CHARACTERIZATION OF SYMBIOTIC MYCOFLORA OF NON-NATIVE AMBROSIA BEETLES

By

CRAIG CHRISTOPHER BATEMAN

A THESIS PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF MASTER OF SCIENCE

UNIVERSITY OF FLORIDA

2014

© 2014 Craig Christopher Bateman

To my parents

ACKNOWLEDGMENTS

I am greatly thankful to have received the guidance, support, time and patience of my committee chair Jiri Hulcr, and committee members Jason Smith and Matthew

Smith. My committee has helped me grow as a scientist and person, while also providing welcome and enjoyment as a member of the lab and as a friend. For discussions, technical aids, and help navigating various analyses, I would like to thank the members of the Hulcr lab; Martin Kostovcik, Caroline Storer, Andrew Johnson, and

Polly Harding, and members of the Forest Pathology lab; Adam Black, Tyler Dreaden, and Marc Hughes. I would like to thank Martin Sigut for his help extracting fungi from the black twig borer and for help with statistical analyses. I am also thankful for the advice and help on Fusarium from Kerry O’Donnell and Stacy Sink. For use of quarantined research facilities, I thank Dr. Timothy Schubert. For help with beetle collections, I am thankful for Dr. Paul Kendra and Chris Gibbard. I would also like to thank collaborators Wang Bo and Wisut Sittichaya for which work overseas would have not been possible.

4

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS ...... 4

LIST OF TABLES ...... 7

LIST OF FIGURES ...... 8

ABSTRACT ...... 9

CHAPTER

1 THE MYCANGIAL SYMBIONT OF THE BLACK TWIG BORER,Xylosandrus compactus (COLEOPTERA: CURCULIONIDAE, SCOLYTINAE) IS SPATIALLY SEGREGATED...... 11

Introduction ...... 11 Methods ...... 14 Sampling and Isolation ...... 14 DNA Sequence Acquisition and Phylogenetic Analysis ...... 15 Statistical Analysis ...... 17 Results ...... 17 Fungi Associated with Xylosandrus compactus ...... 17 The Two Dominant Fungal Associates are Spatially Segregated during Transport ...... 18 Discussion ...... 19

2 FUNGI ASSOCIATED WITH EXOTIC XYLEBORINI AMBROSIA BEETLES IN FLORIDA ...... 27

Introduction ...... 27 Methods ...... 30 Beetle Collections and Fungal Isolations ...... 30 DNA Sequence Acquisition and Analysis ...... 31 Results ...... 32 Fungi Associated with Exotic Xyleborini ...... 32 DNA Sequencing and Phylogenetic Analysis ...... 33 Discussion ...... 34

3 PREDICTING FUTURE INVADERS; ARE UNKNOWN PATHOGENS HIDING IN ASIA? ...... 44

Introduction ...... 44 The “Wait-and-see” Approach ...... 46 Preventing the Next Tree Epidemic ...... 47 Methods ...... 49

5

Study Sites and Beetle Collection ...... 49 Fungal Isolation ...... 49 Molecular Identification and Phylogenetic Analysis ...... 50 Pathogenicity Test ...... 52 Results ...... 53 Fungal Isolations and Identification ...... 53 Pathogenicity Test ...... 54 Discussion ...... 54

LIST OF REFERENCES ...... 61

BIOGRAPHICAL SKETCH ...... 68

6

LIST OF TABLES

Table page

1-1 Primers used for PCR and DNA sequencing ...... 23

1-2 Average number of fungal colonies from different parts of Xylosandrus compactus...... 23

1-3 Prevalence a of fungi isolated from different parts of Xylosandrus compactus. ... 24

1-4 Presence (+)/absence (-) of fungi isolated from beetle galleries from three localities in Gainesville, FL ...... 25

2-1 Prevalence and abundance of fungi isolated from exotic Xyleborini ...... 38

2-2 Combined reports of Xyleborini fungal symbionts in the literature and present study ...... 38

3-1 Prevalence and abundance of fungi associated with Asian bark beetles sampled...... 58

3-2 Mean number of weeks seedlings displayed resin production in loblolly pine genotype 1 inoculated with Asian fungi...... 58

3-3 Mean number of weeks seedlings displayed resin production in loblolly pine genotype 2 inoculated with Asian fungi...... 59

3-4 Mean number of weeks seedlings displayed resin production in slash pine inoculated with Asian fungi...... 59

3-5 Mean vertical xylem lesion length in loblolly pine genotype 1 inoculated with Asian fungi...... 59

3-6 Mean vertical xylem lesion length in loblolly pine genotype 2 inoculated with Asian fungi...... 59

3-7 Mean vertical xylem lesion length in slash pine inoculated with Asian fungi...... 60

3-8 Mean vertical phloem lesion length in loblolly pine genotype 1 inoculated with Asian fungi...... 60

3-9 Mean vertical phloem lesion length in loblolly pine genotype 2 inoculated with Asian fungi...... 60

3-10 Mean vertical phloem lesion length in slash pine inoculated with Asian fungi. ... 60

7

LIST OF FIGURES

Figure page

1-1 Maximum likelihood-based identification of Xylosandrus compactus mycangial isolate of Ambrosiella xylebori with Raffaelea lauricola used as an outgroup...... 25

1-2 Principal Correspondence Analysis (PCA) ordination biplot with histogram of explained cummulative variation. PCA biplot shows beetle parts colonized by different fungi...... 26

2-1 Maximum likelihood-based identification of Xylosandrus amputatus mycangial isolate Ambrosiella beaveri with Microascus cirrosus used as an outgroup ...... 39

2-2 Maximum likelihood-based identification of Xylosandrus amputatus mycangial isolate Ambrosiella beaveri with Raffaelea lauricola used as an outgroup ...... 40

2-3 Maximum likelihood-based identification of the Xyleborinus andrewesi mycangial isolate Raffaelea subalba with Ambrosiella xylebori used as an outgroup ...... 41

2-4 Maximum likelihood-based identification of the Xyleborinus andrewesi mycangial isolate Raffaelea subalba with Ambrosiella xylebori used as an outgroup ...... 42

2-5 Cladogram of main Xyleborini clades corresponding to myangium type and patterns in symbiont communnities inferred from literature and the present study...... 43

3-1 Symptoms of Ophiostoma ips on Pinus taeda seedling stems during disease development (left) and after destructive sampling (lateral bisection, right)...... 58

8

Abstract of Thesis Presented to the Graduate School of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Master of Science

CHARACTERIZATION OF SYMBIOTIC MYCOFLORA OF NON-NATIVE AMBROSIA BEETLES

By

Craig Christopher Bateman

August 2014

Chair: Jiri Hulcr Co-chair: Jason Smith Major: Forest Resources and Conservation

In virtually every forest habitat, ambrosia beetles (Coleoptera: Curculionidae:

Scolytinae, Platypodinae) plant and maintain symbiotic gardens inside dead or dying trees. Some non-native ambrosia beetles aggressively attack live trees and can damage tree crops, lumber, and native woody plant taxa by introducing phytopathogenic ambrosia fungi. Non-native ambrosia beetle introductions have become increasingly frequent in the United States, yet there is almost nothing known about their fungal symbionts. To determine the identity and diversity of symbiotic ambrosia fungi, exotic beetles were collected in Florida and their fungi were isolated.

Results support the hypothesis that some beetles carry highly specific monocultures of fungi, while others have a diverse community of symbionts that may be traded with other ambrosia beetles. These results are now being used to test hypotheses and models explaining the evolution of pathogenicity within ambrosia fungi and invasion ability within beetle-fungus complexes.

Pathogenicity displayed by ambrosia fungi was also investigated in Asia, where symbiotic fungi of beetles were isolated and evaluated for pathogenicity to American

9

trees to determine if new pathogen invasions can be effectively predicted and prevented. The fungal isolates were tested for ability to cause disease in the two most economically important pines in the southeast U.S. (Pinus taeda and P. elliottii) by inoculation in a quarantine facility. The most virulent beetle-associated fungus tested was Ophiostoma ips; an established pathogen of pines. This project demonstrates the feasibility of assessing the invasion potential of not-yet-established insects and fungi and will offer a route for regulatory agencies to effectively protect pines, one of the most valuable commodities in the southeast.

10

CHAPTER 1 THE MYCANGIAL SYMBIONT OF THE BLACK TWIG BORER,Xylosandrus compactus (COLEOPTERA: CURCULIONIDAE, SCOLYTINAE) IS SPATIALLY SEGREGATED.

Introduction

Bark and ambrosia beetles (Coleoptera: Curculionidae, Scolytinae) are normally widespread colonizers of dead or dying wood, but some species attack apparently healthy trees (Kuhnholz et al. 2001, Hulcr and Dunn 2011). This behavior is most common in invasive ambrosia beetle species and poses serious risks to tree crops, lumber industries, landscape ornamentals and native woody plants through structural damage as well as the spread of plant pathogens. Although ambrosia fungi often play a greater role in contributing to tree death or damage than the beetle, traditionally the fungal symbionts have been understudied and most remain undescribed. Information on identity of these fungi is critical for understanding the composition and function of beetle-symbiont communities. The specificity and promiscuity of beetle-fungus communities may determine the evolution of pathogenicity and invasion ability (Carrillo et al. 2013). This idea requires investigation in order to develop methods of control that reduce the economic and ecological impacts of these invasive symbiont complexes.

Xylosandrus compactus, the black twig borer, is one of the most studied ambrosia beetle species worldwide due to its long history as a pest of healthy trees

(Daehler and Dudley 2002). The female bores into and cultivates symbiotic fungal gardens in the pith of young stems on apparently healthy trees, causing death from the terminal ends of the branch to the initial entrance hole (Hara and Beardsley Jr 1976,

Wood 1982). Well over 200 tree species are recorded as hosts to X. compactus (Ngoan et al. 1976). Many commercial crops are impacted, with substantial losses recorded in avocado (McClanahan 1951), mango, tea (Kaneko et al. 1965), coffee (Brader 1964),

11

and cocoa (Hara and Beardsley Jr 1976). Nurseries can be heavily impacted, mostly by aesthetic damage to ornamentals and orchids (Dekle and Kuitert 1968, Ngoan et al.

1976). In island ecosystems such as Hawaii, endangered plants are threatened by this exotic pest (Nishida and Evenhuis 2000). Despite the diverse and far reaching impacts of X. compactus, many basic characteristics of this species remain unknown. This is especially true concerning the identity and stability of its fungal symbiont community.

Many different fungi have been described as associates of X. compactus; but three have been reported most consistently: Ambrosiella xylebori, A. macrospora, and

Fusarium solani (Brader 1964, Arx and Hennebert 1965, Batra 1967, Bhat and

Sreedharan 1988). A. xylebori has also been described in association with other beetle species, including taxa that are congeneric (e.g., Xylosandrus crassiusculus (Gebhardt et al. 2005)) and those that are more distantly related (e.g., Corthylus columbianus

(Batra 1967)). Morphological species recognition frequently fails within Ambrosiella and

Fusarium due to their phenotypic similarity. Thus, the reports of Ambrosiella macrospora from a broad range of beetles, including Ips acuminatus infesting Pinus spp. in Europe (Francke-Grosmann 1952, Batra 1967) and X. compactus (Muthappa and Venkatasubbaiah 1981), needs to be confirmed using DNA sequence data.

Fusarium solani is a name given to a complex of over 45 morphologically cryptic species (O'Donnell et al. 2008). Members of the F. solani complex (hereafter referred to as the FSSC) have been reported in association with many distantly related beetles, including Xyleborus ferrugineus (Baker and Norris 1968), Hypothenemus hampei (Rojas and Morales-Ramos 1999), Scolytodes unipunctatus (Kolarik and Hulcr 2008),

Xylosandrus compactus (Ngoan et al. 1976) and several other genera (Hulcr and

12

Cognato 2010). A clade of true ambrosia fungi was recently discovered within the FSSC that appear to have coevolved with ambrosia beetles (the Ambrosia Fusarium Clade

(AFC)), while many other cases of Fusarium reported from ambrosia beetles are likely environmental contaminants rather than true symbionts (Kasson et al. 2013). This finding raises the question whether Fusarium species reported from X. compactus belong to the AFC clade, represent a new ambrosial lineage, or are environmental contaminants.

In order to characterize the symbiosis between X. compactus and its fungal associates for use in future research on the symbiosis, this study was initiated to:

1. Quantitatively characterize the fungal community in terms of diversity and abundance distribution.

2. Determine whether different fungal symbionts co-occur in beetle mycangia during transport or are spatially separated to different beetle body parts

13

Methods

Sampling and Isolation

Fungi were isolated from live adult female beetles and infested plant material during April-May of 2013. Beetles in flight were collected from ethanol baited traps at three localities in Gainesville, Florida: NW 12th Avenue (GPS 29° 39' 47.48", -82° 20'

2.92"), Northeast Park (GPS 29° 39' 53.32", -82° 19' 15.88"), and McCarty Woods Park

(GPS 29° 38' 43.16", -82° 20' 42.37"). Five adult X. compactus beetles from each site were used for fungal isolation.

Beetles captured in flight were stored on lightly moistened paper tissue at 15° C for up to three days prior to fungus extraction. Only females were examined because

Xyleborini males lack mycangia; they are flightless and do not leave their natal galleries to establish new colonies. Fungi were isolated from four locations on the beetle: the surface, mycangium, and from within the anterior (head and pronotum) and posterior

(mesonotum to abdomen) halves of the beetle. The surface of adult females was washed by vortexing in a 1 mL sterile solution of 1% Tween 80 (Sigma Chemical Co, St.

Louis, MO, USA) and phosphate buffer saline (PBS) and then was serially diluted prior to plating. Following the wash, beetles were vortexed at 2100 rpm for 15 seconds in

1mL sterile PBS and allowed to dry on tissue paper. Live beetles were then secured onto paraffin wax using minuten pins (Bioquip, Rancho Dominguez, CA). Minuten pins were used to pry the pronotum away from the mesonotum, revealing the mycangium

(Francke 1967). Under 60X magnification (Fisher Scientific Stereomaster), the fungal mass was transferred into a 2 mL microcentrifuge tube containing 0.5 mL PBS using a flame sterilized 000 insect pin (Bioquip). Beetles were aseptically severed between pronotum and mesonotum. Each half of the beetle was then crushed in separate

14

microcentrifuge tubes containing .5 or 1 mL sterile PBS using plastic pestles (Fisher

Scientific, Suwanee, GA).

1/10, 1/100, and 1/1000 serial dilutions were made from each sample (i.e., surface wash, head, mycangium and abdomen). Dilutions were plated on Potato

Dextrose Agar (Difco Labs, Detroit, MI) amended with 1.4% additional agar (Fisher

Scientific, Fairlawn, NJ). Plates were stored in an incubator in the dark at 25° C for up to two weeks.

Fungal colony forming units (CFUs) were morphotyped and relative abundances calculated. Plates were monitored every 2-3 days during the two week incubation and photographed to ensure consistent assignment of morphotypes. Morphotype designations were confirmed by retrospective comparisons of pure cultures and by sequencing up to three portions of the nuclear ribosomal DNA (rDNA) and four protein encoding genes.

DNA Sequence Acquisition and Phylogenetic Analysis

Pure cultures between 1-2 weeks old were used for DNA extraction. Mycelium

(10-20 µl) was added to 20 uL of Sigma-Aldrich Extraction Solution (Extract-N-Amp, St.

Louis, MO) in 0.25 mL PCR tubes, which were then incubated in a thermocycler

(Eppendorf Mastercycler) at 96° C for 10 minutes. After incubation, 20 µL of 3% BSA

(Fermentas) was added to each tube, vortexed for 5 sec and then centrifuged at 2000 rpms for 30 seconds. The upper half of this solution was used as genomic DNA template for PCR.

Partial sequences were obtained from three nuclear ribosomal RNA (rDNA) encoding regions, and portions of four protein coding genes, from up to five isolates of each morphotype. A three-locus typing scheme, employing portions of the ITS+LSU

15

rDNA, EF-1α and RPB2 (O’Donnell et al. 2008), was used to identify a Fusarium isolate as a novel species within the F. solani species complex, which was designated FSSC

45-a. Arabic numbers and lowercase Latin letters are used to identify, respectively, phylospecies and haplotypes within the FSSC. Isolate Fusarium sp. FSSC 45-a was accessioned in the ARS Culture Collection as NRRL 62797 where it is available upon request (http://nrrl.ncaur.usda.gov/).

PCR reactions contained a final volume of 25 µL: 1 µL of template DNA, 1 µL of forward and reverse primer (10 µM), 0.125 µl of Taq (Takara®), 2.5 µL PCR buffer (15 mM MgCl2), 2 µL dNTP mix (2.5 mM each dNTP), and 20 uL sterile water. PCR and sequencing primers are listed in Table 1-1. PCR products were purified using Exosap-

IT, following the manufacturer’s protocols (USB Corp., Cleveland, OH). PCR reactions were performed in an Eppendorf Mastercycler Pro following published cycling parameters (Kim et al. 2004).

DNA sequencing was performed on an ABI PRISM 377XL or 3130XL automated

DNA sequencer by the Interdisciplinary Center for Biotechnology Research, University of Florida, Gainesville. Forward and reverse sequences were assembled into contigs in

Geneious (Version 7.0).

BLAST queries of the NCBI GenBank database were performed with each sequence to identify the fungi that were cultured. Final taxonomic identities were assigned by phylogenetic comparisons to congeneric sequences available in Genbank.

Maximum likelihood phylogenetic analyses were performed using PhyML (Guindon et al. 2010) with 1,000 bootstrap replicates and the TrNef+I model of substitution, which was chosen by jModelTest (Guindon and Gascuel 2003, Darriba et al. 2012).

16

Statistical Analysis

Principal component analysis (PCA) in CANOCO 5.0 (Ter Braak and Smilauer

2012) was used to describe fungal community structure. The twelve most commonly isolated species, which comprised 53 samples, were included in the PCA. Four species found in fewer than five of the beetles sampled (i.e., Epicoccum nigrum, Mycosphaerella sp., Penicillium communae) were excluded from the analysis. The effect of two variables on fungal species distribution was explored: (1) beetle part as a factor with 4 levels (head, abdomen, mycangium, surface wash) and (2) individual beetles. These were included as supplementary variables to avoid affecting the ordination (Lepš and

Šmilauer 2003). The factor of beetle individual did not add significant explanatory power to the analysis (Forward step selection analysis, p = 0.142) and was therefore excluded from the final analysis.

Results

Fungi Associated with Xylosandrus compactus

A total of 14 fungal morphotypes were isolated from adult X. compactus females captured in flight in north central Florida (Table 1-2, 1-3, and 1-4). Morphotypes could not be accurately placed at a species-level within the genus Fusarium, which was the most prevalent morphotype. The Fusarium morphotype was comprised of at least three different members of Fusarium solani species complex (FSSC) based on phylogenetic analysis of a three-locus dataset (O'Donnell et al. 2008). Fusarium was recovered from

93.3% of beetles sampled (mean CFUs per beetle = 11,131, sd = 19,244, Nbeetles =

15). The second most common fungus isolated was identified as Ambrosiella xylebori based on a 100% match to SSU and LSU rDNA sequence data from the holotype strain of A. xylebori CBS 110.61 (Figure 1-1, GenBank accessions AY858659 and EU984294,

17

respectively). ITS sequences for this isolate were not available in public databases, but were generated in this study and deposited in Genbank. Ambrosiella xylebori was isolated from 73.3% of all adult beetles sampled (mean CFU = 12,347, sd = 25,835,

Nbeetles = 15).

A yeast species in the genus Filobasidium was recovered from 66.7% of the beetles (mean = 8,934, sd = 25,489, Nbeetles = 15), and Acremonium from 53.3% of the beetles (mean = 1,024, sd = 2,968, Nbeetles = 15). CFU counts of these two fungi were appreciable but their association with specific body parts was inconsistent (Tables

1-2 and 1-3). The 10 remaining fungal species were isolated from less than 50% of the beetles (Table 1-2 and 1-3). DNA typing of fungi from galleries provided qualitative confirmation that the most frequent beetle associates were also present in active galleries (Table 1-4).

The Two Dominant Fungal Associates are Spatially Segregated during Transport

The PCA suggested that the two most common fungi isolated differ significantly in their association with beetle body parts (Figure 1-2). Specifically, there appears to be a significant connection between Ambrosiella and beetle mycangium, and between

Fusarium and beetle surface. Also, PC1 and PC2 axes separated samples from surface+abdomen in the opposite direction from head+mycangium, suggesting that different regions of the body (surface+abdomen, head+mycangium) each shares distinct communities of fungi during beetle migration. The first two axes captured 51.7% of the variability in the species data, (PC1: 31.9%, PC2: 19.8%).

18

Discussion

Results of this study established that fungi in both the genera Ambrosiella and

Fusarium are the two most common fungal associates of Xylosandrus compactus in north central Florida. Ambrosiella was isolated from two-thirds of the mycangia sampled, while Fusarium was recovered primarily from the body surface and to a lesser extent from beetle abdomens and mycangia (Table 1-2 and 1-3). Although difficulties in assigning morphotypes limited resolution of symbiont specificity to a genus level association, the results nonetheless elucidate previously unknown characteristics of the symbiosis. This is the first record of an ambrosia beetle that is associated with spatially segregated fungi. This study was also the first to demonstrate that Fusarium species associated with X. compactus are not likely to be a part of the Ambrosia Fusarium

Clade (AFC, Kasson et al. 2013), and that X. compactus is the first xyleborine ambrosia beetle consistently associated with Fusarium species that is not a member of the AFC.

Although our samples were limited to Florida, reports of Ambrosiella and Fusarium associated with X. compactus from Hawaii (Daehler and Dudley 2002, Kuo 2010) and

Japan (Hayato 2007) suggest these associations may be distributed worldwide.

The role that Fusarium plays in the ecology and life history of X. compactus, is currently unknown but deserves further study. Since it is carried on the surface and probably in the gut, but much less frequently in the mycangium, we speculate that the association is probably evolutionary recent and less specific than that between beetles and coevolved nutritional symbionts. Members of the genus Fusarium were also abundant in the gallery. During the course of this study, Fusarium spp. were typically isolated from heavily stained areas around galleries in twigs attacked by X. compactus

(Table 1-4). Two hypotheses may explain this pattern. (1) Fusarium spp. are

19

opportunists that sporulate prolifically and is competitive in the gallery. (2) Fusarium spp. are so prevalent because they are phytopathogens and suppress defense mechanisms in the twig. The difficulty of distinguishing phoretic opportunists from beneficial mutualists has been discussed extensively within the scolytine-fungus literature (Six and Wingfield 2011).

Our discovery that Ambrosiella is the most prevalent fungus in the mycangium of

X. compactus suggests they are engaged in a close-knit nutritional symbiosis. This finding appears to corroborate reports of this association from around the world

(Daehler and Dudley 2002, Hayato 2007, (Muthappa and Venkatasubbaiah 1981)), including the original description of Ambrosiella from X. compactus infesting coffee in

Ivory Coast (Brader 1964). Portions of the partial SSU and LSU rDNA sequences of A. xylebori in Florida were identical to those from the holotype strain (CBS 110.61). At present, broader geographic taxon sampling and more informative marker loci are needed to understand exactly where the symbiosis originated and whether symbiont fidelity is maintained worldwide.

It is unclear why Ambrosiella xylebori was not recovered from 26.7% of the X. compactus sampled. The absence of Ambrosiella from some beetle samples may be a methodological artifact or due to interactions between the mycangium and fungus.

Mycangia are thought to provide important nutrients for growth and storage of fungi during transport to new substrates (Francke-Grosmann 1967). We observed that mycangia of beetles taken from galleries always appeared empty, while those caught in flight typically appeared filled with fungus. Fungi inhabiting the mycangium may need time to grow sufficiently, or rely on cues from the beetle before inoculation and growth in

20

a new gallery (Kinuura 1995). The beetles that did not yield Ambrosiella may have been captured before the symbiotic fungus had sufficiently colonized the mycangium. The high variability in past studies in reporting Ambrosiella from X. compactus may be due to inconsistent methodology, or due to development-induced variation in symbiont abundance.

A yeast in the genus Filobasidium (Kwon-Chung 1977, Pan et al. 2011) was also recovered from more than half of the beetles as well as from a gallery. Laboratory rearing experiments are needed to assess what role, if any, F. uiguttatum, Acremonium sp. and the 10 other less frequently sampled fungi play in the life history of X. compactus.

Our study revealed that isolation from different X. compactus body parts yielded completely different symbiont communities. However, the dominant fungi from all body parts were also found in beetle galleries. This finding suggests that mechanisms may exist that restrict these symbionts to different parts of the beetle body, or that they are adapted for transport on beetles differently. For example, the abundant Fusarium spores may easily adhere to the beetle surface and/or survive passage through the gut, while Ambrosiella appears to be selected for or is competitive within the mycangium.

Secretions in the mycangium may provide an advantage to specialized coevolved symbionts (i.e., Ambrosiella).

One of the goals of this study was to establish a reliable approach to culture- based analyses of ambrosia beetle symbionts that considers insect anatomical and ontogenetic complexity. Based on these results, the traditional approach of processing whole beetles may not be appropriate for capturing the fine-scale community structure

21

of the symbionts. This and many such beetle-fungal symbioses are ecologically and economically destructive, and need to be studied carefully to elucidate theoretical and practical aspects of their biology. Future studies that focus on, the rapid emergence of fungal pathogenicity and virulence in formerly harmless ambrosia beetles are needed to devise effective methods for the detection, control and eradication of these exotic pests.

22

Table 1-1. Primers used for PCR and DNA sequencing

Primer Locus Gene product Primer sequence (5'-3') a PCR S b Reference name

NS4 CTTCCGTCAATTCCTTTAAG • • (White et al. 1990) Internal ITS

transcribed ITS1 TCCGTAGGTGAACCTGCGG • • (White et al. 1990) rDNA spacer

ITS4 TCCTCCGCTTATTGATATGC • • (White et al. 1990) Nuclear LSU (Vilgalys and Hester ribosomal large NL1 CTTGGTCATTTAGAGGAAGTAA • • rDNA 1990, O'Donnell 1992) subunit (28S) (Vilgalys and Hester LR3 CCGTGTTTCAAGACGGG • •

1990) (Vilgalys and Hester LROR ACCCGCTGAACTTAAGC • •

1990) Translation

EF-1α elongation EF1 ATGGGTAAGGARGACAAGAC • (O’Donnell et al. 1998) factor 1α

EF2 GGARGTACCAGTSATCATG • (O’Donnell et al. 1998)

EF3 GTAAGGAGGASAAGACTCACC • (O'Donnell et al. 2008)

EF22T AGGAACCCTTACCGAGCTC • (O’Donnell et al. 1998) RNA

RPB1 polymerase Fa CAYAARGARTCYATGATGGGWC • (Hofstetter et al. 2007) largest subunit

G2R GTCATYTGDGTDGCDGGYTCDCC • (O'Donnell et al. 2010)

R8 CAATGAGACCTTCTCGACCAGC • (O'Donnell et al. 2010)

F5 ATGGGTATYGTCCAGGAYTC • (O'Donnell et al. 2010)

F6 CTGCTGGTGGTATCATTCACG • (O'Donnell et al. 2010)

F7 CRACACAGAAGAGTTTGAAGG • (O'Donnell et al. 2010)

F8 TTCTTCCACGCCATGGCTGGTCG • (O'Donnell et al. 2010)

R9 TCARGCCCATGCGAGAGTTGTC • • (O'Donnell et al. 2010) RNA polymerase

RPB2 5f2 GGGGWGAYCAGAAGAAGGC • • (Reeb et al. 2004) second largest subunit

7cr CCCATRGCTTGYTTRCCCAT • • (Liu and Hall 2004)

7cf ATGGGYAARCAAGCYATGG • • (Liu and Hall 2004)

11ar GCRTGGATCTTRTCRTCSACC • • (Liu and Hall 2004) (O'Donnell and benA β- tubulin T1 AACATGCGTGAGATTGTAAGT • •

Cigelnik 1997) (Glass and Donaldson BT2B ACCCTCAGTGTAGTGACCCTTGGC • •

1995) a D = A, G or T; R = A or G; S = C or G; W = A or T; Y = C or T. b S= used in sequencing

23

Table 1-2. Average number of fungal colonies from different parts of Xylosandrus compactus. Species Surface Head Mycangium Abdomen Acremonium sp. 177.4 206.7 133.3 281.3 Ambrosiella spp. 0 333.4 1944.4 2252.5 Cladosporium sp. 163.3 124 63.3 257.3 Epicoccum nigrum 7.3 66.7 0 0 Filobasidium sp. 5415.6 1372.6 89.3 70.3 Fusarium spp. 2565 381.3 780.6 1206 Fusarium verticillioides 74.7 0 0 6.7 Meira sp. 1333 0 0 0 Mycosphaerella sp. 0 26.7 0 0 Penicilium minioluteum 161 66.7 0 6.7 Penicillium communae 0 13.3 0 0 Pestalotiopsis sp. 0.66 108.6 3691 80 Phaeoacremonium scolyti 0 0 0 146.7 Phialemonium sp. 18129 7491 4933 11621

Table 1-3. Prevalence a of fungi isolated from different parts of Xylosandrus compactus. Total Species Surface Head Mycangium Abdomen Prevalence Acremonium sp. 46.7 13.3 6.7 26.7 53.3 Ambrosiella spp. 0 53.3 66.7 26.7 73.3 Cladosporium sp. 26.7 33.3 26.7 26.7 40 Epicoccum nigrum 13.3 6.7 0 0 20 Filobasidium sp. 40 33.3 33.3 20.0 73.3 Fusarium spp. 80 33.3 20 66.7 93.3 Fusarium verticillioides 20 0 0 6.7 20 Meira sp. 6.7 0 0 0 6.7 Mycosphaerella sp. 0 6.7 0 0 6.7 Penicilium minioluteum 13.3 6.7 0 6.7 20 Penicillium communae 0 6.7 0 0 6.7 Pestalotiopsis sp. 6.7 26.7 13.3 13.3 46.7 Phaeoacremonium scolyti 0 0 0 6.7 6.7 Phialemonium sp. 20.0 13.3 6.7 26.7 13.3 a Percent samples yielding fungus species

24

Table 1-4. Presence (+)/absence (-) of fungi isolated from beetle galleries from three localities in Gainesville, FL: NW = NW 12th Avenue, NP = Northeast Park, MWP = McCarty Woods Park, only species isolated at least once are included.

Species NW NP MWP

Ambrosiella spp. + + - Filobasidium sp. + - - Fusarium spp. + - + Fusarium verticillioides + - +

Phialemonium sp. - + -

Figure 1-1. Maximum likelihood-based identification of Xylosandrus compactus mycangial isolate of Ambrosiella xylebori with Raffaelea lauricola used as an outgroup. Ingroup sequences are from members of the genera Ceratocystis and Ambrosiella (Microascales) Relationships were inferred from combined SSU+LSU rDNA sequence data. Numbers next to nodes are maximum likelihood bootstrap values based on 1,000 bootstrap replicates.

25

Figure 1-2. Principal Correspondence Analysis (PCA) ordination biplot with histogram of explained cummulative variation. PCA biplot shows beetle parts colonized by different fungi.

26

CHAPTER 2 FUNGI ASSOCIATED WITH EXOTIC XYLEBORINI AMBROSIA BEETLES IN FLORIDA

Introduction

In recent years, exotic wood boring beetles have become some of the most common and the most economically and ecologically detrimental insects being introduced into the United States. Included in some of these groups of invasive beetles are ambrosia beetles (Coleoptera: Curculionidae, Scolytinae), which are commonly transported across the globe and are more likely than native beetles to cause harm to naïve forest ecosystems (Hulcr & Dunn 2011). Fungi associated with ambrosia beetles are also likely to be introduced with the arrival of each new beetle, but unfortunately most of these fungi remain mostly undetected and undescribed.

Most scolytine beetles exist in some kind of nutritional relationship with fungi

(Batra 1963, Beaver 1989). In ambrosia beetles and fungi, the association is obligate.

Larvae and adult beetles must consume fungi to survive, and in turn the beetles relocate their fungal partners to new hosts. Most ambrosia fungi are comprised of the anamorphic ascomycete genera Ambrosiella, Raffaelea, Fusarium, Geosmithia, and

Dryadomyces (Batra 1963, Harrington et al. 2010, Kasson et al. 2013). An earlier hypothesis suggests that fungal symbiont taxa are highly specific and coevolved partners of beetles, with only one fungus per beetle (Francke-Grosmann 1967, Mueller

2005). More detailed analyses indicate that many ambrosia beetles have a community of symbionts, and that fungal symbionts can be lost or incorporated from the environment (Batra 1966, Carrillo et al 2013, Kostovcik et al. in press). Some ambrosia beetles have been recorded to share a fungal symbiont. In some instances, this occurs through mycokleptism (Hulcr and Cognato 2010). In other cases, it may be the fungus

27

or beetle that initiates the new relationship, and the mechanisms behind this are unknown. Either organism may be influenced to initiate new symbiotic relationships depending on many temporal and spatial conditions in which the association exists. This could lead to a varying community of fungal symbionts, especially in beetles that have broad distributions. To understand this process, the symbionts of many exotic ambrosia beetles in the United States require investigation. The beetles included in this study do not form an exhaustive list of exotic ambrosia beetles in the United States, but provide a snapshot of diversity, and from multiple clades, of exotic beetles already established.

Many exotic ambrosia beetles continue to enter the United States each year and for most of these species there are no data on their ecological role or their symbiotic fungi in the non-native environment. No fungal associates have yet been recorded from

Dryoxylon onoharaensum, which has been observed attacking Acer (Bright and

Rabaglia 1999) and Populus spp. (Ghandi et al. 2010), but has caused very minimal damage since its discovery in 1990 (Atkinson et al. 1990).

Xylosandrus amputatus is another beetle that is not yet known to attack healthy native trees in the U.S.; its fungal associates remain unknown (Kajimura 1998).

Harrington (2014) reported A. beaveri, the symbiont of C. mutilatus (another recently exotic ambrosia beetle in the U.S., Six et al. 2009), to be growing inside the galleries of

X. amputatus but did not report fungi isolated from the beetle’s mycangium.

Another Asian ambrosia beetle, Xyleborinus andrewesi, was first detected in

North America in 2009. This beetle has not yet been found in live trees in the U.S., and its associated fungi are unknown (Okins and Thomas 2009). These two beetles (X.

28

amputatus and X. andrewesi) are relatively recent introductions, so their impacts thus far are uncertain.

Comparatively, the ambrosia beetle Cyclorhipidion bodoanum has been present in the U.S. for a much longer period (Wood 1982). Although C. bodoanum seems to prefer dead or dying trees like most ambrosia beetles, it has been implicated as a possible vector of the oak pathogen Phytopthora ramorum (McPherson et al 2010).

McPherson (2010) identified Hypocrea lixii, Mucor racemosus f. racemosus, Trametes viride, Emericellopsis sp., Ophiostoma sp., Pleosporaceae sp. and Trichoderma sp. as possible associates of C. bodoanum (McPherson et al. 2013). The wide variety and inconsistent association of these fungi is likely due to the methods of this study. Entire beetles were crushed and plated, which has much greater potential to yield a mix of saprobic contaminants, opportunists, and endosymbionts. This hypothesis is supported by the fact that H. lixii and M. racemosus were also isolated from sapwood that was not colonized by beetles. Trametes fungi are white-rotting wood decay fungi which have no known anamorphs and unlikely symbionts of ambrosia beetles. Trichoderma spp. are common fungal contaminants in all environmental samples. The remaining fungal species isolated could not be described to the species level. With the exception of the

Ophiostoma sp., all other fungi isolated are unlikely nutritional ambrosia symbionts. It is especially likely that this Ophiostoma sp. is the primary symbiont, since Ophiostoma stenoceras has been reported as a symbiont of C. bodoanum in an earlier study

(Gebhardt et al. 2005).

As evident from the numerous ambrosia beetle introductions into the U.S., some beetles are able to become widespread or attack living trees, while others are not. A

29

greater understanding of the diversity and specificity of their symbiotic fungi will help elucidate mechanisms that govern invasion and pathogen success. These studies are critical because new, potentially dangerous exotic beetles continue to enter the country and old introductions continue to degrade native ecological and agricultural systems.

Methods

Beetle Collections and Fungal Isolations

Fungi were isolated from living, adult foundress female beetles caught in flight using ethanol or freshly cut wood as lures. Beetle dissections took place under a dissection microscope in an ethanol-sterilized environment. Prior to beetle dissection, all beetles were individually surface washed in a 1 mL sterile solution of 1% Tween 80

(Sigma Chemical Co, St. Louis, MO, USA) and PBS, which was then was serially diluted prior to plating. For direct isolation of fungi from beetle mycangia, different dissection procedures were performed on all specimens depending on the location of the mycangium inside the different beetle species. Beetles with mycangia that were not known (Dryoxylon onoharaensum) or beetles with mandibular mycangia (Euwallacea spp.) were aseptically separated into three parts using a scalpel prior to maceration and dilution: head, thorax, and abdomen. Beetles with mesonotal mycangia (Xylosandrus amputatus) were fixed onto paraffin wax using sterilized 000 insect pins so that the pronotum could be pried away from the abdomen. Fungi were then directly transferred from the mesonotal mycangium to 1 mL of sterile PBS prior to serial dilution. The head+pronotum and abdomen (elytra removed) were also macerated and serially diluted. Beetles inferred to have metanotal mycangia (Xyleborinus andrewesi (Francke-

Grosmann 1967, Beidermann et al. 2013)), were aseptically sectioned into three parts using a sterile scalpel: head+pronotum, mesonotum+metanotum, and abdomen.

30

Serial dilutions of 1/10, 1/100, and 1/1000 were made from each sample (location of beetle body). Dilutions were plated on Potato Dextrose Agar amended with 1.4% additional agar. After isolation and dilution of fungi from beetles, fungi were stored in an incubator in the dark at 25° C for up to two weeks to allow for elucidation of colony characteristics.

Morphotypes were assigned based on macromorphology (ie, color, size comparison/growth rate, texture) for all fungal colony forming units (CFUs) occurring in more than one sample. Plates were monitored every 2-3 days during the two-week incubation and photographed to ensure consistent assignment of morphotypes.

Morphotype designations were confirmed by retrospective comparisons of pure cultures and by sequencing portions of the nuclear ribosomal DNA (rDNA).

DNA Sequence Acquisition and Analysis

Mycelia from 1-2 week old pure cultures were used for DNA extraction. 10-20 µl of mycelium was added to 20 uL of Sigma-Aldrich Extraction Solution (Extract-N-Amp,

St. Louis, MO) in 0.25 mL PCR tubes, which were then incubated (Eppendorf

Mastercycler) at 96° C for 10 minutes. After incubation, 20 µL of 3% BSA (Fermentas) was added to each tube, vortexed for 5 seconds and then centrifuged at 2000 rpms for

30 seconds. The upper half of this solution was used as genomic DNA template for

PCR.

PCR reactions contained a final volume of 25 µL: 1 µL of template DNA, 1 µL of forward and reverse primer (10 µM), 0.125 µl of Taq (Takara®), 2.5 µL PCR buffer (15 mM MgCl2), 2 µL dNTP mix (2.5 mM each dNTP), and 20 uL sterile water. PCR and sequencing primers are listed in Table 1-1. PCR products were purified using Exosap-

IT, following the manufacturer’s protocols (USB Corp., Cleveland, OH). PCR reactions

31

were performed in an Eppendorf Mastercycler Pro following published cycling parameters (Kim et al. 2004).

DNA sequencing was performed on an ABI PRISM 377XL or 3130XL automated

DNA sequencer by the Interdisciplinary Center for Biotechnology Research, University of Florida, Gainesville. Forward and reverse sequences were assembled into contigs in

Geneious (Version 7.0). Partial sequences were obtained from up to three nuclear ribosomal RNA (rDNA) encoding regions and up to five isolates per morphotype.

BLAST queries of the NCBI GenBank database were performed with each sequence to identify the fungi that were cultured. Final taxonomic identities were assigned by phylogenetic comparisons to congeneric sequences available in Genbank.

Maximum likelihood (ML) phylogenetic analyses for fungi associated with Xylosandrus amputatus and Xyleborinus andrewesi were performed using PhyML (Guindon et al.

2010) with 1,000 bootstrap replicates. Analyses performed on fungi isolated from X. amputatus utilized a TIM2ef+I substitution model for b-tubulin trees, and a TrNef+G model for LSU trees. The TrN+G model of substitution was used for fungi associated with X. andrewesi. All substitution models used in ML analyses were chosen by jModelTest (Guindon and Gascuel 2003, Darriba et al. 2012).

Results

Fungi Associated with Exotic Xyleborini

Isolations of fungi from the mycangia of four species of exotic ambrosia beetles yielded one highly abundant and prevalent filamentous fungus morphotype with each ambrosia beetle species sampled. No consistent filamentous or yeast fungi were isolated from Dryoxylon onorenhaesum. Seven morphotypes of filamentous fungi were isolated from Xylosandrus amputataus, but only one morphotype was recovered from

32

both adult beetle mycangia and a gallery, and in high frequency and abundance (Figure

2-1). Similarly, six morphotypes were isolated from X. andrewesi, but only one was found to be consistently associated with the beetle’s mycangia or abdomen, where symbionts have been reported to be transported in this beetle genus (Francke-

Grosmann 1967, Beidermann et al. 2013). In all Euwallacea validus beetles sampled, three morphotypes of filamentous fungi were recovered, with two being found on 100% of beetles sampled. Euwallacea interjectus specimens sampled also yielded three filamentous fungal morphotypes, with one morphotype recovered from all beetles.

DNA Sequencing and Phylogenetic Analysis

Blast queries of the NCBI GenBank database indicated that partial sequences generated from the large subunit (LSU) rDNA and the β-tubulin gene from the most abundant filamentous fungus morphotype associated with Xylosandrus amputatus was most similar to Ambrosiella beaveri CBS 121751 with 99% similarity in overlapping regions (Genbank accessions EU825650, EU825656). After alignment with all closely related Ambrosiella and Ceratocystis spp., ML phylogenetic analysis of LSU and β - tubulin sequences indicated A. beaveri and the fungus isolate from X. amputatus formed a monophyletic group (Figures 2-2, 2-3).

The most abundant morphotype isolated from Xyleborinus andrewesi was found to be most similar to Raffaelea subalba (isolate C2224, GenBank accession EU177467) based on BLAST queries using partial sequences of LSU rDNA. Sequence data from other loci was unavailable for R. subalba in GenBank. Maximum likelihood phylogenetic analysis using LSU sequences revealed the isolate to form a monophyletic group with

Raffaelea subalba (Figure 2-4).

33

Two different filamentous fungi isolated from Euwallacea interjectus and E. validus were found to be members of the Fusarium solani species complex (FSSC) and were subsequently found to be part of a novel clade of ambrosia fungi within the FSSC by congruence between all three methods of phylogenetic inference using a four-locus data set (described in Kasson et al. 2013).

In addition to a novel Fusarium sp. being described from E. validus, another filamentous fungus isolated from 100% of beetles sampled was identified. Initial BLAST queries of LSU sequence data indicated this second filamentous fungus to be 99% similar to Raffaelea subfusca (isolate C2335, GenBank accession EU177450) in overlapping regions.

Discussion

Fungal isolations from beetle mycangia and identification of consistent fungal associates through sequencing and phylogenetic analyses performed here have helped to identify symbiotic fungi associated with four exotic ambrosia beetles in the United

States. Phylogenetic analyses of LSU and B-tubulin genes indicate that Ambrosiella beaveri is strongly associated with Xylosandrus amputatus, Raffaelea subalba with

Xyleborinus andrewesi, and R. subfusca and Fusarium sp. with Euwallacea validus.

It is interesting that each of these three fungi has been previously reported in association with other ambrosia beetle species, especially in the case of Xylosandrus amputatus. All beetles within the genus Xylosandrus have been previously appeared to have strict symbioses with only one Ambrosiella species, which is apparently not maintained in symbiosis with any other beetle species. However, in addition to A. beaveri being reported as a symbiont of X. amputatus in the present study, Six (2009) reported A. beaveri to be dominant fungus found in the mycangia of Cnestus mutilatus.

34

If Ambrosiella species are indeed specific to only one beetle species, it is possible that

A. beaveri is associated with only one of beetle species in Asia where both beetles are native, and that symbiont communities have changed following the introduction into a non-native territory. Although if Xylosandrus and Ambrosiella are tightly co-evolved, it is also more likely that symbiont host switches will occur with closely related beetles and fungi. Reports of host switching in insect-fungus symbioses are more common in closely related species (Mueller and Gerardo 2002, Nobre et al. 2011). Indeed, C. mutilatus is closely related to X. amputatus and both have similar mesonotal mycangia (Cognato et al. 2011, Dole and Cognato 2010).

Another explanation for the apparent symbiont sharing is that the loci used in identification of Ambrosiella spp. are unable to resolve cryptic species-level identities.

Other clades of hyperdiverse fungi, such as many Fusarium spp., require the use of six or more loci to differentiate species (O’Donnell et al. 2008). Further studies need to incorporate more genetic loci in species descriptions and phylogenetic analyses.

Furthermore, it’s unknown how diverse the genus of Ambrosiella is, as the vast majority of ambrosia beetles have never had their fungal communities characterized. The lack in studies investigating beetle communities also makes it more difficult to establish how often fungi are traded or shared among different beetle species.

Two other symbionts of beetles reported here, R. subalba and R. subfusca associated with X. andrewesi and E. validus respectively, were also originally reported to be a symbiont of another beetle, Xyleborus glabratus (Harrington et al. 2010). Many additional Raffaelea spp. have been recovered from diverse fungal communities within the mycangium of X. glabratus. Due to limited geographic sampling and insufficient

35

inclusion of most ambrosia beetle species in symbiont surveys, it’s again unclear if these relationships uphold across the distribution of the beetle, and if exotic beetles are changing native beetle symbiont communities.

Although results from the present study are limited in geographic scope, they still provide contribute to a broader understanding of patterns in symbiont communities across Xyleborini ambrosia beetles. Clearer patters only emerge when combining other culture-based symbiont reports from the literature. A summary of these reports is provided in Table 2-2, and visualized on a cladogram in Figure 2-5. Only recent studies incorporating molecular markers in fungal identification are included in Table 2-2.

These combined reports of symbiont surveys suggest that the number and identity of symbiotic fungi associated with the beetles varies substantially, and appears to relate to beetle clade. The fidelity and diversity of beetle-fungus relationships may be influenced by mycangium type, which is specific to each beetle clade. Mandibular mycangia, which often host diverse and fluid communities of fungi, may be more permissive to the addition of new symbionts because any fungus traveling through the beetle mouth (either through wood boring or consumption) has a chance of entering the mycangium. Mesonotal mycangia, which are well protected underneath the pronotum, may be less prone to accessory fungal contamination.

The diverse ambrosia fungi may also be differently adapted to surviving in different mycangia. Ambrosiella species, which are only consistently found in mesonotal mycangia, may be more co-evolved and rely on specific conditions or nutrients only found in their corresponding mycangium. Raffaelea species, which are also found in a variety of mycangium types, may be less co-evolved with any single beetle clade and

36

possess traits that allow them to be easily disseminated and acquired by new ambrosia beetle species. However, inferences of co-evolutionary relationships are dependent on comprehensive taxon sampling of symbionts and dating of divergence times which are still unavailable for ambrosia beetles and fungi.

Future studies need to utilize both culture and culture-independent sampling of symbiont communities to correct for culture bias, allow for higher throughput analyses, and more precisely characterized communities. Experimental evidence also needs to establish roles of symbionts and how the beetles and fungi interact in a natural landscape to determine what selective pressures are at play in the symbiosis. The results in the present study can be used in future research by directing taxon sampling efforts, and by establishing the current symbionts of exotic ambrosia beetles for future investigation of temporal and geographic variation in symbiont communities.

37

Table 2-1. Prevalence and abundance of fungi isolated from exotic Xyleborini Prevalence Mean CFU Stand. Beetle species Fungal isolate on beetle (%) recovereda Dev. Euwallacea interjectus Fusarium sp. 100% 2,622.2 2,095.38

Euwallacea validus Fusarium sp. 100% 373.2 312.88

Raffaelea subfusca 100% 278.8 283.47

Xylosandrus amputatus Ambrosiella beaveri 100% 1,088 1,495.7

Xyleborinus andrewesi Raffaelea subalba 86% 1,513 2,922.9 a = recovered from location beetle known to transport fungi

Table 2-2. Combined reports of Xyleborini fungal symbionts in the literature and present study Cumulative Mycangium Core fungal Beetle # of Reference type associates associates Mandibular Xyleborus spp. Raffaelea spp. (some- Harrington et al. 2010, times Fusarium sp., Kasson et al. 2013, Carrillo et >6 others) al. 2013, Ploetz et al. 2011, Beidermann et al. 2009

Mandibular Euwallacea spp. Fusarium spp. (some- Kasson et al. 2013, Present times Raffaelea sp., >2 study others)

Metanotal Xyleborinus spp. Raffaelea spp. Biedermann at al. 2013, 1-2 /abdomen Present study

Mesonotal Xylosandrus spp. Ambrosiella spp. Harrington & Fraedrich 2010, 1 Six et al. 2009, Present study

38

Figure 2-1. Maximum likelihood-based identification of Xylosandrus amputatus mycangial isolate Ambrosiella beaveri with Microascus cirrosus used as an outgroup. Ingroup sequences are from species of Ceratocystis and Ambrosiella (Microascales) Relationships were inferred from β-tubulin sequence data. Numbers next to nodes are maximum likelihood bootstrap values based on 1,000 bootstrap replicates.

39

Figure 2-2. Maximum likelihood-based identification of Xylosandrus amputatus mycangial isolate Ambrosiella beaveri with Raffaelea lauricola used as an outgroup. Ingroup sequences are from species of Ceratocystis and Ambrosiella (Microascales) Relationships were inferred from LSU rDNA sequence data. Numbers next to nodes are maximum likelihood bootstrap values based on 1,000 bootstrap replicates.

40

Figure 2-3. Maximum likelihood-based identification of the Xyleborinus andrewesi mycangial isolate Raffaelea subalba with Ambrosiella xylebori used as an outgroup. Ingroup sequences are from species of Raffaelea (Ophiostomatales). Relationships were inferred from LSU rDNA sequence data. Numbers next to nodes are maximum likelihood bootstrap values based on 1,000 bootstrap replicates.

41

Figure 2-4. Maximum likelihood-based identification of the Xyleborinus andrewesi mycangial isolate Raffaelea subalba with Ambrosiella xylebori used as an outgroup. Ingroup sequences are from species of Raffaelea (Ophiostomatales). Relationships were inferred from β-tubulin sequence data. Numbers next to nodes are maximum likelihood bootstrap values based on 1,000 bootstrap replicates.

42

Figure 2-5. Cladogram of main Xyleborini clades corresponding to myangium type and patterns in symbiont communnities inferred from literature and the present study.

43

CHAPTER 3 PREDICTING FUTURE INVADERS; ARE UNKNOWN PATHOGENS HIDING IN ASIA?

Introduction

Forest ecosystems across the world are rapidly undergoing changes in form and function as a result of non-indigenous (or non-native, exotic) species. The arrival of exotic insects and fungi in many cases has initiated a shift in forest compositions and extirpation or extinctions of tree species or their associated communities of organisms

(Lovett et al. 2006 , Gandhi and Herms 2010, Brokerhoff et al. 2006). The near removal of American chestnut (Castanea dentata) in North America due to Chestnut blight, caused by the fungus Cryphonectria parasitica from Asia, has been accompanied by changes in forest composition, increased tree diversity, disruptions in ecosystem nutrient cycling (Loo 2008) and the loss of American chestnut-associated insect species

(Lynne Rieske-Kinney, pers. comm. 2014). Similarly, exotic insects such as the exotic emerald ash borer, asian longhorned beetle, and hemlock wooly adelgid have begun eliminating hardwood and conifer species from north and eastern North America, all likely to bring ecosystem damage that has not yet been quantified. Except for habitat destruction, exotic insects and fungi form the largest immediate threat to biodiversity on earth (Simberloff et al. 2005).

Unintentional introductions of insects and fungi have been increasingly frequent following the course of international trade and travel (Wingfield et al. 2010). New arrivals of exotic pests are unlikely to decrease since trade and travel are expected to increase, resources limit more stringent trade regulations, and policy compliance is difficult to control. Slowing the rate of introduction is difficult due to cryptic and pervasive qualities of insects and fungi, and the ability of many species to reproduce asexually. Forest

44

insects and pathogens are most commonly carried to new regions on nursery stock, wood packing materials, produce, and seed (Haack 2001). Wood boring insects are especially difficult to control due to their protection from external conditions, such as pesticide treatments, by traveling inside wood. Wood boring bark and ambrosia beetles

(Coleoptera: Scolytinae) are among the most commonly intercepted exotic forest pests

(Haack 2006, Aukema et al. 2010). Although most of these introductions are harmless, a few exotic bark and ambrosia beetles or their associated fungi have been disproportionately detrimental forest pests.

Alliances between bark and ambrosia beetles (Curculionidae: Scolytinae,

Platypodinae) and their associated fungi (members of Ophiostomatales, Microascales, and Hypocreales) are currently driving tree mortality in ecologically and economically important tree species across the world. In some cases, trees have undergone massive selection events and regional extirpations as a result of the introduced fungi and beetles

(Kendra et al. 2013, Ploetz el al. 2013). Until recently, plant pathogenic fungi identified as associates of bark beetles have been found to be antagonistic to the survival of the beetle (Six and Wingfield 2011). New forest diseases are increasingly caused by non- native fungi that appear nutritionally beneficial to the ambrosia beetles. These disease- causing fungal mutualists have only been shown to act as pathogens in non-native environments. For example, fungi associated with Xyleborus glabratus (the redbay ambrosia beetle), Pityophthorus juglandis, and Platypus quercivorus are all thought to be non-native in the region where they cause harm (Ploetz et al. 2013). Beetles may also display novel characteristics in non-native regions, such as attacking live trees rather than dead ones (Hulcr & Dunn 2011). The lack of preliminary data indicating how

45

a species will behave in a non-native region makes it almost impossible for government agencies to respond in time to prevent establishment of a pest species.

The “Wait-and-see” Approach

A number of government programs are designed to reduce and prevent damages caused by invasive alien species. The Early Detection and Rapid Response program

(EDRR, U.S. Forest Service) and Cooperative Agricultural Pest Survey (CAPS, Florida

Department of Agriculture and Consumer Services) continually monitor for new and existing invasive species, most heavily in ports of entry where exotic organisms are most likely to become introduced (Rabaglia et al. 2008). These programs rely on previous information on the potential for species to become invasive, such as those generated by the USDA APHIS Offshore Pest Information System (OPIS) or through ecological niche modeling studies. These programs have been successful at preventing ecological and economic destruction from invasive pests, but have been unable to prevent new invasive species introductions when too little was known about the organism before its arrival. The USDA-OPIS is often unable to predict which exotic beetles and fungi will be destructive because they rely upon existing information on the organisms in their native territory, which is usually either unavailable or not relevant for beetles/fungi which act differently in non-native regions. It is not fiscally feasible to prevent every beetle introduction because bark and ambrosia beetle introductions are so common and because most of these introductions appear to be harmless. The beetles and fungi that do cause harm to trees are only discovered to cause disease after they are already established. This “wait-and-see” approach was employed for a number of damaging introduced species, including the redbay ambrosia beetle (RAB).

The RAB was detected by the EDRR program more than a year before it was

46

associated with mass-mortality of redbays (Fraedrich et al. 2008). The period of time in between introduction and disease occurrence provides a newly introduced species to establish a population and disperse to new areas—effectively making it impossible for eradication. This is especially true for some groups of bark and ambrosia beetles, which can establish a population with a single individual. Thus, the only opportunity to prevent the establishment of invasive beetles and fungi is through prior knowledge of which species are able to cause significant harm.

Preventing the Next Tree Epidemic

There have been many approaches used to assess invasion or pathogen ability in non-native species, but such strategies rely on previous knowledge of exotic species.

Some studies have focused on generating new data for predictions where prior information was lacking on a potential new pest species. One example is the use of plant “sentinel” individuals, which are typically established in botanical gardens outside their native range and monitored for novel interactions between the host and pests/pathogens (Britton et al. 2010). These transplant studies can provide valuable new information about the ability for non-native species to cause damage to native plants but require extensive long-term sampling and are limited in scope. For example, diseases can be difficult to diagnose and characterize when a whole slew of new organisms are interacting with the non-native host, especially when novel abiotic variables such as climate may cause additional stress to the sentinel species. Host- pathogen interactions may be further shadowed by the presence of predators in the non-native habitat, or absence of native predators. A more systematic approach is necessary to test direct effects of pests on their hosts as a counterpart to these sentinel studies.

47

A new approach, proposed in this study, is to evaluate interactions between native hosts and exotic pests/pathogens in a controlled, quarantined environment before the exotic species is introduced. These pre-invasion diagnostic studies may provide exclusive information on the ability for a single exotic species to cause harm to a native species, which is exactly what regulatory agencies need to make decisions for eradication and prevention of newly arriving exotic species. The present study was established as a proof-of-concept approach for evaluating the potential of fungi associated with bark and ambrosia beetles in Asia to act as pathogens to American trees. Three beetles (Xyleborus pinicola, Tomicus minor, and Ips chinensis) were selected as potentially invasive based on their occurrence in climates and habitats similar to those in the southeastern United States, and their specificity to pines, which are the most ecologically and economically important tree species in the Southeast

(Beeson 1930). Xyleborus pinicola was of particular interest due its ability to establish a population with a single individual, and its unusual host specificity when compared to close beetle relatives. All three beetle species were also chosen because of their close associations to fungi, indicating a high likelihood that they have specific fungal associates that may be transported with the beetle during an introduction to a new region.

Specifically, this study was initiated to test the following questions which pre- formulated hypotheses could not be made:

1. Is evaluating fungi associated with non-native bark and ambrosia beetles a feasible approach for assessing pathogen potential before their introduction into a non-native region?

2. Are Xyleborus pinicola, Tomicus minor, and Ips chinensis carrying fungi that are pathogenic to trees that are native to the southeastern United States?

48

3. Are there life history traits of beetle vectors that allow prediction of whether the beetles are carrying fungal tree pathogens?

Methods

Study Sites and Beetle Collection

Fungi used in pathogenicity tests were isolated from beetles collected in China and Thailand during May-July of 2013. Tomicus minor was collected from several dead

Pinus yunnanensis bolesin Kunming, Yunnan, China (25.068895, 102.762013). Ips chinensis was collected from a dying P. armandii stand in Chuxiong, Yunnan, China.

Xyleborus pinicola and I. chinensis were collected from the bole of a dying P. kesiya in

Hot, Chiang Mai, Thailand (18.144611, 98.319944). Adult beetles were identified using

Wood’s key to genera (1986), Kirkendall’s key to Tomicus (2008), and Yin (1984).

Voucher specimens of Tomicus spp. were also confirmed by Massimo Faccoli (personal comm. 2013) who originally aided in delimited Tomicus spp. in Yunnan, China

(Kirkendall 2008). Specimens of I. chinensis were confirmed by Anthony Cognato

(personal comm. 2013) who is currently working on the systematics of this group.

Beetles were stored at 10-15° C for up to 48 hours before fungal isolation.

Fungal Isolation

Fungi were isolated from beetle galleries, and adult, pupal, and larval life stages depending on availability. Beetle galleries were sampled by surface disinfesting wood with ethanol and scraping of gallery walls into phosphate buffer saline (PBS) prior to serial dilution.

Beetle dissections took place either in a laboratory biosafety hood (Chinese specimens) or inside an ethanol-sterilized plastic bin with holes cut in one side to function as a field-isolation chamber (Thai specimens). All life stages of beetles were

49

surface washed in a 1 mL sterile solution of 1% Tween 80 (Sigma Chemical Co, St.

Louis, MO, USA) and PBS, which was then was serially diluted prior to plating. Larvae and pupae of all species were individually macerated in PBS prior to serial dilution.

Adult T. minor and I. chinensis specimens were sectioned into three parts prior to maceration and dilution: the head, thorax, and abdomen. Heads of adult X. pinicola foundress females aseptically removed prior to maceration and dilution. Only heads were included for this species because beetles in this genus contain mycangia within the mandibles.

Serial dilutions of 1/10, 1/100, and 1/1000 times were made from each sample

(location of beetle body). Dilutions were plated on Potato Dextrose Agar amended with

1.4% additional agar and 2% streptomycin. Plates were stored in an incubator in the dark at 25° C for up to two weeks.

Morphotypes were designated based on macromorphology for all fungal colony forming units (CFUs) found in more than one sample. Plates were monitored every 2-3 days during the two-week incubation period and photographed to ensure consistent assignment of morphotypes. Morphotype designations were confirmed by retrospective comparisons of pure cultures and by sequencing portions of the nuclear ribosomal DNA

(rDNA) and four protein-encoding genes.

Molecular Identification and Phylogenetic Analysis

One- to two-week old pure cultures were used for DNA extraction. 10-20 µl of mycelium was added to 20 uL of Sigma-Aldrich Extraction Solution (Extract-N-Amp, St.

Louis, MO) in 0.25 mL PCR tubes, which were then incubated in a thermocycler

(Eppendorf Mastercycler) at 96° C for 10 minutes. After incubation, 20 µL of 3% BSA

(Fermentas) was added to each tube, vortexed for 5 sec and then centrifuged at 2000

50

rpms for 30 seconds. The upper half of this solution was used as genomic DNA template for PCR.

Morphotype designations were evaluated using both RAPD-PCR and Sanger sequencing. Each RAPD PCR contained a final volume of 25 µL: 1 µL of template DNA,

2 µL of M13 primer (10 µM), 0.125 µl of Taq (Takara®), 2.5 µL PCR buffer (15 mM

MgCl2), 2 µL dNTP mix (2.5 mM each dNTP), and 20 µL sterile water. Amplification was performed in an Eppendorf Mastercycler Pro following parameters in (Arnau et al.

1994). RAPD-PCRs were run on a 1% agarose gel for 90 minutes at 110V. Each well contained 2 µL of SyBr Green dye and either 9 µL template or 10 µL ladder.

Basic PCR reactions used in sequencing consisted of a final volume of 25 µL: 1

µL of template DNA, 1 µL of forward and reverse primer (10 µM), 0.125 µl of Taq

(Takara®), 2.5 µL PCR buffer (15 mM MgCl2), 2 µL dNTP mix (2.5 mM each dNTP), and 20 µL sterile water. PCR and sequencing primers are listed in Table 1-1. PCR products were purified using Exosap-IT, following the manufacturer’s protocols (USB

Corp., Cleveland, OH). PCR reactions were performed in an Eppendorf Mastercycler

Pro following published cycling parameters (Kim et al. 2004).

DNA sequencing was performed on an ABI PRISM 377XL or 3130XL automated

DNA sequencer by the Interdisciplinary Center for Biotechnology Research, University of Florida, Gainesville. Forward and reverse sequences were assembled into contigs in

Geneious (Version 7.0). Partial sequences were obtained from up to three nuclear ribosomal RNA (rDNA) encoding regions and portions of four protein-encoding genes, from up to five isolates of each morphotype.

51

BLAST queries of the NCBI GenBank database were performed with each sequence to identify the fungi that were cultured. Final taxonomic identities were assigned by phylogenetic comparisons to congeneric sequences available in Genbank.

Maximum likelihood phylogenetic analyses were performed using PhyML (Guindon et al. 2010) with 1,000 bootstrap replicates.

Pathogenicity Test

The most common and abundant fungi (Ophiostoma sp., Ophiostoma ips,

Raffaelea sp., and Geosmithia sp.) isolated from each beetle species surveyed was used to create spore suspensions for inoculation into pine seedlings. Spore suspensions were made from sterile water and 10-14 day-old fungal cultures plated on

Potato Dextrose Agar (Difco). Spore concentrations reflected the highest total CFU count isolated from a beetle (Table 3-1). The concentration of spore suspensions were calculated using a hemocytometer.

Three pine genotypes were tested for pathogenicity, including two Pinus taeda families, and one P. elliottii family. Pine families were provided by Weyerhaeuser. Pines were stored in a quarantined greenhouse facility at the Division of Plant Industry (DPI),

Department of Agriculture and Consumer Services in Gainesville, Florida. The trees were maintained under natural light conditions, watered as necessary, and kept under a night-day temperature regime averaging at 27° C. Treatments were arranged in a completely randomized design with eight trees per fungus treatment and pine genotype combination. Each seedling stem was at least 40 cm from neighboring seedlings

A single hole was drilled at a downward angle (approx. 45 degrees) into the xylem of each seedling using a 1.98 mm (5/64 inch) drill bit. Holes were made within the basal 30 cm of the stem and were up to 4 cm deep. The diameter of the stem at the

52

inoculation site was an average of 1.4 cm. Spore suspensions were pipetted into the xylem in 50 uL aliquots. Wound sites were wrapped in parafilm immediately following inoculation.

Seedlings were monitored weekly for signs of mortality and disease development

(including foliar discoloration, wilting, resinosis, and mortality). After 10 weeks, seedlings were destructively sampled. Bark was carefully scrapped away using a knife to allow for measurement of phloem necrosis or cankers. A razor was used to cut vertical sections of the inner stem to allow for measurement of stain in the xylem (Figure

3-1). Stem sections were stored for up to two days at 10° C prior to attempted re- isolation of inoculated fungi. Stained and un-stained portions of the xylem above and below the inoculation point were surface-disinfested and plated on PDA (Difco) incubated at 25° C.

Differences in disease expression (resinosis, lesion size) was tested using a one- tailed t-test, analysis of variance (ANOVA) and Fisher’s (LSD) test.

Results

Fungal Isolations and Identification

Genotyping isolates by RAPD-PCR and Sanger sequencing narrowed fungal isolates to five morphotypes that were isolated from at least 75% of respective beetles sampled (Table 3-1). BLAST queries and phylogenetic analyses identified the five isolates as Ophiostoma sp., O. ips, Raffaelea sp., and Geosmithia sp. (Dreaden et al. in review). Isolates of O. ips from both China and Thailand were assayed in this study based on a 1% divergence in B-tubulin gene sequences between the two cultures.

53

Pathogenicity Test

Across all pine genotypes before destructive sampling, resin production at the inoculation site was the only symptom of disease development that was significantly different compared to the control for some treatments. Only Ophiostoma ips and O. sp. displayed significantly longer periods of resinosis compared to the control (Tables 3-2 to

3-4). Destructive sampling after 10 weeks showed that lesions in the xylem and phloem of O. ips and O. sp. were significantly longer compared to controls across all pine families, except in loblolly pine family 2, which displayed no difference in xylem lesion length when inoculated with Ophiostoma sp. (Tables 3-5 to 3-10). There were no differences observed in pathogenicity between O. ips genotypes from China and

Thailand when compared to controls. Re-isolation of all Ophiostoma isolates from the xylem of Pinus taeda were successful.

Raffaelea sp. and Geosmithia sp. displayed no significant differences in symptoms during the period allowed for disease development (Tables 3-2 to 3-4).

Phloem and xylem lesion lengths in Raffaelea and Geosmithia treatments were not found to be different to controls by Fisher’s LSD test (Tables 3-5 to 3-10).

No mortality occurred during the experiment. No significant wilting or foliar discoloration was observed during the experiment in any treatment. Except for resin production and small cankers at wound sites, all seedlings appeared healthy and formed calluses at the end of the experiment.

Discussion

This study validates the feasibility of rapidly assessing pathogen potential in cryptic, symbiotic organisms that are from non-native regions and not yet established outside their native range. The approach implemented here demonstrated that two

54

beetle-associated Ophiostoma spp. isolated from Ips chinensis and one from Xyleborus pinicola were able to cause disease symptoms in Pinus taeda or P. elliotti, while two other beetle-associated fungi from Asia did not cause noticeable symptoms of disease.

Results presented here suggest that none of the fungi tested are able to act as deadly vascular wilt pathogens, and thus they are unlikely to cause a significant number of tree deaths in North American forests.

The pathogenicity of Ophiostoma ips inferred from resin production and xylem lesions in the present study is consistent with previous studies that have shown that O. ips is able to act as a pathogen on American trees, and has already been introduced into North America (Kim et al. 2003, Lim et al. 2006). Ophiostoma ips has also been documented in association with many species of bark beetle species in Europe, Asia, and North America. Greater taxon sampling and experimental evidence is needed to establish how diverse O. ips is across large geographic scales, how it maintains association to beetles, and whether it is transported as a nutritional symbiont or plays some other role.

The Ophiostoma sp. was observed to be less pathogenic compared to O. ips based on less significant differences resin production and length of lesions when compared using Fisher’s LSD test. Ophiostoma sp. was also found to be less prevalent than O. ips on Ips chinensis suggesting it may be less strictly associated or play some other role in the symbiosis. Unless the Ophiostoma sp. is able to cause further damage to trees through mass-attack of beetles, it is unlikely that the introduction of this fungus species would cause any significant harm to Pinus spp. in the southeastern United

States. In cases where invasive bark beetles have caused widespread damage to trees,

55

it has been typically the action of the beetle mass attack or a highly virulent fungus in causing tree death, rather than weakly-pathogenic associated fungi (Six and Wingfield

2011). Nonetheless, future studies should include treatments with multiple inoculation sites to test the response of trees to mass attacks of beetle-introduced fungi, and may benefit from the assay of more tree host species.

The two fungi that did not induce significant symptoms of disease, Raffaelea sp., and Geosmithia sp., are probably harmless to living trees. The vast majority of species within these genera are harmless to trees, and the isolates included in this study are not closely related to disease causing species (Kostovcik et al. in prep, Dreaden et al. in prep). One caveat to this finding is that symptoms may be differently expressed in older trees, or take longer than 10 weeks to be expressed. However, all other vascular wilt pathogens capable of killing large numbers of trees are able kill even seedlings and are able to do so in under 12 weeks (Fraedrich et al. 2008, Townsend et al. 1995, Yadeta and Thomma 2013). For this reason, Raffaelea sp. and Geosmithia sp. are still unlikely to be mass tree-killing vascular wilt pathogens. In addition, these fungi were not recovered from stained sapwood above and below inoculation sites. This indicates they may have difficulty growing inside living, healthy xylem whereas vascular wilt pathogens would be expected to readily spread via healthy xylem.

The low to absent pathogenicity observed in the present study suggests that none of the fungi found associated with Tomicus minor, Xyleborus pinicola, or Ips chinensis warrants a high threat rating. However, we need to document the fungal communities of these same beetles across wider areas in Asia in order to determine whether other potentially dangerous fungi might be present in other parts of the range of

56

these beetles, and how specific the associations seen in this study hold true across geographically disparate locations. Combining ecological understanding from experimental and exploratory research of invasive species will become increasingly important as new pests and pathogens become established around the world. Trees provide profound structuring to native ecosystems and have specific coevolved assembles which become disrupted following the introduction of an invasive pathogen

(Thomson 2005). Preventative approaches to management of invasive species are ideal for saving valuable economic and ecological resources such as trees used in forestry.

The current reactive approach is too permissive to pests and pathogens that have not been characterized. The pre-invasion diagnostic presented in this study offers government regulatory agencies the opportunity to increase the focus of their prevention and eradication efforts by generating necessary information on those species that are likely to become invasive.

57

Figure 3-1. Symptoms of Ophiostoma ips on Pinus taeda seedling stems during disease development (left) and after destructive sampling (lateral bisection, right).

Table 3-1. Prevalence and abundance of fungi associated with Asian bark beetles sampled. Prevalence Prevalence in Spore amount used in Beetle species Fungal isolate on beetle (%) gallery (%) inoculation experiment Ips chinensis Ophiostoma sp. 75 75 2,000 Ophiostoma ips 100 86 1,900 Tomicus minor Geosmithia sp. 100 72 1,220 Xyleborus Ophiostoma ips 60 80 2,500 pinicola Raffaelea sp. 100 80 3,500

Table 3-2. Mean number of weeks seedlings displayed resin production in loblolly pine genotype 1 inoculated with Asian fungi. Treatment Mean sd p-value a Control 1.5 1.069 Ophiostoma sp. 1.75 2.0529 0.383 Ophiostoma ips 3.12 1.1259 0.005 Ophiostoma ips 2.75 1.983 0.0727 Raffaelea sp. 1.75 1.488 0.353 Geosmithia sp. 1.25 1.1649 0.6692 a corresponds to Welch’s one-tailed t-test

58

Table 3-3. Mean number of weeks seedlings displayed resin production in loblolly pine genotype 2 inoculated with Asian fungi. Treatment Mean sd p-value a Control 0.88 0.641 Ophiostoma sp. 2.5 1.6903 0.0158 Ophiostoma ips 4.5 2.0702 3.27E-05 Ophiostoma ips 3.75 1.9226 0.0037 Raffaelea sp. 1.37 0.744 0.0861 Geosmithia sp. 1.75 1.3887 0.06858 a corresponds to Welch’s one-tailed t-test

Table 3-4. Mean number of weeks seedlings displayed resin production in slash pine inoculated with Asian fungi. Treatment Mean sd p-value a Control 2.5 2.7255 Ophiostoma sp. 3.5 1.9272 0.2063 Ophiostoma ips 5.5 2.3905 0.0743 Ophiostoma ips 5 2.8284 0.0467 Raffaelea sp. 2.5 2.0701 0.5 Geosmithia sp. 4 2.2039 0.1235 a corresponds to Welch’s one-tailed t-test

Table 3-5. Mean vertical xylem lesion length in loblolly pine genotype 1 inoculated with Asian fungi. Treatment Mean sd p-value a group Control 0.83125 0.3788 a Ophiostoma sp. 1.775 0.6041 0.0014 b Ophiostoma ips 3.513 0.6443 2.52E-07 c Ophiostoma ips 3.7875 0.9046 5.09E-06 c Raffaelea sp. 1.371 0.1704 0.0023 ab Geosmithia sp. 1.1375 0.4983 0.1181 a a corresponds to Welch’s one-tailed t-test b corresponds to Fisher’s LSD ANOVA

Table 3-6. Mean vertical xylem lesion length in loblolly pine genotype 2 inoculated with Asian fungi. Treatment Mean sd p-value a group b Control 0.6571 0.1718 a Ophiostoma sp. 2.2375 0.7782 0.0002 b Ophiostoma ips 4.075 1.416 2.67E-05 c Ophiostoma ips 4.7 1.2375 1.10E-04 c Raffaelea sp. 1.3313 0.4605 0.0019 a Geosmithia sp. 1.1375 0.3889 0.01004 a a corresponds to Welch’s one-tailed t-test b corresponds to Fisher’s LSD ANOVA

59

Table 3-7. Mean vertical xylem lesion length in slash pine inoculated with Asian fungi. Treatment Mean sd p-value a group b Control 0.975 0.3196 a Ophiostoma sp. 1.9285 0.4535 0.00039 a Ophiostoma ips 3.95 2.1105 0.00148 b Ophiostoma ips 3.9375 0.8433 0.00255 b Raffaelea sp. 1.2 0.239 0.6744 a Geosmithia sp. 1.1375 0.3159 0.3238 a a corresponds to Welch’s one-tailed t-test b corresponds to Fisher’s LSD ANOVA

Table 3-8. Mean vertical phloem lesion length in loblolly pine genotype 1 inoculated with Asian fungi. Treatment Mean sd p-value a group b Control 0.4875 0.3613 a Ophiostoma sp. 1.1375 0.30207 0.0008 b Ophiostoma ips 1.9375 0.5926 4.15E-05 c Ophiostoma ips 2.4857 0.7151 5.54E-05 d Raffaelea sp. 0.7714 0.2563 0.0504 ab Geosmithia sp. 0.6625 0.4274 0.1959 ab a corresponds to Welch’s one-tailed t-test b corresponds to Fisher’s LSD ANOVA

Table 3-9. Mean vertical phloem lesion length in loblolly pine genotype 2 inoculated with Asian fungi. Treatment Mean sd p-value a group b Control 0.5687 0.3673 a Ophiostoma sp. 1.0428 0.2878 0.007 a Ophiostoma ips 1.6125 0.6556 0.0012 b Ophiostoma ips 2 0.7071 0.0002 b Raffaelea sp. 0.7313 0.0704 0.1281 a Geosmithia sp. 0.575 0.3694 0.4867 A a corresponds to Welch’s one-tailed t-test b corresponds to Fisher’s LSD ANOVA

Table 3-10. Mean vertical phloem lesion length in slash pine inoculated with Asian fungi. Treatment Mean sd p-value a group b Control 0.5714 0.3352 a Ophiostoma sp. 0.8625 0.0876 0.0312 ab Ophiostoma ips 1.55 0.4504 2.00E-04 b Ophiostoma ips 1.6429 0.48599 3.00E-04 b Raffaelea sp. 0.7875 0.2531 0.0954 ab Geosmithia sp. 0.625 0.324 0.3795 a a corresponds to Welch’s one-tailed t-test b corresponds to Fisher’s LSD ANOVA

60

LIST OF REFERENCES

Alamouti, M., S., Tsui, C. K., & Breuil, C. (2009). Multigene phylogeny of filamentous ambrosia fungi associated with ambrosia and bark beetles.Mycological research, 113(8), 822-835.

Arnau, J., Housego, A. P., & Oliver, R. P. (1994). The use of RAPD markers in the genetic analysis of the plant pathogenic fungus Cladosporium fulvum.Current genetics, 25(5), 438-444.

Atkinson, T. H., Rabaglia, R. J., & Bright, D. E. (1990). Newly detected exotic species of Xyleborus (Coleoptera: Scolytidae) with a revised key to species in eastern North America. The Canadian Entomologist, 122(01), 93-104.

Aukema, J. E., McCullough, D. G., Von Holle, B., Liebhold, A. M., Britton, K., & Frankel, S. J. (2010). Historical accumulation of nonindigenous forest pests in the continental United States. Bioscience, 60(11), 886-897.

Baker, J. M., & Norris, D. M. (1968). A complex of fungi mutualistically involved in the nutrition of the ambrosia beetle Xyleborus ferrugineus. Journal of Invertebrate Pathology, 11(2), 246-250.

Banno, I., Mikata, K., & Kodama, K. (1983). Ascomycetous yeasts isolated from galleries of the ambrosia beetles in Japan. Trans. Mycol. Soc. Jpn, 24, 441-450.

Batra, L. R. (1967). Ambrosia fungi: a taxonomic revision, and nutritional studies of some species. Mycologia, 59, 976-1017.

Beeson, C. F. C. (1930). The biology of the genus Xyleborus, with more new species. Indian Forest Records, 14(pt. 10), 209-272.

Bhat, S. S., & Sreedharan, K. (1988). Association of Ambrosiella xylebori Brader, with the shot-hole borer Xylosandrus compactus Eichhoff, a pest of robusta coffee. Journal of Coffee Research, 18(1), 54-57.

Brader, L. (1964). Étude de la relation entre le scolyte des rameaux du caféier Xyleborus compactus Eichh.(X. morstatti Hag), et sa plante-hôte. H. Veenman & Zonen NV.

Bright, D. E., & Rabaglia, R. J. (1999). Dryoxylon, a new genus for Xyleborus onoharaensis Murayama, recently established in the southeastern United States (Coleoptera: Scolytidae). The Coleopterists' Bulletin, 333-337.

Britton, K. O., White, P., Kramer, A., & Hudler, G. (2010). A new approach to stopping the spread of invasive insects and pathogens: early detection and rapid response via a global network of sentinel plantings. New Zealand Journal of Forestry Science, 40.

61

Carrillo, D., Duncan, R. E., Ploetz, J. N., Campbell, A. F., Ploetz, R. C., & Peña, J. E. (2014). Lateral transfer of a phytopathogenic symbiont among native and exotic ambrosia beetles. Plant Pathology, 63(1), 54-62.

Cognato, A. I., Hulcr, J., Dole, S. A., & Jordal, B. H. (2011). Phylogeny of haplo–diploid, fungus‐growing ambrosia beetles (Curculionidae: Scolytinae: Xyleborini) inferred from molecular and morphological data. Zoologica Scripta,40(2), 174-186.

Daehler, C. C., & Dudley, N. I. C. K. (2002). Impact of the black twig borer, an introduced insect pest, on Acacia koa in the Hawaiian Islands. MICRONESICA- AGANA-, 35, 35-53.

Darriba, D., Taboada, G. L., Doallo, R., & Posada, D. (2012). jModelTest 2: more models, new heuristics and parallel computing. Nature Methods, 9(8), 772-772.

Dole, S. A., & Cognato, A. I. (2010). Phylogenetic revision of Xylosandrus Reitter (Coleoptera: Curculionidae: Scolytinae: Xyleborina). Proceedings of the California Academy of Sciences, 61(7), 451.

Faccoli, M., Frigimelica, G., Mori, N. I. C. O. L. A., Petrucco Toffolo, E., Vettorazzo, M., & Simonato, M. (2009). First record of Ambrosiodmus (Hopkins, 1915) (Coleoptera: Curculionidae, Scolytinae) in Europe. Zootaxa,2303, 57-60.

Francke-Grosmann, H. (1952). Über die Ambrosiazucht der beiden Kiefernborkenkäfer Myelophilus minor Htg. und Ips acuminatus Gyll.

Francke-Grosmann, H. (1967). Ectosymbiosis in wood-inhabiting insects.symbiosis, 2, 141-205.

Gandhi, K. J., & Herms, D. A. (2010). Direct and indirect effects of alien insect herbivores on ecological processes and interactions in forests of eastern North America. Biological Invasions, 12(2), 389-405.

Gandhi, K. J., Cognato, A. I., Lightle, D. M., Mosley, B. J., Nielsen, D. G., & Herms, D. A. (2010). Species composition, seasonal activity, and semiochemical response of native and exotic bark and ambrosia beetles (Coleoptera: Curculionidae: Scolytinae) in northeastern Ohio. Journal of economic entomology, 103(4), 1187- 1195.

Gebhardt, H., Weiss, M., & Oberwinkler, F. (2005). Dryadomyces amasae: a nutritional fungus associated with ambrosia beetles of the genus Amasa (Coleoptera: Curculionidae, Scolytinae). Mycological research,109(6), 687-696.

Glass, N. L., & Donaldson, G. C. (1995). Development of primer sets designed for use with the PCR to amplify conserved genes from filamentous ascomycetes. Applied and Environmental Microbiology, 61(4), 1323-1330.

62

Guindon, S., Dufayard, J. F., Lefort, V., Anisimova, M., Hordijk, W., & Gascuel, O. (2010). New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Systematic biology,59(3), 307-321.

Guindon, S., & Gascuel, O. (2003). A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Systematic biology, 52(5), 696-704.

Haack, R. A. (2001). Intercepted Scolytidae (Coleoptera) at US ports of entry: 1985– 2000. Integrated Pest Management Reviews, 6(3-4), 253-282.

Hara, A. H., & Beardsley Jr, J. W. (1976). The biology of the black twig borer, Xylosandrus compactus (Eichhoff), in Hawaii. Proceedings of the Hawaiian Entomological Society, 23(1), 55-70.

Harrington, T. C., Aghayeva, D. N., & Fraedrich, S. W. (2010). New combinations in Raffaelea, Ambrosiella, and Hyalorhinocladiella, and four new species from the redbay ambrosia beetle, Xyleborus glabratus. Mycotaxon,111(1), 337-361.

Harrington, T. C., M. R. Kazmi, A. M. Al-Sadi, and S. I. Ismail. 2014. Intraspecific and intragenomic variability of ITS rDNA sequences reveals taxonomic problems in Ceratocystis fimbriata sensu stricto. Mycologia 106:224-242. doi:10.3852/13- 189

Hill, D. S. (1987). Agricultural insect pests of temperate regions and their control. CUP Archive.

Hofstetter, V., Miadlikowska, J., Kauff, F., & Lutzoni, F. (2007). Phylogenetic comparison of protein-coding versus ribosomal RNA-coding sequence data: a case study of the (). Molecular phylogenetics and evolution, 44(1), 412-426.

Hulcr, J., & Cognato, A. I. (2010). Repeated evolution of crop theft in fungus‐farming ambrosia beetles. Evolution, 64(11), 3205-3212.

Hulcr, J., & Dunn, R. R. (2011). The sudden emergence of pathogenicity in insect– fungus symbioses threatens naive forest ecosystems. Proceedings of the Royal Society B: Biological Sciences, 278(1720), 2866-2873.

Kaneko, T., Tamaki, Y., & Takagi, K. (1965). Preliminary report on the biology of some Scolytid beetles, the tea root borer, Xyleborus germanus Blanford, attacking tea roots, and the tea stem borer, Xyleborus compactus Eichhoff, attacking tea twigs. Japanese Journal of Applied Entomology and Zoology, 9(1), 23-28.

63

Kasson, M. T., O’Donnell, K., Rooney, A. P., Sink, S., Ploetz, R. C., Ploetz, J. N., ... & Geiser, D. M. (2013). An inordinate fondness for Fusarium: Phylogenetic diversity of fusaria cultivated by ambrosia beetles in the genus Euwallacea on avocado and other plant hosts. Fungal Genetics and Biology, 56, 147-157.

Kendra, P. E., Montgomery, W. S., Niogret, J., & Epsky, N. D. (2013). An uncertain future for American Lauraceae: A lethal threat from redbay ambrosia beetle and laurel wilt disease (A review). American Journal of Plant Sciences,4, 727.

Kim, J. J., Kim, S. H., Lee, S., & Breuil, C. (2003). Distinguishing Ophiostoma ips and Ophiostoma montium, two bark beetle‐associated sapstain fungi. FEMS Microbiology Letters, 222(2), 187-192.

Kim, J. J., Lim, Y. W., Wingfield, M. J., Breuil, C., & Kim, G. H. (2004). Leptographium bistatum sp. nov., a new species with a Sporothrix synanamorph from Pinus radiata in Korea. Mycological research, 108(6), 699-706.

Kinuura, H. (1995). Symbiotic fungi associated with ambrosia beetles. Japan Agricultural Research Quarterly, 29, 57-57.

Kirkendall, L. R., Faccoli, M., Ye, H. (2008). Description of the Yunnan shoot borer, Tomicus yunnanensis Kirkendall & Faccoli sp. n.(Curculionidae, Scolytinae), an unusually aggressive pine shoot beetle from southern China, with a key to the species of Tomicus. Zootaxa, 1819, 25-39.

Kolařík, M., & Hulcr, J. (2009). Mycobiota associated with the ambrosia beetle Scolytodes unipunctatus( Coleoptera: Curculionidae, Scolytinae). Mycological research, 113(1), 44-60.

Kühnholz, S., Borden, J. H., & Uzunovic, A. (2001). Secondary ambrosia beetles in apparently healthy trees: adaptations, potential causes and suggested research. Integrated Pest Management Reviews, 6(3-4), 209-219.

Kuo, H. C. (2010). Plant-fungus-beetle interactions: Case studies in Hawaiian endemic Xyleborus species and the black twig borer.

Lepš, J., & Šmilauer, P. (2003). Multivariate analysis of ecological data using CANOCO. Cambridge university press.

Lim, Y. W., Alamouti, S. M., Kim, J. J., Lee, S., & Breuil, C. (2004). Multigene phylogenies of Ophiostoma clavigerum and closely related species from bark beetle‐attacked Pinus in North America. FEMS microbiology letters, 237(1), 89- 96.

Liu, Y. J., & Hall, B. D. (2004). Body plan evolution of ascomycetes, as inferred from an RNA polymerase II phylogeny. Proceedings of the National Academy of Sciences of the United States of America, 101(13), 4507-4512.

64

Masuya, H. (2007). Note on the dieback of Cornus florida caused by Xylosandrus compactus. Bulletin of the Forestry and Forest Products Research Institute, 6.

McClanahan HS. (1951). Grove Inspection Department. Florida State Plant Board, Biennial Report 18: 41-42.

McPherson, B. A., Wood, D. L., Erbilgin, N., & Bonello, P. (2010, April). Ambrosia Beetles and Their Associated Fungi Appear to Accelerate Mortality in Phytophthora ramorum-Infected Coast Live Oaks. In Sudden Oak Death Fourth Science Symposium (p. 347).

McPherson, B. A., Erbilgin, N., Bonello, P., & Wood, D. L. (2013). Fungal species assemblages associated with Phytophthora ramorum-infected coast live oaks following bark and ambrosia beetle colonization in northern California. Forest Ecology and Management, 291, 30-42.

Morales-Ramos, J. A., Rojas, M. G., Sittertz-Bhatkar, H., & Saldaña, G. (2000). Symbiotic relationship between Hypothenemus hampei (Coleoptera: Scolytidae) and Fusarium solani (Moniliales: Tuberculariaceae). Annals of the Entomological Society of America, 93(3), 541-547.

Mueller, U. G., Gerardo, N. M., Aanen, D. K., Six, D. L., & Schultz, T. R. (2005). The evolution of agriculture in insects. Annual Review of Ecology, Evolution, and Systematics, 563-595.

Muthappa, B. N., & Venkatasubbaiah, P. (1981). Association of Ambrosiella macrospora with Xylosandrus compactus, the shot-hole borer of robusta coffee in India. Journal of coffee research, 11(2).

Ngoan, N. D., Wilkinson, R. C., Short, D. E., Moses, C. S., & Mangold, J. R. (1976). Biology of an introduced ambrosia beetle, Xylosandrus compactus, in Florida. Annals of the Entomological Society of America, 69(5), 872-876.

Nishida, G. M., & Evenhuis, N. L. (2000). Arthropod pests of conservation significance in the Pacific: A preliminary assessment of selected groups.Invasive species in the Pacific: A technical review and draft regional strategy, 115.

O'Donnell, K. (1992). Ribosomal DNA internal transcribed spacers are highly divergent in the phytopathogenic ascomycete Fusarium sambucinum (Gibberella pulicaris). Current genetics, 22(3), 213-220.

O'Donnell, K., & Cigelnik, E. (1997). Two Divergent Intragenomic rDNA ITS2 Types within a Monophyletic Lineage of the Fungus Fusarium Are Nonorthologous. Molecular phylogenetics and evolution, 7(1), 103-116.

O’Donnell, K., Kistler, H. C., Cigelnik, E., & Ploetz, R. C. (1998). Multiple evolutionary origins of the fungus causing Panama disease of banana: concordant evidence

65

from nuclear and mitochondrial gene genealogies.Proceedings of the National Academy of Sciences, 95(5), 2044-2049.

O'Donnell, K. (2000). Molecular phylogeny of the Nectria haematococca-Fusarium solani species complex. Mycologia, 919-938.

O'Donnell, K., Sutton, D. A., Rinaldi, M. G., Sarver, B. A., Balajee, S. A., Schroers, H. J., R. C. Summerbell, Robert, V. A.,Crous, P.W., Zhang, N., & Geiser, D. M. (2010). Internet-accessible DNA sequence database for identifying fusaria from human and animal infections. Journal of clinical microbiology, 48(10), 3708-3718.

O'Donnell, K., Sutton, D. A., Fothergill, A., McCarthy, D., Rinaldi, M. G., Brandt, M. E., Zhang, H., & Geiser, D. M. (2008). Molecular phylogenetic diversity, multilocus haplotype nomenclature, and in vitro antifungal resistance within the Fusarium solani species complex. Journal of Clinical Microbiology, 46(8), 2477-2490.

Oliveira, C. M., Flechtmann, C. A., & Frizzas, M. R. (2008). First record of Xylosandrus compactus (Eichhoff)(Coleoptera: Curculionidae: Scolytinae) on soursop, Annona muricata L.(Annonaceae) in Brazil, with a list of host plants.The Coleopterists Bulletin, 62(1), 45-48.

Pennacchio, F., Santini, L., & Francardi, V. (2012). Bioecological notes on Xylosandrus compactus (Eichhoff)(Coleoptera Curculionidae Scolytinae), a species recently recorded into Italy. Redia, 95, 67-77.

Rabaglia, R., Duerr, D., Acciavatti, R., & Ragenovich, I. (2008). Early detection and rapid response for non-native bark and ambrosia beetles. US Department of Agriculture Forest Service, Forest Health Protection.

Reeb, V., Lutzoni, F., & Roux, C. (2004). Contribution of RPB 2 to multilocus phylogenetic studies of the euascomycetes (Pezizomycotina, Fungi) with special emphasis on the -forming Acarosporaceae and evolution of polyspory. Molecular phylogenetics and evolution, 32(3), 1036-1060.

Simberloff, D., Parker, I. M., & Windle, P. N. (2005). Introduced species policy, management, and future research needs. Frontiers in Ecology and the Environment, 3(1), 12-20.

Six, D. L., & Wingfield, M. J. (2011). The role of phytopathogenicity in bark beetle- fungus symbioses: a challenge to the classic paradigm. Annual review of entomology, 56, 255-272.

Thompson, J. N. (2005). The geographic mosaic of coevolution. University of Chicago Press.

Townsend, A. M., Bentz, S. E., & Johnson, G. R. (1995). Variation in response of selected American elm clones to Ophiostoma ulmi. Journal of environmental horticulture, 13, 126-126.

66

Vilgalys, R., & Hester, M. (1990). Rapid genetic identification and mapping of enzymatically amplified ribosomal DNA from several Cryptococcus species.Journal of Bacteriology, 172(8), 4238-4246.

Von Arx, J. A., & Hennebert, G. L. (1965). Deux champignons ambrosia.Mycopathologia, 25(3), 309-315.

White, T. J., Bruns, T., Lee, S. J. W. T., & Taylor, J. W. (1990). Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. PCR protocols: a guide to methods and applications, 18, 315-322.

Wood, S. L. (1982). The bark and ambrosia beetles of North and Central America (Coleoptera: Scolytidae), a taxonomic monograph. The bark and ambrosia beetles of North and Central America (Coleoptera: Scolytidae), a taxonomic monograph.

Yadeta, K. A., & Thomma, B. P. (2013). The xylem as battleground for plant hosts and vascular wilt pathogens. Frontiers in plant science, 4.

Zhang, N., O'Donnell, K., Sutton, D. A., Nalim, F. A., Summerbell, R. C., Padhye, A. A., & Geiser, D. M. (2006). Members of the Fusarium solani species complex that cause infections in both humans and plants are common in the environment. Journal of Clinical Microbiology, 44(6), 2186-2190.Von Arx, J. A., & Hennebert, G. L. (1965). Deux champignons ambrosia Mycopathologia, 25(3), 309-315.

67

BIOGRAPHICAL SKETCH

Craig grew up exploring the forests of southern Michigan, where his passion for understanding nature around him was his focus since he can remember. After deeply exploring the world of birds and ornithology as a teenager, he soon learned that insects and fungi abounded with more interesting mysteries and his passions shifted to the field of entomology as a student at Michigan State University (MSU). Before receiving his

Bachelor of Science from MSU, his research focused on interactions between mosquitos and microbes under the guidance of Drs. Michael Kaufman and Edward

Walker. His growth and experience as a scientist during his undergraduate career gave him new interests in symbioses, leading him to investigate the relationships between ambrosia beetles, fungi, and trees at the University of Florida, where he received his

Master of Science working with Jiri Hulcr.

68