and Fusarium Comparative

Genomics, a platform designed for Fusarium (http://genomics.fusariumdb.org/) (Park et al.

2011).

Three significant contributions using comparative genomics that enhanced our understanding about Fusarium oxysporum are described below. A landmark comparative genomic study was conducted by Ma et al. (2010) on F. verticillioides, F. graminearum and F. oxysporum f. sp. lycopersici (FOL) at the Broad Institute. Their study revealed that the FOL genome is partitioned into core and conditionally dispensable or lineage-specific chromosomes

(LS). LS regions are rich in transposons, pathogenicity genes and repetitive elements. The genes present on LS had different phylogenetic profiles (ancestry) than Fusarium housekeeping genes.

This was the first study to suggest genome level horizontal gene transfer as a mode of transfer for pathogenicity genes between F. oxysporum isolates resulting in new pathogenic strains.

Van Dam et al. (2016) conducted a large scale whole genome analysis on 59 isolates belonging to five F. oxysporum formae speciales (ff. spp. cucumerinum, niveum, melonis, radicis-cucumerinum and lycopersici) of which the first four infect members of the Curcurbita genus. They identified a total of 104 effectors and indicated that effector profiles of different formae speciales infecting the same hosts tend to be similar when compared to other formae speciales. Strains belonging to different lineages within the same forma specialis tend to share identical effector profiles. 27

A similar study conducted by Williams et al. (2016) on three legume infecting F. oxysporum formae speciales (ff. spp. medicaginis, ciceris and pisi) revealed the presence of genomes segregated into core and lineage specific (LS) regions. The research concluded that effectors are usually located in LS regions. Each forma specialis had unique effectors, indicating convergent evolution and an absence of a common ancestor. Yet, a small set of effectors were conserved among all the legume-infecting formae speciales implying their role in host specificity

Secreted in Xylem Gene Family

Secreted in Xylem (SIX) gene family were the first effectors identified in Fusarium oxysporum f. sp. lycoperisci (FOL) using a combination of proteomic and in-silico approaches, along with gene knock-out and Agrobacterium-based assays (Rep et al. 2004; Schmidt et al.

2013). The SIX1 protein was first identified by proteome analysis of xylem sap of a tomato plant infected by F. oxysporum f. sp. lycopersici (Rep et al. 2004). So far, fourteen proteins have been identified in FOL, and the majority of SIX proteins are located on lineage specific chromosome

14 (Ma et al. 2010; Schmidt et al. 2013). SIX proteins are generally cysteine rich with 300 amino acids, an N-terminal signal and non-autonomous transposable elements (MITEs) present at the promoter sites for the majority of these SIX genes (Houterman et al. 2007; Lievens et al. 2009;

Ma et al. 2010; Rep et al. 2004; Schmidt et al. 2013). Besides FOL, homologs of SIX genes are present in other formae speciales, including Fusarium oxysporum ff. spp. cepae, canariensis, cubense, pisi and vasinfectum, with SIX gene profiles distinct from one another (Chakrabarti et al. 2011; Fraser-Smith et al. 2014; Laurence et al. 2015; Lievens et al. 2009; Meldrum et al.

2012; Taylor et al. 2016).

Although SIX1 through SIX6 genes have been characterized, their biological functional roles and the plant host proteins they interact with are still unknown. In the tomato and FOL 28

pathosystem, there are three resistant tomato cultivars, I-1, I-2, I-3, and subsequently, three races of FOL (1, 2, and 3) have been identified. SIX1 (AVR3) was the first SIX gene to be identified in xylem sap of I-3 resistant tomato cultivar (Rep et al. 2004). On susceptible plants, SIX1 is required for pathogenicity, is expressed during initial colonization and requires the presence of living cells (van der Does et al. 2008). A transcriptional factor for SIX gene expression (SGE1) is required for its regulation and contains eight cysteine residues (Houterman et al. 2007; Michielse et al. 2009b; Rep et al. 2004). The FOL SIX4 is an avirulence gene recognized by the I-1 tomato cultivar. It is present only in race 1 isolates of FOL. The protein is secreted during colonization

(Houterman et al. 2007). SIX3, also called AVR2, has a signal peptide with two cysteine residues. AVR2 is recognized by the I-2 tomato cultivar, and on resistant cultivars, it is required for pathogenicity and expressed in xylem vessels during the early stages of root colonization

(Houterman et al. 2007). SIX5, which has six cysteine residues, and SIX3 share a 1365bp upstream region. Additionally, a shared transcriptional regulator SGE1 controls SIX3 and SIX5.

Both genes physically interact with each other and are required for I-2 mediated resistance.

Deletion of SIX5 leads to loss of pathogenicity implying its role as an effector (Ma et al. 2015).

The combination of SIX3 and SIX5 is extensively present in FOL isolates (Lievens et al. 2009).

SIX6 is an exception to the general rule that SIX genes are specific to Fusarium.

Homologs of SIX6 are present in Colletotrichum orbiculare and C. higginsiantum (Gan et al.

2013; Kleemann et al. 2012). SIX6 consists of eight cysteine residues, requires the presence of living cells and its protein is needed for full virulence in FOL. It suppresses I-2 mediated cell death in Nictoniana benthamiana (Gawehns et al. 2014). SIX1 (AVR3), SIX3 (AVR2), and SIX4

(AVR1) act as avirulence genes and are recognized by R proteins in resistant tomato cultivars.

SIX1, SIX3, SIX5 and SIX6 gene knock out and transformation experiments have proven their

29

role in virulence and are classified as effectors on susceptible tomato lines (Gawehns et al, 2014;

Lievens et al. 2009; Ma et al. 2015; Rep et al. 2004).

The majority of SIX genes are single copy and located on chromosome 14 with the exception of SIX8. In the FOL 4287 genome, two variants of SIX8 were observed, SIX8a and

SIX8b, with nine identical copies of SIX8a and four copies of SIX8b present (Schmidt et al.

2013). While the majority of SIX8 copies are present on chromosome 14, a few are present on chromosomes 3 and 6 (Schmidt et al. 2013). Using a proteomic approach, SIX8a was identified in xylem sap of infected tomato plant, whereas SIX8b was absent (Schmidt et al. 2013). There are three SIX8 variants observed in F. oxysporum f. sp. cubense, and their presence varies among the different races (1, 2 and 4 (ST4 and TR4)) (Fraser-Smith et al. 2014).The function of SIX8 remains unknown.

Objectives

In the USA, palms are extensively planted as ornamentals. In Florida, Fusarium wilt caused by FOC and FOP has severe consequences because the disease is always lethal. Florida is the only geographical location where both palm-infecting formae speciales have been reported and coexist. The present research focuses on diagnostics, genomics and population studies of these two Fusarium wilt pathogens. The dissertation is divided into three main objectives, which comprise individual chapters. Chapter 5 summarizes the results from all three objectives and suggests further direction for research.

1. The first objective (Chapter 2) is to design a diagnostic primer pair for Fusarium wilt pathogens (FOC and FOP) isolated from ornamental palms in Florida.

2. The second objective (Chapter 3) is to identify candidate pathogenicity factors, by employing comparative genomics, which are unique or conserved across ornamental

30

palm-infecting formae speciales and can explain the aggressiveness of FOP compared with FOC.

3. The third objective (Chapter 4) is to identify the SIX gene profiles of FOP and FOC based on the 14 SIX genes identified previously from F. oxysporum f. sp. lycopersici.

31

CHAPTER 2 DIAGNOSTIC PRIMER PAIR TO IDENTIFY FUSARIUM WILT PATHOGENS ON ORNAMENTAL PALMS IN FLORIDA

Introduction

Palms are economically important ornamentals in Florida. In 2010, the retail sale value of palms was equal to all other trees (deciduous, evergreen, flowering and fruit) sold in Florida

(Hodges et al. 2011). However, as with any plant, they are prone to various nutritional deficiencies and diseases. Fusarium wilt caused by Fusarium oxysporum is one of the important fungal diseases threatening ornamental palms in Florida (Elliott et al. 2010; Plyler et al. 2000).

Fusarium oxysporum is classified into more than 120 formae speciales based on the specific host plants infected (Michielse and Rep 2009). Currently, there are four Fusarium wilt diseases on palms. F. oxysporum f. sp. albedinis (FOA) causes date palm wilt on Phoenix dactylifera and is observed in Algeria, Morocco and Mauritania (EPPO 2003). F. oxysporum f. sp. eleidis (FOE) causes oil palm wilt on Elaeis guineensis and is reported from countries in Africa and localized areas in Brazil and Ecuador (Flood 2006). Canary Island date palm wilt, caused by F. oxysporum f. sp. canariensis (FOC), is present worldwide on Phoenix canariensis including Australia (Priest and Letham 1996), Argentina (Palmucci 2005), the Canary Island Islands (Hernandez-Hernandez et al. 2010), France (Mercier and Louvet 1973), Greece (Elena 2005), Japan (Arai and

Yamamoto1977), Spain (Fernandez et al. 1997) and USA (Feather et al. 1989). In the USA, FOC has been reported on Phoenix canariensis in California (Feather et al. 1989), Florida (Simone

2004), Louisiana (Singh et al. 2011), Nevada (Wang and McKie 2007), South Carolina and

Texas (Elliott et al. 2011). FOC has also been documented on Phoenix reclinata in Florida

(Elliott 2015) and California (California Department of Agriculture). Fusarium wilt of queen

(Syagrus romanzoffiana) and Mexican fan (Washingtonia robusta) palms is caused by F.

32

oxysporum f. sp. palmarum (FOP), which is widespread in Florida (Elliott et al. 2010). Except for one report from one location in Texas (Giesbrecht et al. 2013), this pathogen is not known anywhere else in the world. FOP has a wider host range than most Fusarium oxysporum formae speciales, as it not only causes wilt diseases on queen and Mexican fan palms, but also on

Phoenix canariensis and x Butyagrus nabonnandii ( Elliott 2011; Elliott et al. 2010; 2017). It has also been detected in wilt symptomatic specimens of Bismarckia nobilis and Phoenix reclinata in

Florida (Elliott, personal observation).

Unlike the rest of the world where palms are grown, Florida has two F. oxysporum formae speciales affecting palms, FOC and FOP, with Phoenix canariensis and possibly P. reclinata afflicted by both formae speciales. These two Fusarium wilt pathogens primarily cause a canopy infection with similar symptoms. A one-sided necrosis of the leaf blade is first observed on older leaves with a brown stripe on the rachis or petiole associated with the necrotic leaflet tissue. There is corresponding internal discoloration of the petiole or rachis, and it is this tissue that is used to isolate the pathogen. The symptoms progress upwards through the palm canopy until the palm is killed. Once symptoms are expressed, FOP can kill susceptible palm hosts within a few weeks to several months, as compared to FOC, which usually takes a year to kill the palm (Elliott et al. 2010; Plyler et al. 2000). If the pathogen becomes incorporated into the soil, then infection can occur via the root system, with similar symptoms observed in the canopy. Currently, there are no treatments available to cure these Fusarium wilt diseases (Elliott et al. 2010; Plyler et al. 2000). Disease management relies heavily on prevention and reducing the spread of inoculum (Elliott et al. 2010; Plyler et al. 2000).

It is crucial to identify palm Fusarium wilt pathogens rapidly in order to implement management strategies to prevent the spread of disease. Traditional methods to identify palm 33

Fusarium wilt pathogens are based on symptomology, isolation of the pathogen from the infected palm tissue, and growing isolates on specific media in order to observe colony and spore morphology, both microconidia and macroconidia (Leslie and Summerell 2006). These methods need expertise and are laborious. Additionally, nonpathogenic Fusarium oxysporum and other

Fusarium species are occasionally isolated from symptomatic palm tissue, at times, from the same symptomatic palm tissue (Elliott et al. 2010). These traditional methods are limiting as these Fusarium species might share similar colony and spore morphology, leading to misidentification (Zhang et al. 2005).

PCR-based assays are being used to detect and identify pathogens, as they are fast, reliable and specific, allowing one to differentiate closely related species or even races within a species (Edel et al. 1995; Glass and Donaldson 1995; Li and Hartman 2003; Nayaka et al. 2011).

Designing a diagnostic tool to identify F. oxysporum formae speciales is challenging due to the presence of polyphyletic lineages and possibly races within some formae speciales (Plyler et al.

2000; Suga et al. 2013). There is a PCR-based assay for detecting FOC that uses the primer pair

HK66 and HK67, which is based on the beta-glucosidase gene (Plyler et al. 1999). However, frequent false positives are obtained because this primer pair can amplify Fusarium proliferatum, a common saprophyte associated with ornamental palm tissue (Elliot et al. 2010; Vitoreli 2008).

This problem was solved for detecting FOC by developing a duplex PCR based assay (Vitoreli et al. 2009)

Molecular markers that are routinely used in identifying Fusarium are the partial translation elongation factor 1–alpha (EF-1α), β-tubulin, and calmodulin genes (Baayen et al.

2000; O’Donnell et al. 1998; 2009; Wulff et al. 2010). EF-1α gene is an ideal choice to differentiate Fusarium because it is a single locus gene and contains adequate polymorphism to 34

distinguish Fusarium species and formae speciales of F. oxysporum (Geiser et al. 2004; Nayaka et al. 2011; Wulff et al. 2010).

The first objective of this study was to design a primer pair for a PCR-based assay for rapid, accurate identification of Fusarium wilt pathogens on ornamental palms. The initial goal was for identification of FOP only, but as the study progressed, it became clear that the primer pair could be used for FOC and possibly for FOA and FOE also. The second objective was to evaluate the newly designed primer pair using three different polymerases and associated PCR buffers and two types of template DNA with a collection of 62 pathogenic and nonpathogenic

Fusarium isolates that had been recovered from symptomatic palm tissue in the past ten years in

Florida. Additional template DNA was obtained of FOC, FOA and FOE and were evaluated accordingly.

Material and Methods

Isolates

Table 2-1 is a list of all isolates used in this study. Sixty-two isolates comprised of FOP (18),

FOC (12), F. proliferatum (5), Fusarium incarnatum-equiseti species complex (FIESC) (4), F. solani (3), F. sacchari (1), F. concentricum (3), and nonpathogenic F. oxysporum (16) had been obtained from palm tissues over a ten-year period, primarily from Florida, but also California,

South Carolina and Texas. Most, but not all of these isolates, were obtained from the petiole of palms with Fusarium wilt symptoms. Half of the nonpathogenic F. oxysporum isolates were obtained from roots, leaflets or trunk tissue. All isolates were single-spored and stored at room temperature as mycelial agar plugs in cryogenic vials containing sterile deionized water at the

University of Florida-IFAS, Fort Lauderdale Research and Education Center. Additionally, genomic DNA of 12 isolates was acquired from USDA Agricultural Research Service Culture 35

Collection (NRRL), Peoria, IL and California Department of Agriculture (Table 2-1). These included F. oxysporum ff. spp. albedinis (3), canariensis (4) and elaeidis (5).

Genomic DNA and EF-1α Sequencing

Each isolate from the Florida collection was revived from storage by placing a single agar plug on 1/5 strength potato dextrose agar (1/5 PDA) (Difco Laboratories, Detroit, MI) and grown for 5 days at 28°C. Five agar plugs from 1/5 PDA were transferred to a conical flask with 20 ml of 1/5 strength potato dextrose broth (1/5 PDB) and incubated for 5 days at 28°C without shaking. DNA was extracted from fungal mycelial tissue using Gentra Puregene Tissue Kit

(Qiagen, Valencia, CA) according to the manufacturer instructions. The resulting DNA pellet was suspended in 50 µl of molecular grade water (HyCloneTM, Fisher Scientific, Pittsburg, PA) and quantified using a QubitTM fluorometer (Thermo Fisher Scientific, Pittsburg, PA). The DNA was stored at 4°C for further use. All DNA, either from the Florida collection or provided by other agencies, was assessed for quality and identity by amplifying with ef1 and ef2 primers

(Geiser et al. 2004), which amplify a portion of the EF-1α gene, sequencing the amplicons and comparing sequences with those deposited in GenBank using BLASTn. Purification of PCR products was performed on spin columns using the Wizard® PCR Preps DNA Purification

System (Promega Corp. Madison, WI). Sanger sequencing in both directions was conducted at

University of Florida’s Interdisciplinary Center for Biotechnology (UF-ICBR), (Gainesville,

Florida) with the same primers.

Primer Design

A multiple sequence alignment of 12 EF-1α sequences of Fusarium oxysporum f. sp. palmarum from GenBank (GQ154454, GQ154455, GQ154458, GQ154459, GQ154460,

GQ154462, GQ154463, GQ154464, GQ154465, GQ154468, GQ154469, and GQ154457) was 36

conducted to generate a consensus sequence (Clustal W, Lasergene, DNASTAR, Madison, WI).

Primer-BLAST created five primer pairs based on this consensus sequence (Ye et al. 2012).

Primer Stat (Sequence Manipulation Suite, 2000) (Stothard 2000) analyzed each primer pair based on GC clamp, hairpin formation, melting temperature and self-annealing criteria. A primer set (PFW-F and PFW-R) satisfied all the parameters and was selected for further study. Primers were synthesized by Sigma-Aldrich (St. Louis, MO). The melting temperature for the PFW primer pair was determined by a gradient PCR.

Genomic DNA Amplification with PFW Primer Pair

Genomic DNA of all isolates listed in Table 2-1 was used for the PCR assay using the

PFW primer pair. PCR was performed using three different sources of polymerases: Clontech

(Takara Bio USA, Mountain View, CA), New England Biolabs (NEB) (Ipswich, MA) and

Sigma-Aldrich (Sigma) (St. Louis, MO). These three PCR kits were selected as they were commonly used in plant diagnostic settings. Each PCR was carried out in a 50 µl reaction mixture brought to volume with sterile deionized water. The NEB PCR assay was set up with a final concentration of 1X PCR standard Taq Reaction Buffer (10 mM Tris–HCl, 50 mM KCl, 1.5 mM MgCl2), 200 µM of each dNTP, 1 µM of each primer and 1U of NEB Taq DNA polymerase. The Clontech PCR final concentration consisted of 25 µl of 2X TerraTM PCR Direct

TM Buffer (400 µM of each dNTP; 2 mM MgCl2), 1 µM of each primer and 1.25U of Terra PCR

Direct Polymerase. The Sigma PCR final concentration consisted of 25 µl of 2X ReadyMixTM

Taq Reaction Mix (1.5U Taq polymerase, 10 mM Tris-HCl, 50 mM KCl, 1.5 mM MgCl2,

0.001% gelatin, and 200 µM of each dNTP) with 1 µM of each primer. Approximately, 10 ng of genomic DNA served as the template for each reaction. Each PCR assay included a negative control consisting of sterile deionized water.

37

PCR parameters for all the three polymerase sources were: initial denaturation at 94°C for 2 min, followed by 30 cycles of denaturation at 94°C for 45 sec, annealing at 65°C for 90 sec, and extension at 72°C for 2 min, with a final extension step at 72°C for 10 min. An Eppendorf

Thermocycler (Eppendorf AG, Foster City, CA) was used for all PCR assays. PCR products were examined by 1.5% agarose gel electrophoresis in Tris borate EDTA (TBE) buffer with ethidium bromide staining. Purification of selected PCR products was performed on spin columns using Wizard® PCR Preps DNA Purification System (Promega Corp. Madison, WI).

Sanger sequencing in both directions was conducted at UF-ICBR (Gainesville, Florida) with the same primers.

DNA Amplification with Agar Plug and PFW Primer Pair

The PFW primer pair was tested in a PCR assay using a single colonized agar plug (1 mm in width and 3-5 mm in length) as DNA template instead of genomic DNA to eliminate the need to grow the fungus in liquid culture and extract DNA. A total of 26 isolates were grown on

1/5 PDA for 5 days at 28°C: F. concentricum (PLM57D, PLM62D), Fusarium oxysporum ff. spp canariensis (PLM696A, PLM706A, PLM724A, PLM754A) and palmarum (PLM153B,

PLM249A, PLM510A, PLM541D, PLM619A, PLM632A, PLM814A, PLM853A), F. proliferatum (PLM196B, PLM782A), F. sacchari (PLM 700D), F. solani (PLM280D,

PLM332C), F. incarnatum–equiseti species complex (PLM187D, PLM 261D) and Fusarium oxysporum (PLM64A, PLM182C, PLM 428A, PLM 818A, PLM851A). PCR assays were performed using Clontech, NEB and Sigma polymerases and PCR parameters as described above. Each PCR assay included a negative control consisting of sterile deionized water.

38

DNA Amplification with the FOA Primer Pair

A PCR-based screening was performed on selected representative isolates to test the specificity of F. oxysporum f. sp. albedinis primer pair TL3/FOA28. Genomic DNA of 21 isolates were used in this assay including three isolates of Fusarium oxysporum (PLM208A,

PLM818A, PLM820A), five isolates each of F. oxysporum ff. spp canariensis (PLM221B,

PLM387B, PLM511A, PLM588A, PLM908A), eleidis (NRRL22543, NRRL36358,

NRRL36359, NRRL38313, NRRL38314) palmarum (PLM153B, PLM249A, PLM344E,

PLM632A, PLM710A) and albedinis (NRRL26622, NRRL38288, NRRL38298). The latter were used as positive controls. PCR assays were set up with a total 50 µl volume with final concentration of 1X PCR standard Taq Reaction Buffer (10 mM Tris –HCl, 50 mM KCl, 1.5 mM MgCl2), 200 µM of each dNTP, 1 µM of each primer and 1U of NEB Taq DNA polymerase. Approximately 10 ng of genomic DNA served as the template for each reaction.

Each PCR assay included a negative control consisting of sterile deionized water. PCR was performed using cycling conditions as described by Fernandez et al. (1998) with initial denaturation of 4 mins at 95°C followed by 30 cycles of denaturation at 30 sec at 92°C, annealing at 62°C for 30 sec and extension for 45 sec at 72°C with a final extension step of 15 min at 72°C. All the PCR products were analyzed for the presence or absence of PCR product using 1.5% agarose gel electrophoresis in TBE buffer with GelRedTM staining (Biotium,

Hayward, CA).

Phylogenetic Analysis

Sequences derived for EF-1α from PCR assays using ef1 and ef2 primers (Geiser et al

2004) and sequences from the amplicons derived from the PFW primer pair were queried against known sequences in GenBank using BLASTn to confirm their identity. A phylogenetic analysis

39

with MEGA version 7 (Kumar et al. 2016) was conducted independently for the two data sets: 35 sequences obtained with EF-1α primers and 35 sequences obtained with PFW primers. A multiple sequence alignment was performed with MUSCLE using the default parameters (Edgar

2004). The phylogenetic tree was constructed by the Maximum Parsimony method using the heuristic search option Tree-Bisection-Reconnection (TBR) branch swapping algorithm with

1000 random addition sequences with 1000 bootstrap replications. The consistency index (CI) and the retention index (RI) were also calculated. All the gaps and missing data were eliminated from the analysis. Unrooted EF-1α and PFW based phylogenetic trees were constructed to infer evolutionary relationships between isolates (Kinene et al. 2016).

Results

Genomic DNA and EF-1α Sequencing

Using genomic DNA for all 74 isolates, a portion of the EF-1α gene was amplified for all

74 isolates using ef1 and ef2 primers. Resulting amplicons were sequenced. Sequences were queried against GenBank using BLASTn to confirm their identity (Table 2-1).

Primer Design

The newly evaluated primer pair was designated PFW-F (5`-

CCATCGTCAATCCCGACCA-3’) and PFW-R (5`-GGAGCGTCTGAGTGATATGTTAG-3’).

The resulting amplicon size is 569 bp. The melting temperature for the PFW primer is 65°C as determined by a gradient PCR.

Genomic DNA Amplification with PFW Primer Pair

Using genomic DNA, the PFW primer pair amplified all isolates of F. oxysporum ff. spp albedinis (3), canarienis (16), elaedis (5), and palmarum (18) across all three polymerases and

40

PCR buffer mixes evaluated (Table 2-1 and Figure 2-1). Of the 16 nonpathogenic Fusarium oxysporum isolates tested, Clontech TerraTM PCR Direct polymerase amplified all of them

(Table 2-1). However, only 12 and 8 isolates were amplified by the NEB and Sigma polymerases, respectively (Table 2-1). Genomic DNA of other Fusarium species in the study group were not amplified with the PFW primer pair with any of the polymerases tested (Table 2-

1 and Figure 2-1).

DNA Amplification with Agar Plug and PFW Primer Pair

Using a colonized agar plug, Clontech TerraTM Direct polymerase, but not the NEB or

Sigma polymerases, amplified all Fusarium oxysporum (PLM64A, PLM182C, PLM 428A,

PLM818A, PLM851A), f. sp. canariensis (PLM696A, PLM706A, PLM724A, PLM754A) and f. sp. palmarum (PLM153B, PLM249A, PLM510A, PLM541D, PLM619A, PLM632A,

PLM814A, PLM853A) (Table 2-1 and Figure 2-2). As with genomic DNA, no amplification was observed for F. proliferatum, F. solani, F. incarnatum–equiseti species complex, F. concentricum and F. sacchari with any of the evaluated polymerases (Table 2-1 and Figure 2-2).

DNA Amplification with FOA Primer Pair

The Fot1 transposable element based primer pair TL3/FOA28 amplified only the genomic DNA of F. oxysporum f. sp. albedinis (NRRL26622, NRRL38288, NRRL38298), resulting in a 400 bp PCR product.

Phylogenetic Analysis

An alignment of the 35 EF-1α data set resulted in 600 characters in which 21 were parsimony informative sites. An unrooted consensus tree generated by Maximum Parsimony analysis indicated there are four well-established clades (Figure 2-3). Clade 1 is divided into

41

FOP and FOA subgroups. All FOP sequences, irrespective of palm host and geographical location, clustered within the same clade, but two single nucleotide polymorphism within FOP are reflected in two isolates (PLM320B and PLM676A), which were isolated from different palm hosts. A single point mutation, distinct from other FOP sequences, was observed in PLM632A.

No variation was observed in the three FOA sequences, and they clustered in a single subgroup.

Variation was observed within nonpathogenic Fusarium oxysporum sequences. The majority of F. oxysporum sequences clustered in clade 2. F. oxysporum sequences PLM51A

(leaflets), PLM208A (petiole), PLM476C (petiole) and PLM851A (roots) had identical EF-1α sequences, even though they were recovered from different palm hosts and sampling tissues.

Clade 3 is represented by FOE and two nonpathogenic F. oxysporum isolates. The F. oxysporum isolates PLM568A (roots) and PLM271A (petiole) were more similar to FOE than other F. oxysporum. All FOC isolates clustered together in clade 4 with 100% sequence similarity, irrespective of the location or palm host. No polymorphism was observed among the FOC isolates suggesting a monophyletic lineage. Among the formae speciales that infect palm species,

FOC appeared more distinct with at least ten polymorphic sites compared with FOA, FOE and

FOP.

The phylogenetic tree based on the PFW primer pair sequenced amplicons resulted in 431 characters with 13 parsimony informative sites. The PFW phylogenetic tree yielded similar topology to the EF-1α phylogenetic tree, with the exception of clade 1 (Figure 2-4). In clade 1, subgrouping of FOA and FOP was not observed. The unique patterns of PLM320B, PLM676A and PLM632A observed in the EF-1α tree were also displayed in the PWF tree. The same clade groupings observed with EF- 1α were observed for FOE, FOC and nonpathogenic F. oxysporum isolates (Figure 2-4). 42

Discussion

Fusarium oxysporum consists of both nonpathogenic and pathogenic isolates with the latter classified into formae speciales based on the host plant pathogenicity. Each forma specialis is confined to a limited host range, normally comprised of only one or two host plant species.

FOP is an exception, as it has a wide host range, infecting palms belonging to Phoenix,

Washingtonia and Syagrus genera within the family Arecaceae. Diseases caused by both FOP and FOC are lethal and result in pathogen-infested soils that are unsuited for planting susceptible palm hosts. Canary Island date palm wilt caused by FOC is spread mainly by pruning tools, whereas windblown spores and pruning tools are thought to spread Fusarium wilt on queen and

Mexican fan palms caused by FOP (Elliott et al. 2010). With no fungicide treatments available, timely diagnosis can prevent the spread of the diseases by implementation of phytosanitary measures, primarily palm removal. Furthermore, both FOP and FOC occur in Florida, and their host ranges continue to expand, with P. canariensis and possibly P. reclinata being affected by both pathogens.

Currently, while there is a primer pair (HK66/HK67) used for detection of FOC (Plyler et al. 1999), it does not detect FOP (Elliott, unpublished results), and it also often yields false positives if the fungus isolate is F. proliferatum (Vitoreli 2008). To solve this latter problem, a duplex PCR-based assay is used that requires the FOC primer pair and a primer pair for F. proliferatum (Vitoreli et al. 2009). Another forma specialis specific primer pair TL3/FOA28, based on Fot1 transposon, detects only FOA (Fernandez et al. 1998; results reported herein).

There are no published FOE-specific primers, although they were reported to be in development

(Rusli 2012).

43

The present study objective was to design primers for a PCR-based assay, which would be fast, cost-effective and reliable for identifying Fusarium wilt pathogens of palm occurring in

Florida. Because the diseases caused by FOC and FOP are lethal, for initial management purposes (i.e. removal), it is not important to know which pathogen, FOC or FOP, is present. It was also desirable to have a PCR based assay that could use mycelial agar plugs as DNA template to eliminate the DNA extraction process. To evaluate the primers developed from multiple sequence alignment, a collection of Fusarium species isolated in Florida from symptomatic Fusarium wilt palms, but not necessarily from petiole or rachis tissue (leaflets, roots and trunk) was used, with the main intent to include comprehensively all the Fusarium species that are expected to be encountered in Florida plant diagnostic labs.

Genomic DNA of all the evaluated isolates of FOC and FOP were amplified using the three polymerases tested and the PFW primer pair developed. Along with FOC and FOP, nonpathogenic F. oxysporum were also amplified but not consistently with every polymerase. F. proliferatum, F. solani, F. concentricum, F. sacchari and F. incarnatum-equiseti species complex failed to amplify with any of the polymerases evaluated, indicating that the PFW primers are specific for F. oxysporum. This is not surprising because the primers are based on the

EF-1α gene. Nonpathogenic F. oxysporum isolates may be recovered from symptomatic palm tissues. However, for greater accuracy with the PCR based assay, the type of tissue sampled is important, and perhaps, the polymerase used for the PCR reaction. As both FOC and FOP primarily cause a canopy infection, the petiole is the ideal tissue to sample (Elliott et al. 2010).

Only 8 of 16 nonpathogenic F. oxysporum isolates from Florida were recovered from petiole tissue, whereas all but one of 25 FOC and FOP isolates from Florida were obtained from petiole tissue. Because the PFW primers amplify both FOP and FOC, if it is necessary to identify the

44

forma specialis, sequencing of the PCR amplicon derived from the PFW primer, and subsequent

BLAST analysis will distinguish between FOC and FOP (Figure 2-5).

The three PCR buffers and polymerase kits used in this study, Clontech, NewEngland

Biolabs and Sigma-Aldrich, were choosen as they are premade ready-to-use buffers. Moreover, at the time of this research, SigmaReady mix was routinely used for diagnositic purposes at the

Plant Diagnostic Clinic (University of Florida, Gainesville, Florida). Clontech was selected based on its ability to amplify DNA from crude samples, such as colonized agar plugs, fungal spores or other fungal tissue. New England Biolab PCR kit was included in the study as it was the normal PCR kit used in the Elliott laboratory, where this study was conducted. While the type of polymerase and/or associated buffer mix did not affect PCR results for FOC or FOP when using genomic DNA, it did affect PCR results of the 16 nonpathogenic F. oxysporum isolates. All were amplified by the Clontech polymerase, but not all were amplified by NEB and

Sigma polymerases. The disparity among polymerases may be due to more than the polymerase itself, as there is a difference in the buffer mixes. Magnesium (Mg2+) is a vital cofactor needed for DNA polymerase, and lower concentration leads to higher specificity (Lorenz 2012). The

Mg2+ concentration in the buffer mixes used with the NEB and Sigma polymerases (1.5 mM

MgCl2) was lower compared to the Clontech polymerase buffer mix (2.0 mM MgCl2) leading to more positive PCR results with the Clontech polymerase. However, the Clontech polymerase has properties that can be exploited if one wants to use mycelial agar plugs as the DNA template for the PCR assay. The Clontech TerraTM PCR Direct Polymerase, according to the manufacturer,

“is a novel enzyme developed for direct amplification from tissue samples, crude extracts, and dirty templates.” This probably explains why DNA of F. oxysporum, FOP and FOC could be

45

amplified using agar plugs with the Clontech polymerase but not with the NEB and Sigma polymerases.

PFW primers also amplified FOA and FOE, the formae speciales that cause Fusarium wilts on date and oil palms, respectively. However, these two pathogens are not present in the

USA, and thus far have limited geographic ranges. Nonetheless, the PFW primers could be useful for detection of FOA and FOE in those countries where those diseases occur. While sequencing the PCR amplicon using the PFW primer will identify FOE, the sequences of FOP and FOA are the same (Figure 2-4). This is not surprising considering the previously shown close relationship (Elliott et al. 2010) and results in this study (Figure 2-3).

This study points out that there is a need for additional markers that can separate F. oxysporum formae speciales, especially those associated with palm species. Host specific or pathogenicity genes, like the family of secreted in xylem (SIX) genes, are currently being explored to help solve this problem (Suga et al. 2013). In Florida, confirming the presence of

Fusarium wilt pathogens is necessary for monitoring the continued expansion of the host range of FOC and FOP and to aid in decisions regarding palm removal and nursery quarantines. As both FOP and FOC continue to expand their geographic range, the PFW primer pair may be useful for other states where palms are part of the landscape or grown in nurseries.

While the PFW primers are not specific for FOC and FOP, they are specific for F. oxysporum, and this eliminates false positives if other Fusarium species are isolated. False positives can also be reduced by using only symptomatic petiole tissue for isolation purposes. If necessary, sequencing the amplicons resulting from the PCR-based assay using the PFW primers will separate FOP from FOC. Previously, a second PCR assay using the EF-1α primers (ef1 and

46

ef2) would have had to be completed for their identification. Thus, best diagnostic practices for identifying palm Fusarium wilt pathogens in Florida include: 1) only isolate from Fusarium wilt symptomatic petiole tissue; 2) use the PFW primer pair in a PCR assay; and 3) if necessary, to distinguish between FOP and FOC, sequence amplicons derived from the PCR assay (Figure 2-

5).

47

Table 2-1. Fusarium isolates used in the study, indicating isolate, palm host, geographic location and tissue sampled. Genomic DNA or a fungal colonized agar plug was used as template for PCR using PFW primers and three different polymerases. Isolatea Palm host Geographic Sampled Genomic DNA Agar plug amplification locationb tissue amplificationc

Clontech NEB Sigma Clontech NEB Sigma Fusarium oxysporum f. sp. albedinis (FOA) NRRL 26622 Phoenix sp. Morocco Unknown + + + NTd NT NT NRRL 38288 Phoenix dactylifera - Unknown + + + NT NT NT NRRL 38298 Phoenix dactylifera Algeria Unknown + + + NT NT NT F. oxysporum f. sp. canariensis (FOC) PLM221B Phoenix canariensis Florida Petiole + + + NT NT NT PLM224A Phoenix sylvestris Florida Petiole + + + NT NT NT PLM385B Phoenix canariensis Texas Petiole + + + NT NT NT PLM387B Phoenix reclinata Florida Petiole + + + NT NT NT PLM511A Phoenix canariensis South Petiole + + + NT NT NT Carolina PLM588A Phoenix canariensis Florida Petiole + + + NT NT NT PLM696A Phoenix reclinata Florida Petiole + + + + - - PLM706A Phoenix canariensis California Petiole + + - + - - PLM724A Phoenix canariensis Florida Petiole + + + + - - PLM754A Phoenix canariensis Florida Petiole + + + + - -

48

Table 2-1. Continued Isolatea Palm host Geographic Sampled Genomic DNA Agar plug amplification locationb tissue amplificationc

Clontech NEB Sigma Clontech NEB Sigma PLM776A Phoenix reclinata Florida Petiole + + + NT NT NT PLM908A Phoenix canariensis California Petiole + + + NT NT NT 1522180 Phoenix canariensis California Unknown + + + NT NT NT USNVEPO6194335 Phoenix canariensis Nevada Unknown + + + NT NT NT NT NT 370P06144091-E Phoenix reclinata California Unknown + + + NT NT NT NT NT F. oxysporum f. sp. elaeidis (FOE) NRRL 22543 Elaeis guineensis Surinam Unknown + + + NT NT NT NRRL 38313 Elaeis guineensis Brazil Unknown + + + NT NT NT NRRL 38314 Elaeis guineensis Brazil Unknown + + + NT NT NT NRRL 36358 Elaeis guineensis Zaire Unknown + + + NT NT NT NRRL 36359 Elaeis guineensis Zaire Unknown + + + NT NT NT F. oxysporum f. sp. palmarum (FOP) PLM153B Syagrus romanzoffiana Florida Petiole + + + + - - PLM231D Syagrus romanzoffiana Florida Petiole + + + NT NT NT PLM249A Washingtonia robusta Florida Petiole + + + + - -

49

Table 2-1. Continued Isolatea Palm host Geographic Sampled Genomic DNA Agar plug amplification locationb tissue amplificationc

Clontech NEB Sigma Clontech NEB Sigma PLM320D x Butyagrus nabonnandii Florida Petiole + + + NT NT NT PLM344E Syagrus romanzoffiana Florida Petiole + + + NT NT NT PLM409A Washingtonia robusta Florida Petiole + + + NT NT NT PLM510A Phoenix canariensis Florida Petiole + + + + - - PLM541D Syagrus romanzoffiana Florida Petiole + + + + - - PLM619A Bismarckia nobilis Florida Petiole + + + + - - PLM632A Washingtonia robusta Florida Petiole + + + + - - PLM676A Phoenix canariensis Florida Petiole + + + NT NT NT PLM710A Washingtonia robusta Florida Petiole + + + NT NT NT PLM741A Syagrus romanzoffiana Florida Trunk + + + NT NT NT PLM764A Phoenix reclinata Florida Petiole + + + NT NT NT PLM779A Washingtonia robusta Florida Petiole + + + NT NT NT PLM814A Bismarckia nobilis Florida Petiole + + + + - - PLM853A xButygrus nabonnandii Florida Petiole + + + + - - PLM920 Washingtonia robusta Texas Petiole + + + NT NT NT

50

Table 2-1. Continued Isolatea Palm host Geographic Sampled Genomic DNA Agar plug amplification locationb tissue amplificationc

Clontech NEB Sigma Clontech NEB Sigma Fusarium oxysporum (FO) PLM51A Syagrus romanzoffiana Florida Leaflets + + - NT NT NT PLM182C Syagrus romanzoffiana Florida Petiole + + + + - - PLM208A Syagrus romanzoffiana Florida Petiole + + + NT NT NT PLM218A Syagrus romanzoffiana Florida Petiole + + + NT NT NT PLM260C Washingtonia robusta Florida Petiole + - - NT NT NT PLM271A Syagrus romanzoffiana Florida Petiole + + - NT NT NT PLM384B Washingtonia robusta Florida Petiole + + - NT NT NT PLM428A Syagrus romanzoffiana Florida Petiole + + + + - - PLM476C Syagrus romanzoffiana Florida Petiole + + + NT NT NT PLM568B Phoenix sylvestris Florida Roots + + + NT NT NT PLM818A Phoenix sylvestris Florida Roots + + + + - - PLM820A Phoenix canariensis Florida Trunk + + + NT NT NT PLM848A Washingtonia robusta Florida Roots + - - NT NT NT PLM850A Phoenix dactylifera Florida Roots + - + NT NT NT PLM851A Roystonia regia Florida Roots + + - + - -

51

Table 2-1. Continued Isolatea Palm host Geographic Sampled Genomic DNA Agar plug amplification locationb tissue amplificationc

Clontech NEB Sigma Clontech NEB Sigma F. concentricum PLM57D Syagrus romanzoffiana Florida Petiole ------PLM190B Syagrus romanzoffiana Florida Petiole - - - NT NT NT F. incarnatum–equiseti species complex PLM187D Syagrus romanzoffiana Florida Petiole ------PLM261D Washingtonia robusta Florida Petiole ------PLM836A Bismarckia nobilis Florida Petiole - - - NT NT NT PLM880A Syagrus romanzoffiana Florida Petiole - - - NT NT NT F. proliferatum PLM89A Syagrus romanzoffiana Florida Petiole - - - NT NT NT PLM196B Syagrus romanzoffiana Florida Petiole ------PLM782A Syagrus romanzoffiana Florida Petiole ------PLM841A Washingtonia robusta Florida Roots - - - NT NT NT F. sacchari PLM700A Syagrus romanzoffiana Florida Petiole ------

52

Table 2-1. Continued Isolatea Palm host Geographic Sampled Genomic DNA Agar plug amplification locationb tissue amplificationc

Clontech NEB Sigma Clonetech NEB Sigma F. solani PLM280D Ravenea rivularis Florida Trunk ------PLM281C Ravenea rivularis Florida Petiole - + - NT NT NT PLM332C Ravenea rivularis Florida Petiole ------aNRRL= Genomic DNA received from USDA Agricultural Research Service Culture Collection; Non PLM numbers of F. oxysporum f. sp. canariensis indicates genomic DNA received from California Department of Agriculture. bState is listed, unless the isolate is from outside the USA. cClontech=Terra™ PCR Direct Polymerase; NEB=New England Biolabs Taq Polymerase; Sigma=Sigma-Aldrich ReadyMix™ Taq Polymerase. dNT=not tested.

53

Figure 2-1. Comparison of genomic DNA amplification using three different polymerases with the PFW primer pair as viewed on a 1.5% agarose gel. Lanes 3-10 (A), 11-19 (B), 21- 28 (C) represent PCR amplification performed using Clonetech PCR Direct, New England Biolabs, and Sigma ReadyMix polymerases, respectively. Lanes 3, 4, 12, 13, 21 and 22 are Fusarium oxysporum f. sp. canariensis (PLM224A; PLM754A). Lanes 8, 9, 17, 18, 26 and 27 are Fusarium oxysporum f. sp. palmarum (PLM 632A, PLM541D). Lanes 5, 14 and 23 are F. proliferatum (PLM196B). Lanes 6, 15 and 24 are Fusarium incarnatum-eqisetti species complex (PLM 261D). Lanes 7, 16 and 25 are Fusarium solani (PLM332C). Lanes 10, 19 and 28 are water controls. Lanes 2, 11, 20, and 29 are blank. Lanes 1 and 30 are 1 kb ladder.

54

Figure 2-2. Comparison of DNA amplification using mycelial agar plug as template for PCR with three different polymerases and PFW primer pair as viewed on a 1.5% agarose gel. Lanes 1 and 30 are 1kb ladder. Lanes 2-10 (A), 11-19 (B), 20-28 (C) represent PCR amplification performed using Clonetech PCR Direct, New England Biolabs and Sigma ReadyMix polymerase, respectively. Lanes 2, 11 and 20 are Fusarium oxysporum f. sp. canariensis (PLM696A). Lanes 3, 12 and 21 are F. proliferatum (PLM196B). Lanes 4, 13 and 22 are Fusarium incarnatum-eqisetti species complex (PLM 261D). Lanes 5, 14 and 23 are F. solani (PLM280D). Lanes 6, 15 and 24 are Fusarium oxysporum (PLM428A). Lanes 7, 16 and 25 are Fusarium oxysporum f. sp. palmarum (PLM153B). Lanes 8, 17 and 26 are F. sacchari (PLM700A). Lanes 9, 18 and 27 are F. concentricum (PLM57D). Lanes 10, 19 and 28 are water controls. Lane 29 is blank.

55

Figure 2-3. Unrooted consensus tree inferred from 10 most parsimonious trees using the Maximum Parsimony method based on a portion of the translation elongation factor (EF-1α) gene. Branches corresponding to partitions reproduced in less than 70% trees are collapsed. The percentage of parsimonious trees in which the associated taxa clustered together are shown next to the branches. The tree was obtained using the Tree-Bisection-Reconnection algorithm in which the initial trees were obtained by the random addition of sequences (1000 replicates). The tree is drawn to scale, with branch lengths calculated using the average pathway method. Isolates are listed with numbers that correspond to Table 2-1.

56

Tree: EF-1α CI=0.95 RI=0.99 Tree length=30

CLADE 1

CLADE 2

CLADE 3

CLADE 4

Fusarium oxysporum = Fusarium oxysporum f. sp. albedinis =

Fusarium oxysporum f. sp. canariensis = Fusarium oxysporum f. sp. elaeidis =

Fusarium oxysporum f. sp. palmarum =

57

Figure 2-4. Unrooted consensus tree inferred from 10 most parsimonious trees using the Maximum Parsimony method based on the PFW primer. Branches corresponding to partitions reproduced in less than 70% trees are collapsed. The percentage of parsimonious trees in which the associated taxa clustered together are shown next to the branches. The tree was obtained using the Tree-Bisection-Reconnection algorithm in which the initial trees were obtained by the random addition of sequences (1000 replicates). The tree is drawn to scale, with branch lengths calculated using the average pathway method. Isolates are listed with numbers that correspond to Table 2- 1.

58

Tree: PFW primer CI=0.93 RI=0.98 Tree length=21

CLADE 1

CLADE 2

CLADE 3

CLADE 4

Fusarium oxysporum = Fusarium oxysporum f. sp. albedinis =

Fusarium oxysporum f. sp. canariensis = Fusarium oxysporum f. sp. elaeidis =

Fusarium oxysporum f. sp. palmarum =

59

Symptompatic Palm Tissue Petiole or Rachis

Isolation of the Pathogen; Morphological Characterization

Agar Plug or Genomic DNA

PCR with PFW Primers

Positive Amplification

Fusarium oxysporum

If necessary, sequence amplicon and conduct BLAST analysis

Fusarium oxysporum f.sp. canariensis Fusarium oxysporum f. sp. palmarum

Figure 2-5. Flowchart of the proposed diagnostic method for identification of Fusarium wilt pathogens on ornamental palms.

60

CHAPTER 3 GENOME CHARACTERIZATION OF Fusarium oxysporum f. sp. palmarum

Introduction

The Fusarium oxysporum species complex is a grouping of closely related asexual, ubiquitous, soil-borne fungi. While, the majority of isolates are non-pathogenic endophytes and saprophytes, several lineages are plant pathogens that can cause either vascular wilt, or bulb, foot and root rot diseases (Michielse and Rep 2009). Vascular wilt results in severe damage on a wide range of economically important plants, affecting both dicots and monocots (Michielse and Rep

2009). Fusarium oxysporum is ranked fifth in the list of top ten fungal plant pathogens of scientific and economic importance (Dean et al. 2012). Collectively, pathogenic members have a broad host range, but individual members exhibit a high degree of host specificity (Armstrong and

Armstrong 1981; Gordon and Martyn 1997) and are classified into formae speciales based on their host range. Each forma specialis is usually polyphyletic (Baayen et al. 2000; O’ Donnell et al. 1998), but a few exhibit monophyletic lineages apparently arising only once in nature

(Schmidt et al. 2013). Studies suggest that most formae speciales may have evolved multiple times through mutations and transposition or acquisition of new traits by horizontal gene transfer

(Baayen et al. 2000; O’Donnell et al. 1998).

Fusarium oxysporum f. sp. palmarum (FOP) and Fusarium oxysporum f. sp. canariensis

(FOC) cause lethal Fusarium wilt on ornamental palms. FOC affects Phoenix canariensis predominantly but can also affect Phoenix reclinata, whereas FOP affects a broad host range that includes Washingtonia robusta, Syagrus romanzoffiana, x Butyagrus nabonnandii and P. canariensis (Elliott 2015; Elliott et al. 2010; 2017; Giesbrecht et al. 2013; Simone 2004). Unlike,

61

FOC which has a worldwide distribution, FOP appears to have originated in and is restricted to

Florida. Regardless of the forma specialis, both of these palm pathogens elicit similar symptoms including typical one-sided wilt and necrosis of leaves with a reddish or brown stripe on the petiole or rachis and discoloration of vascular bundles in the same area as the stripe. With both

FOC and FOP, the palm ultimately dies. The prominent difference between FOC and FOP is accelerated disease progression by FOP, where infected palms are dead within a few weeks to a few months (Elliott et al. 2010). Disease progression is much slower in FOC, requiring a year or two for the infected palms to die (Simone 2004). EF-1α gene based phylogeny suggests that FOC and FOP are two independent lineages that have convergently adapted to palms (Elliott et al. 2010 and this study). Currently, with no treatments or curative treatments available, early detection of the pathogen is key to implementing management strategies and prevent the spread of disease.

For successful host compatibility, fungal pathogens usually secrete a set of pathogenicity factors which are defined in numerous ways. Pathogenicity genes are defined as “essential factors required uniquely or redundantly by the pathogen to cause various aspects of the disease on a host” (Idnurm and Howlett 2001). Pathogenicity factors can be categorized in multiple ways.

They can be broadly classified based on function, including genes involved in autophagy, cell wall degradation, maintaining cell wall integrity, cell signaling, detoxification, fungal nutrition, secondary metabolite biosynthesis and transcription regulation (Idnurm and Howlett 2001).

Pathogenicity factors can also be grouped into basic and specialized.

Studies on Fusarium oxysporum formae speciales reveal a dynamic genome structure partitioned into core and lineage-specific (accessory or dispensable) regions (Ma et al. 2013;

2010; Rep and Kistler 2010). The core genome consists of essential housekeeping genes conserved across F. oxysporum and not involved in pathogenesis (Ma et al. 2010). However, 62

lineage-specific regions are enriched with pathogenicity associated proteins, secretory proteins, secondary metabolites, transposons and repetitive elements, which vary among F. oxysporum formae speciales (Schmidt et al. 2013; Williams et al. 2016).

Basic pathogenicity factors such as cell wall degrading enzymes (e.g., cutinases, cellulases, polygalaturonases, pectate lyases and xylanases) and proteins involved in cellular signaling pathways, including RAS proteins, and mitogen activated protein kinases (MAPK) are conserved and present in Fusarium along with other fungal pathogens (Ma et al. 2013).

The cell wall degrading enzymes are essential as they facilitate fungal adhesion to the cuticle, crossing the cell wall barrier and forming initial infection structures (Beckman 1987; Di

Pietro et al. 2003). Though they are crucial for the infection process, reverse genetic approaches are generally unable to implicate these enzymes directly in virulence, due to their functional redundancy. Signal transduction pathways play a major role in establishing infection by sensing and adapting to plant environment. The signaling genes, like transcription factors, have a pleiotropic effect of cellular processes along with roles in pathogenesis (Xu 2000)

Specialized pathogenicity factors, defined as effectors, include cysteine-rich proteins or secondary metabolites that are secreted into the host to manipulate host metabolism and their immune responses (Presti et al. 2015). These effectors are highly expressed in planta. Typically, effectors reside on lineage-specific genomic regions and are dynamically changing through processes such as horizontal gene transfer, transposon activity, direct mutation and deletion resulting with gain or loss of these effector genes (De Wit et al. 2016; Rep and Kistler 2010).

Moreover, these genes are host specific and might be involved in host adaptation with few or no close homologs with proximal genera (Coleman et al. 2009).

63

Pathogenic Fusarium species are known to produce secondary metabolites (SM) such as polyketides and nonribosomal peptides that have various functions including pigments, alkaloids and toxins that have a role in pathogenesis. SM are structurally and biologically diverse, and the modes of action for most are unknown. When produced in association with a host, SM can have multiple effects on plants including suppression of host defenses, promotion of plant tissue chlorosis or necrosis and induction of host programmed cell death (Brown and Proctor 2013).

Some well-known SM produced by F. oxysporum includes beauvericin, enniatins, fusaric acid, moniliformin, naphthazarins and sambutoxin (Bacon et al. 1996; Herrmann et al. 1996; Kim and

Lee 1994; Rabie et al. 1982; Thrane 2001). In Fusarium, as in most fungi, genes involved in secondary metabolism are physically clustered together. A secondary metabolite cluster includes a biosynthetic gene that is responsible for the core secondary metabolite synthesis, enzymes that modify the core metabolite, transporters that can move the metabolites across the membrane, and specific transcription factors that co-regulate the expression of these genes (Keller et al. 2005).

Identification of SM clusters through genome sequencing predicts higher SM diversity than estimates based on biochemical characterization under laboratory or host interaction conditions

(Bok et al. 2006; Chiang et al. 2011; Hertweck 2009).

Recently, specialized proteinceous effectors referred to as Secreted In Xylem (SIX) proteins were the first identified genuine effectors that are specific to F. oxysporum f. sp. lycopersici (FOL) (Houterman et al. 2009). The SIX gene family is comprised of at least 14 different proteins which are proposed to be involved in colonization and promote the infection process. Besides FOL, SIX genes are present in various F. oxysporum formae speciales, with very few homologs present outside of F. oxysporum. Schmidt et al. (2013) indicated presence of two transposons, upstream (Mimp) and downstream (Fot5), always present within close proximity of

64

SIX genes. Gene knock out experiments have indicated SIX1, SIX3, SIX5 and SIX6 are required for full virulence and characterized as functional effectors in FOL (Gawehns et al. 2014;

Houterman et al. 2009; Ma et al. 2015; Rep et al. 2004). All the SIX gene effectors are regulated by the SIX gene expression (SGE1) transcription factor (Michielse et al. 2009b). Additionally

SIX1, SIX3 and SIX4 are avirulence factors (defeated effectors), which are recognized by R proteins in resistant tomato cultivars resulting in disease resistance (Houterman et al. 2009, 2008;

Rep 2005; Rep et al. 2004). SIX1 and SIX3 have dual functions depending upon the tomato cultivar (Rep et al.2004). Along with FOL, homologs of SIX genes are present in other formae speciales, including Fusarium oxysporum ff. spp. cepae, canariensis, cubense, pisi and vasinfectum, with SIX gene profiles distinct from one another (Chakrabarti et al. 2011; Fraser-

Smith et al. 2014; Laurence et al. 2015; Lievens et al. 2009; Meldrum et al. 2012; Taylor et al.

2016).

Traditionally, the interaction between pathogenicity factors and host are studied through biochemical methods involving protein isolation and purification and reverse genetic approaches such as target gene insertions, deletions and RNA silencing (Hammond and Keller 2005;

Kadotani et al. 2003). Such studies provide valuable information about biological function; however, they are limited to a one gene at a time approach or a few selected model organisms.

Moreover, gene knockout studies are challenging to execute if the effector genes are present in multiple copies (Di Pietro et al. 2003). Low sequencing costs have resulted in an expansion of fungal plant pathogenic genome sequences. The extensive data obtained from sequencing facilitates large-scale comparative genomic studies to identify effector profiles, establishes the molecular framework for host-pathogen interactions, and provide clues into their effector origin and evolution (Guyon et al. 2014; Krijger et al. 2014; Soanes et al. 2008; Williams et al. 2016). In

65

addition, comparative genomics assist in classifying genes that are conserved among species and reveal whether genes are unique to a particular species. As a genus, Fusarium has a large number of sequenced genomes and serves as a model plant pathogenic genus to understand host-plant interactions and their evolutionary history.

The objectives of the present study include: a) to sequence and assemble genomes of FOC

(PLM386B) and FOP (PLM249A); b) to compare genes that are involved in pathogenicity between FOC and FOP genomes with other publicly available Fusarium genomes; c) to compare predicted proteomes of FOC and FOP and determine if observed genetic differences can provide hypotheses for future research to explain the increased host range and virulence of FOP over

FOC.

Material and Methods

Isolates

FOP isolate PLM249A (NRRL53543) isolated from Washingtonia robusta and FOC isolate PLM386B isolated from Phoenix canariensis were selected for genome sequencing. These isolates were obtained from Fusarium wilt symptomatic plants and Koch’s postulates were completed previously to demonstrate their pathogenicity (Elliott et al., 2010; Elliott personal communication). The isolates were maintained as mycelial agar plugs in sterile deionized water in cryogenic vials at room temperature at the University of Florida, Fort Lauderdale Research and

Education Center.

DNA Extraction and Whole Genome Sequencing

The FOC and FOP isolates were revived from storage by placing a single agar plug on 1/5 strength potato dextrose agar (1/5 PDA) (Difco Laboratories, Detroit, MI) and grown for 10 days

66

at 28°C. Ten agar plugs from 1/5 PDA were transferred to a conical flask with 25 ml of full strength potato dextrose broth and incubated for 10 days on a shaker. The fungal mycelial mats were harvested using vacuum filtration through filter paper. The harvested mycelium was frozen using liquid nitrogen and dried by lyophilization. The lyophilized tissue was grounded into a fine powder using a mini bead beater (Cole-Parmer; Vernon Hills, IL) in a 2 ml Eppendorf tube.

Twenty grams of fungal mycelium powder was re-suspended in cell lysis buffer, and DNA was extracted using Gentra Puregene Tissue Kit (Qiagen, Valencia, CA) according to the manufacturer’s instructions. The resulting DNA pellet was suspended in 70 µl of molecular grade water (HyCloneTM, Fisher Scientific, Pittsburg, PA). The purified DNA was treated with 5 units of RNase at 37°C and incubated for 1 hr to degrade RNA (Invitrogen, Thermo Fisher Scientific,

Pittsburg, PA). Genomic DNA was quantified using a Nanodrop (Thermo Fisher Scientific,

Pittsburg, PA), and DNA quality was assessed on 1% agarose gel with ethidium bromide staining and visualized under UV light.

Genome Sequencing

Ion torrent TM (4.4) sequencing was performed on high quality genomic DNA of FOC and FOP at the Interdisciplinary Center for Biotechnology (University of Florida, Gainesville).

Whole genome sequencing resulted in raw reads of 5,814,032 bp and 6,305,746 bp for FOC and

FOP with an average read length of 312 bp and 311 bp, respectively. This provided a total of

1,643 Mbp and 1,691 Mbp of Q20 bases for FOC and FOP, respectively. The raw sequences were analyzed for their quality by FastQC by providing per base and sequence quality score, GC content, and sequence length distribution. De novo genome assembly was done by SPAdes (3.5)

(Russian Academy of Science; St. Petersburg, Russia) with different kmer sizes (K21, 33, 55, 77,

99 and 127). The assembly generated contigs and scaffolds. Contigs less than 0.5 kb were

67

filtered by Artemis (Sanger Institute; Oxford, England) and removed from further analysis. The

FOC and FOP scaffolds were screened for the presence of contaminants using BLAST suite.

GeneMarkTM (Georgia Institute of Technology; Atlanta, GA) was used to predict genes and map the gene sequences to the contigs as “Gene Transfer Format” (.gft) files. These files were read by

Artemis to generate protein fasta files and were verified manually for correct predicted protein translation and presence of exons and stop codons. Functional annotations on FOC and FOP proteomes were performed using Blast2GO (BioBam, Spain), which provides BLAST similarity matches, GO ontology mapping with annotation and InterProScan protein domains. Additional functional annotations, Pfam (database of protein families; Version 31.0) and EuKaryotic

Orthologous group (KOG), which predicts conserved protein function were performed on FOC and FOP proteomes using webMGA (http://weizhong-lab.ucsd.edu).

Benchmarking Universal Single Copy Orthologs

The quality of predicted gene calls from the genome assembly was assessed by

Benchmarking Universal Single-Copy Orthologs (BUSCO) (http://busco.ezlab.org/) based on evolutionary gene content using 1428 universal single copy orthologs on a local server (Simao et al. 2015).

Comparative Genomics to Identify Orthologous Gene Families

Fourteen annotated and closely related Fusarium proteomes were downloaded from the

Broad Institute/MIT (ftp://ftp.broadinstitute.org/pub/annotation/fungi/fusarium/) (Table 3-1). To establish conserved and unique gene families among F. oxysporum formae speciales and within palm-infecting formae speciales FOC and FOP. Two different approaches were used. First, predicted proteins of all 16 genomes were analyzed using “all against all” to find orthologous clusters using OrthoMCL (http://orthomcl.org/orthomcl/) (Li et al. 2003). This program is based 68

on NCBI BLAST and Marko Cluster Algorithm (MCL) (Enright et al. 2002). Analysis was performed on a local server using threshold e-value of 10-5 with percent match length equal or higher than 50% of the query length. Orthologous genes that are present among the Fusarium genomes and genes with no paired orthologs are listed as clusters. Additionally, Reciprocal

Smallest Distance (RSD) was used to find ortholog pairs between two genomes (query and subject) (Wall et al. 2003). This was performed using FOP vs. each of the remaining 15 Fusarium genomes with only one query and subject at any given time. Gene family was classified as expanded only when the genome had twice the number of genes than the average genes present in the cluster.

Identification of Secretory Proteins

The hallmark of canonical secretory proteins is the presence of an N-terminal signal peptide. Signalp 4.1 (http://www.cbs.dtu.dk/services/SignalP) (Petersen et al. 2011) was used to identify the N-terminal sequence of each protein in the predicted proteomes of FOC and FOP on a local server with default settings. The FOP secretome is categorized into two groups based on orthologous relationships estabilished using RSD with FOC: a) secretory proteins that are conserved in palm infecting formae speciales (FOC and FOP) designated as SPIFS; b) secretory proteins that are specific to F. oxysporum f. sp. palmarum (FOP) when compared to FOC designated as SSFOP.

Detection of Carbohydrate Active Enzymes

Enzymes that are involved in cell wall degradation and carbohydrate metabolism were classified in FOC and FOP along with the rest of the Fusarium proteomes.The proteomes were queried with dbCAN HMM 5.0 (Database for Carbohydrate Active Enzyme ANnotation) based

69

on a hidden markov model (HMM) to identify protein domains that in involved in carbohydrate metabolism (http://csbl.bmb.uga.edu/dbCAN/) (Yin et al. 2012).

Detection of Secondary Metabolite Clusters

Secondary Metabolite Unique Region Finder (http://www.jcvi.org/smurf/index.php)

(SMURF) (Khaldi et al. 2010), a web based tool, was used to predict secondary metabolite (SM) clusters. SMURF predicts clusters using Pfam and TigrFam (HMM) domain models. SMURF input files include predicted proteome fasta file and gene coordinate file for each of the 16

Fusarium genomes. The output files for each genome included biosynthetic genes associated with either one of four classes of SM and a second file consisted of physically clustered genes that are predicted to function enzymatically as a transporter or a regulator of each SM gene cluster along with their annotations. The core biosynthetic genes of SM were classified into one of four categories: dimethylally tryptophan synthase (DMAT), polyketide synthase (PKS), non- ribosomal peptide synthase (NRPS) and hybrid (NRPS-PKS). To determine if putative secondary metabolite clusters of FOP were conserved among Fusarium genomes or unique to FOP, a threshold requiring the presence of at least 80% of the SM cluster genes exhibiting synteny within the compared genome was used.

Secreted in Xylem (SIX) Genes in FOC and FOP

To identify putative SIX gene effectors, predicted proteomes and nucleotide contigs of

FOC and FOP were queried. All the GenBank submitted SIX protein sequences were downloaded and formatted as a BLAST database. On a local server, Local–BLAST (BlastP) was performed using proteomes and scaffolds (BlastN) as query to identify homologs with at least a similarity value of 70% and e-value of 10-3. The potential SIX gene candidates were then identified for the presence of signal peptide by SignalP. Further, SIX gene data set alignments 70

using all identified Fusarium sequences were made by MUSCLE (Edgar et al. 2004). These protein alignments was used as queries to identify SIX gene homologs in FOC and FOP proteomes using HMMER (http://hmmer.org/) with default settings (Finn et al. 2015).

Detection of Virulence Genes in FOP and FOC

Virulence genes that have been experimentally verified to have pathogenicity phenotypes are catalogued in Pathogen-Host Interaction–base (PHI-base) (http://www.phi-base.org) (Urban et al. 2017). The collection of 4460 virulence proteins from PHI-base version 4.3 were formatted as a database and Local-BLAST (BlastP) was performed using the proteomes of FOC and FOP as queries to identify putative virulence genes. Proteins with a 90% or higher similarity match and e-value less than 0.001 were selected for further analysis. The putative virulence proteins were categorized based on the phenotype of the experimental mutants: lethal, loss of pathogenicity, and reduced virulence. Putative virulence genes of FOC and FOP were then evaluated across the F. oxysporum formae speciales for presence of orthologs using OrthoMCL.

Results

Genome Sequencing

Using the total number of base pairs in final assembly as a prediction of genome size, the assembled FOC and FOP genomes had at least 40X sequence coverage (Table 3-2). The genome of FOC is assembled into 941 scaffolds with an estimated size of 46.0 Mb. The FOP genome is assembled into 1269 scaffolds with an estimated size of 46.9 Mb. The genome assembly sizes of

FOC and FOP are similar to F. oxysporum f. sp. cubense (46.55 Mb) but smaller than the 54-60

Mb range of F. oxysporum f. sp. melonis and F. oxysporum f. sp. lycopersici (Ma et al. 2014; Ma et al. 2010). Proteome prediction analysis of FOC and FOP genome assemblies resulted in

71

15,528 and 15,974 predicted proteins respectively; slightly less than the published 16,634 -

20,033 range predicted for the three ff. spp.compared above (Table 3-2).

Benchmarking Universal Single Copy Orthologs

Benchmarking Universal Single Copy Orthologs (BUSCO) analysis of the FOC and FOP genomes recovered all but 1 and 2 genes respectively, of the 1438 genes in the BUSCO fungal set

(Table 3-3). For both FOC and FOP, ~97% single copy orthologs were recovered indicating good quality and integrity of the predicted proteins from the assembled FOC and FOP genomes (Table

3-3).

Proteome Analysis of FOC and FOP

The prediction of proteomes for F. oxysporum f. sp. palmarum resulted in 15,974 proteins of which 6,801 were either hypothetical or uncharacterized with no functional annotation based on BlastP descriptors. The predicted proteome of F. oxysporum f. sp. canariensis resulted in

15,528 proteins of which 6,272 proteins were either hypothetical or uncharacterized. The FOP proteome characterization based on broad functional annotation categories are indicated in (Table

3-3). Reciprocal Smallest Distance (RSD) analysis predicted single copy orthologs; 13,661 were orthologous proteins that were shared between the FOC and FOP proteomes (86-88% of predicted proteins from each genome, respectively). Eukaryotic Orthologous Groups (KOG) based protein annotation of the FOP proteome revealed that 10,678 genes have functionally annotated orthologs, with the highest number of proteins involved in cell signaling, secondary metabolites and post translational modification (Figure 3-1). The Pfam domain/family annotation of FOP proteome encoded a high number of cytochrome p450, transporters, transcription factors and transposons (Table 3-4). Within the transporters, major facilitator superfamily (PF07690) and the

ATP binding cassette superfamily (ABC) (PF00005, PF00664, PF01061) were abundant in FOP 72

(Table 3-4). In both FOC and FOP, the most abundant transcriptional families were Zn(2)-Cys(6) binuclear cluster domain (PF00172) followed by Zinc finger C2H2 type (PF00096) (Table 3-4).

Compared with FOC, the FOP proteome had a slightly higher number of catalases, peroxidases, cysteine rich secretory proteins, cytochrome p450, LysM domains, fungal trichothecene efflux proteins and phopshopantetheine attachment sites. However, a fungal specific domain (CFEM), fungal hydrophobins, heme peroxidases, necrosis producing proteins (NPP1) and putative necrosis inducing proteins (Hce2) were present in similar copy numbers in FOC and FOP proteomes (Table 3-5).

Identifying Orthologous Families among Fusarium Genomes

A total of 15,528 and 15,974 predicted proteins from the FOC and FOP genomes, respectively, (this study) and the annotated proteomes of 14 Fusarium genomes that were obtained from the Fusarium Comparative Genome Project were compared using OrthoMCL to determine the presence of common and unique gene family clusters. The publicly curated genomes included 2 Fusarium species, 10 F. oxysporum formae speciales, one F. oxysporum human pathogen isolate and a F. oxysporum biocontrol (nonpathogenic) isolate (Table 3-1). From

25,069 homologous gene clusters identified by OrthoMCL across all 16 Fusarium genomes, a total of 8,876 gene clusters were shared among all 16 Fusarium genomes, suggesting conserved functions. Within these conserved clusters a total 9,782 FOP proteins were present, accounting for

~60% of the predicted FOP proteome. Expansion of gene families pertaining to FOP was observed in only seven clusters: OrthoMCL99, OrthoMCL391, OrthoMCL635, OrthoMCL668,

OrthoMCL1194, OrthoMCL1640, and OrthoMCL3129 (Table 3-7). All the expanded gene clusters are assigned Pfam domains except OrthoMCL668 (Table 3-8). The common domains that were present in clusters include fatty acid and nonribosomal peptide synthetase (PF00550), AMP-

73

binding domain (PF00501) and NRPS condensation domain (PF00668) in OrthoMCL99. Other dominant domains present were: reverse transcriptase (PF00078) in OrthoMCL391 and

OrthoMCL1194; Tc5 transposon (PF03221) in OrthoMCL635; NAD dependent epimerase

(PF01370) in OrthoMCL1640 and GMC oxidoreducatse (PF00732) in OrthoMCL3129.

Interestingly, only 37 single copy gene clusters were specific to FOC and FOP proteomes.

Of the 12 clusters with Pfam domains, the most common are transporters and transcription factors. A total of 396 and 511 specific singleton clusters were found to be unique to the FOC and the FOP predicted proteome, respectively, compared to other Fusarium genomes.

Secretome Analysis

SignalP was utililized to predict the secretome of FOC and FOP. This analysis identified

1488 (FOC) and 1531 (FOP) secretory proteins, which accounts for ~ 9.6% of each total proteome.

The primary goal of further analysis was to identify unique secretory genes of FOP that may be associated with host specificity or aggressiveness. RSD was used to identify orthologous proteins within FOC and FOP and those unique to each genome. The FOP secretory proteins were classified into two categories. The first category, named secretome of palm infecting formae species (SPIFS), consists of 1374 proteins which are conserved between FOC and FOP. Of these

1374 proteins, 1144 proteins were conserved in 12 Fusarium oxysporum formae speciales and the

F. oxysporum biocontrol and F. oxysporum human pathogen isolates. The remaining 230 of

SPIFS secretory proteins are present in FOC and FOP, but their presence varied among the other

F. oxysporum formae speciales. The second category contained predicted secreted proteins specific to FOP (SSFOP). It included 157 proteins which are present in FOP but absent in FOC.

74

Within these 157 proteins, 21 of these secretory proteins are unique to FOP compared with the remaining Fusarium genomes.

Among the 157 proteins of SSFOP, 83 had no Pfam domain/family assigned and were categorized as hypothetical proteins based on BlastP descriptors with no additional functional annotations. The 74 SSFOP proteins with at least one assigned Pfam domain included carbohydrate binding domains, carbohydrate active enzymes, CFEM (a fungal specific protein family with at least 8 cysteine residues), cytochrome p450, cysteine rich secretory proteins, fungal hydrophobins, GLEYA domain, LysM domains, NACHT domains, proteases, secretory lipases, and tannase and feruloyl esterases (Table 3-6).

The 21 SSFOP proteins that are not found in any other studied Fusarium genomes include class V chitinase, cytochrome p450, lyases involved in fructose and mannose metabolism, lipase, tannase and tyrosinase. Among the 21 SSFOP are three interesting proteins: FOP2_14658, which contains a lipocalin domain found in proteins such as kievitone hydratase, and two FOP2_15929 and FOP2_15930 proteins that contain a phosphopantetheine domain, which is involved in fatty acid, polyketide and antibiotic synthesis. About 40% of these proteins are hypothetical proteins.

Carbohydrate Active Enzymes of FOC and FOP

Carbohydrate active enzymes (CAZymes) were identified with the dbCAN tool (Yin et al .

2012) using predicted proteomes of FOC and FOC. A total of 920 (FOC) and 967 (FOP)

CAZymes were predicted. The predicted CAZymes of FOC were divided into six different categories. These include 125 auxiliary activity (AA), 95 carbohydrate binding modules (CBM),

193 carbohydrate esterases (CE), 363 glycoside hydrolases (GH), 118 glycoside transferases (GT) and 26 pectate lyases (PL). The predicted CAZymes for FOP include 135 FOP auxiliary activity

75

(AA), 106 carbohydrate binding modules (CBM), 199 carbohydrate esterases (CE), 383 glycoside hydrolases (GH), 119 glycoside transferases (GT) and 25 pectate lyases (PL) (Figure 3-2).

Comparative analysis of CAZymes among the 16 Fusarium genomes indicate that F. graminearum (708) and F. verticillioides (878) have fewer predicted CAZymes than F. oxysporum genomes (1,040). Interestingly, F. oxysporum biocontrol strain Fo47 had the maximum number of predicted CAZymes (1210).

A comprehensive analysis of selected CAZymes subfamilies that are involved in cell wall degradation was performed (Table 3-9). CAZymes subfamilies which are involved in plant polysaccharide degradation of cellulose include (GH1, GH3, GH5, GH6, GH7, GH12 and GH45), hemicellulose (GH10, GH11, GH27, GH30, GH35, GH36, GH43, GH51, GH54, GH62, GH67,

GH74, GH115) and pectin degrading enzymes (GH28, GH78, PL1, PL3, PL4, PL9, PL11).These are uniformly distributed cross all Fusarium genomes (Table 3-9). No major expansion of any

CAZymes family was observed in FOC and FOP genomes or between monocot and dicot pathogens nor among F. oxysporum formae speciales.

Detection of Secondary Metabolite Clusters

Fusarium species produce low molecular weight molecules such as pigments, antibiotics and metabolites that can contribute to pathogenicity (Ma et al. 2013). Secondary metabolite unique regions finder (SMURF) identified a total of 39 and 38 putative secondary metabolite core biosynthesis genes resulting in 33 and 30 SMURF clusters within FOP and FOC, respectively.

The genes that were associated with gene clusters were predicted to function in synthesis, transport and regulation of SM. Compared to other Fusarium genomes, FOC and FOP along with

76

F. oxysporum ff. spp. lycoperisci, Fo5176 and melonis have a higher number of predicted SM gene clusters (Figure 3-3).

The 33 FOP SM clusters were distributed into four categories: 3 dimethylallyltryptophan synthases (DMAT), 16 nonribosomal peptide synthases (NRPS), 11 polyketide synthases (PKS) and 3 hybrid NRPS-PKS. Comparative genomic analysis using FOC and the other 14 Fusarium genomes was performed to identify which SM clusters are unique to FOP. Comparative genomics indicated three non-ribosomal peptide synthase (NRPS) and one polyketide synthase (PKS) clusters unique to FOP (Table 3-10). Along with the biosynthetic gene, SM clusters generally include cytochrome p450, transporters, transcriptional factors and several hypothetical or uncharacterized proteins. Only six of 33 secondary metabolite clusters have been identified using

BlastP designators and Pfam annotations, including bacitracin, bivakerin, enniatin, mycocerosic acid and two siderophores. The products of the remainder of the secondary metabolite clusters are unknown.

The OrthoMCL99 cluster, one of seven expanded gene familes of FOP, consists of 13 FOP proteins compared to an average of six in all other Fusarium genome (Table 3-7). Six of the thirteen proteins (FOP2_06155, FOP2_07629, FOP2_11142, FOP2_12983, FOP2_14907, and

FOP2_15585) are core biosynthesis genes belonging to non-ribosomal peptide synthase (NRPS).

The remaining seven proteins within the cluster have domains related to phosphopantetheine attachment sites and condensation domains. The BlastP descriptors for seven of these 13 proteins are related to nonribosomal peptide synthase of Fusarium proliferatum.

77

Secreted in Xylem (SIX) Gene Family

The FOC and FOP proteomes were screened for the presence of SIX protein homologs using the GenBank curated SIX proteins from Fusarium oxysporum formae speciales. A local-

BLAST (BlastP) on the FOP and FOC predicted proteomes, queried with all available SIX proteins from NCBI GenBank and their alignments using a cutoff of 90% similarity and a minimum e-value of 0.001, predicted the presence of SIX7 and SIX10 in FOC and SIX9 in FOP

(Table 3-11). Screening of FOC and FOP nucleotide scaffolds with BlastN resulted in the further detection of SIX12 in FOC and of SIX8 in FOP. Only the SIX10 from FOC (FOC3_15235) had a predicted signal peptide. The SIX12 is the only protein with no N- terminal signal peptide

(Schmidt et al. 2013; Van Dam et al. 2016). Additionally, Schmidt et al. (2013) indicated that

SignalP failed to identify N-terminal signal peptide in SIX5, SIX7, SIX8 and SIX10 due to presence of short exon in the proteins. PCR screening for all SIX genes (SIX1-14), based on published primers (Taylor et al. 2016; Lievens et al. 2009), confirmed the presence of the SIX 7,

SIX10 and SIX12 genes in FOC and SIX8 and SIX9 in FOP (see Chapter 4). Each of the SIX gene orthologs identified in FOC and FOP assemblies were determined to be single copy genes based on local BlastN/BlastP and OrthoMCL analysis.

Virulence Genes

As a final assessment to determine if the FOC and FOP genomes encoded conserved proteins with functionally defined roles in virulence, genes from the pathogen-host interactions (PHI Base) database were used to query FOP and FOC genomes by BlastP. The analysis identified a total of

237 and 223 putative virulence associated genes in FOC and FOP, respectively (Table 3-12). A total of 114 FOP proteins were related to pathogenicity and are classified arbitrarily according to their putative function in the disease process. A total of 21 putative virulence associated genes

78

that were well characterized in Fusarium oxysporum are present in FOC and FOP, including argininosuccinate lyase (ARG1), Class II chitin synthase (CHS2), Chaperon like protein (CHS7),

Carboxy,cis-cis muconate cyclase (CMLE), Dnj1, F-box protein (Fbp1), G-protein α-subunit

(FGA1), G protein α-subunit (FGA2), Mitogen-activated protein kinase (Fmk1), MAP kinase

(FoBck1), putative a-1,6-mannosyltransferase (FoOCH1), MAP kinase pathway (FoSlt2), protein kinase (foSNF1), Mitochondrial carrier (FOW1), transcription factor (FOW2), F-box protein

(FRP1), fungal transcription factor (FTF2), β-1,3 Glucanosyltransferase (GAS1), transcription factor for SIX gene expression (SGE1), and BAH/PHD-containing transcription factor (Snt2)

(Table 3-13). The 114 FOP putative virulence gene orthologs are found in Fusarium graminearum (68), F. oxysporum (21), F. verticillioides (6) and Verticillium dahlia (1). A putative virulence regulating gene (ARSEF 2860) involved in Ras/MAPK signaling pathway identified in Beauveria bassiana was present in palm-infecting formae speciales (FOC and FOP) but absent in other Fusarium genomes. A transcription factor present in F. graminearum

(FgSW16) has an ortholog only in FOP and was absent in the remaining Fusarium oxysporum genomes.

Discussion

In Florida, Fusarium wilt caused by F. oxysporum f. sp. canariensis (FOC) and F. oxysporum f. sp. palmarum (FOP) is killing ornamental palms, and there are no fungicidal treatments currently available. These two formae speciales are similar in several aspects, yet they are distinct. FOP appears to be more virulent and has an expanded host range when compared to

FOC. In the present study, we sequenced and performed a de-novo draft assembly for FOC and

FOP and performed a comparative analysis to determine if unique gene content in the FOP genome could provide insight into its aggressiveness.

79

The genome size of FOP (46.9 Mb) is slightly larger than FOC (46.0 Mb). Indeed, the

FOP genome was predicted to encode 446 more genes than FOC. The initial hypothesis for the larger genome size of FOP was the presence of novel genes or an expansion of gene families compared to FOC. Initial proteome analysis of FOP revealed enriched copies of cytochrome p450, transcription factors, transporters and transposons with suggested roles in pathogen adaption to the host environment, import/export of metabolites and regulation, but these difference were not significant enough to explain the difference in genome size.

Although there are more than 100 known transporter families, we focused only on three transporter families for their importance in fungal pathogenesis. ATP-binding cassette (ABC) are active transporters which can transport a wide range of compounds such as sugar, heavy metals, and secondary metabolites. Along with ABC transporters, major facilitator superfamily (MSF) was the most abundant transporter in FOC and FOP proteomes in comparative accordance with the FOL genome (Ma et al. 2010). MSF is also the common transporter present in secondary metabolite clusters and might be involved in secondary metabolite efflux and secretion of endogenously produced toxins. Third subfamily, pleiotropic drug resistance (PDR) protein transporters, are involved in tolerance to plant secondary metabolites and toxic compounds such as cycloheximide. PDR transporters are present in similar range in FOC and FOP (20) but there are fewer compared to FOL (25) (Ma et al. 2010).

Transcription factors (TF) are extensively studied in fungi and known to regulate both essential functions such as survival and reproduction along with roles in pathogenesis. Fungal

Zn(2)-Cys(6) binuclear cluster domain is the main TF family present in abundance in ascomycetes and plays an important role in primary and secondary metabolism and drug resistance (Todd et al. 2014). Proteomes of FOC and FOP have expanded copies of Zn2Cys6 80

domains (~400) compared to FOL and F. oxysporum f. sp. cubense which are in the range of

~350 (Guo et al. 2014; Ma et al. 2010). However, other TF families, such as bZIP and zinc finger

C2H2 type and Myb DNA binding domains, are fewer in copy number in FOC and FOP as compared to FOL. Additionally, the host-pathogen interaction database (PHI Base) had identified 30 TF that have role in a virulence. Among them are three well characterized homolog transcriptional factors in FOP genome, including FOW2 present in F.oxysporum f. sp. melonis and FTF1 in F.oxysporum f. sp. phaseoli and SGE1 in F.oxysporum f. sp. lycopersici which are extensively studied for their role in pathogenicity (Brown et al. 2014; Michielse et al. 2009b;

Ramos et al. 2007). The FOW2 and FTF2 genes conclude Zn(II)2 Cys6 type transcription regulators (Michielse et al. 2009a).

In reponse to pathogens, plant cells produce a variety of antifungal and defense compounds along with oxidative burst to combat pathogens. Cytochrome p450s play an essential role in fungal secondary metabolism and detoxification of toxic compounds (Cresnav and Petric

2011). Cytochrome p450s were present in higher abundance in FOP (190) than F. oxysporum f. sp. cubense (170) and FOC (155). FOP has a slightly higher copy number of catalase and peroxidase-encoding genes [FOP (28) vs. FOC (22) vs F. oxysporum f. sp. cubense (25)], and

FOP might have the edge in combating oxidative stress and detoxification of host defenses

(Welinder 1991).

F. oxysporum genomes can be rich in transposable elements (TEs) with some genomes containing as much as 4% of the genome content as TEs (Daboussi and Capy 2003; Ma et al.

2010). TEs are known to cause phenotypic changes by causing duplications, translocations, and deletions. In the absence of meiosis, TEs can play a major role in generating F. oxysporum genetic diversity. Annotation with Pfam predicted a higher number of transposons in FOP (49) 81

versus FOC (28). The expanding host range of FOP could be derived from transposons to some degree in that proximity of transposons to the vicinity of fungal effectors genes may help in rapid host adaptation to new hosts in response to disease pressure (Van de Wouw and Howlett 2010).

OrthoMCL analysis indicated 90% of the proteomes are highly conserved between FOC and FOP with only 511 clusters specific to FOP. Since the goal was to identify the genes which may explain FOP disease aggressiveness, we focused only on two clusters types. First, gene clusters representing protein families that are expanded in FOP. Secondly, gene clusters that are specific to only palm-infecting formae speciales (FOC and FOP). OrthoMCL revealed the presence of at least seven gene families that are expanded in FOP when compared to the other examined 15 Fusarium genomes. Of the Pfam domains assigned to these gene clusters, orthoMCL99 was interesting due to presence of domains involved in secondary metabolites, with six of 13 proteins being nonribosomal peptide synthetase biosynthetic enzymes and the remaining seven related to nonribosomal peptide synthetases. OrthoMCL391and

OrthoMCL1194 contained domains for reverse transcriptase, which is involved in retrotransposons replication. OrthoMCLcluster 635 has three proteins with Tn5 transposon domain and OrthoMCL3129 is involved in oxidase reductase activity.

Additionally, 37 single copy gene clusters were specific to palm-infecting formae speciales. Twenty-seven percent of these clusters were annotated as hypothetical proteins with no assigned Pfam domain, and the remaining 73% were associated with putative transcription factors and enzymes, which are not novel proteins but extra copies of proteins already present in other F. oxysporum genomes. The hypothetical proteins present in the single copy OrthoMCL clusters that are specific to palm-infecting formae speciales need to be further annotated to evaluate their role in host adaptability to palms and narrow the choices for functional analysis. 82

To narrow the pool of putative FOP pathogenicity genes, we focused on predicted secretory proteins, as they play essential roles in pathogenicity including cell wall degradation, acquiring nutrients, suppressing host defenses and cell signaling by responding to the host environment (Giraldo and Valent 2013). FOP secretory proteins constituted roughly 10% of the proteome.

Members of FOP secretome were classified into two categories. Secretory proteins that are conserved between FOC and FOP (SPIFS) and secretory proteins that are specific to FOP

(SSFOP). To achieve our goal to identify lineage-specific secretory proteins, we focused on the

157 SSFOP proteins. Out of these, only 73 proteins were assigned with at least one Pfam domain. The BlastP designators for the majority of these proteins are extra copies of catalase- peroxidases, lipases, cytochrome P450, LysM domains, kinases, pectate lyases, small-secreted protiens, tannase, tyrosinases and hypothetical proteins. Interesting, FOP2_14658 encodes kievitone hydratase. The kievitone hydratase was first detected in F. solani f. sp. phaseoli (Li et al. 1995). This enzyme neutralizes a phytoalexins kievitone produced by beans suggesting its role in virulence. No homology of this gene was found in other F. oxysporum isolates (Li et al.

1995).

We hypothesize that 84 FOP secretory genes with no homology are likely to be involved in host specificity. In plant pathogenic fungi, proteins lacking known domains generally are cysteine rich, are required for stabilization of their proteins and might contribute to pathogenicity

(Sperschneider et al. 2015). Moreover, such proteins help the pathogen to avoid recognition by different hosts. However, host specificity may also include divergence of orthologous proteins that might differ in their interaction with host proteins or through the regulation of conserved orthologous proteins. Transcriptomic characterization of the host-pathogen interaction dynamic 83

may help to further narrow the list of potential virulence effectors but will be challenging due to the woody nature of the palm host tissue.

Important proteins in the SPIFS category present in multiple copies with known roles in fungal pathogenicity include CFEM, cysteine-rich secretory proteins, cytochrome p450, fungal hydrophobins, LysM domains, necrosis-inducing proteins (NPP1) and putative necrosis-inducing factor (Ecp2). The CFEM proteins are fungal-specific cysteine-rich domains (8 residues)

(Kulkarni et al. 2003). A well known example of CFEM domain proteins is Pth11 in

Magnaporthe grisea, which has important roles in fungal pathogenesis relating to appressorium function (Kou et al. 2016). The number of FOC, FOP and F. oxysporum f. sp. cubense CFEM domains (20 each) are within similar range as FOL (25) and F. oxysporum f. sp. ciceris (19)

(William et al. 2016). The FOP proteome revealed the presence of five copies of necrosis- inducing protein NPP1 domain. The NPP1 domains are present in necrosis and ethylene inducing proteins, which are called necrosis inducing-like proteins (NLPs). The NLPs were first identified in Fusarium oxysporum and are involved in pathogenicity in dicot but not monocot hosts (Bailey

1995). Reverse genetic studies on pathogens which infect monocot hosts and possess NLPs genes, including Mycosphaerella graminicola (1), Magnaporthe oryzae (4) and Verticillium dahliae (1), indicated NLPs are not major virulence determinants on monocot hosts (Feng et al.

2014; Veit et al. 2001). Although NLP encoding genes are present in monocot infecting pathogens, the presence of five copies in FOC and FOP, as well as F.oxysporum f. sp. cubense, indicates they may have functional consequence in these two pathogens adaption to palms or their aggressiveness. They should be experimentally evaluated to understand their role in palm- infecting formae speciales.

84

The Ecp2 necrosis-inducing factor is an effector gene first identified in the tomato pathogen Cladosporium fulvum with a conserved function on dicot and monocot hosts. The Ecp2 proteins are involved in pathogenicity by binding to host target cells and later induce cell necrosis (Lauge et al. 1997). Two homologs are present in F.graminearum, which are upregulated during the early infection stage, suggesting their role in fungal pathogenesis; three homologs are present in F. oxysporum f. sp.cubense, FOC, FOL and FOP (Lu and Edwards

2016; this study).

LysM domains suppress chitin-triggered immunity by protecting fungal hyphae against host defenses or sequestering cell wall fragments from host detection (Kombrink and Thomma

2013). Gene expression studies on F. oxysporum f. sp. medicaginis proteins indicated proteins with LysM domains are upregulated during infection suggesting their contribution in pathogenesis (Williams et al. 2016). FOP has nine copies of LysM domains compared to six copies in FOC.

To address fundamental differences between FOC and FOP, analysis of carbohydrate active ezymes (CAZymes), screening of secondary metabolite and SIX effector genes were performed. CAZymes analysis was performed on all 16 Fusarium genomes. Comparative genomics of CAZymes indicates that CAZymes families were conserved among Fusarium genomes, and this study revealed the absence of duplication or novel CAZymes gene families in

FOP. Palm-infecting formae speciales FOC and FOP had a lower range of predicted CAZymes

(~940) than other F. oxysporum formae speciales (~1000). Important cell wall degrading enzymes such as cellulases, cutinases, pectin lyases and xylanases are highly conserved across the Fusarium genomes and may contribute to pathogenicity across F. oxysporum. Plant host did not have an effect on CAZymes profiles, but monocot pathogens (F. oxysporum ff. spp. 85

canariensis, cubense and palmarum) tend to have a lower range of CAZymes compared to dicot pathogens that might be attributed to cell wall structure (Zhao et al. 2014). The biocontrol F. oxsyporum isolate (Fo47) has the highest and human pathogen isolate had the lowest CAZymes prediction, indicating saprophytes possess a high number of CAZymes as noted previously (Zhao et al. 2013).

Secondary metabolites (SM) are produced by fungi in response to plant interaction, environmental stress and development and play a vital role in pathogenesis (Fountain et al.

2014). SMURF annotation predicted 33 SM clusters in FOP, of which 27 are shared with FOC.

NRPS and PKS are multi-modular enzymes frequently present in fungal genomes. FOP had the higher potential nonribosomal peptide synthetases (NRPS) (20 genes) compared to an average of

17 in Fusarium oxysporum. DMAT and hybrid NRPS-PKS clusters are conserved in both palm- infecting formae speciales. However, FOP has four NRPS and two polyketide syntheses (PKS) that are absent in FOC. Comparative genomics indicates three of these NRPS are unique to FOP and absent from all other Fusarium genomes and might produce novel secondary metabolites.

Shared SM clusters with FOC included two siderophores and one each for the production of bikaverin, enniatin, and mycocerosic acid; however, the products of the majority of SM clusters and their function are unknown. The bikaverin cluster is well characterized in Fusarium fujikuroi, which is involved in mycelial pigmentation (Wiemann et al. 2009). The expansion of an OrthoMCL99 consisting of protein domains involved in secondary metabolism indicates a higher potential of SM production in FOP compared to FOC. The two siderophores produced by

NRPS along with putative virulence gene siderophore biosynthesis gene (SID1) identified by

PHI-base are involved in iron uptake and essential for virulence in several fungi like Cochilobus heterostrophus, C. miyabeanus and F. graminearum (Oide et al. 2006). Together we can

86

conclude that a higher potential for secondary metabolite production might play a role in the expanded and more aggressive nature of FOP host-plant interactions compared to FOC.

A total of 114 FOP genes homologous to those curated by the pathogen-specific PHI-

Base database were identified as virulence associated genes. Twenty-one of these virulence associated genes are well studied and previously identified in F.oxysporum genomes (review by

Michielse and Rep 2009). The putative virulence-associate genes are mainly involved in signal transduction, fungal cell wall integrity, transcriptional regulators and detoxification, which are mostly related to Fusarium species. Peroxisomes are involved in β-oxidation of fatty acids, and detoxification like reactive oxidative species (ROS) are known to play a role in virulence in pathogenic fungi such as Colleotrichum lagenarum, Magnophthora grisea and F. oxysporum.

The F. oxysporum possess four different PEX genes including PEX1, PEX10, PEX12, and

PEX26, which are required for host infection (Michielse and Rep 2009). But the predicted proteomes of FOC and FOP indicate the presence of only PEX12 plus two virulence associated genes PEX5 and PEX6. PEX5 and PEX6 have homologs in F. gramineaurm (Min et al. 2015). A functional annotation of PEX5 and PEX6 needs to be explored to determine if these virulence genes are specific to monocots.

The final category of pathogenicity factors that were examined for comparison between

FOP and FOC were the F. oxysporum specific effectors called secreted in xylem (SIX) genes.

These are well studied in F. oxysporum, known for their xylem colonization and for specific roles in host colonization. Whole genomic study indicated the presence of SIX8 and SIX9 homologs in FOP but not in FOC. Likewise SIX7, SIX10, and SIX12 were found to be present in

FOC but not in FOP. That these two palm-infecting formae speciales have different SIX profiles

87

is consistent with their convergent evolution for the pathogenicity on palm (Elliott et al.2010; this study).

F. oxysporum formae speciales that infect multiple hosts within the same family tend to have similar SIX gene profile. For example, four cucurbit infecting formae speciales (F. oxysporum ff. spp. cucumerinum, niveum, melonis, radicis-cucumerinum) share similar SIX gene profiles (SIX1, SIX6, SIX13) and three legume-infecting formae speciales (F. oxysporum ff. spp. ciceris, medicaginis and pisi) have distinct profiles, but share SIX13 gene among them (Van

Dam et al. 2016; Williams et al. 2016). SIX profiles of palm-infecting formae speciales are distinct. This lack of similarity in their SIX profiles suggests that FOP and FOC have gained host specificity to palms through convergent evolution consistent with phylogenetic studies (Elliott et al. 2010; this study). Biological functions for SIX7, SIX10, and SIX12 are unknown. SIX genes present in other formae speciales indicate that they might not be the only determinants of host specificity but complement other pathogenic genes like the secretory proteins that are unique to

FOP (Jelinksi et al. 2017).

At least eight different copies and three variants of SIX8 are observed in F. oxysporum f. sp. lycopersici and F. oxysporum f. sp. cubense isolates (Schmidt et al. 2013; Fraser-Smith et al.

2014), but our study indicates the presence of a single copy of SIX8 in FOP. Functional annotations should be performed to provide insights into SIX8 interactions with host proteins.

These comparative genomic results indicate that palm-infecting formae speciales of F. oxysporum share a large amount of gene content and conserved pathogenicity factors. The genome size, CAZymes and homologs of putative virulence genes (PHI-base) all suggest that

FOC and FOP genomes share more similarities with F. graminearum and F. verticilliodes

88

genomes than other F. oxysporum formae speciales. FOP is distinct due to the presence of four novel secondary metabolites, the presence of SIX8 and SIX9, higher copy number of detoxifying enzymes, and expansion of transposons. We hypothesize a greater role for secondary metabolites in FOP virulence compared to FOC. Targeting three novel SM biosynthetic gene clusters in FOP for product identification and requirements for pathogenicity, aggressiveness or host range is warranted. Additional studies should include transcript based protein annotations, improvements of genome assembly and a more comprehensive study of the FOC genome structure to shed greater light on the evolution of palm host adaptation.

89

Table 3-1. List of the Fusarium genomes used in the study. Fusarium genome NRRL# Strain Host(s) Locus Source prefix F. oxysporum f. sp. canariensis PLM386B Phoenix canariensis FOC3 This study Phoenix reclinata F. oxysporum f. sp. palmarum 53543 PLM249A Syagrus romanzoffiana FOP2 This study Washingtonia robusta Phoenix canariensis X Butyagrus nabonnandii F. graminearum 31084 PH-1 Triticum FGSG Broad Institute F. verticillioides 20956 7600 Zea mays FVEG Broad Institute F. oxysporum f. sp. cubense 54006 II5 Musa FOIG Broad Institute F. oxysporum f. sp. conglutinans 54008 PHW815 Brassica/Arabidopsis FOPG Broad Institute F. oxysporum f. sp. melonis 26406 Cucurbita FOMG Broad Institute F. oxysporum f. sp. lycopersici 4287 Lycopersicum FOXG Broad Institute F. oxysporum f. sp. lycopersici 54003 MN25 Lycopersicum FOWG Broad Institute F. oxysporum f. sp. pisi 37622 HDV247 Pisum FOVG Broad Institute F. oxysporum f. sp. raphani 54005 PHW815 Raphanus/Arabidopsis FOQG Broad Institute F. oxysporum f. sp. radicis-lycopersici 26381 CL57 Lycopersicum FOCG Broad Institute F. oxysporum f. sp. vasinfectum 25433 Gossypium FOTG Broad Institute F. oxysporum 32931 Human FOYG Broad Institute F. oxysporum 54002 FO47 Biocontrol FOZG Broad Institute F. oxysporum 5176 Fo5176 Brassica FOXB Broad Institute

90

Table 3-2. Genomic statistics of F. oxysporum f. sp. canariensis and F. oxysporum f. sp. palmarum with selected Fusarium oxysporum formae speciales. Genome features Fusarium oxysporum formae speciales canariensisa palmaruma cubenseb melonisc lycopersicib

Strain PLM386B PLM249A II5 26406 4287 Genome Size (Mb) 46.0 46.9 46.55 54.03 59.9 Coverage 41 42 100 55 Scaffolds 941 1269 418 1146 114 Predicted genes 15,528 15,974 16,634 17,735 20,033 aThis study. bMa et al. 2010. cMa et al. 2014.

91

Table 3-3. Comparison of functional annotations among Fusarium oxysporum f. sp. canariensis and Fusarium oxysporum f. sp. palmarum. Parameters Prediction F. oxysporum f. F. oxysporum f. tools sp. canariensis sp. palmarum Total genes GeneMark 15,528 15,974 Secretory proteins SignalP 1,488 1,531 Single copy orthologs BUSCO 1401 1408 Carbohydrate active enzymes (CAZymes) dbCAN 920 967 Secondary metabolite backbone genes SMURF 38 39 Hypothetical proteins BLASTP 6,272 6,801 Transposons PFam 25 49 Virulence associated genes PHI-Base 237 223

92

Table 3-4. Important proteins grouped by Pfam domain/families based on Fusarium oxysporum f.sp. canariensis (FOC) and Fusarium oxysporum f. sp. palmarum (FOP) predicted proteomes. Protein categories FOC FOP Transporters: Major facilitator superfamily 635 634

ATP binding cassette superfamily (ABC) 257 256

CDR ABC transporters 22 20

Transcription factor families: bZIPS transcription factor 14 12

Zn(2)-Cys(6) binuclear cluster 412 404

Zinc Finger C2H2 type 55 58

Transposons: Hat family C-terminal dimerization region 5 18

MULE transposase 7 6

Retro-transposon 1 2

Transposase 1 2

Tc5 transposase 12 18

Transposase IS4 3 3

93

Table 3-5. Comparison of important PFAM domains/families between Fusarium oxysporum f. sp. canariensis (FOC) and Fusarium oxysporum f. sp.palmarum (FOP). Pfam domain/family FOC FOP Catalase (PF00199) 7 9

Peroxidase (PF00141) 8 12

Heme Peroxidase (PF03098) 7 7

CFEM (PF05730) 20 20

Cysteine rich secretory proteins (PF00188) 5 8

Cytochrome p450 (PF00067) 155 190

Fungal hydrophobin (PF06766) 3 3

LysM domain (PF01476) 6 9

Necrosis inducing proteins (PF-5630) 5 5

Putative necrosis inducing factor (PF14856) 3 3

Fungal trichothecene efflux (TRI 12) (PF06609) 59 62

Phosphopantetheine attachment site (PF00550) 53 62

94

Table 3-6. The common Pfam domains that are present in the secretome of Fusarium oxysporum f. sp. palmarum. PFAM domains Name of the domain Domain description PF10528 GLEYA domain carbohydrate-binding domain that is found in fungal adhesins

PF00150 Cellulase degradation of cellulose

PF05730 CFEM domain eight cysteines with proposed roles in fungal pathogenesis

PF00187 Chitin recognition protein protein domain found in carbohydrate active enzymes

PF00188 Cysteine-rich secretory protein family proteins rich in cysteine

PF00067 Cytochrome P450 involved in the oxidative metabolism of a high number of natural compounds

PF01185 Fungal hydrophobin low-molecular-weight, cysteine-rich, hydrophobic proteins

PF11951 Fungal specific transcription factor domain transcription factors

PF06985 Heterokaryon incompatibility protein (HET) heterokaryon incompatibility proteins

PF01476 LysM domain involved in binding peptidoglycan in bacteria and chitin in eukaryotes

PF07690 Major Facilitator Superfamily transporter

PF05729 NACHT domain apoptosis protein

PF00544 Pectate lyase enzyme involved in the maceration and soft rotting of plant tissue

PF00550 Phosphopantetheine attachment site biosynthesis of metabolites such as polyketides, and nonribosomal peptides

PF03583 Secretory lipase secreted during the infection cycle of these pathogens

PF01822 WSC domain involved in carbohydrate binding (consists of eight cysteine residues)

95

Table 3-7. Fusarium oxysporum f. sp. palmarum expanded gene clusters based on orthoMCL comparison with 15 Fusarium genomes. Fusarium Genome 99a 391 635 668 1194 1640 3129

F. graminearum 9 0 0 1 0 1 1 F. verticillioides 7 0 0 5 0 1 1

Fusarium oxysporum f. sp. canariensis 6 9 5 2 2 1 1 f. sp. palmarum 13 9 6 6 7 4 4 f. sp. cubense 5 3 2 1 0 1 1 f. sp. conglutinans 6 4 0 3 2 2 1 f. sp. lycopersici (NRRL4287) 5 6 15 5 5 2 2 f. sp. lycopersici(MN25) 5 5 1 2 1 3 1 f. sp. pisi 7 4 5 2 4 1 1 f. sp. radices-lycopersici 5 1 2 2 0 3 2 f. sp. raphani 6 4 4 4 5 1 1 f. sp. vasinfectum 6 5 1 6 4 3 1

F. oxysporum Fo5176 5 4 1 4 3 2 1 F. oxysporum (Human) 1 1 3 1 0 1 1 F. oxysporum (Fo47) 5 5 2 2 3 3 3 aNumbers refer to OrthoMCL cluster.

96

Table 3-8. Expanded gene families of Fusarium oxysporum f. sp. palmarum with genes and their Pfam domain description. OrthoMCL clusters FOP2 proteins Pfam domains Description of the Pfam domains OrthoMCL99 FOP2_06155 PF00501,PF00550,PFOO668 AMP-binding, Condensation,PP-binding FOP2_07629 PF00501,PF00550,PFOO668 AMP-binding, Condensation,PP-binding FOP2_11142 PF00501,PF00550,PFOO668 AMP-binding, Condensation,PP-binding FOP2_12983 PF00501,PF00550,PFOO668 AMP-binding, Condensation,PP-binding FOP2_14907 PF00501,PF00550,PFOO668 AMP-binding, Condensation,PP-binding FOP2_15585 PF00501,PF00550,PFOO668 AMP-binding, Condensation,PP-binding FOP2_15739 PF00550,PFOO668 Condensation,PP-binding FOP2_15785 PF00501,PFOO668 AMP-binding, Condensation, FOP2_15786 PF00501,PFOO668 AMP-binding, Condensation FOP2_15827 PF00668 Condensation FOP2_15842 PF00501 AMP-binding FOP2_15843 PF00501 AMP-binding FOP2_15850 PF00668 Condensation

OrthoMCL391 FOP2_14572 na na FOP2_14573 na na FOP2_14946 PF00078 Reverse transcriptase FOP2_15603 PF00078 Reverse transcriptase FOP2_15630 PF00075 RNase H FOP2_15715 PF00078 Reverse transcriptase FOP2_15720 PF00078 Reverse transcriptase FOP2_15754 PF00078 Reverse transcriptase FOP2_15805 PF00075 RNase H

OrthoMCL635 FOP2_14599 PF03184 DDE superfamily endonuclease FOP2_14668 PF05225 helix-turn-helix, Psq domain FOP2_15241 PF03184 DDE superfamily endonuclease FOP2_15257 PF03221 Tc5 transposase DNA-binding domain FOP2_15659 PF03221 Tc5 transposase DNA-binding domain FOP2_15688 PF03221 Tc5 transposase DNA-binding domain

97

Table 3-8. Continued OrthoMCl clusters FOP2 proteins Pfam domins Description of the Pfam domains OrthoMCL668 FOP2_15181 na na FOP2_15463 na na FOP2_15476 na na FOP2_15539 na na FOP2_15658 na na FOP2_15763 na na

OrthoMCL1194 FOP2_15083 PF14529 Endonuclease-reverse transcriptase FOP2_15084 na na FOP2_15351 PF00487 Fatty acid desaturase FOP2_15353 PF00078 Reverse transcriptase FOP2_15456 na na FOP2_15751 PF00078 Reverse transcriptase FOP2_15781 na na

OrthoMCL1640 FOP2_05604 PF00501, PF01370 AMP-binding enzyme ;NAD dependent FOP2_14980 PF00501 epimerase/dehydratase family FOP2_14981 PF01202, PF01370 AMP-binding enzyme FOP2_12623 PF01370, PF07993 Shikimate kinase; NAD dependent epimerase/dehydratase family NAD dependent epimerase/dehydratase family; Male sterility protein

OrthoMCL3129 FOP2_14008 PF00732; PF05199 GMC oxidoreductase,GMC oxidoreductase FOP2_14526 PF00732,PF05199 GMC oxidoreductase,GMC oxidoreductase FOP2_14992 PF00732,PF05199 GMC oxidoreductase,GMC oxidoreductase FOP2_14997 PF00732,PF05199 GMC oxidoreductase, GMC oxidoreductase

na= No pfam domains are assigned.

98

Table 3-9. Carbohydrate active enzymes (CAZymes) families involved in cellulose, hemicellulose and pectin degradation in Fusarium genomes.

m

cillioides

4287 Type of

polysaccharide almaru

Subfamily p canariensis conglutinans cubense lycopersici NRRL lycopersici MN25 melonis radicis lycopersici pisi raphani vasinfectum Fo5167 Fo47 Fohuman F.graminearum F.verti Cellulose GH1 6 6 6 6 6 12 6 6 7 5 6 5 6 5 3 6 GH3 33 28 36 27 34 64 37 36 38 39 30 32 51 32 22 25 GH5 22 22 21 22 20 42 22 22 23 23 23 31 21 22 14 19 GH6 1 1 1 1 1 2 1 1 1 1 1 1 1 1 1 1 GH7 3 3 3 3 3 6 3 3 3 3 3 2 6 3 2 3 GH12 5 5 6 4 5 8 5 5 7 4 5 5 6 5 4 4 GH45 1 1 2 1 1 2 1 1 1 1 1 2 2 1 1 1 Hemicellulose GH10 5 5 5 5 5 10 6 6 5 3 5 3 6 5 5 4 GH11 3 3 3 3 3 6 3 3 3 3 3 3 3 3 2 3 GH30 2 2 1 2 2 6 3 2 2 2 1 1 2 2 0 2 GH74 3 2 2 2 2 4 2 5 3 7 2 5 5 5 4 2 GH27 3 4 3 3 7 6 4 3 3 3 5 3 6 2 2 1 GH36 3 3 3 3 3 4 3 3 3 3 2 2 3 3 2 3 GH35 4 6 8 6 3 12 6 7 9 4 5 5 9 4 3 3 GH67 3 2 2 2 4 4 2 3 3 3 1 2 3 3 1 3 GH115 3 3 3 3 3 2 3 3 3 3 3 1 4 3 3 2 GH43 30 28 31 31 32 66 38 35 35 34 34 28 44 29 18 25 GH51 2 3 3 3 3 6 3 3 3 3 3 2 3 3 2 2 GH54 1 1 1 1 1 2 1 1 1 1 1 2 1 1 1 1 GH62 1 1 1 1 1 2 1 1 1 1 1 1 1 1 1 1 Pectin GH28 12 11 13 12 30 12 11 10 12 10 13 12 13 12 7 9 GH78 16 16 13 16 36 15 18 18 16 18 17 14 20 15 6 9 PL1 12 12 13 12 20 12 12 12 12 12 11 13 12 13 9 11 PL3 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 PL4 3 3 3 3 3 3 3 3 3 3 3 3 4 3 3 3 PL9 2 2 2 2 2 2 2 2 4 2 2 2 1 2 1 2 PL11 1 1 1 1 2 1 1 1 1 0 0 1 0 1 0 0 99

Table 3-10. Description of Fusarium oxysporum f. sp. palmarum secondary metabolite gene clusters and its orthologous genes present Fusarium oxysporum genomes that were studied by ORTHOMCL. Backbone genes Gene Genes within Genomes with 80% of the genes BlastP annotation of key enzyme prediction the cluster present within each cluster based on OrthoMCL FOP2_12645 NRPS 23 Absent in all formae specialesa Nonribosomal peptide synthetase FOP2_15055 NRPS 3 Absent in all formae speciales Nonribosomal peptide synthetase FOP2_12623 NRPS-Like 31 Absent in all formae speciales NRPS like enzyme FOP2_08395 NRPS-Like 7 Absent in f. sp. canariensis Linear gramicidin synthase subunit D Present only in Fo human; f. sp. radicis-lycopersici FOP2_11653 PKS 12 Absent in f. sp. canariensis, Polyketide synthetase Absent in ff. spp. cubense, raphani, vasinfectum FOP2_01243 PKS-Like 16 Absent in f. sp canariensis Polyketide synthetase Absent in ff. spp. cubense, lycopersici, raphani, vasinfectum FOP2_02302 DMAT 28 Present in all formae speciales Tryptophan dimethylallyltransferase FOP2_12957 DMAT 17 Present in all formae speciales Tryptophan dimethylallyltransferase FOP2_05180 HYBRID 17 Present in all formae speciales Hypothetical protein FOP2_07398 HYBRID 12 Present in all formae speciales Hypothetical protein FOP2_07629 NRPS 20 Present in all formae speciales HC toxin synthetase FOP2_11142 NRPS 12 Present in all formae speciales Nonribosomal peptide synthetase FOP2_12983 NRPS 11 Present in all formae speciales Enniatin synthetase FOP2_14907 NRPS 3 Present in all formae speciales Catechol-O-methyltransferase FOP2_01077 NRPS-Like 25 Present in all formae speciales Dihydrodipicolinate synthetase FOP2_02433 NRPS-Like 10 Present in all formae speciales Linear gramicidin synthase subunit D FOP2_03507 NRPS-Like 22 Present in all formae speciales Linear gramicidin synthase subunit D FOP2_03593 NRPS-Like 28 Present in all formae speciales Hypothetical protein FOP2_05604 NRPS-Like 7 Present in all formae speciales Hypothetical protein FOP2_05997 PKS 26 Present in all formae speciales Hypothetical protein FOP2_07345 PKS 18 Present in all formae speciales Hypothetical proteins FOP2_09560 PKS 21 Present in all formae speciales Bikaverin cluster-polyketide synthase

100

Table 3-10. Continued Backbone genes Gene Genes within Genomes with 80% of the genes BlastP annotation of key enzyme prediction the cluster present within each cluster based on OrthoMCL FOP2_10185 PKS 5 Present in all formae speciales Hypothetical protein FOP2_11153 PKS 2 Present in all formae speciales Hypothetical protein FOP2_11564 PKS-Like 4 Present in all formae speciales 3-Oxoacyl-[acyl-carrier-] synthease II FOP2_04697 DMAT 11 Absent only in f. sp. conglutinans Tryptophan dimethylallyltransferase FOP2_04711 HYBRID 9 Absent only in ff. spp. raphani, pisi Hypothetical protein and lycopersici FOP2_02013 NRPS-Like 17 Absent only in f. sp. melonis, Fo5176 Ferrichrome siderophore peptide synthetase FOP2_09508 NRPS-Like 15 Absent only in f. sp. conglutinans, Hypothetical protein Fo5176 FOP2_06155 NRPS 20 Absent only in f. sp. lycopersici Bacitracin synthetase FOP2_00261 PKS 21 Absent only in Fo human Lovastatin nonaketide synthase FOP2_12602 PKS 19 Absent only in f. sp. raphanin Polyketide synthetase FOP2_14401 PKS 2 Absent only in ff. spp. conglutinans, Mycocerosic acid synthetase lycopersici race 3, radicis-lycopersici, raphani, pisi and Fo5176 aFusarium species, formae speciales or isolate refers to Table 3-1.

101

Table 3-11. Secreted in Xylem (SIX) genes recovered from Fusarium. oxysporum f. sp. canariensis and F. oxysporum f. sp. palmarum genome assemblies using multiple bioinformatics approaches. Fusarium oxysporum Secreted in xylem (SIX) genes Genes/ Contigs f. sp. canariensis SIX7 FOC3_15236

SIX10 FOC3_15235

SIX12 Contig 209

f. sp. palmarum SIX9 FOP2_15139

SIX8 Contig 532

Table 3-12. Putative virulence gene distribution in Fusarium oxysporum f. sp. canariensis and F. oxysporum f. sp palmarum predicted by PHI-Base. Fusarium oxysporum Virulence Lethal/Loss of Reduced Unaffected associated genes pathogenicity virulence pathogenicity/Mixed f. sp. canariensis 237 34 93 110 f. sp. palmarum 223 22 82 109

102

Table 3-13. Putative virulence associated genes orthologs identified in predicted proteome of Fusarium oxysporum f. sp. palmarum by PHI-Base, classified based on role in infection process. Gene Name or Putative function Fungus Adapation to the environment ARG1 Argininosuccinate lyase F.oxysporum f.sp.melonis CMLE 3-carboxy- cis,cis-muconate lactonizing enzyme F.oxysporum f.sp.lycopersici Dnj1 Compatible interaction with the host. F.oxysporum f.sp.lycopersici Cell wall intergrity CHS2 Chitin Synthase F.oxysporum f.sp.lycopersici CHS7 Chitin Synthase F.oxysporum f.sp.lycopersici FoOCH1 A putative a-1,6-mannosyltransferase F.oxysporum f.sp.cubense gas1 Putative β-1,3-Glucanosyltransferase Fusarium oxysporum Signaling cascade FGA1 G alpha protein subunit F.oxysporum f.sp.cucumericum FGA2 G alpha protein subunit F.oxysporum f.sp.cucumericum Fmk1 Mitogen activated protein kinases Fusarium oxysporum FoBck1 MAP Kinase pathway F.oxysporum f.sp.cubense FoSlt2 MAP Kinase pathway F.oxysporum f.sp.cubense foSNF1 protein kinase Fusarium oxysporum FSR1 Fungal virulence and sexual mating Fusarium verticillioides FOW1 Mitochondrial carrier protein Fusarium oxysporum Fbp1 F-box protein F.oxysporum f.sp.lycopersici FRP1 F-box protein F.oxysporum f.sp.lycopersici ARSEF2860 Ras/MAPK signaling pathway Beauveria bassiana Transcription Regulators CTF2 Putative Zn finger transcription factor Fusarium oxysporum FTF2 Fusarium transcription factor. Fusarium oxysporum FOW2 Putative Zn finger transcription factor Fusarium oxysporum SGE1 Transcriptional factor F.oxysporum f.sp.lycopersici Snt2 BAH/PHD-containing transcription factor F.oxysporum f.sp.melonis ZIF2 B-ZIP transcriptional factor Fusarium graminearum

103

Table 3-13. Continued Gene Name or putative function Fungus Peroxisomal function PEX5 Peroxisomal targeting signal receptor Fusarium graminearum PEX6 Peroxisomal biogenesis factor 6 Fusarium graminearum Unclassified SID1 Siderophore biosynthetic gene Fusarium graminearum

104

Transcription,Translation and post translation Secondary metabolites biosynthesis,… Intracellular trafficking, secretion, and… Signal transduction mechanisms Lipid transport and metabolism General function prediction only Energy production and conversion

DNA chromatin and replication FOC Coenzyme transport and metabolism FOP KOG Classes KOG Cell wall/membrane/envelope biogenesis Defense mechanisms Cell division and cell cycle control Carbohydrate transport and metabolism Amino acid transport and metabolism

0 500 1000 1500 2000 2500 Number of genes

Figure 3-1. Distribution of Fusarium oxysporum f. sp.canariensis (FOC) and F.oxysporum f. sp. palmarum (FOP) proteomes into sub categories of Eukaryotic Orthologous Groups (KOG) gene families.

105

F. oxysporum Fo47 F. oxysporum human F. oxysporum Fo5176 F. oxysporum f. sp. lycopersici… F. oxysporum f. sp. lycopersici… F. oxysporum f. sp. pisi F. oxysporum f. sp. vasinfectum Auxilary activites

genomes genomes F. oxysporum f. sp .raphani Carbohydrate binding modules F. oxysporum f. sp. conglutinans Carbohydrate esterases F. oxysporum f. sp. radicis… Glycoside hydrolases

F. oxysporum f. sp. melonis Glycoside transferase Fusarium Fusarium F. oxysporum f. sp. cubense Polysaccharide lyases F. oxysporum f. sp. canariensis F. oxysporum f. sp. palmarum F. verticillioides F. graminearum

0 200 400 600 800 1000 1200 1400 Number of the genes

Figure 3-2. The carbohydrate active enzymes (CAZymes) class distribution in Fusarium genomes.

106

F. oxysporum strain FO47 F. oxysporum strain human F. oxysporum f. sp. lycopersici race 2 F. oxysporum strain FO5176 F. oxysporum f. sp. lycopersici race 3 F. oxysporum f. sp. pisi

genomes genomes F. oxysporum f. sp. vasinfectum F. oxysporum f. sp .raphani F. oxysporum f. sp. conglutinans F. oxysporum f. sp. melonis F. oxysporum f. sp. cubense Fusarium F. oxysporum f. sp. radicis lycopersici F. oxysporum f. sp. palmarum F. oxysporum f. sp. canariensis

0 5 10 15 20 25 30 35 40 45

Secondary metabolite biosynthesis genes

DMAT Hybrid NRPS-PKS NRPS PKS

Figure 3-3. SMURF secondary metabolite (SM) backbone gene prediction of Fusarium proteomes used in this study. The biosynthetic genes of SM were classified into one of the four categories: dimethylally tryptophan synthase (DMAT), polyketide synthase (PKS), non-ribosomal polyketide synthase (NRPS) and hybrid (NRPS-PKS).

107

CHAPTER 4 SCREENING FOR SECRETED IN XYLEM (SIX) GENES IN FUSARIUM WILT PATHOGENS OF ORNAMENTAL PALMS

Introduction

Fusarium oxysporum is an asexual, common, soil-borne fungus. It is a species complex comprised of pathogenic and nonpathogenic isolates, including saprophytes and endophytes

(Gordan and Martyn 1997). The plant pathogenic F. oxysporum cause crown rots, root rots or xylem colonizing vascular wilt diseases (Leslie and Summerell 2006). Collectively, F.oxysporum has a broad host range affecting important economic plants such as wheat, maize, oil palm, tomato, pea, watermelon, banana, date palm, strawberry and ornamentals, including Canary

Island date palm, queen palm, carnation and chrysanthemum plants (Leslie and Summerell 2006;

Michielse and Rep 2009). Pathogenic strains of F. oxysporum are classified into >120 formae speciales based on host specificity (Armstrong and Armstrong 1981; Michielse and Rep 2009).

Typically, each forma specialis causes disease on one or a few host species. Additionally, a few formae speciales are further divided into races based on the cultivar specificity within the host species.

Differentiating these formae speciales is complicated, as nonpathogenic Fusarium oxysporum may be co-isolated from symptomatic tissues and some formae speciales consist of different lineages. Traditionally, plant-pathogenic F. oxysporum (discriminating formae speciales/races) are identified by pathogenicity assays which are time consuming and laborious.

Characterizing formae speciales based on colony pigmentation and morphology of macroconidia and microconidia are of limited use as they are morphologically similar (Baayen et al. 2000;

Lievens et al. 2008). Currently, PCR assays and subsequent sequencing of amplicons is conducted using primers for housekeeping genes such as the translation elongation factor (EF-

108

1α) and the internal transcribed spacer region (ITS) for diagnostics in Fusarium wilt disease.

However, they lack resolution to distinguish all the formae speciales successfully (Recorbet et al.

2003; Chapter 2 of this dissertation). There is a need for reliable pathogenicity based markers that can differentiate formae speciales as well as nonpathogenic F. oxysporum.

Fungal plant pathogens secrete several categories of proteins into the host to block host defenses and promote infection (Rep 2005). These proteins are defined as effectors, which are secreted only during host pathogen interactions and are presumed to be pathogen-host specific.

The Secreted in Xylem (SIX) gene family is a potential candidate for markers to differentiate F. oxysporum formae speciales as these proteins have little or no homology with any other proteins and encode small-cysteine rich secretory proteins, which promote infection and colonization in the host (Houterman et al. 2007; Schmidt et al. 2013). However, SIX protein homologs are found in Fusarium foetens (SIX1) (Laurence et al. 2015) and Colletotrichum orbiculare and C. higginsianum (SIX6) (Gan et al. 2013; Kleemann et al. 2012).

SIX proteins were first studied from the xylem sap of a tomato plant infected by F. oxysporum f. sp. lycopersici (FOL). So far, fourteen SIX genes have been identified in FOL by proteomic and reverse translational approaches (Houterman et al. 2008, 2009; Rep et al. 2004;

Schmidt et al. 2013). The biological function for the majority of SIX genes is largely unknown.

Only SIX1 through SIX6 gene have been characterized. Based on gene knock-out experiments

SIX1, SIX3, SIX5 and SIX6 are known to play a direct role in virulence in FOL in tomato and are classified as effectors (Gawehns et al. 2014; Houterman et al. 2008, 2009; Michielse et al. 2015;

Rep et al. 2004). In contrast, SIX1 and SIX3 play a dual role along with SIX4 as avirulence proteins, which are recognized by R proteins and result in resistance against the FOL pathogen in susceptible tomato cultivars (Houterman et al. 2008, 2009; Rep et al. 2004). In FOL, SIX 1

109

through 14 genes are located on lineage specific chromosomes (Chr 3, Chr 6 and Chr 14), which are enriched with transposon and pathogenicity associated genes (Ma et al. 2010; Schmidt et al.

2013). Studies have revealed that SIX genes are present in mini clusters on lineage specific chromosomes (LS) of FOL, and SIX1, SIX2, SIX3 and SIX5 genes are regulated by a specific transcription factor SGE1 (SIX gene expression 1) (Michielse et al. 2015; Schmidt et al. 2013).

Further, the majority of SIX genes present on LS chromosomes are associated with MITES, Class

II type transposons, both upstream and downstream from the SIX gene (Schmidt et al. 2013).

SIX gene homologs have been detected in other F. oxysporum formae species by various approaches and prominent SIX profiles are listed in Table 4-1. Studies have explored the possibility of SIX genes as genetic markers to distinguish F.oxysporum formae speciales

(Chakrabarti et al. 2011; Fraser-Smith et al. 2014; Laurence et al. 2015; Lievens et al. 2009;

Meldrum et al. 2012). Presence of SIX1, SIX2, SIX4 and SIX5 can effectively separate FOL from other formae speciales (Lievens et al. 2009). Within FOL, SIX4 is present only in race 1 and absent in race 2 and 3 isolates (Lievens et al. 2009). Race 2 and 3 isolates of FOL can be separated based upon sequence polymorphism of the SIX3 sequence (Lievens et al. 2009). In F. oxysporum f. sp. cubense, the presence of SIX8 in race 4 distinguishes it from race 1 and 2 isolates, and SIX8 sequence polymorphism can separate tropical (TR4) from subtropical (ST4) isolates within race 4 (Fraser-Smith et al. 2014). These two examples demonstrate that screening of SIX genes can facilitate the identification of different formae speciales and their races.

Fusarium wilt on ornamental palms in Florida is a lethal disease caused by Fusarium oxysporum f. sp. canariensis (FOC) and F. oxysporum f. sp. palmarum (FOP). FOC causes a vascular wilt on Phoenix canariensis and P. reclinata, whereas FOP affects Washingtonia robusta, Syagrus romanzoffiana, x Butyagrus nabonnandii, and Phoenix canariensis (Elliott

110

2015; Elliott et al. 2010; 2011; 2017; Plyler et al. 2000). While FOC and FOP share many similarities, including pathogenicity on P. canariensis, FOP has a wider palm host range, causes a more aggressive disease and is primarily localized to Florida, whereas FOC is found world- wide and restricted primarily to one palm species (Elliott, 2015; Elliott et al. 2010; 2011; 2017;

Giesbrecht et al. 2013). Currently they are no treatments available to cure or eradicate either pathogen. Early detection of the pathogen is key to develop and implement management strategies.

Although no population studies have been conducted on FOP, a few studies have been completed on FOC populations to understand genetic diversity. These studies indicate that the majority of the FOC isolates (71%) belong to a single VCG group (0240) and mitochondrial

DNA markers revealed the presence of only four haplotypes (Gunn and Summerell 2002; Plyler et al. 2000). Laurence et al. (2015) conducted a phylogenic study of Australian and international

FOC isolates based on a portion of the translation elongation factor gene (EF-1α). This study indicated the presence of at least two different lineages and concluded FOC was polyphyletic.

The objectives of this study were: 1) to screen for the presence of SIX gene orthologs in

FOC and FOP genomes using multiple bioinformatic approaches; 2) to screen FOC and FOP isolates for the presence of SIX genes using published PCR primers to identify population structures and their close homologs; 3) to compare FOC and FOP SIX gene phylogenies with EF-

1α gene phylogeny to determine if SIX genes were vertically or horizontally transmitted; and 4) to evaluate SIX gene profiles as genetic markers to differentiate these two formae speciales.

111

Material and Methods

Isolates

A total of 47 FOC and FOP isolates plus one FOL race 3 isolate were used in this study

(Table 4-2). FOC and FOP isolates were recovered from petioles of palm trees with Fusarium wilt from different locations within Florida (40), California (2), Texas (2) and South Carolina (1) in a span of twelve years (2004-2016). The remaining two isolates (CA4, CA5) were obtained as genomic DNA from the California Deprtment of Agriculture, with CA4 from California and

CA5 from Nevada. Nineteen isolates were FOC and twenty-eight isolates were FOP. They were selected based on their palm host and geographical location (Table 4- 2). All the isolates were single spored and characterized morphologically and molecularly. The FOL race 3 (PLM905A) was provided by Dr. Garry Vallard from University of Florida-IFAS, Gulf Coast Research and

Education Center.

DNA Extraction and Quality Assessment

Except for the two isolates where DNA was provided, study group isolates were revived from storage by placing a single agar plug on potato dextrose agar (1/5 PDA) (Difco

Laboratories, Detroit, MI) and grown for 5 days at 28°C. Five agar plugs (5 mm) from 1/5 PDA were transferred to a conical flask with 20 ml of full strength potato dextrose broth (PDB) (Difco

Laboratories, Detroit, MI) and incubated for 5 days at 28°C without shaking. After incubation, mycelium was harvested and DNA was extracted using Gentra Puregene Tissue Kit (Qiagen,

Valencia, CA) according to the manufacturer’s instructions. The resulting DNA pellet was suspended in 50 µl of molecular grade water (HyCloneTM, Fisher Scientific, Pittsburg, PA) and quantified using a QubitTM fluorometer (Thermo Fisher Scientific, Pittsburg, PA). The DNA was stored at 4°C for further use.

112

Molecular Characterization by EF-1α Gene

A translation elongation factor (EF-1α) gene based PCR was performed on each isolate in the study group as described by O’Donnell et al. (1998) to assess the DNA quality and construct a phylogenetic tree in order to compare with SIX gene phylogenies.

Screening of SIX Genes from Assembled Genomes

To identify putative effector SIX genes in genomes, predicted proteomes of FOC and

FOP genomes, generated as described in Chapter 3, were used. All the GenBank submitted SIX protein sequences were downloaded and formatted in a database. On a local server, the Basic

Alignment Search Tool (BLASTp) was employed using FOP and FOC genomes as query to identify homologs with a similarity value of 70% and e-value of 10-3. The potential SIX gene candidates were then identified for the presence of signal peptide by SignalP (Petersen et al.

2011). Further, GenBank protein sequence alignments were used to screen SIX gene homologs using HMMER (Finn et al. 2015), a hidden markov model, in FOC and FOP assembled genomes. Additionally, un-assembled scaffolds of FOC and FOP genomes were also screened using GenBank SIX gene data sets using BLASTn.

Detection of SIX Genes by PCR

All the isolates listed in Table 4-2 were screened for presence or absence of SIX1 through

SIX14 genes using published primers. A FOL isolate PLM 905A (race 3) was used as positive control for screening of all the SIX1-14 genes except SIX4 because it is present only in FOL race

1. To determine the annealing temperature for each primer set, a gradient PCR was performed using all SIX primers with PLM905A (FOL), PLM778A (FOP) and PLM588A (FOC). Table 4-3 indicates the primers, their annealing temperatures and amplicon size, along with their reference source. Each PCR was carried out in a 50 µl reaction mixture brought to volume with sterile

113

deionized water. All reactions were carried out using New England Biolabs (NEB) Taq polymerase (Ipswich, MA). The reaction consisted of 25 µl of 1X PCR standard Taq Reaction

Buffer (10 mM Tris–HCl, 50 mM KCl, 1.5 mM MgCl2), 200 µM of each dNTP, 1 µM of each primer and 1U of NEB Taq DNA polymerase with 20-50 ng of genomic DNA template. Each

PCR assay included a negative control consisting of sterile deionized water.

An Eppendorf Thermocycler was used for all PCR assays (ThermoFisher Scientific,

Waltham, MA). PCR products were examined by 1.5% agarose gel electrophoresis in Tris borate

EDTA (TBE) buffer with GelRed TM (Biotium, Hayward, CA). A subset of FOC and FOP amplified PCR products were purified using spin columns Wizard® PCR Preps DNA

Purification System (Promega Corp. Madison, WI), and Sanger sequencing in both forward and reverse direction was conducted by Genewiz (South Plainfield, NJ).

Phylogenetic Analysis

Sequences derived from PCR assays using EF-1α gene (O’Donnell et al. 1998) (ef1 and ef2 primers) and for each SIX gene set were aligned and manually edited using Mega 7.0 (Kumar et al. 2016). The aligned consensus sequences were then queried for identity match using

BLASTn to confirm their identity and their close homologs. Multiple sequence alignment of EF-

1α and SIX data sets were performed using MUSCLE using the default parameters (Edgar 2004).

Phylogenetic trees were generated using Maximum Likelihood (ML) to understand the phylogenetic relatedness among the data sets. ML analysis for EF-1α gene was constructed using

Kimura 2-parameter with Gamma distribution. The ML phylogenetic tree were assessed with

1,000 boot strap replications. FOC EF-1α sequences from Laurence et al. (2015), a study of

Australian and non-Australian isolates, were also included in the phylogenetic tree with the sequence designation RBG. Selected representatives of F. oxysporum formae speciales are also

114

included in the phylogenetic trees. They were downloaded from Fungi DB (Fungi & Oomycete

Genomic Resource) (www fungidb.org/fungidb/).

Phylogenetic trees were constructed for SIX1, SIX7, SIX8, SIX9, SIX10 and SIX12 datasets only as these were the only SIX genes present in FOC and FOP. For ML trees their substitutions methods and rate of substitution were analyzed for each SIX dataset as follows:

SIX1, Kimura 2-parameter with uniform distribution; SIX7 and SIX8, Kimura 2-parameter with non-uniformity of evolutionary sites with Gamma distribution (G); SIX9 Kimura 2-parameter with invariable sites (I); SIX10, Jukes–Cantor model with Gamma distribution; and SIX12, Jukes

Cantor model with uniform distribution of sites. SIX data set phylogenetic trees included SIX homolog representatives of other F. oxysporum formae speciales and sequences from the

Laurence et al. (2015) study were also included with sequence designation RBG for FOC phylogenetic trees.

Results

Molecular Characterization by EF-1α Gene

Study group isolates along with PLM905A (FOL) were amplified with EF-1α primers, resulting in a single band size of 650 bp. The ML phylogenetic tree for the partial EF-1α gene for

FOC isolates belonging to Florida, non-Australian (RBG) and one Australian isolate (RBG2978) clustered within one clade, whereas the remaining Australian FOC (RBG) isolates clustered in at least two different clades (Figure 4-1). FOP isolates were grouped in a single clade. However, within this clade, a sub-cluster representing isolates PLM140B, PLM320B, PLM336D,

PLM351B, PLM676A, PLM741A, PLM745A and PLM791A were observed. All these sequences had a two nucleotide transition mutation difference from the remaining FOP sequences as observed in Elliott et al. (2010). This subgroup of sequences came from diverse

115

palm hosts and geographical locations. Clades belonging to FOC and FOP were well established and distinct.

Screening of SIX Genes from Assembled Genomes

FOC and FOP proteomes were screened for presence of SIX protein homologs using the

GenBank curated SIX proteins from Fusarium oxysporum formae speciales. Genome sequencing screening revealed the presence of SIX7, SIX10 and SIX12 in PLM386B (FOC) and SIX8 and

SIX9 in PLM249A (FOP) (Table 4-4).

Detection of SIX Genes by PCR

FOL genome-based published primers for SIX genes detected the presence of SIX1, SIX7,

SIX10 and SIX12 homologs in FOC isolates and SIX8, SIX9 and SIX10 homologs in FOP isolates. SIX1 is variable within FOC, being detected by PCR in 16 of the 19 isolates. SIX10 was variable in FOP being detected by PCR in 6 of the 28 FOP isolates. Resulting amplicons were sequenced to verify gene identity. Nonspecific PCR products were observed for SIX3 and SIX5 primers in some of the FOC and FOP isolates, and these PCR products were also sequenced. The

BLASTn analysis for these nonspecific products indicated they were not related to any SIX genes. The SIX2, SIX3, SIX4, SIX5, SIX6, SIX11, SIX13 and SIX14 genes were not detected in any of FOC and FOP isolates.

SIX1

Sixteen FOC isolates yielded an amplicon using SIX1 primers, but the remaining three isolates including the genome-sequenced isolate (PLM386B) and two other isolates (PLM724A and PLM754A) failed to amplify with the primers (Table 4-2 and Figure 4-8). The absence of

SIX1 sequence from the FOC-PLM386B genome sequence was verified by enhanced bioinfomatic tools on the genome sequence. The SIX1 gene was not amplified in any FOP

116

isolate. The SIX1 gene is conserved in all 16 isolates, irrespective of the palm host and geographical location (Figure 4-2). The SIX1 gene in FOC shared a high level of nucleotide similarity with F.oxysporum f. sp. cubense (95-100%).

The SIX1 phylogenetic tree (Figure 4-2) had 38 taxa with 201 nucleotides with 31 parsimony informative sites. The FOC isolates from our study (USA) along with Australian and international strains are highly conserved as shown by grouping within one clade. A single base variation was observed in two Australian isolates (RBG4219 and RBG2236).

SIX7

SIX7 primer amplicons were obtained from all FOC isolates, but not from FOP isolates

(Table 4-2 and Figure 4-9). The FOC sequences were highly conserved except for two isolates

The SIX7 ML phylogenetic tree was constructed using 39 taxa with 660 nucleotides and had 84 parsimony informative sites (Figure 4-3).

PLM221B and PLM386B, isolated from P.canariensis in Florida, varied significantly from the other FOC isolates which were clustered in one clade (Figure 4-3). The SIX7 sequence of PLM221B and PLM386B was most similar to f. sp. cubense (96%), whereas similarity for remaining FOC isolates had identity matches with f. sp. lini (98%).

SIX10

SIX10 was present in all the FOC isolates in this study (Table 4-2 and Figure 4-12) and no sequence variation was observed (Figure 4-4). The nearest SIX10 homolog is f. sp. lycopersici with 95% nucleotide similarity identity. Six of 28 FOP isolates (PLM320B, PLM510A,

PLM560A, PLM596A, PLM741A, and PLM791A) were also amplified with the SIX10 primers.

The FOP sequence PLM510A, PLM560A, PLM741A and PLM791A had an identical nucleotide

117

match with FOC SIX10 sequences (Figure 4-4). These six isolates were recovered from different palm hosts and geographical locations within Florida.

The SIX10 phylogenetic tree (Figure 4-4) was constructed using 33 taxa and 461 nucleotides and had 29 parsimony informative sites. The ML tree showed that FOC isolates, irrespective of geographical location in the world, clustered within the same clade, with the exception of RBG2236 from Australia. The FOP SIX10 seqeunce from FOP isolates PLM510A,

PLM560A, PLM741A, and PLM791A also clustered within the FOC clade.

SIX12

The SIX12 gene is present in all FOC isolates, but sequence polymorphism was observed and appears to be based on host and geographical location (Table 4-2; Figure 4-13). No FOP isolates were amplified with the SIX12 primers. The SIX12 ML phylogenetic tree for FOC isolates resulted in two distinct and well established clades. The first clade included PLM221B,

PLM386B, PLM588A, PLM601A and PLM605A, isolates recovered from Phoenix canariensis from Florida (Figure 4-5). The second clade included non-Florida FOC isolates [CA4 (CA), CA5

(NV), PLM385 (TX), PLM511A (SC) and PLM706A (CA)], PLM224B recovered from P. sylvestris in Florida and PLM387B and PLM696A recovered from P. reclinata in Florida (Figure

4-5). However, two isolates (PLM183C and PLM760A) recovered from P. canariensis in Florida also clustered with these non-Florida FOC or non-P. canariensis isolates. Based on the nucleotide sequences, 26 single nucleotide polymorphisms (SNPs) were observed, leading to 15 nonsynonymous substitutions between the two FOC groups. Close homologs of the P. canariensis Florida isolates (bottom group in Figure 4-5) are more similar to f. sp. narcissi

(94%), whereas close homologs to the non-Florida isolates were f. sp. lini (98%). The SIX12 dataset included 26 taxa with 381 nucleotides with 30 parsimony informative sites.

118

To determine if FOC isolates are subjected to diversifying or purifying selection, a

Tajima’s Test for Neutrality was performed using Mega 7.0 (Kumat et al. 2015; Tajima 1989).

The Tajima’s test stastitic was positive indicating D= 2.130246.

SIX8 and SIX9

The SIX8 and SIX9 homologs are present only in FOP; they are highly conserved and present in all isolates (Table 4-2, Figure 4-10 and Figure 4-11). They are not present in FOC.

SIX9 close homolog is f. sp. lycopersici (91%), whereas SIX8 close homolog is f. sp. lycopersici

(90%). The nucleotide identity between SIX8 FOP and f. sp. cubense is 86% (SIX8a) and 84%

(SIX8b), with amino acid similarity of 80% and 75%, respectively. This indicates that SIX8 FOP is very divergent from f. sp. cubense SIX8a and SIX8b.

SIX8 dataset included 24 taxa and 252 nucleotides with 41 parsimony informative sites

(Figure 4-6). SIX9 dataset included 21 taxa with 265 nucleotides and 100 parsimony informative sites (Figure 4-7). For SIX8 and SIX9 datasets, FOP isolates grouped in one clade (Figures 4-6 and 4-7).

Discussion

The Fusarium oxysporum species complex is an asexual, cosmopolitan fungus with a wide host range. However, each subspecies (forma specialis) infects only a few host species. F. oxysporum formae speciales are known for their partitioned genome into core and lineage- specific regions (Ma et al. 2010). Lineage-specific regions are enriched with pathogenicity related genes, transposons and repetitive elements (Ma et al. 2010). The Secreted in Xylem (SIX) gene family secretes at least 14 proteins into the plant hosts and are presumed to be involved in the disease process. SIX genes have been largely studied in Fusarium oxysporum f. sp. lycopersici (FOL), a tomato infecting forma specialis, using bioinformatic, proteomic and

119

mutagenesis approaches. Extensive research on FOL (Ma et al. 2010), revealed that the majority of SIX genes are located in lineage-specific chromosomes (especially Chr14) and can be transferred between formae speciales via horizontal gene transfer. Besides F. oxysporum, SIX6 homologs are present in Collelotrichum orbiculare and C. higginsiantum (Gan et al. 2013;

Kleemann et al. 2012), and SIX1 is present in Fusarium foetens (Laurence et al. 2015).

The present study attempted to detect and survey putative SIX genes in the palm-infecting formae speciales of FOC and FOP populations using extensive genomic and PCR-based screening. FOC causes Fusarium wilt primarily on Phoenix species, whereas FOP can infect palms belonging to genera Phoenix, Sygarus and Washingtonia. To evaluate if palm-infecting formae speciales have shared SIX effector profiles, isolates selected for the study represented various palm hosts and geographical locations within the continental USA. Three (SIX1, SIX7, and SIX10) of the four SIX genes that were identified in this FOC population were also reported in a previous study (Laurence et al. 2015) using Australian and international (Japan, USA and

Spain) strains recovered from P. canariensis.

This is the first study to indicate the presence of SIX12 in FOC. Significant sequence variation was observed in SIX12, with two distinct sequence types (Florida and Non-Florida/Non

P. canariensis hosts), which do not correspond to any other formae speciales. The consistent sequence patterns seem to correlate with palm host and location, but two Florida P. canariensis isolates did cluster in the non-Florida group disrupting this pattern. SIX7, SIX10 and SIX12 were present in all FOC isolates. Schmidt et al. (2013), using in-silico approaches on FOL 4287 genome, demonstrated that SIX genes are usually present as mini-clusters on LS chromosomes with SIX7, SIX10 and SIX12 forming one such cluster. Research to determine if this is alos true for FOC should be conducted.

120

Presence of distinct clades in the SIX12 phylogeny of FOC indicate that Florida isolates may have attained SIX12 through multiple horizontal events. Our results with SIX12 supports an earlier hypothesis that FOC was introduced into Florida through multiple independent sources and high diversity is expected among Floridia isolates when compared to California and Nevada

(Plyler et al. 2000). To determine if any of sequence variants are in selective pressure, the

Tajima’s test for neutrality was performed. The test statistic D was a positive value, suggesting a balancing selection for SIX12 alleles (Tajima 1989). Variation within the two groups of the FOC

SIX12 gene suggests their rapid adaption to host environment or to evade host defenses and recognition (Plissonneau et al. 2017). Evaluation of more USA isolates (different palm hosts and locations) for SIX12 may shed light on the population structure and movement of the pathogen within the continental USA. Evaluation of non-USA isoaltes for presence of SIX12 should also be conducted.

This is the first study to reveal presence of SIX genes in the FOP population. SIX8 and

SIX9 genes were found in all FOP isolates and are highly conserved, suggesting an essential role in pathogenicity. As revealed in other F. oxysporum formae speciales, presence of conserved SIX genes within populations may be attributed to a single ancestor origin or horizontal gene transfer

(Fraser-Smith et al. 2014; Laurence et al. 2015; Ma et al. 2010; Taylor et al. 2016).

In FOL, at least eight copies of SIX8a and four copies of SIX8b are present on multiple lineage specific chromosomes (Chr3, Chr6 and Chr14) (Schmidt et al. 2013). The F. oxysporum f. sp. cubense isolates possess three sequence variants of SIX8, including SIX8, SIX8a, SIX8b

(Fraser-Smith et al. 2014). Our study indicates the absence of any polymorphism in SIX8 among

FOP isolates. Preliminary FOP genome assembly suggests presence of a single copy of SIX8.

Currently the biological function of SIX8 is unknown. Exploring the molecular basis and

121

interaction between SIX8 protiens and host proteins will shed light on the effect of single copy vs multiple copies of SIX8 and their effect on FOP pathogenicity.

Interestingly, SIX10 gene is present in 6 of 28 FOP isolates that were studied. More over, the sequences were identical to SIX10 seqeunces from FOC isolates. The six isolates belonged to various palm hosts and various geographical locations within Florida. This uneven distribution suggests that these isolates may have acquired SIX10 through a horitzontal gene transfer from

FOC or from the same ancestor as FOC rather than independently. The possibility of horizontal gene transfers can result in accumulation of newer SIX genes combination, resulting in emergence of newer formae speciales with varying fitness levels. Pathogenicity tests need to be conducted with these FOP isolates to evaluate their disease severity and host specificity.

EF-1α gene phylogeny of FOC reveals a monophyletic lineage for USA and non–

Australian isolates, but as a formae species, it is polyphyletic with at least two different lineages, with Australian isolates being more diverse. The SIX1 and SIX10 based phylogenies suggests a monophyletic nature for FOC isolates, but the SIX7 and SIX12 phylogeny suggests a polyphyletic lineage within USA isolates. For FOP, EF-1α and SIX genes phylogenies indicates the presence of a clonal lineage.

Detection of SIX genes in this study was based on multiple bioinformatics approaches and PCR-based assays. Absence of remaining SIX genes by PCR screening in FOC and FOP populations or in the draft genomes can be attributed to either sequence variation in priming sites or absence of SIX orthologs. We suspect the latter because this study is complemented by whole genome sequencing data. We utilized both predicted proteomes and scaffolds using GenBank proteins sequences and alignments as queries to detect SIX gene homologs (Chapter 3).

122

This study finds that FOC and FOP have distinct SIX gene profiles, even though they have with an overlapping host range. This is contrary to studies done on cucurbit and legume infecting formae speciales populations. A study conducted with four cucurbit infecting formae speciales (cucumerinum, melonis, niveum, and radicis-cucumerinum) revealed that they share a similar SIX effector profile (SIX1, SIX6 and SIX13) with minimum sequence variation among them (Van Dam et al. 2016). In legume infecting formae speciales (medicaginis, ciceris and pisi), even though each formae speciales had a diverse SIX effector profile, SIX13 effector was shared among all three formae speciales (Williams et al. 2016). With palm-infecting formae speciales, no shared SIX effectors were present in FOC and FOP population with the lone exception of six FOP islates with SIX10. Nevertheless, conservation of SIX genes was observed within forma specialis regardless of palm host. Phylogenies based on EF-1α along with SIX effectors profiles indicate that FOC and FOP have evolved independently.

SIX genes detected in FOC and FOP are present in other formae speciales found in

Florida, specifically FOL. This may indicate these genes are not host specific but may have evolutionarily conserved function (Thatcher et al. 2012). Nevertheless, each forma specialis has a unique SIX effector profile, which might complement other pathogenic genes that determine host specificity and pathogenicity (Jelinski et al. 2016). Because each infecting forma specialis has a distinct SIX effector profile, they are potential candidates for genetic markers that can be explored for diagnostic purposes to identify Fusarium wilt on ornamental palms without the need of pathogencity studies. This population scan of SIX gene effectors will provide insights into the evolution patterns of these formae speciales. In future, functional based assays using reverse genetic approaches are needed to understand the biological function and their role in pathogenicity for SIX effectors.

123

Table 4-1. Previous studies on prominent Fusarium oxysporum formae speciales plus F. foetens with their SIX gene profiles elucidated by various approaches. Fusarium oxysporum Host SIX genes Reference Formae speciales f. sp. lycopersici 4287 Tomato SIX1-SIX3, SIX5-SIX14 Houtermann et al. 2007; Lievens et al. 2009 f. sp. lycopersici (race 1) SIX1-SIX14 Ma et al. 2010; Schmidt et al. 2013

f. sp. vasinfectum Cotton SIX6 Chakrabarti et al. 2011

F.oxysporum (Fo5176) Wild cabbage SIX1, SIX4, SIX8, and SIX9 Thatcher et al. 2012

f. sp. betae Sugar beet SIX1 and SIX6 Covey et al. 2014

f. sp. canariensis (RBG2978) Canary Island date SIX1, SIX7 and SIX10 Laurence et al. 2015 palm f. sp. cubense (E421A) Banana SIX1, SIX8 and SIX13 Taylor et al. 2016 f. sp. cubense (II5) SIX1,SIX7, and SIX8 Meldrum et al. 2012 f. sp. cepae Onion SIX3, SIX5, SIX7, SIX9-10, Taylor et al. 2016 SIX12, and SIX14 f. sp. dianthi (R207) Carnation SIX7, SIX10, SIX12 and SIX13 Taylor et al. 2016

f. sp. freesia Freesia SIX7, SIX10,SIX12-13,and SIX14 Taylor et al. 2016 (NRRL26990) f. sp. melonis Melon SIX1, SIX6 and SIX13 Van Dam et al. 2016

f. sp. lini (FOLIN) Linseed SIX7, SIX10, SIX12 and SIX13 Taylor et al. 2016

f. sp. lini (FRL11811) SIX1, SIX7 and SIX10 Laurence et al. 2015

f. sp. narcissi (FOXN7) Daffodil SIX7, SIX9, SIX10 and SIX12 Taylor et al. 2016 f. sp. pisi (FOP1) (race 1) Pea SIX7, SIX10, SIX11, SIX12, and Taylor et al. 2016 SIX13 f. sp. phaseoli (ATCC90245) Pinto bean SIX6, SIX8 and SIX11 Taylor et al. 2016

Fusarium foetens (NRRL31046) Begonia sp. SIX1 Laurence et al. 2015

124

Table 4-2. Isolates used in the study and PCR based presence or absence of Secreted in Xylem (SIX) genes. Isolate Palm host Location Secreted in xylem (SIX) genes

1 2 3 4 5 6 7 8 9 10 11 12 13 14

Fusarium oxysporum f. sp. canariensis (FOC)

PLM183C Phoenix canariensis Florida + - - - - - + - - + - + - -

PLM221B Phoenix canariensis Florida + - - - - - + - - + - + - -

PLM224A Phoenix sylvestris Florida + - - - - - + - - + - + - -

PLM385B Phoenix canariensis Texas + - - - - - + - - + - + - -

PLM386B Phoenix canariensis Florida ------+ - - + - + - -

PLM387B Phoenix reclinata Florida + - - - - - + - - + - + - -

PLM511A Phoenix canariensis South Carolina + - - - - - + - - + - + - -

PLM588A Phoenix canariensis Florida + - - - - - + - - + - + - -

PLM601A Phoenix canariensis Florida + - - - - - + - - + - + - -

PLM605A Phoenix canariensis Florida + - - - - - + - - + - + - -

PLM696A Phoenix reclinata Florida + - - - - - + - - + - + - -

PLM706A Phoenix canariensis California + - - - - - + - - + - + - -

PLM724A Phoenix canariensis Florida ------+ - - + - NT - -

PLM754A Phoenix canariensis Florida ------+ - - + - NT - -

PLM760A Phoenix canariensis Florida + - - - - - + - - + - + - -

PLM776A Phoenix reclinata Florida + - - - - - + - - + - NT - -

CA4 Phoenix reclinata California + - - - - - + - - + - + - -

CA5 Phoenix canariensis Nevada + - - - - - + - - + - + - -

125

Table 4-2. Continued Isolate Palm host Location Secreted in xylem (SIX) genes

1 2 3 4 5 6 7 8 9 10 11 12 13 14

PLM908A Phoenix canariensis California + - - - - - + - - + - + - - Fusarium oxysporum f. sp. palmarum (FOP)

PLM119C Syagrus romanzoffiana Florida ------+ + - - - - -

PLM140B Syagrus romanzoffiana Florida ------+ + - - - - -

PLM153B Syagrus romanzoffiana Florida ------+ + - - - - -

PLM181C Syagrus romanzoffiana Florida ------+ + - - - - -

PLM195C Syagrus romanzoffiana Florida ------+ + - - - - -

PLM246B Syagrus romanzoffiana Florida ------+ + - - - - -

PLM249A Washingtonia robusta Florida ------+ + - - - - -

PLM256D Washingtonia robusta Florida ------+ + - - - - -

PLM258B Washingtonia robusta Florida ------+ + - - - - -

PLM265D Washingtonia robusta Florida ------+ + - - - - -

PLM320B x Butyagrus nabonnandii Florida ------+ + + - - - -

PLM336D Washingtonia robusta Florida ------+ + - - - - -

PLM344E Syagrus romanzoffiana Florida ------+ + - - - - -

PLM351B Syagrus romanzoffiana Florida ------+ + - - - - -

PLM510A Phoenix canariensis Florida ------+ + + - - - -

PLM560A Washingtonia robusta Florida ------+ + + - - - -

PLM596A Washingtonia robusta Florida ------+ + + - - - -

126

Table 4-2. Continued Isolate Palm host Location Secreted in xylem (SIX) genes

1 2 3 4 5 6 7 8 9 10 11 12 13 14

PLM619A Bismarckia nobilis Florida ------+ + - - - - -

PLM676A Phoenix canariensis Florida ------+ + - - - - -

PLM729A Syagrus romanzoffiana Florida ------+ + - - - - -

PLM741A Syagrus romanzoffiana Florida ------+ + + - - - -

PLM745B Syagrus romanzoffiana Florida ------+ + - - - - -

PLM764A Phoenix reclinata Florida ------+ + - - - - -

PLM778A Washingtonia robusta Florida ------+ + - - - - -

PLM791A Phoenix canariensis Florida ------+ + + - - - -

PLM845A Bismarckia nobilis Florida ------+ + - - - - -

PLM853A x Butyagrus nabonnandii Florida ------+ + - - - - -

PLM920 Washingtonia robusta Texas ------+ + - - - - -

Fusarium oxysporum f. sp. lycopersici (race 3) (FOL)

PLM905A Solanum lycopersicum Florida + + + - + + + + + + + + + + NT=NOT TESTED.

127

Table 4-3. PCR primer pairs that were used in this study, with their annealing temperature and amplicon size. Gene Primer sequence (Forward/Reverse) Annealing Amplicon Reference temperature size SIX1 ATGGTACTCCTTGGCGCCCTC/TGACAATGCGACCACGCCTCG 60.0°C 260bp Meldrum et al. 2012

SIX2 CAACGCCGTTTGAATAAGCA /TCTATCCGCTTTCTTCTCTC 55.5°C 749bp Lievens et al. 2009

SIX3 CCAGCCAGAAGGCCAGTTT/GGCAATTAACCACTCTGCC 55.5°C 608bp Lievens et al. 2009

SIX4 TCAGGCTTCACTTAGCATAC/GCCGACCGAAAAACCCTAA 55.5°C 967bp Lievens et al. 2009

SIX5 ACACGCTCTACTACTCTTCA/GAAAACCTCAACGCGGCAAA 55.5°C 667bp Lievens et al. 2009

SIX6 CTCTCCTGAACCATCAACTT/CAAGACCAGGTGTAGGCATT 59.1° 793bp Lievens et al. 2009

SIX7 CATCTTTTCGCCGACTTGGT/CTTAGCACCCTTGAGTAACT 55.5°C 862bp Lievens et al. 2009

SIX8 TCGCCTGCATAACAGGTGCCG/TTGTGTAGAAACTGGACAGTCGA 58.0°C 250bp Meldrum et al. 2012 TGC

SIX9 CTTCTAGCAGTTGTAGCCAC/GTACGCCATTGACGCAAG 56.0°C 258bp Laurence et al. 2015

SIX10 AAAAAGCAGGCTCCATGAAGCTCTTGTGGTTG/AGAAAGCTGGGT 55.5°C 509bp Laurence et al. 2015 CCTACTTAGACCTGGTAATTGTT SIX11 GATGTTCTCCAAAGCCATCC/AGAATGCCACTCGGTGTGA 56.0°C 400bp Laurence et al. 2015

SIX12 CTAACGAAGTGAAAAGAAGTCCTC/GCCTCGCTGGCAAGTATTTG 58.0°C 449bp Taylor et al. 2016 TT SIX13 CCTTCATCATCGACAGTACAACG/ATCAAACCCGTAACTCAGCTCC 58.0°C 1027bp Taylor et al. 2016

SIX14 TTGCCACCTATGCATACCG/TCCACATTCCTAAGCGAACC 56.0°C 400bp Laurence et al. 2015

128

Table 4-4. Secreted in Xylem (SIX) genes recovered from Fusarium oxysporum f. sp. canariensis and Fusarium oxysporum f. sp. palmarum draft genomes using multiple bioinformatics approaches. Fusarium oxysporum forma specialis Secreted in xylem (SIX) genes Genes/ Contig f .sp. canariensis SIX7 FOC3_15236

SIX10 FOC3_15235

SIX12 Contig 209

f. sp. palmarum SIX8 FOP2_15139

SIX9 Contig 532

129

Figure 4-1. EF-1α based phylogenetic tree of Fusarium oxysporum f. sp. canariensis (FOC) and F. oxysporum f. sp. palmarum (FOP) including various formae speciales inferred by Maximum Likelihood. Tree is constructed based on the Kimura 2-parameter model with Gamma distribution. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. All positions with less than 95% site coverage were eliminated. The tree is rooted with Fusarium foetens. FOC isolates from Laurence et al. (2015) are designated with RBG identifier. Sequences obtained from Fungi DB and GenBank are designated with unique isolate identifier and description.

130

PLM140B PLM351B PLM791A PLM741A PLM745B PLM336D PLM676A PLM320B PLM119C PLM181C PLM195C PLM246B PLM-249A PLM256D FOP PLM258B PLM265D PLM344E PLM510A PLM560A PLM596A PLM619A PLM729A PLM764A PLM778A PLM845A PLM853A PLM920 PLM-153B FJ985399 Fusarium oxysporum f. sp. cepae strain NRRL 38481 EF056781 Fusarium oxysporum f. sp. radicis-cucumerinum AF246847 Fusarium oxysporum f. sp. gladioli strain NRRL26989 KM893891.1 RBG4215 AUSTRALIA AF008489 Fusarium oxysporum f. sp. cubense NRRL 25607 FJ790386 Fusarium oxysporum f. sp. lycopersici strain 14844 FJ790389 Fusarium oxysporum f. sp. lycopersici strain FOL-MM66 HM057316 Fusarium oxysporum f. sp. radicis-lycopersici strain NRRL 26788 FJ790384 Fusarium oxysporum f. sp. lycopersici strain 281 AF008504 Fusarium oxysporum f. sp. melonis NRRL 26406 AF246889 Fusarium oxysporum f. sp. lini strain NRRL28929 FJ985359 |Fusarium oxysporum f. sp. pisi strain NRRL 37616 FJ985277 Fusarium oxysporum f. sp. vasinfectum strain NRRL 25424 KM893890.1 RBG4223 AUSTRALIA KM893889.1 RBG4221 AUSTRALIA KM893888.1 RBG4220 AUSTRALIA KM893887.1 RBG4219 AUSTRALIA KM893886.1 RBG4216 AUSTRALIA KM893885.1 RBG2236 AUSTRALIA KM893884.1 RBG1776 AUSTRALIA FJ985330 Fusarium oxysporum f. cubense strain NRRL 36118 KM893882.1 RBG1757 AUSTRALIA KM893883.1 RBG1780 AUSTRALIA KM893881.1 RBG1504 AUSTRALIA PLM221B FL PLM224A FL PLM385B TX CA4 CA PLM386B FL PLM387B FL PLM511A SC PDC-4701 LA FOC PLM588A-FL PLM601A-FL PLM696A-FL PLM706A CA PLM754A-FL PLM760A-FL PLM776A-FL PLM908A CA KM893870.1 RBG2242 CANARY ISLAND SPAIN PLM605A FL CA5 CA KM893869.1 RBG2241 JAPAN KM893867.1 RBG2243 CANARY ISLAND SPAIN KM893866.1 RBG2978 AUSTRALIA KM893865.1 RBG2244 FLORIDA PLM724A FL PLM183C FL AF008485.1 NRRL26035 CANARY ISLAND SPAIN AY320074.1 Fusarium foetens strain NRRL31046

131

Figure 4-2. SIX1 based phylogenetic tree of Fusarium oxysporum f. sp. canariensis (FOC) by Maximum Likelihood based on the Kimura 2-parameter model. Tree with highest log likelihood (-659.2835) is shown. The percentage of trees in which the associated taxa clustered together is shown next to the branches. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. All the FOC isolates from Laurence et al. (2015) have a RBG number. PLM905A is F. oxysporum f. sp. lycopersici.

132

KM893922.1 RBG2242 CANARY ISLAND SPAIN KM893896.1 RBG1504 AUSTRALIA KM893898.1 RBG2243 CANARY ISLAND SPAIN KM893900.1 RBG4220 AUSTRALIA KM893901.1 RBG2241 JAPAN KM893904.1 RBG2240 FLORIDA KM893906.1 RBG1776 AUSTRALIA KM893908.1 RBG4221 AUSTRALIA KM893909.1 RBG4219 AUSTRALIA KM893897.1 RBG2236 AUSTRALIA KM893912.1 RBG1780 AUSTRALIA KM893913.1 RBG2244 FLORIDA KM893914.1 RBG4216 AUSTRALIA KM893915.1 RBG1757 AUSTRALIA FOC KM893916.1 RBG2978 AUSTRALIA PLM760A SIX1 FL PLM605A SIX1 FL PLM601A SIX1 FL PLM387B SIX1 FL PLM385B SIX1 TX PLM224A SIX1 FL CA4 SIX1 CA PLM706A SIX1 CA PLM696A SIX1 FL PLM511A SIX1 SC CA5 SIX1 NV KM893917.1 RBG4223 AUSTRALIA KM893910.1 Fusarium oxysporum f. cubense RBG5183 KM893911.1 Fusarium oxysporum f. cubense RBG5181 KM893907.1 Fusarium oxysporum f. cubense RBG5182 KM893920.1 Fusarium oxysporum f. sp. lini FRL11810 KM893918.1 Fusarium oxysporum f. sp. lini FRL11811 KR855720.1 Fusarium oxysporum f. sp. conglutinans strain FRL1578 PLM905A SIX1 KR855719.1 Fusarium oxysporum f. sp. medicaginis strain FRL1582 KM893919.1 Fusarium foetens RBG4179 KR855716.1 Fusarium oxysporum f. sp. lycopersici strain RBG5768 KM893921.1 Fusarium oxysporum f. sp. lycopersici FRL4273

133

KM893936.1 RBG4215 AUSTRALIA PLM605A SIX7 FL PLM760A SIX7 FL PLM601A SIX7 FL PLM588A SIX7 FL PLM183C SIX7 FL PLM754A SIX7 FL PLM511A SIX7 SC PLM706A SIX7 CA PLM696A SIX7 FL PLM387B SIX7 FL FOC PLM385B SIX7 TX PLM224A SIX7 FL CA5 SIX7 NV CA4 SIX7 CA KM893939.1 RBG4223 AUSTRALIA KM893938.1 RBG4221 AUSTRALIA KM893933.1 RBG2243 CANARY ISLAND SPAIN KM893932.1 RBG2242 CANARY ISLAND SPAIN KM893930.1 RBG2240 FLORIDA KM893931.1 RBG2241 JAPAN KM893935.1 RBG2978 AUSTRALIA KM893937.1 RBG4220 AUSTRALIA KM893934.1 RBG2244 FLORIDA KM893929.1 Fusarium oxysporum f. sp. lini FRL11811 KP964970.1 Fusarium oxysporum f. sp. dianthi R207 KP964972.1 Fusarium oxysporum f. sp. narcissi FOXN139 KP964969.1 Fusarium oxysporum f. sp. pisi isolate FOP1 FJ755836.1 Fusarium oxysporum f. sp. lycopersici KM893940.1 Fusarium oxysporum f. sp. lycopersici FRL4273 GQ268954.1 Fusarium oxysporum f. sp. lycopersici BFOL-51 KP964969.1 Fusarium oxysporum f. sp. pisi FOP1 PLM905 SIX7 FOL KP964968.1 Fusarium oxysporum f. sp. cepae FUS2 FOC PLM386B SIX7 FL PLM221B SIX7 FL KM503196.1 Fusarium oxysporum f. cubense STR4 KM893925.1 Fusarium oxysporum f. cubense RBG5143 KM893924.1 Fusarium oxysporum f. cubense RBG4029 KM893926.1 Fusarium oxysporum f. cubense RBG5181 KM893923.1 Fusarium oxysporum f. cubense RBG3607

Figure 4-3. SIX7 based phylogenetic tree of Fusarium oxysporum f. sp. canariensis (FOC) by Maximum Likelihood method based on the Kimura 2-parameter model. Tree with the highest log likelihood (-1686.8738) is shown. The percentage of trees in which the associated taxa clustered together is shown next to the branches. A discrete Gamma distribution was used to model evolutionary rate differences among sites (5 categories (+G, parameter = 0. 0.2358). The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. All the FOC sequences from Laurence et al. (2015) have a RBG number. Other F. oxysporum formae speciales SIX7 gene sequences were obtained from GenBank. PLM905A is F.oxysporum f.sp.lycopersici (FOL).

134

KM893959.1 RBG4223 AUSTRALIA FOC PLM511ASIX10 SC CA5SIX10 NV PLM560ASIX10 PLM791ASIX10 PLM741ASIX10 FOP PLM510ASIX10 CA4SIX10 CA PLM696ASIX10 FL PLM387BSIX10 FL PLM386BSIX10 FL PLM385BSIX10 TX PLM224ASIX10 FL KM893945.1 RBG1757 AUSTRALIA KM893946.1 RBG1776 AUSTRALIA KM893947.1 RBG1780 AUSTRALIA KM893949.1 RBG2240 FLORIDA KM893950.1 RBG2241 JAPAN FOC KM893951.1 RBG2242 CANARY ISLAND SPAIN KM893952.1 RBG2243 CANARY ISLAND SPAIN KM893953.1 RBG2244 FLORIDA KM893954.1 RBG2978 AUSTRALIA KM893955.1 RBG4215 AUSTRALIA KM893956.1 RBG4219 AUSTRALIA KM893958.1 RBG4221 AUSTRALIA KM893957.1 RBG4220 AUSTRALIA KM893948.1 RBG2236 AUSTRALIA KM893943.1 Fusarium oxysporum f. sp. lycopersici isolate FRL4273 KP964985.1 Fusarium oxysporum f. sp. narcissi isolate FOXN139 KP964981.1 Fusarium oxysporum f. sp. dianthi isolate R207 KP964983.1 Fusarium oxysporum f. sp. pisi isolate FOP1 KP964987.1 Fusarium oxysporum f. sp. cepae isolate FUS2 KP964982.1 Fusarium oxysporum f. sp. lini isolate FOLIN

Figure 4-4. SIX10 based phylogenetic tree of Fusarium oxysporum f. sp. canariensis (FOC) along with F.oxysporum f. sp.palmarum (FOP) using the Maximum Likelihood method based on the Jukes-Cantor model. The evolutionary tree with the highest log likelihood (-1075.4471) is shown. The percentage of trees in which the associated taxa clustered together is shown next to the branches. A discrete Gamma distribution was used to model evolutionary rate differences among sites (5 categories (+G, parameter = 0.1578)). All positions with less than 95% site coverage were eliminated. FOC isolates from Laurence et al. (2015) have a RBG number. SIX10 homolog sequences from other F.oxysporum formae speciales were obtained from GenBank.

135

CA4 SIX12 CA PLM908A-SIX12 CA PLM760A-SIX12 FL PLM183C-SIX12 FL PLM706A SIX12 CA PLM696A SIX12 FL FOC PLM511A SIX12 SC PLM387B SIX12 FL PLM385B SIX12 TX PLM224A SIX12 FL CA5 SIX12 NV KP964992.1 Fusarium oxysporum f. sp. lini isolate FOLIN KP964993.1 Fusarium oxysporum f. sp. pisi isolate FOP1 PLM905A SIX12 KU710369.1 Fusarium oxysporum f. sp. lycopersici isolate race 3 KC701450.1 Fusarium oxysporum f. sp. lycopersici strain Fol007 KP964996.1 Fusarium oxysporum f. sp. cepae isolate FUS2 KP964991.1 Fusarium oxysporum f. sp. dianthi isolate R207 KP964995.1 Fusarium oxysporum f. sp. narcissi isolate FOXN139 KP964990.1 Fusarium oxysporum f. sp. narcissi isolate FOXN7 KP964994.1 Fusarium oxysporum isolate NRRL26988 PLM588A SIX12 FL PLM601A SIX12 FL PLM386B SIX12 FL FOC PLM221B SIX12 FL PLM605A-SIX12 FL

Palm Hosts Phoenix canariensis Phoenix reclinata Phoenix slyvestris

Figure 4-5. SIX12 based phylogenetic tree of Fusarium oxysporum f. sp. canariensis (FOC) inferred by Maximum Likelihood based on the Jukes-Cantor model. Tree with the highest log likelihood (-702.8494) is shown. The percentage of trees in which the associated taxa clustered together is shown next to the branches. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. All positions with less than 95% site coverage were eliminated. Sequences from other F. oxysporum formae speciales SIX12 gene were obtained from GenBank. PLM905A is F. oxysporum f. sp. lycopersici.

136

PLM920ASIX8 PLM336ASIX8 PLM334ESIX8 PLM246BSIX8 PLM320BSIX8 PLM510ASIX8 PLM619ASIX8 FOP PLM676ASIX8 PLM741ASIX8 PLM764ASIX8 PLM778ASIX8 PLM791ASIX8 PLM845ASIX8 PLM249ASIX8 PLM853ASIX8 KP964974.1 Fusarium oxysporum f. sp. phaseoli ATCC90245 FJ755837.1 Fusarium oxysporum f. sp. lycopersici PLM905ASIX8 KF548064.1 Fusarium oxysporum f. cubense 22615 Six8b KF548063.1 Fusarium oxysporum f. cubense 155 Six8a JF957144.1 Fusarium oxysporum f. cubense 24371 JF957143.1 Fusarium oxysporum f. cubense W95-182 JF957145.1 Fusarium oxysporum f. cubense 23707 KP964975.1 Fusarium oxysporum f. cubense E421A

Figure 4-6. SIX8 based phylogenetic tree of Fusarium oxysporum f. sp. palmarum (FOP) by Maximum Likelihood method based on the Kimura 2-parameter model. The tree with the highest log likelihood (-11935.1554) is shown. The percentage of trees in which the associated taxa clustered together is shown next to the branches. A discrete Gamma distribution was used to model evolutionary rate differences among sites (5 categories (+G, parameter = 0.2898)). The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. All positions with less than 95% site coverage were eliminated. Sequences from other F.oxysporum formae speciales SIX8 gene were obtained from GenBank. PLM905A is F. oxysporum f. sp. lycopersici.

137

PLM246B2SIX9 PLM320BSIX9 PLM510ASIX9 PLM619ASIX9 PLM676ASIX9 PLM741ASIX9 FOP PLM778ASIX9 PLM791ASIX9 PLM845ASIX9 PLM853ASIX9 PLM920SIX9 PLM249ASIX9 KR855731.1 Fusarium oxysporum f. sp. lycopersici strain RBG5768 PLM905SIX9FOL KR855725.1 Fusarium oxysporum f. sp. niveum strain RBG5771 KR855724.1 Fusarium oxysporum f. sp. passiflorae strain RBG6380 KR855738.1 Fusarium oxysporum strain RBG5850 KP964977.1 Fusarium oxysporum f. sp. cepae isolate PG KP964980.1 Fusarium oxysporum f. sp. dianthi isolate R207 KP964979.1 Fusarium oxysporum f. sp. narcissi isolate FOXN7

Figure 4-7. SIX9 based phylogenetic tree of Fusarium oxysporum f. sp. palmarum (FOP) inferred by Maximum Likelihood method based on the Kimura 2-parameter model. The tree with the highest log likelihood (-968.1485) is shown. The percentage of trees in which the associated taxa clustered together is shown next to the branches. A discrete Gamma distribution was used to model evolutionary rate differences among sites (5 categories (+G, parameter = 0.9779)). The rate variation model allowed for some sites to be evolutionarily invariable ([+I], 35.9609% sites). The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. All positions with less than 95% site coverage were eliminated. Sequences from other F.oxysporum formae speciales SIX9 gene were obtained from GenBank. PLM905A is F. oxysporum f. sp. lycopersici.

138

Figure 4-8. Genomic DNA amplification of Fusarium oxysporum f. sp. canariensis (FOC) isolates using SIX1 primer pair as viewed on 1.5 % agarose gel stained with GelRed. Lanes 1and 26 are 1 kb ladder. Lanes 2 to 20 represent FOC isolates (PLM183C, PLM 221B, PLM224A, PLM385B, PLM386B, PLM387B, PLM511A, PLM588A, PLM601A, PLM605A, PLM696A, PLM706A, PLM724A, PLM754A, PLM760A, PLM776A, CA4, CA5 and PLM908A). Lane 22 is a positive control PLM 905A (Fusarium oxysporum f. sp. lycopersici). Lane 21, 23 and 24 are negative water controls and lane 25 is blank.

139

Figure 4-9. Genomic DNA amplification of Fusarium oxysporum f. sp. canariensis (FOC) and F. oxysporum f.sp.palmarum (FOP) isolates using SIX7 primer pair as viewed on 1.5 % agarose gel stained with GelRed. Lanes 1 and 25 are 1 kb ladder. Lanes 2-11 represent different FOC isolates (PLM183C, PLM221B, PLM38B, PLM386B, PLM511A, PLM588A, PLM601A, PLM605A, PLM696A and PLM908A). Lanes 12- 21 are FOP isolates (PLM119C, PLM140B, PLM249A, PLM320B, PLM510A, PLM741A, PLM778A, PLM845A, PLM853A, and PLM920). Lane 23 is blank. Lane 23 is a positive control PLM 905A (Fusarium oxysporum f. sp. lycopersici). Lane 24 is a negative water control.

140

Figure 4-10. Genomic DNA amplification of Fusarium oxysporum f. sp. palmarum (FOP) isolates using SIX8 primer pair as viewed on 1.5 % agarose gel stained with GelRed. Lanes marked as 1 and 25 1 kb ladder. Lanes 2-22 are FOP isolates (PLM119C, PLM140B, PLM153B, PLM181C, PLM195C, PLM246B, PLM249A, PLM320B, PLM336D, PLM351B, PLM510A, PLM560A, PLM 596A, PLM619A, PLM676A, PLM 745A, PLM764A, PLM778A, PLM845A, PLM853A, PLM920). Lane 24 is positive control PLM905A (Fusarium oxysporum f. sp. lycopersici). Lane 23 is a negative water control.

141

Figure 4-11. Genomic DNA amplification of Fusarium oxysporum f. sp. palmarum (FOP) and F. oxysporum f.sp. canariensis (FOC) isolates using SIX9 primer pair as viewed on 1.5 % agarose gel stained with GelRed. Lanes 1 and 25 are 1 kb ladder. Lanes 2-11 are FOP isolates (PLM119C, PLM140B, PLM249A, PLM320B, PLM510A, PLM741A, PLM77A, PLM845A, PLM853A and PLM920). Lanes 12-21 are FOC isolates (PLM183C, PLM221B, PLM385B, PLM386B, PLM511A, PLM588A, PLM601A, PLM605A, PLM696A and PLM908A). Lanes 22 and 23 are positive control PLM905A (Fusarium oxysporum f. sp. lycopersici). Lane 24 is a negative water control.

142

Figure 4-12. Genomic DNA amplification of Fusarium oxysporum f. sp. canariensis (FOC) and F. oxysporum f. sp. palmarum (FOP) isolates using SIX10 primer pair as viewed on 1.5 % agarose gel stained with GelRed. Lanes 1 and 27 are 1 kb ladder. Lanes 2-12 are FOC isolates (PLM183C, PLM221B, PLM385B, PLM386B, PLM511A, PLM588A, PLM601A, PLM605A, PLM696A and PLM908A). Lanes 13 and 24 are blank. Lanes 14-19 are FOP isolates that are positive for SIX10 (PLM320B, PLM510A, PLM 560A, PLM596A, PLM741A and PLM 791A). Lanes 20-23 are FOP isolates that were not amplified by SIX10 primers (PLM119C, PLM 249A, PLM920, PLM853A). Lane 25 is a negative control and lane 26 is positive control PLM905A (Fusarium oxysporum f. sp. lycopersici).

143

Figure 4-13. Genomic DNA amplification of Fusarium oxysporum f. sp. canariensis (FOC) isolates using SIX12 primer pair as viewed on 1.5 % agarose gel stained with GelRed. Lanes 1 and 26 are 1 kb ladder. Lanes 2-11are FOC isolates (PLM183C, PLM 221B, PLM 385B, PLM386B, PLM511A, PLM588A, PLM601A, PLM605A, PLM696A and PLM908A). Lane 12 is blank. Lanes 13-22 are Fusarium oxysporum f. sp. palmarum (FOP) isolates (PLM119C, PLM140B, PLM249A, PLM320B, PLM510A, PLM741A, PLM778A, PLM845A, PLM853A and PLM920). Lane 23 is a negative water control. Lanes 24 and 25 are positive control PLM905A (Fusarium oxysporum f. sp. lycopersici).

144

CHAPTER 5 RESEARCH SUMMARY

Fusarium oxysporum f. sp. palmarum (FOP) and Fusarium oxysporum f. sp. canariensis

(FOC) cause lethal Fusarium wilt of palms in Florida. While FOC and FOP cause similar diseases, FOP has a wider host range and appears to be more aggressive than FOC. While FOC has a worldwide presence, FOP orginated and is present primarly in Florida. Currently there are no treatments available. The focus of the current study is to address the diagnostic needs of

Fusarium wilt pathogens on ornamental palms and lay a basic foundation in the molecular underpinning that can explain the expanding host range and aggressiveness of FOP.

Traditional detection and identification of FOP involves PCR amplification and sequencing with EF-1α primers. The first objective was to design an efficient diagnostic method to differentiate palm Fusarium wilt pathogens from other Fusarium species often associated with palms. An EF-1α based primer pair (PFW) was designed using NCBI Primer BLAST. To validate the specificity of the PFW primer set, 70 known isolates of FOP (18), FOC (18), F. solani (3), F. proliferatum (5), F. concentricum (3), F. incarnatum-equiseti species complex (4),

F. sacchari (2) and non-pathogenic Fusarium oxysporum (17) were selected and PCR was performed using three different sources of Taq polymerase. The PFW amplicon was ~ 569 bp, and all amplicons were sequenced. All FOP and FOC isolates, which had been obtained from symptomatic petiole or rachis tissue, were amplified. Twelve of 17 non-pathogenic F. oxysporum yielded an amplicon. However, 50% of these 12 isolates were recovered from roots and trunk.

No other Fusarium species were amplified with PFW primers. We suggest that symptomatic petiole or rachis is the only tissue to be used for isolation of palm Fusarium wilt pathogens.

Confirmation of the specific forma specialis (FOP vs FOC) will still require sequencing the PWF amplicon.

145

To understand the FOP expanded host range and its disease severity, we sequenced, assembled and annotated the genomes of FOC and FOP. Comparative genomics of FOP with 14

Fusarium genomes including FOC was performed to identify common and unique gene families that can explain FOP aggressiveness. The genome sizes of FOP and FOC are 46.9 Mbp and 46.0

Mbp, respectively, and the predicted proteomes consist of 15,528 (FOC) and 15,974 (FOP) proteins. Predicted proteomes of FOP and FOC were subjected to multiple functional annotations such as carbohydrate active enzymes, secretory proteins, secondary metabolites proteins family database (Pfam) and putative virulence genes. Our results indicate that 90% of the predicted FOP proteome is well conserved within FOC. The predicted proteomes of FOP and FOC share conserved pathogenic factors such as cell wall degrading enzymes, necrosis inducing proteins, cysteine rich secretory protiens, transcription factors and putative virulence associated genes.

The predicted proteome of FOP has higher secondary metabolite potential than FOC. Our studies indicated presence of at least three secondary metabolic clusters relating to nonribosomal peptide synthase (NRPS) that are unique to FOP. The orthoMCL 99 cluster, one of expanded gene family in FOP, consists of 13 proteins that are related to NRPS, indicating the possible role of toxins in FOP pathogenicity.

Along with large scale screening of predicted proteomes of FOC and FOP, we addressed the presence of Secreted in Xylem (SIX) gene family, and evaluated their presence in 47 isolates comprised of FOC (19) and FOP (28). Our results indicated that FOC has SIX1, SIX7, SIX10 and

SIX12 genes, whereas FOP has SIX8, and SIX9. However three FOC isolates did not have SIX1 gene, and six FOP isolates also had SIX10 gene, with sequences matching those of FOC SIX10.

Long term goals include designing lineage specific diagnostic markers that can differentiate the palm-infecting formae speciales that cause Fusarium wilt on ornamental palms

146

(FOC and FOP). A more comprehensive study of FOC genome structure should be conducted to identify unique and common genes and to develop hypotheses concerning differences in pathogenic specificity and palm host adapation. We hope this study will lay a foundation by describing the candidate FOP pathogenicity genes.

147

LIST OF REFERENCES

Agrios, G. N. 2005. Plant Pathology, 5th ed. Elsevier Academic Press, Burlington, MA.

Arai, K., and Yamamoto, A. 1977. New Fusarium disease of Canary Island Date Palm in Japan. Bull. Fac. Agric, Kagoshima Univ. 27:31-37.

Armstrong, G.M., and Armstrong, J.K. 1981. Formae speciales and races of Fusarium oxysporum causing wilt diseases. Pages 391-399 in: Fusarium: Diseases, Biology and . P.E. Nelson, T.A. Toussoun and R.J. Cook, eds. The Pennysylvania State University Press, University Park.

Baayen, R.P., O’Donnel, K., Bonants, P.J.M., Cigelnik, E., Kroon, L.P.N.M., Roebroeck, E.J.A., and Waalwijk, C. 2000. Gene genealogies and AFLP analyses in the Fusarium oxysporum complex identify monophyletic and non-monophyletic formae speciales causing wilt and rot disease. Phytopathology 90:891-900.

Bacon, C.W., Porter, J.K., Norred, W.P., and Leslie, J.F. 1996. Production of fusaric acid by Fusarium species. Appl. Environ. Microbiol. 62:4039–4043.

Bailey, B.A.1995. Purification of a protein from culture filtrates of Fusarium oxysporum that induces ethylene and necrosis in leaves of Erythroxylum coca. Phytopathology 85:1250– 1255.

Beckman, C.H. 1987. The Nature of Wilt Diseases of Plants. APS Press, St.Paul, MN.

Bok, J.W., Chung, D., Balajee, S.A., Marr, K.A., Andes, D., Nielsen, K.F., and Keller, N.P. 2006. GliZ, a transcriptional regulator of gliotoxin biosynthesis, contributes to Aspergillus fumigatus virulence. Infect. Immun. 74:6761–6768.

Booth, C. 1971. The Genus Fusarium. Commonwealth Mycological Institute, Kew, Surrey, UK.

Broschat, T.M., and Black, R.J. 2016. Ornamental Palms for South Florida. University of Florida, Gainesville, FL. ENH21. Online publication. http://edis.ifas.ufl.edu/ep009.

Brown, D.W., Busman, M., and Proctor, H.R. 2014. Fusarium verticillioides SGE1 is required for full virulence and regulates expression of protein effector and secondary metabolite biosynthetic genes. Mol. Plant-Microbe Interact. 27:809-823.

Brown, D.W., and Proctor, R.H. 2013. Diversity of polyketide synthases in Fusarium. Pages 143-164 in: Fusarium: Genomics, Molecular and Cellular Biology. D.W. Brown and R. H. Proctor, eds. Caister Academic Press, Norfolk, UK.

Calero-Nieto, F., Di Pietro, A., Roncero, M.I.G., and Hera, C. 2007. Role of the transcriptional activator XlnR of Fusarium oxysporum in regulation of xylanase genes and virulence. Mol. Plant-Microbe Interact. 20:977-985.

148

Casadevall, A., and Pirofski, L. 1999. Host-pathogen interactions: Redefining the basic concepts of virulence and pathogenicity. Infect. Immun.67:3703–3713.

Chakrabarti, A., Rep, M., Wang, B., Ashton, A., Dodds, P., and Ellis, J. 2011. Variation in potential effector genes distinguishing Australian and non-Australian isolates of the cotton wilt pathogen Fusarium oxysporum f.sp. vasinfectum. Plant Pathol. 60:232–243.

Chiang, Y.M., Chang, S.L., Oakley, B.R., and Wang, C.C.C. 2011. Recent advances in awakening silent biosynthetic gene clusters and linking orphan clusters to natural products in microorganisms. Curr. Opin. Chem. Biol. 15:137–143.

Christenhusz, M.J.M., and Byng, J.W. 2016. The number of known plants species in the world and its annual increase. Phytotaxa 261:201–217.

Coleman,J.J., Rounsley, S.D., Rodriguez-Carres, M., Kuo, A., Wasmann, C.C., Grimwood, J., Schmutz, J., Taga, M., White, G.J., Zhou, S., Schwartz, D.C., Freitag, M., Ma, L.J., Danchin, E.G., Henrissat, B., Coutinho, P.M., Nelson, D.R., Straney, D., Napoli, C.A., Barker, B.M., Gribskov, M., Rep, M., Kroken, S., Molnár, I., Rensing, C., Kennell, J.C, Zamora, J., Farman, M.L., Selker, E.U., Salamov, A., Shapiro, H., Pangilinan, J., Lindquist. E., Lamers, C., Grigoriev, I.V., Geiser, D.M., Covert, S.F., Temporini, E., and Vanetten, H.D. 2009. The genome of Nectria haematococca: contribution of supernumerary chromosomes to gene expansion. PLoS Genet. 5:e1000618. doi:https://doi.org/10.1371/journal.pgen.1000618.

Covey, P.A., Kuwitzky, B., Hanson, M., and Webb, K.M. 2014. Multilocus analysis using putative fungal effectors to describe a population of Fusarium oxysporum from sugar beet. Phytopathology 104:886–896.

Crešnar, B. and Petrič, S. 2011. Cytochrome P450 enzymes in the fungal kingdom. Biochim Biophys Acta 1814:29–35. doi:10.1016/j.bbapap.2010.06.020.

Daboussi, M.J., and Capy, P. 2003. Transposable elements in filamentous fungi. Annu. Rev. Microbiol. 57:275-299.

Dean, R., Van Kan, J.A.L., Pretorius, Z.A., Hammond- Kosack, K.E., Di Pietro, A., Spanu, P.D., Rudd, J.J., Dickman, M., Kahmann, R., Ellis, J., and Foster, G.D. 2012. The top 10 fungal pathogens in molecular plant pathology. Mol. Plant Pathol. 13:414–430.

De Wit, P.J.G.M., Testa, A.C., and Oliver, R.P. 2016. Fungal plant pathogenesis mediated by effectors.Microbiol.Spectra 4:6. doi: 10.1128/microbiolspec.FUNK-0021-2016.

Di Pietro, A., Madrid, M.P., Caracuel, Z., Delgado-Jarana, J., and Roncero, M. I. G. 2003. Fusarium oxysporum: exploring the molecular arsenal of a vascular wilt fungus. Mol. Plant Pathol. 4:315–325. doi:10.1046/j.1364-3703.2003.00180.x.

Downer, A.J., Hodel, D.R., Mathews, D.M., and Pittenger, D.R. 2013. Effect of fertilizer nitrogen source on susceptibility of five species of field grown palms to Fusarium oxysporum f.sp. canariensis. Palms 57:89-92.

149

Duyvesteijn, R.G., Van Wijk, R., Boer, Y., Rep, M., Cornelissen, B.J., and Haring, M.A. 2005. Frp1 is a Fusarium oxysporum F-box protein required for pathogenicity on tomato. Mol. Microbiol. 57:1051-1063.

Edel, V., Steinberg, C., Avelange, I., Laguerre, G., and Alabouvette, C. 1995. Ribosomal DNA- targeted oligonucleotide probe and PCR assay specific for Fusarium oxysporum. Phytopathology 85:579–585.

Edgar, R.C. 2004. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32:1792-1797.

Elena, K. 2005. Fusarium wilt of Phoenix canariensis: first report in Greece. Plant Pathol. 54:244.

Elliott, M.L. 2011. First report of Fusarium wilt caused by Fusarium oxysporum f. sp. palmarum on Canary Island date palm in Florida. Plant Dis. 95:356.

Elliott, M.L. 2015. First report of Fusarium wilt caused by Fusarium oxysporum f. sp. canariensis on Phoenix reclinata in Florida. Plant Dis. 99:887.

Elliott, M.L., Broschat, T.K., Uchida, J.Y., and Simone, G.W., eds. 2004. Compendium of Ornamental Palm Diseases and Disorders. APS Press, St. Paul, MN.

Elliott, M.L., Des Jardin, E.A., Harmon, C.L., and Bec, S. 2017. Confirmation of Fusarium wilt caused by Fusarium oxysporum f. sp. palmarum on x Butyagrus nabonnandii (mule palm) in Florida. Plant Dis. 101:381.

Elliott, M.L., Des Jardin, E.A., O’Donnell, K., Geiser, D.M., Harrison, N.A., and Broschat, T.K. 2010. Fusarium oxysporum f. sp. palmarum, a novel forma specialis causing a lethal disease of Syagrus romanzoffiana and Washingtonia robusta in Florida. Plant Dis. 94:31- 38.

Elliott, M.L., Honeycutt, E., West, J., and Franklin, P. 2011. First report of Fusarium wilt caused by Fusarium oxysporum f. sp. canariensis on Canary Island date palm in Texas and South Carolina. Plant Dis. 95:358.

Enright, A.J., Van Dongen, S., and Ouzounis, C.A. 2002. An efficient algorithm for large-scale detection of protein families. Nucleic Acids Res. 30:1575–1584.

EPPO. 2003. Fusarium oxysporum f. sp. albedinis. EPPO Bull. 33:265-269.

Feather, T.V., Ohr, H.D., Munnecke, D.E., and Carpenter, J.B. 1989. The occurrence of Fusarium oxysporum on Phoenix canariensis, a potential danger to date production in California. Plant Dis. 73:78-80.

Feng, B. Z., Zhu, X. P., Fu, L., Lv, R.F., Storey, D., Tooley, P., Zhang, X. G. 2014. Characterization of necrosis-inducing proteins in Phytophthora capsici. BMC Plant Biol.14:126. doi: 10.1186/1471-2229-14-126.

150

Fernandez, D., Ouinten, M., Tantaoui, A., Geiger, J.P., Daboussi, M.J., and Langin, T. 1998. Fot1 Insertions in the Fusarium oxysporum f. sp. albedinis genome provide diagnostic PCR targets for detection of the date palm pathogen. Appl. Environ. Microbiol. 64:633– 636.

Fernandez, D., Plyler, T.R., and Kistler, H.C. 1997. The Phoenix spp. pathogens Fusarium oxysporum f. sp. albedinis and F. oxysporum f. sp. canariensis are distinct genetic entities as evidenced by molecular markers. Abstract. Proceedings of the 10th Congress of the Mediterranean Phytopathological Union, Montpellier, France.

Finn, R.D., Clements, J., Arndt, W., Miller, B.L., Wheeler, T.J., Schreiber, F., Bateman, A., and Eddy, S.R. 2015. HMMER web server: 2015 update. Nucleic Acids Res. 43:W30–W38.

Flood, J. 2006. A review of Fusarium wilt of oil palm caused by Fusarium oxysporum f. sp. elaeidis. Phytopathology 96:660-662.

Fountain, J. C., Scully, B. T., Ni, X., Kemerait, R. C., Lee, R. D., Chen, Z.Y., and Guo, B. 2014. Environmental influences on maize-Aspergillus flavus interactions and aflatoxin production. Front. Microbiol. 5:40. doi: 10.3389/fmicb.2014.00040.

Fraser-Smith, S., Czislowski, E., Meldrum, R.A., Zander, M., O'Neill, W., Balali, G.R., and Aitken, E.A.B. 2014. Sequence variation in the putative effector gene SIX8 facilitates molecular differentiation of Fusarium oxysporum f. sp. cubense. Plant Pathol. 63:1044– 1052.

Gan, P., Ikeda, K., Irieda, H., Narusaka, M., O’Connell, R. J., Narusaka, Y., Takano, Y., Kubo, Y., and Shirasu, K. 2013. Comparative genomic and transcriptomic analyses reveal the hemibiotrophic stage shift of Colletotrichum fungi. New Phytol. 197:1236-1249.

Gawehns, F., Houterman, P., Ichou, F.A., Michielse, C., Hijdra, M., Cornelissen, B., Rep, M., and Takken, F.L.W. 2014. The Fusarium oxysporum effector Six6 contributes to virulence and suppresses I-2-mediated cell death. Mol. Plant-Microbe Interact. 27:336– 348.

Geiser, D.M., Jimenez-Gasco, M.M., Kang, S., Makalowska, I., Veeraraghavan, N., Ward, T.J., Zhang, N., Kuldau, G.A., and O’Donnell, K. 2004. FUSARIUM-ID v. 1.0: A DNA sequence database for identifying Fusarium. Eur. J. Plant Pathol. 110:473-479.

Giesbrecht, M., McCarthy, M., Elliott, M.L., and Ong, K.L. 2013. First report of Fusarium oxysporum f. sp. palmarum in Texas causing Fusarium wilt of Washingtonia robusta Plant Dis. 97:1511.

Giraldo, M.C., and Valent, B. 2013. Filamentous plant pathogen effectors in action. Nature Rev. Microbiol. 1:800–814.

Glass, N.L., and Donaldson, G.C. 1995. Development of primer sets designed for use with the PCR to amplify conserved genes from filamentous ascomycetes. Appl. Environ. Microbiol. 61:1323–1330.

151

Gordon, T.R., and Martyn, R.D. 1997. The evolutionary biology of Fusarium oxysporum. Annu. Rev. Phytopathol. 35:111-128.

Grigoriev, I.V., Nordberg, H., Shabalov, I., Aerts, A., Cantor, M., Goodstein, D., Kuo, A., Minovitsky, S., Nikitin, R., Ohm, R.A., Otillar, R., Poliakov, A., Ratnere, I., Riley, R., Smirnova, T., Rokhsar, D., and Dubchak, I. 2012. The genome portal of the Department of Energy Joint Genome Institute. Nucleic Acids Res. 40:D26–D32.

Gunn, L.V., and Summerell, B.A. 2002. Differentiation of Fusarium oxysporum isolates from Phoenix canariensis (Canary Island date palm) by vegetative compatibility grouping and molecular analysis. Australas. Plant Pathol. 31:351-358.

Guo, L., Han, L., Yang, L., Zeng, H., Fan, D., Zhu, Y., Feng, Y., Wang, G.,Peng, C., Jiang, X., Zhou, D., Ni, P., Liang, C., Liu, L., Wang, J., Mao, C., Fang, X., Peng, M., and Huang, J. 2014. Genome and transcriptome analysis of the fungal pathogen Fusarium oxysporum f. sp.cubense causing banana vascular wilt disease.PLoS ONE 9: e95543. http://doi.org/10.1371/journal.pone.0095543.

Guyon, K., Balague, C., Roby, D., and Raffaele, S. 2014. Secretome analysis reveals effector candidates associated with broad host range necrotrophy in the fungal plant pathogen Sclerotinia sclerotiorum. BMC Genomics 15:336. doi: 10.1186/1471-2164-15-336.

Hammond, T.M., and Keller, N.P. 2005. RNA silencing in Aspergillus nidulans is independent of RNA dependent RNA polymerases. Genetics 169:607–617.

Hernandez-Hernandez, J., Espino, A., Rodriguez-Rodriguez, J.M., Perez-Sierra, A., Leon, M., Abad-Campos, P., and Armengol, J. 2010. Survey of diseases caused by Fusarium spp. on palm trees in the Canary Islands. Phytopathol. Mediterr. 49:84-88.

Herrmann, M., Zocher, R., and Haese, A. 1996. Enniatin production by Fusarium strains and its effect on potato tuber tissue. Appl. Environ. Microbiol. 62:393-398.

Hertweck, C. 2009. The biosynthetic logic of polyketide diversity. Angew. Chem. Int. Ed. 48:4688–4716.

Hodges, A.W., Stevens, T.J., Rahmani, M., and Khachatryan, H. 2011. Economic contribution of the Florida Environmental Horticulture Industry in 2010. Online publication. http://www.fred.ifas.ufl.edu/pdf/economic-impact analysis/Economic_Contributions_Florida_Environmental_Horticulture_Industry_2010.p df.

Houterman, P.M., Cornelissen, B.J., and Rep, M. 2008. Suppression of plant resistance gene- based immunity by a fungal effector. PLoS Pathog. 4:e1000061. doi:10.1371/journal.ppat.1000061.

Houterman, P.M., Ma, L., Van Ooijen, G., De Vroomen, M.J., Cornelissen, B.J., Takken, F.L., and Rep, M. 2009. The effector protein Avr2 of the xylem-colonizing fungus Fusarium oxysporum activates the tomato resistance protein I-2 intracellularly. Plant J. 58:970–978.

152

Houterman, P.M., Speijer, D., Dekker, H.L., de Koster, C.G., Cornelissen, B.J., and Rep, M. 2007. The mixed xylem sap proteome of Fusarium oxysporum infected tomato plants. Mol. Plant Pathol. 8:215–821.

Idnurm, A. and Howlett, B. J. 2001. Pathogenicity genes of phytopathogenic fungi. Mol. Plant Pathol. 2:241–255.

Imazaki, I., Kurahashi, M., Lida, Y., and Tsuge, T. 2007. Fow2, a Zn(III)2cys6-type transcription regulator, controls plant infection of the vascular wilt fungus Fusarium oxysporum. Mol. Microbiol. 63:737-753.

Jain, S., Akiyama, K., Takata, R., and Ohguchi, T. 2005. Signaling via the G protein alpha subunit FGA2 is necessary for pathogenesis in Fusarium oxysporum. FEMS Microbiol. Lett. 243:165-172.

Jelinski, N.A., Broz, K., Jonkers, W., Ma, L.J., and Kistler, H.C. 2017. Effector gene suites in some soil isolates of Fusarium oxysporum are not sufficient predictors of vascular wilt in tomato. Phytopathology 107:842-851.

Jiménez-Gasco, M.M., Milgroom, M.G., and Jiménez-Díaz, R.M. 2002. Gene genealogies support Fusarium oxysporum f. sp. ciceris as a monophyletic group. Plant Pathol. 51:72– 77.

Jonkers, W., Andrade Rodrigues, C.D., and Rep, M. 2009. Impaired colonization and infection of tomato roots by the Δfrp1 mutant of Fusarium oxysporum correlated with reduced CWDE gene expression. Mol. Plant-Microbe Interact. 22:507-518.

Kadotani, N., Nakayashiki, H., Tosa, Y., and Mayama, S. 2003 RNA silencing in the phytophatogenic fungus Magnaporthe oryzae. Mol. Plant-Microbe Interact. 16:769–776.

Keller, N.P., Turner, G., and Benett, J.W. 2005. Fungal secondary etabolism-from biochemistry to genomics. Nat.Rev.Microbiol.3:937-947.

Khachatryan, H., and Hodges, A.W. 2012. Florida Nursery Crops and Landscaping Industry Economic Impacts Situation, and Outlook. University of Florida, Gainesville. FL. Online publication. http://edis.ifas.ufl.edu/fe946.

Khaldi, N., Seifuddin, F.T., Turner, G., Haft, D., Nierman, W.C., Wolfe, K.H., and Fedorova, N.D. 2010. SMURF: genomic mapping of fungal secondary metabolite clusters. Fungal Genet Biol. 47:736–41.

Kim, J.C., and Lee, Y.W. 1994. Sambutoxin, a new mycotoxin produced by toxic Fusarium isolates obtained from rotten potato tubers. Appl. Environ. Microbiol. 60:4380-4386.

Kinene, T., Wainaina, J., Maina, S., and Boykin, L.M. 2016. Rooting Trees, Methods for. Pages 489-493 in: Encyclopedia of Evolutionary Biology, vol. 3. R.M. Kliman, ed. Oxford Academic Press, UK.

153

Kistler, H.C. 1997. Genetic diversity in the plant-pathogenic fungus Fusarium oxysporum. Phytopathology 87:474-479.

Kistler, H.C, Rep, M., and Ma, L.J. 2013. Structural dynamics of Fusarium genomes. Pages 31- 43 in: Fusarium: Genomics, Molecular and Cellular Biology. D.W.Brown and R.H. Proctor, eds. Caister Academic Press, Norfolk, UK.

Kleemann, J., Rincon-Rivera, L.J., Takahara, H., Neumann, U., Ver Lorenvan Themaat, E., van der Does, H.C., Hacquard, S., Stuber, K., Will, I., Schmalenbach, W., Schmelzer, E., and O’Connell, R.J. 2012. Sequential delivery of host-induced virulence effectors by appressoria and intracellular hyphae of the phytopathogen Colletotrichum higginsianum. PLoS Pathog. 8:e1002643.doi:10.1371/journal.ppat.1002643.

Kombrink, A., and Thomma, B.P.H.J. 2013. LysM effectors: secreted proteins supporting fungal life. PLoS Pathog. 13:e1003769. https://doi.org/10.1371/journal.ppat.1003769.

Kou, Y., Tan, Y. H., Ramanujam, R. and Naqvi, N. I. 2016. Structure–function analyses of the Pth11 receptor reveal an important role for CFEM motif and redox regulation in rice blast. New Phytol. 214:330–342.

Krijger, J.J., Thon, M.R., Deising, H.B., and Wirsel, S.G. 2014. Compositions of fungal secretomes indicate a greater impact of phylogenetic history than lifestyle adaptation. BMC Genomics. 15:722. doi: 10.1186/1471-2164-15-722.

Kulkarni, R. D., Kelkar, H. S. and Dean, R. A. 2003. An eight-cysteine-containing CFEM domain unique to a group of fungal membrane proteins. Trends Biochem. Sci. 28:118– 121.

Kumar, S., Stecher, G., and Tamura, K. 2016. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33:1870-1874.

Larini, K., Perez-Espinosa, A., Pineda, M., and Ruiz-Rubio, M. 1996. Purification and characterization of tomatinase from Fusarium oxysporum f. sp. lycopersici. Appl. Environ. Microbiol. 62:1604-1609.

Laugé, R., Joosten, M.H.A.J., van den Ackerveken, G.F.J.M., van den Broek, H.W.J., and de Wit, P.J.G.M. 1997. The in planta-produced extracellular proteins ECP1 and ECP2 of Cladosporium fulvum are virulence factors. Mol. Plant-Microbe Interact. 10:725-734.

Laurence, M.H., Summerell, B.A., and Liew, E.C.Y. 2015. Fusarium oxysporum f. sp. canariensis: evidence for horizontal gene transfer of putative pathogenicity genes. Plant Pathol. 64:1068–1075.

Leslie, J.F., and Summerell, B.A. 2006. The Fusarium Laboratory Manual. Blackwell Publishing, Ames, IA.

154

Li, D., Chung, K.R., Smith, D.A., and Schardli,C.L. 1995.The Fusarium solani gene encoding kievitone hydratase, a secreted enzyme that catalyzes detoxification of a bean phytoalexin. Mol.Plant Microbe Interact.8:388-397.

Li, L., Stoeckert, C.J., and Roos, D.S. 2003. OrthoMCL: identification of ortholog groups for eukaryotic genomes. Genome Res.13: 2178-2189. doi:10.1101/gr.1224503.

Li, S., and Hartman, G.L. 2003. Molecular detection of Fusarium solani f. sp. glycines in soybean roots and soil. Plant Pathol. 52:74–83.

Lievens, B., Houterman, P.M., and Rep, M. 2009. Effector gene screening allows unambiguous identification of Fusarium oxysporum f. sp. lycopersici races and discrimination from other formae speciales. FEMS Microbiol. Lett. 300:201–215.

Lievens, B., Rep, M., and Thomma, B.P. 2008. Recent developments in the molecular discrimination of formae speciales of Fusarium oxysporum. Pest. Manag. Sci. 64:781– 788.

Ling, J., Zhang, J., Zeng, F., Cao, Y., Xie, B., and Yang, Y. 2016. Comparative genomics provide a rapid detection of Fusarium oxysporum f. sp. conglutinans. J. Integr. Agric. 15:822-831.

Lorenz, T.C. 2012. Polymerase Chain Reaction: Basic protocol plus troubleshooting and optimization strategies. J. Vis. Exp. 63:e3998. DOI:10.3791/3998.

Lu, S., and Edwards, M.C.2016. Genome-wide analysis of small secreted cysteine-rich proteins identifies candidate effector proteins potentially involved in Fusarium graminearum wheat interactions. Phytopathology 106:166-176.

Lucas, J.A. 1998. Plant Pathology and Plant Pathogens 3rd ed. Blackwell Publishing Ltd. Oxford. UK.

Ma, L.J., Geiser, D.M., Proctor, R.H., Rooney, A.P., O'Donnell, K., Trail, F., Gardiner, D.M., Manners, J.M., and Kazan, K. 2013. Fusarium pathogenomics. Annu. Rev. Microbiol. 67:399-416.

Ma, L.J., Houterman, P.M., Gawehns, F., Cao, L., Sillo, F., Richter, H., Clavijo-Ortiz, M.J., Schmidt, S.M., Boeren, S., Vervoort, J., Cornelissen, B.J., Rep, M., and Takken, F.L. 2015. The AVR2–SIX5 gene pair is required to activate I-2-mediated immunity in tomato. New Phytol. 208:507–18.

Ma, L.J., Shea, T., Young, S., Zeng, Q., and Kistler, H.C. 2014. Genome sequence of Fusarium oxysporum f. sp. melonis strain NRRL 26406, a fungus causing wilt disease on melon. Genome Announc. 2:e00730–14. doi 10.1128/genomeA.00730-14.

Ma, L.J., van der Does, H.C., Borkovich, K.A., Coleman, J.J., Daboussi, M.J., Di Pietro, A., Dufresne, M., Freitag, M., Grabherr, M., Henrissat, B., Houterman, P.M., Kang, S., Shim, W-B. Woloshuk C., Xie, X., Xu, J.R., Antoniw, J., Baker, S.E., Bluhm, B.H., Breakspear,

155

A., Brown, D.W., Butchko, R.A.E., Chapman, S., Coulson, R., Coutinho, P.M., Danchin, E.G.J., Diener, A., Gale, L.R., Gardiner, D.M, Goff, S., Hammond-Kosack, K.E., Hilburn, K., Hua-Van, A., Jonkers, W., Kazan, K., Kodira, C.D., Koehrsen, M., Kumar, L., Lee, Y-H., Li, L., Manners, J.M., Miranda-Saavedra, D., Mukherjee, M., Park, G., Park, J., Park, S-Y., Proctor, R.H., Regev, A., Ruiz-Roldan, M.C., Sain, D., Sakthikumar, S., Sykes, S., Schwartz, D.C., Turgeon, B.G., Wapinski, I., Yoder, O., Young, S., Zeng, Q., Zhou, S., Galagan, J., Cuomo, C.A., Kistler, H.C., and Rep, M. 2010. Comparative genomics reveals mobile pathogenicity chromosomes in Fusarium. Nature. 464:367–373.

Meerow, A.W. 1994. Betrock’s Guide to Landscape Palms. Betrock Information Systems, Davie, FL.

Meldrum, R.A., Fraser-Smith, S., Tran-Nguyen, L.T.T., Daly, A.M., and Aitken, E.A.B. 2012. Presence of putative pathogenicity genes in isolates of Fusarium oxysporum f. sp. cubense from Australia. Australas. Plant Pathol. 41:551–557.

Mercier, S., and Louvet, J. 1973. Recherches sur les fusarioses. X. Une fusariose vasculaire (Fusarium oxysporum) du palmier des Canaries (Phoenix canariensis). Ann. Phytopathol. 5:203-211.

Michielse, C.B., and Rep, M. 2009. Pathogen profile update: Fusarium oxysporum. Molecular Plant Pathol. 10:311–324.

Michielse, C.B., Studt, L., Janevska, S., Sieber, C.M.K., Arndt, B., Espino, J.J., Humpf, H.-U., Güldener, U., and Tudzynski, B. 2015. The global regulator FfSge1 is required for expression of secondary metabolite gene clusters but not for pathogenicity in Fusarium fujikuroi. Environ Microbiol. 17:2690–2708.

Michielse, C.B., van Wijk, R., Reijnen, L., Cornelissen, B.J., and Rep, M. 2009a. Insight into the molecular requirements for pathogenicity of Fusarium oxysporum f. sp. lycopersici through large-scale insertional mutagenesis. Genome Biol. 10:R4. doi: 10.1186/gb-2009- 10-1-r4.

Michielse, C.B., van Wijk, R., Reijnen, L., Manders, E.M.M., Boas, S., Olivain C, Alabouvette, C., and Rep, M. 2009b The nuclear protein Sge1 of Fusarium oxysporum is required for parasitic growth. PLoS Pathog. 5:e1000637. doi: https://doi.org/10.1371/journal.ppat.1000637.

Migheli, Q., Balmas, V., Muresu, M., Otgianu, L., and Fresu, B. 2005. First report of Fusarium oxysporum f. sp. canariensis causing Fusarium wilt on Phoenix canariensis in Sardinia, Italy. Plant Dis. 89:773.

Min, K., Son, H., Lee, J., Choi, G.J., Kim, J.C., and Lee, Y.W. 2012. Peroxisome function is required for virulence and survival of Fusarium graminearum. Molecular Plant-Microbe Interact. 25:1617-1627.

156

Nayaka, S.C., Wulff, E.G., Udayashankar, A.C., Nandini, B.P., Niranjana, S.R., Mortensen, C. N., and Prakash, H.S. 2011. Prospects of molecular markers in Fusarium species diversity. Appl. Microbiol. Biotechnol. 90:1625–1639.

Nelson, P.E., Toussoun, T.A. and Marassas, W.F.O. 1983. Fusarium species. An Illustrated Manual for Identification. The Pennsylvania State University Press, University Park, PA.

O’Donnell, K., Gueidan, C., Sink, S., Johnston. R., Crous, P.W., Glenn, A., Riley, R., Zitomer, N., Coyler, P., Waalwijk, C., Van Der Lee, T., Moretti, A., Kang, S., Kim, H.S., Geiser, D.M., Juba, J., Baayen, R.P., Cromey, M.G., Bithel, S., Sutton, D.A., Skovgaard, K., Ploetz, R., Kistler, H.C., Elliott, M., Davis, M., and Sarver, B.A.J. 2009. A two-locus DNA sequence database for typing plant and human pathogens within the Fusarium oxysporum species complex. Fungal Genet. Biol. 46:936-948.

O'Donnell, K., Kistler, H.C., Cigelnik, E., and Ploetz, R.C. 1998. Multiple evolutionary origins of the fungus causing Panama disease of banana: concordant evidence from nuclear and mitochondrial gene genealogies. Proc. Natl. Acad. Sci. USA 95:2044-2049.

Oide, S., Moeder, W., Krasnoff, S., Gibson, D., Haas, H., Yoshioka, K., and Turgeon, B. G. 2006. NPS6, Encoding a nonribosomal peptide synthetase involved in siderophore- mediated iron metabolism, is a conserved virulence determinant of plant pathogenic ascomycetes. Plant Cell. 18:2836–2853.

Okagaki, L.H., Sailsbery, J.K., Eyre, A.W., and Dean, R.A. 2016. Comparative genome analysis and genome evolution of members of the Magnaporthaceae family of fungi. BMC Genomics 17:135. doi: 10.1186/s12864-016-2491-y.

Ospina-Giraldo, M.D., Mullins, E., and Kang, S. 2003. Loss of function of the Fusarium oxysporum SNF1 gene reduces virulence on cabbage and Arabidopsis. Curr. Genet. 44:49-57.

Palmucci, H. E. 2005. Fusarium oxysporum-causal agent of wilt on crops of Phoenix canariensis in Buenos Aires Province, Argentina. New Dis. Rep. 12:5.

Pareja-Jaime, Y., Roncero, M.I.G., and Ruiz-Roldan, M.C. 2008. Tomatinase from Fusarium oxysporum f.sp. lycopersici is required for full virulence on tomato plants. Mol. Plant- Microbe Interact. 21:728-736.

Park, B., Park, J., Cheong, K.C., Choi, J., Jung, K., Kim, D., Lee, Y.H., Ward, T.J., O' Donnell, K., Geiser, D.M., and Kang, S. 2011. Cyber infrastructure for Fusarium: three integrated platforms supporting strain identification, phylogenetics, comparative genomics and knowledge sharing. Nucleic Acids Res. 39:D640–D646.

Petersen, T.N., Brunak, S., von Heijne, G., and Nielsen, H. 2011. SignalP 4.0: discriminating signal peptides from transmembrane regions. Nat. Methods 8:785–786. DOI:10.1038/nmeth.1701.

157

Plissonneau, C., Benevenuto, J., Norfarhan, M.A., Fouché, S. H.F., and Croll, D. 2017. Using population and comparative genomics to understand the genetic basis of effector-driven fungal pathogen evolution. Front. Plant Sci. 8:119. doi:10.3389/fpls.2017.00119.

Plotkin, M.J., and Balick, M.J. 1984. Medical uses of South American palms. J. Ethanopharmacol. 10:157-179.

Plyler, T.R. 1997. Genetic diversity studies on Fusarium oxysporum f. sp. canariensis and the development of a polymerase chain reaction technique for its detection. M.S. Thesis. University of Florida, Gainesville, FL.

Plyler, T.R., Simone, G.W., Fernandez, D., and Kistler, H.C. 1999. Rapid detection of the Fusarium oxysporum lineage containing the Canary Island Date Palm wilt pathogen. Phytopathology 89:407-413.

Plyler, T.R., Simone, G.W., Fernandez, D., and Kistler, H.C. 2000. Genetic diversity among isolates of Fusarium oxysporum f. sp. canariensis. Plant Pathol. 49:155-164.

Presti, L.L., Lanver, D., Schweizer, G., Tanaka, S., Liang, L., Tollot, M., Zuccaro, A., Reissmann, S., and Kahmann, R. 2015. Fungal effectors and plant susceptibility Annu. Rev. Plant Biol. 66:513-545.

Priest, M.J., and Letham, D.B. 1996. Vascular wilt of Phoenix canariensis in New South Wales caused by Fusarium oxysporum. Australas. Plant Pathol. 25:110-113.

Rabie, C.J., Marasas, W.F.O., Thiel. P.G., Lübben, A., and Vleggaar, R. 1982. Moniliformin production and toxicity of different Fusarium species from southern Africa. Appl. Environ. Microbiol. 43:517–521.

Ramos, B., Alves-Santos, F.M., Garcıa-Sanchez, M.A., Martın-Rodrigues, N., Eslava, A.P., and Dıaz-Mınguez, J.M. 2007. The gene coding for a new transcription factor (ftf1) of Fusarium oxysporum is only expressed during infection of common bean. Fungal Genet. Biol. 44:864–876.

Rep, M. 2005. Small proteins of plant-pathogenic fungi secreted during host colonization. FEMS Microbiol. Lett. 253:19–27.

Rep, M., and Kistler, H.C. 2010. The genome organization of plant pathogenicity in Fusarium species. Curr. Opin. Plant Biol. 3:420-426.

Rep, M., Van Der Does, H.C., Meijer, M., Van Wijk, R., Houterman, P.M., Dekker, H.L., De Koster, C.G., and Cornelissen, B.J.C. 2004. A small, cysteine-rich protein secreted by Fusarium oxysporum during colonization of xylem vessels is required for I-3-mediated resistance in tomato. Mol. Microbiol. 53:1373–1383.

Recorbet, G., Steinberg, C., Olivain, C., Edel, V., Trouvelot, S., Dumas-Gaudot, E., Gianinazzi, S., and Alabouvette, C. 2003. Wanted: pathogenesis-related marker molecules for Fusarium oxysporum. New Phytologist. 159:73–92.

158

Rocha, A.L.M., Di Pietro, A., Ruiz-Roldoan, C., and Roncero, M.I.G. 2008. Ctf1 a transcriptional activator of cutinase and lipase genes in Fusarium oxysporum, is dispensable for virulence. Mol. Plant Pathol. 9:293-304.

Roddick, J.G.1977. Subcellular localization of steroidal glycoalkaloids in vegetative organs of Lycopersicon esculentum and Solanum tuberosum. Phytochemistry 16:805-807.

Rusli, M.H. 2012. Detection, control and resistance expression in oil palm (Elaeis guineensis) caused by Fusarium oxysporum f. sp. elaeidis. Ph.D. Dissertation, University of Bath, Bath, UK.

Schmidt, S.M., Houterman, P.M., Schreiver, I., Ma, L., Amyotte, S., Chellappan, B., Boeren, S., Takken, F.L., and Rep, M. 2013. MITEs in the promoters of effector genes allow prediction of novel virulence genes in Fusarium oxysporum. BMC Genomics14:119. doi:10.1186/1471-2164-14-119.

Simão, F.A., Waterhouse, R.M., Ioannidis, P., Kriventseva, E.V., and Zdobnov, E.M. 2015. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics 31:3210-3212.

Simone, G.W. 2004. Fusarium wilt. Pages 17-22 in: Compendium of Ornamental Palm Diseases and Disorders. M. L. Elliott, T. K. Broschat, J. Y. Uchida, and G. W. Simone, eds. APS Press, St. Paul, MN.

Singh, R., Castro, A., Ferrin, D.M., Harris III, R.S., and Olson, B. 2011. First report of Fusarium wilt of Canary Island date palm caused by Fusarium oxysporum f. sp. canariensis in Louisiana. Plant Dis. 95:1192.

Soanes, D.M., Alam, I., Cornell, M., Wong, H.M., Hedeler, C., Paton, N.W., Rattray, M., Hubbard, S.J., Oliver, S.G., and Talbot, N.J. 2008. Comparative genome analysis of filamentous fungi reveals gene family expansions associated with fungal pathogenesis. PLoS ONE 3:e2300. doi:https://doi.org/10.1371/journal.pone.0002300.

Sperschneider, J., Dodds, P.N., Gardiner, D.M., Manners, J.M., Singh, K.B., and Taylor, J.M. 2015. Advances and challenges in computational prediction of effectors from plant pathogenic fungi. PLoS Pathog 11:e1004806. doi:https://doi.org/10.1371/journal.ppat.1004806.

Stothard, P. 2000. The Sequence Manipulation Suite: JavaScript programs for analyzing and formatting protein and DNA sequences. Biotechniques 28:1102-1104.

Summerell, B.A., Kistler, H.C., and Gunn, L.V. 2001. Fusarium wilt of Phoenix canariensis caused by Fusarium oxysporum f. sp. canariensis. Pages 263-270 in: Fusarium: Paul E. Nelson Memorial Symposium. B.A. Summerell, J.F. Leslie, D. Backhouse, W.L. Bryden, and L.W. Burgess, eds. APS Press, St. Paul, MN.

159

Suga, H., Hirayama, Y., Morishima, M., Suzuki, T., Kageyama, K., and Hyakumachi, M. 2013. Development of PCR primers to identify Fusarium oxysporum f. sp. fragariae. Plant Dis. 97:619-625.

Tajima, F. 1989. Statistical method to test for nucleotide mutation hypothesis by DNA polymorphism. Genetics 123:585-595.

Taylor, A., Vágány, V., Jackson, A.C., Harrison, R.J., Rainoni, A., and Clarkson, J.P. 2016. Identification of pathogenicity related genes in Fusarium oxysporum f. sp. cepae. Mol. Plant Pathol. 17:1032–1047.

Thatcher, L.F., Gardiner, D.M., Kazan, K., and Manners, J.M. 2012. A highly conserved effector in Fusarium oxysporum is required for full virulence on Arabidopsis. Mol. Plant-Microbe Interact. 25:180–190.

Thrane, U. 2001. Development in the taxonomy of Fusarium species based on secondary metabolites. Pages 29-49 in Fusarium: Paul E. Nelson Memorial Symposium. A. Summerell, J.F.Leslie, D.Backhouse, W.L.Bryden, and L.W.Burgess, eds. APS Press, St.Paul, MN.

Todd, R.B., Zhou, M., Ohm, R.A., Leeggangers, H.A., Visser, L., and de Vries, R.P. 2014. Prevalence of transcription factors in ascomycete and basidiomycete fungi. BMC Genomics, 15: 214. doi: http://doi.org/10.1186/1471-2164-15-214.

Tomlison, P.B., and Huggett, B.A. 2012. Cell longevity and sustanined primary growth in palm stems. Am. J. Bot. 99:1891-1902.

Urban, M., Cuzick, A., Rutherford, K., Irvine, A., Pedro, H., Pant, R., Sadanadan, V., Khamari, L., Billal, S., Mohanty, S., and Hammond-Kosack, K.E. 2017. PHI-base: a new interface and further additions for the multi-species pathogen-host interactions database. Nucleic Acids Res. 45:D604–D610.

Van Dam, P., Fokkens, L., Schmidt, S.M., Linmans, J.H.J., Kistler, H.C., Ma, L.J., and Rep, M. 2016. Effector profiles distinguish formae speciales of Fusarium oxysporum. Environ. Microbiol. 18:4087–4102. van der Does, H.C., Fokkens, L., Yang, A., Schmidt, S.M., Langereis, L., Lukasiewicz, J.M., Hughes, T.R., and Rep, M. 2016. Transcription factors encoded on core and accessory chromosomes of Fusarium oxysporum induce expression of effector genes. PLoS Genet. 12:e1006401. van der Does, H.C., Lievens, B., Claes, L., Houterman, P.M., Cornelissen, B.J.C., and Rep, M. 2008. The presence of a virulence locus discriminates Fusarium oxysporum isolates causing tomato wilt from other isolates. Environ. Microbiol. 10:1475–1485. van der Does, H.C., and Rep, M. 2007. Virulence genes and the evolution of host specificity in plant-pathogenic fungi. Mol. Plant-Microbe Interact. 20:1175–1182.

160

Van de Wouw, A.P. and Howlett, B.J. 2011. Fungal pathogenicity genes in the age of ‘omics’. Mol. Plant Pathol. 12:507-514.

VanEtten, H.D., Mansfield, J.W., Bailey, J.A., and Farmer, E.E. 1994. Two classes of plant antibiotics: phytoalexins verus ‘phytoanticipins’. Plant Cell 6:1191-1192.

Veit, S., Worle, J.M., Nurnberger, T., Koch, W., Seitz, H.U. 2001. A novel protein elicitor (PaNie) from Pythium aphanidermatum induces multiple defense response in carrot, Arabidopsis, and tobacco. Plant Physiol. 127: 832-841. DOI:10.1104/pp.010350.

Vitoreli, A.M. 2008. Development of a real time Polymerase Chain Reaction diagnostic technique for Fusarium oxysporum f. sp. canariensis on palm species. M.S Thesis. University of Florida, Gainesville, FL.

Vitoreli, A.M., Harmon, C.L., and Harmon, P.F. 2009. Development of a real time PCR diagnostic protocol for Fusarium wilt of palm. Phytopathology 99:S135.

Wall, D.P., Fraser, H.B. and Hirsh, A.E. 2003. Detecting putative orthologs, Bioinformatics 19: 1710-1711.

Wang, S., and McKie, P. 2007. Monitoring Fusarium Wilt of Canary Island date palm. Nevada State Department of Agriculture, Plant Disease Alert. 001.

Welinder, K.G.1992. Superfamily of plant, fungal and bacteria peroxidases. Curr. Opin. Struc. Biol. 2:388-393.

Welinder, K.G. 1992Wiemann, P., Willmann, A., Straeten, M., Kleigrewe, K., Beyer, M., Humpf, H.U. and Tudzynski, B. 2009. Biosynthesis of the red pigment bikaverin in Fusarium fujikuroi: genes, their function and regulation. Mol. Microbiol. 72: 931–946. doi:10.1111/j.1365-2958.2009.06695.x

Williams, A.H., Sharma, M., Thatcher, L.F., Azam, S., Hane, J.K., Sperschneider, J., Kidd, B.N., Anderson,J.P., Ghosh,R., Garg, G., Lichtenzveig, J., Kistler, H.C., Shea, T., Young, S., Buck, S.A.G., Kamphius, L.G., Saxena, R., Pande, S., Ma, L.J., Varshney, R.K., and Singh, K.B. 2016. Comparative genomics and prediction of conditionally dispensable sequences in legume–infecting Fusarium oxysporum formae speciales facilitates identification of candidate effectors. BMC Genomics. 17:191. doi:10.1186/s12864-016- 2486-8.

Wulff, E.G., Sorensen, J.S., Lübeck, M., Nielsen, K.F., Thrane, U., and Torp, J. 2010. Fusarium spp. associated with rice Bakanae: ecology, genetic diversity, pathogenicity and toxigenicity. Environ. Microbiol.12:649–657.

Wunsch, M.J., Baker, A.H., Kalb, D.W., and Bergstrom, G.C. 2009. Characterization of Fusarium oxysporum f. sp. loti forma specialis nov., a monophyletic pathogen causing vascular wilt of birdsfoot trefoil. Plant Dis. 93:58-66.

Xu, J.R. 2000. MAP kinases in fungal pathogens. Fungal Genet.Biol. 31:137-152.

161

Ye, J., Coulouris, G., Zaretskaya, I., Cutcutache, I., Rozen, S., and Madden, T. 2012. Primer- BLAST: A tool to design target-specific primers for polymerase chain reaction. BMC Bioinformatics. 13:134. doi:10.1186/1471-2105-13-134.

Yin, Y., Mao, X., Yang, J., Chen, X., Mao, F., and Xu, Y. 2012. dbCAN: a web resource for automated carbohydrate-active enzyme annotation. Nucleic Acids Res. 40:W445–451. doi:10.1093/nar/gks479.

Zhang, Z., Zhang, J., Wang, Y., and Zheng, X. 2005. Molecular detection of Fusarium oxysporum f. sp. niveum and Mycosphaerella melonis in infected plant tissues and soil. FEMS Microbiol. Lett. 249:39–47.

Zhao, Z, Liu, H, Wang, C, and Xu, J.R. 2013. Comparative analysis of fungal genomes reveals different plant cell wall degrading capacity in fungi. BMC Genomics.14:274-285 DOI:10.1186/1471-2164-14-274.

162

BIOGRAPHICAL SKETCH

Sushma Ponukumati was born in India in 1978. She completed her bachelor’s degree in science in 1998, followed by a master’s degree in microbiology in 2000. She moved to the USA in 2003. In USA, she obtained her master’s degree in biology at Florida Atlantic University in

2007. In 2014, she started her doctorate research with Dr. Monica Elliott in plant pathology focusing on palm trees. Currently she is an adjunct instructor teaching a wide range of biology courses at Broward College located in South Florida.

163