Spatial and temporal variation in mycorrhizal associations in a rare North American orchid

By

Pablo Antonio Tovar, B.S.

A Thesis In

PLANT AND SOIL SCIENCE

Submitted to the Graduate Faculty of Texas Tech University in Partial Fulfillment of the Requirements for the Degree of

MASTER OF SCIENCES

Dr. Jyotsna Sharma Chair of Committee

Dr. Richard Strauss

Dr. Stephen Maas

Dr. Mark Sheridan Dean of the Graduate School

December, 2015

© 2015 Pablo Antonio Tovar

Texas Tech University, Pablo Antonio Tovar, December 2015

ACKNOWLEDGEMENTS

First, I would like to express my sincere gratitude to my advisor Dr. Jyotsna Sharma for her guidance, support, motivation and contribution throughout the development of my master’s degree. In addition, I would like to acknowledge the other members of my advisory committee,

Dr. Richard Strauss and Dr. Stephen Maas for the great assistance and support, and patience for helping me out with the data analyses that were carried out for this research. Moreover, I am thankful to Dr. Madhav Pandey, a former Senior Research Associate in Dr. Sharma's lab, for providing aid with statistical analyses.

I wish to extend my appreciations to the U.S Fish and Wildlife Service for providing funding to Dr. Sharma for this investigation.

To my lab mates, I want to thank you all for your support revising my proposal, multiple times, and for helping me practice my thesis oral presentation. I hope we can work together again in the close future. To my parents, thanks for all the long distance support through this journey.

Being away from home is never easy, and I will keep following your constant advice always.

Finally, I am thankful to my girlfriend Lila for all the encouragement and affection over this process.

ii

Texas Tech University, Pablo Antonio Tovar, December 2015

TABLE OF CONTENTS

ACKNOWLEDGEMENTS ...... ii

LIST OF TABLES ...... vii

LIST OF FIGURES ...... xii

I. INTRODUCTION AND BACKGROUND ...... 1

Introduction ...... 1

Orchid diversity ...... 1

Plant habit ...... 4

Platanthera praeclara Sheviak and Bowles ...... 6

Morphology of praeclara ...... 7

Habitat of Platanthera praeclara ...... 8

Reproduction in Platanthera praeclara ...... 10

Mycorrhizal associations ...... 12

Threats to Platanthera praeclara ...... 15

Why study this taxon? ...... 17

Objectives of the study...... 20

Significance of the study ...... 20

Literature Cited ...... 22

II. MYCORRHIZAL ASSOCIATIONS OF PLATANTHERA PRAECLARA

ACROSS ITS NATURAL DISTRIBUTION ...... 35

Abstract ...... 35

Introduction ...... 37

Materials and methods ...... 41

iii

Texas Tech University, Pablo Antonio Tovar, December 2015

Collection sites ...... 41

Sample collection ...... 43

Fungal identification from root sections ...... 44

PCR and sequencing ...... 45

Cloning ...... 47

OTU clustering ...... 48

Mean pairwise sequence distance (pi, π) ...... 48

Fisher’s exact test ...... 49

Two-way hierarchical ordination ...... 49

Multiple response permutation procedure ...... 50

Diversity curves ...... 50

Nonmetric multidimensional scaling ...... 51

Phylogenetic analyses ...... 51

Results ...... 53

Operational taxonomic unit (OTU) diversity ...... 53

Mean pairwise sequence distance (π) ...... 55

Fisher’s exact test ...... 55

Two-way hierarchical ordination ...... 56

Multiple response permutation procedure ...... 56

Diversity curves ...... 57

Nonmetric multidimensional scaling ...... 57

Phylogenetic analyses ...... 58

Discussion ...... 60

Literature Cited ...... 68

iv

Texas Tech University, Pablo Antonio Tovar, December 2015

III. MYCORRHIZAL ASSOCIATIONS OF PLATANTHERA PRAECLARA IN

RESPONSE TO PRESCRIBED FIRE AND HAYING ...... 114

Abstract ...... 114

Introduction ...... 116

Materials and methods ...... 118

Collection sites ...... 118

Sample collection ...... 119

Environmental data ...... 120

Fungal identification from root sections ...... 121

PCR and sequencing ...... 122

Cloning ...... 123

OTU clustering ...... 124

Mean pairwise sequence distance (pi, π) ...... 125

Fisher’s exact test ...... 125

Two-way hierarchical ordination ...... 126

Nonmetric multidimensional scaling ...... 126

Results ...... 127

Operational taxonomic Units diversity ...... 127

Mean pairwise sequence distance (pi, π) ...... 128

Fisher’s exact test ...... 129

Two-way hierarchical ordination ...... 130

Nonmetric multidimensional scaling ...... 130

Discussion ...... 131

Literature Cited ...... 135 v

Texas Tech University, Pablo Antonio Tovar, December 2015

IV. CONCLUSIONS...... 163

vi

Texas Tech University, Pablo Antonio Tovar, December 2015

LIST OF TABLES

2.1. Number of sampled and number of samples that yielded useable sequences from roots of Platanthera praeclara in 2013 and 2014. Plant roots

were sampled for mycorrhizal diversity studies across 11 sites representing the natural distribution of the species………………………………………… 77

2.2. Soil chemical characteristics of 11 sites from where Platanthera praeclara roots were sampled in 2013 for mycorrhizal analyses. OM, organic matter; ENR, estimated nitrogen release; P1, phosphorus 1 (readily available to plants); P2, phosphorus 2 (readily available to plants plus part of the active reserve in the soil); K, potassium; Mg, magnesium; Ca, calcium; Na, sodium and CEC, cation exchange capability...... 78

2.3. Soil chemical characteristics of 11 sites from where Platanthera praeclara roots were sampled in 2014 for mycorrhizal analyses. OM, organic matter; ENR, estimated nitrogen release; P1, phosphorus 1; P2, phosphorus 2; K, potassium; Mg, magnesium; Ca, calcium; Na, sodium; H, hydrogen and CEC, cation exchange capability.…………………………………………….. 80

2.4. Primer pairs utilized, oligo sequences, and annealing temperatures to amplify the fungal ITS region of the nuclear ribosomal DNA via polymerase chain reaction from the DNA extracted from root fragments of Platanthera praeclara……………………………………………………………………... 82

2.5. Number of root sections (i.e. sequences) representing each of the 19 operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal

fungi at 11 sites in the year 2013. Values inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed……………………………. 83

2.6. Number of root sections (i.e. sequences) representing each of the 19 operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi at 11 sites in the year 2014. Values inside the parenthesis denote the

number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed……………………………. 84

2.7. Number of root sections (i.e. sequences) representing each of the 39 operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi at 11 sites in the combined data set (years 2013 and 2014). Values

inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed……………………...... 85

vii

Texas Tech University, Pablo Antonio Tovar, December 2015

2.8. Mean pairwise sequence distances and standard errors (SE) for all observed fungal ITS sequences from roots of Platanthera praeclara sampled across 11 sites in years 2013 and 2014, and for the combined data set (years 2013 and 2014), separated by fungal family. The values presented within the parenthesis are the numbers of individual ITS sequences (n) composing each group………………………………………………………………………….. 86

2.9. Mean pairwise fungal ITS sequence distances (pi-distance) estimated based on combined transitions and transversions substitution model, within fungal families identified in roots of Platanthera praeclara across 11 sites in North America. Mean pairwise sequence distances from publically available fungal ITS sequences obtained from orchids other than Platanthera praeclara and Nervilia nipponica were calculated by Pandey et al., 2013.………………….. 87

2.10. Statistical significance from the Fisher’s exact test comparing the fungal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) at each site for 2013 below the diagonal. Pairwise geographic distances (kilometers) between population pairs are presented above the diagonal. MB, Manitoba; PC, Pembina Control; PH, Pembina Hay; PF, Pembina Fall fire; PS, Pembina Spring fire; NDG, Graze; NDVK, North Dakota Viking; BU, Bluemound Uphill; BD, Bluemound Downhill; NE, and IA, …………….. 88

2.11. Statistical significance from the Fisher’s exact test comparing the fungal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) at each site for 2014 below the diagonal. Pairwise geographic distances (kilometers) between population pairs are presented above the diagonal. MB, Manitoba; PC, Pembina Control; PH, Pembina Hay; PF, Pembina Fall fire; PS, Pembina Spring fire; NDG, North Dakota Graze; NDVK, North Dakota Viking; BU, Bluemound Uphill; BD, Bluemound Downhill; NE, Nebraska and IA, Iowa.……………. 89

2.12. Statistical significance from the Fisher’s exact test comparing the fungal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) at each site using the combined data set (years 2013 and 2014) below the diagonal. Pairwise geographic distances

(kilometers) between population pairs are presented above the diagonal. MB, Manitoba; PC, Pembina Control; PH, Pembina Hay; PF, Pembina Fall fire; PS, Pembina Spring fire; NDG, North Dakota Graze; NDVK, North Dakota Viking; BU, Bluemound Uphill; BD, Bluemound Downhill; NE, Nebraska and IA, Iowa………………………………………………………………….. 90

viii

Texas Tech University, Pablo Antonio Tovar, December 2015

2.13. Multiple response permutation procedure (MRPP) showing the significance of the variation of the mycorrhizal fungal communities by using abundance data observed at 11 sites of Platanthera praeclara by grouping the sites according to the type of land management (prescribed fire, haying or grazing, or no management). The MRPP was performed with the combined data set (years 2013 and 2014) as well as years 2013 and 2014 separately.…. 91

3.1. Soil chemical characteristics in each prairie management treatment (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) from where Platanthera praeclara roots were sampled in 2013 for mycorrhizal analyses. OM, organic matter; ENR, estimated nitrogen release; P1, phosphorus 1 (readily available to plants); P2, phosphorus 2 (readily available to plants plus part of the active reserve in the soil); K, potassium;

Mg, magnesium; Ca, calcium; Na, sodium; CEC, cation exchange capability; and NO3, nitrate. Please see next page for the remaining table.……………... 139

3.2. Soil chemical characteristics in each prairie management treatment (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) from

where Platanthera praeclara roots were sampled in 2014 for mycorrhizal analyses. Mg, magnesium; Ca, calcium; Na, sodium; CEC, cation exchange capability; NO3, nitrate; S, Sulphur; Zn, zinc; Mn, manganese; Fe, iron; Cu, copper and B, boron and soil textural components (sand, silt and clay).…….. 141

3.3. Number of plants sampled and number of samples that yielded useable sequences from roots of Platanthera praeclara in 2013 and 2014. Plant roots were sampled for mycorrhizal diversity studies across four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) at Pembina Trail Preserve in , USA……….. 143

3.4. Number of root sections (i.e. sequences) representing each of the four operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi in each of the four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in 2013. Values inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed………………………………………………………………………. 144

3.5. Number of root sections (i.e. sequences) representing each of the 20 operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi in each of the four prairie management treatment plots (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in 2014. Values inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was 145 observed………………………...... ix

Texas Tech University, Pablo Antonio Tovar, December 2015

3.6. Number of root sections (i.e. sequences) representing each of the 20 operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi in each of the four prairie management treatment plots (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) using the combined data set (years 2013 and 2014). Values inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed…………… 146

3.7. Mean pairwise sequence distances and standard errors (SE) for all observed fungal ITS sequences from roots of Platanthera praeclara sampled from four experimental prairie management treatments in years 2013 and 2014, and for the combined data set (years 2013 and 2014), separated by fungal family Ceratobasidiaceae and Tulasnellaceae. The values presented within the parenthesis are the numbers of individual ITS sequences (n) composing each group…….……………………………………………………………… 147

3.8. Mean pairwise sequence distances and standard errors (SE) for all observed fungal ITS sequences from roots of Platanthera praeclara in each of four prairie treatment plots (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) for years 2013 and 2014 separated by land management treatment. The values presented within the parenthesis are the numbers of individual ITS sequences (n) composing each group………………………… 148

3.9. Statistical significance from the Fisher’s exact test comparing the fungal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) observed in each of the four prairie management treatment plots (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) for 2013……………………………………….. 149

3.10. Statistical significance from the Fisher’s exact test comparing the mycorrhizal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) observed in each of the four prairie

management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) for 2014………………………………………………. 150

x

Texas Tech University, Pablo Antonio Tovar, December 2015

3.11. Statistical significance from the Fisher’s exact test comparing the fungal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) observed in each of the four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) for the combined data set (years 2013 and 2014)…….. 151

xi

Texas Tech University, Pablo Antonio Tovar, December 2015

LIST OF FIGURES

1.1. An inflorescence of Platanthera praeclara. Photo by Jyotsna Sharma.... 32

1.2. A partial map of the US and Canada showing the natural distribution of Platanthera praeclara. The states where the plant is currently known to be present (orange), the counties where the populations occur (gray) and the counties where roots were collected for the studies (yellow) are highlighted.……………………………………………………………... 33

1.3. A transverse section of a root of Platanthera praeclara showing colonization by fungal pelotons (examples indicated by arrows). The scale-bar represents 200 microns………………………………………. 34

2.1. A partial map of the US and Canada showing the natural distribution of Platanthera praeclara. The states where the plant is currently known to be present (orange), the counties where the populations occur (gray) and the counties where roots were collected for the studies (yellow) are highlighted.……………………………………………………………… 92

2.2. A map of the nuclear fungal ribosomal internal transcribed spacer (ITS) region. The two ITS regions are between the SSU18S and LSU28S ribosomal RNA genes and are separated by the 5.8S rRNA gene. Image taken from http://www.gatc-biotech.com...... 93

2.3. A Platanthera praeclara plant excavated for sampling while keeping the root system intact. Photo on the left was taken by Nancy Sather…… 94

2.4. A whole seedling of Platanthera praeclara along with severed root and tuber...... 95

2.5. Root fragments collected from plants of Platanthera praeclara………. 96

2.6. A transversal section of a root of Platanthera praeclara showing colonization by fungal pelotons in cortical cells (examples indicated by arrows). The scale-bar represents 400 microns………………………… 97

xii

Texas Tech University, Pablo Antonio Tovar, December 2015

2.7. Pie charts showing the distribution of the fungal families Ceratobasidiaceae and Tulasnellaceae observed in roots of Platanthera praeclara at 11 sites in the year 2013. Numeric values represent the individual ITS sequences belonging to each fungal family. of in the year 2013. Numbers in the pie charts represent the amount of ITS sequences from each fungal family in a site……………………………. 98

2.8. Pie charts showing the distribution of the fungal families Ceratobasidiaceae and Tulasnellaceae observed in roots of Platanthera

praeclara at 11 sites in the year 2013. Numeric values represent the individual ITS sequences belonging to each fungal family.…………… 99

2.9. One to four operational taxonomic units in individual plants of Platanthera praeclara (a) and number of individual plants that hosting one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b) across 11 sites in 2013.……………………………………………… 100

2.10. One to four operational taxonomic units in individual plants of Platanthera praeclara (a) and number of individual plants that hosted one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b) across 11 sites in 2014.………… …………………………………... 101

2.11. One to four operational taxonomic units in individual plants of Platanthera praeclara (a) and number of individual plants that hosted one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) across 11 sites (b) for the combined data set (years 2013 and 2014)…… 102

2.12. Two-way hierarchical cluster tree and matrix coding based on abundance of 19 fungal operational taxonomic units (OTUs) observed at 11 sites where roots of Platanthera praeclara were sampled in the year 2013……………………………………………………………….. 103

2.13. Two-way hierarchical cluster tree and matrix coding based on abundance of 19 fungal operational taxonomic units (OTUs) observed at 11 sites where roots of Platanthera praeclara were sampled in the year 2014……………………………………………………………….. 104

xiii

Texas Tech University, Pablo Antonio Tovar, December 2015

2.14. Two-way hierarchical cluster tree and matrix coding based on abundance of 39 fungal operational taxonomic units (OTUs) observed at 11 sites where roots of Platanthera praeclara were sampled in the combined data set (years 2013 and 2014)…………………………….... 105

2.15. Sample based observed (Mao Tau) and rarefaction (Chao 1) cumulative fungal operational taxonomic unit (OTU) abundance curves for Platanthera praeclara. Fungal communities within the roots of plants were determined by sampling plants from 11 sites that host the orchid species.…………………………………………………………………... 106

2.16. Nonmetric multidimensional scaling (NMS) ordination of the mycorrhizal fungal communities within the roots of Platanthera praeclara sampled from 11 sites with correlations to soil variables [Na, Mn, Zn, Ca, OM (organic matter), CEC (cation exchange capability), soluble salts and pH] for the year 2013……………………………….... 107

2.17. Nonmetric multidimensional scaling (NMS) ordination of the fungal communities of Platanthera praeclara sampled from 11 sites with correlations to soil variables (S, B, H, clay and silt) for the year 2014... 108

2.18. Partial Maximum likelihood tree of the fungal family Ceratobasidiaceae constructed with operational taxonomic unit (OTU) sequences observed in Platanthera praeclara roots collected from 11 sites across its natural habitat. Sequences of known orchid mycorrhizal fungi were added to the alignments to place our sequences in a broader context. Scale bar represents estimated number of DNA substitutions per site. Only branch support values over fifty percent are displayed (maximum likelihood / posterior probability). Please see the next page for the remaining tree…. 109

2.19a. Maximum likelihood tree of the fungal family Tulasnellaceae constructed with operational taxonomic unit (OTU) sequences observed in Platanthera praeclara roots collected from eleven populations across its natural habitat. Sequences of known orchid mycorrhizal fungi were added to the alignments to place our sequences in a broader context. Scale bar represents estimated number of DNA substitutions per site. Only branch support values over fifty percent are displayed (maximum likelihood / posterior probability)……………………………………….. 111

xiv

Texas Tech University, Pablo Antonio Tovar, December 2015

2.19b. Maximum likelihood tree of the fungal family Tulasnellaceae constructed with operational taxonomic unit (OTU) sequences observed in Platanthera praeclara roots collected from eleven populations across its natural habitat. Sequences of known orchid mycorrhizal fungi were added to the alignments to place our sequences in a broader context. Scale bar represents estimated number of DNA substitutions per site. Only branch support values over fifty percent are displayed (maximum likelihood / posterior probability)……………………………………….. 112

2.19c. Maximum likelihood tree of the fungal family Tulasnellaceae constructed with operational taxonomic unit (OTU) sequences observed in Platanthera praeclara roots collected from eleven populations across its natural habitat. Sequences of known orchid mycorrhizal fungi were added to the alignments to place our sequences in a broader context. Scale bar represents estimated number of DNA substitutions per site. Only branch support values over fifty percent are displayed (maximum likelihood / posterior probability)……………………………………….. 113

3.1. Satellite image of the arrangement of the four prairie management treatment plots (PC, no management; PS, spring burn; PF, fall burn, and PH, annual haying) where mycorrhizal fungi of Platanthera praeclara were sampled in years 2013 and 2014. Experimental plots were located at Pembina Trail Preserve, Polk County, Minnesota, USA. Image taken from Google Earth.……………………………………………………… 152

3.2. Soil chemical characteristics in each prairie management treatment (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) from where Platanthera praeclara roots were sampled in 2014 for mycorrhizal analyses. Mg, magnesium; Ca, calcium; Na, sodium; CEC, cation exchange capability; NO3, nitrate; S, Sulphur; Zn, zinc; Mn, manganese; Fe, iron; Cu, copper and B, boron and soil textural components (sand, silt and clay)………………………………………... 153

3.3. Pie charts showing the distribution of the fungal families Ceratobasidiaceae and Tulasnellaceae observed in roots of Platanthera praeclara sampled from four experimental prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in the year 2014. Numeric values represent the number of individual ITS sequences belonging to each fungal family……………... 154

xv

Texas Tech University, Pablo Antonio Tovar, December 2015

3.4. One to three operational taxonomic units in individual plants of Platanthera praeclara (a), and number of individual plants hosting one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b), across four experimental prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in 2013…. 155

3.5. One to four operational taxonomic units in individual plants of Platanthera praeclara (a), and number of individual plants hosting one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b), across four experimental prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in 2014…. 156

3.6. One to four operational taxonomic units in individual plants of Platanthera praeclara (a), and number of individual plants hosting one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b), across four experimental prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment). Data represent samples collected from each experimental treatment over two years (2013 and 2014).………………………………………………….. 157

3.7. Two-way hierarchical cluster tree and matrix coding based on abundance of four fungal operational taxonomic units (OTUs) within the roots of Platanthera praeclara representing four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment). Data represent samples collected from each experimental treatment in 2013…………………………………… 158

3.8. Two-way hierarchical cluster tree and matrix coding based on abundance of 15 fungal operational taxonomic units (OTUs) within the roots of Platanthera praeclara representing four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment). Data represent samples collected from each experimental treatment in 2014……………………………………………………….. 159

3.9. Two-way hierarchical cluster tree and matrix coding based on abundance of 20 fungal operational taxonomic units (OTUs) within the roots of Platanthera praeclara representing four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment). Data represent samples collected from each experimental treatment over two years (2013 and 2014).…………………………….. 160 xvi

Texas Tech University, Pablo Antonio Tovar, December 2015

3.10. Nonmetric multidimensional scaling (NMS) ordination of the fungal communities of Platanthera praeclara in 4 land management treatment plots (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) with correlations to soil variables (clay, silt, sand, S, P1, Na, B and pH) for the year 2013……………………………………………. 161

3.11. Nonmetric multidimensional scaling (NMS) ordination of the fungal communities of Platanthera praeclara representing four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying, and PC, no treatment) with correlations to soil physicochemical variables (clay, silt, sand, S, P1, OM and pH) and soil environmental variables (electrical conductivity in June, EC J; electrical conductivity in May, EC M, and volumetric water content in June, VWC J) for the year 2014..…………………………………………………………...... 162

xvii

Texas Tech University, Pablo Antonio Tovar, December 2015

CHAPTER I

INTRODUCTION AND BACKGROUND

Introduction

Orchid diversity

The is one of the largest and arguably the most diverse families of angiosperms. Estimates suggest a range between 20,000 and 35,000 species in 5 subfamilies and approximately 870 genera, occurring in all vegetated continents and some Antarctic islands (Dressler, 1993; Mabberley, 1997; Chase et al., 2003). The highest abundance and distribution occurs towards the tropics, following hotspots of species richness and endemism (Myers et al., 2000; Swarts & Dixon, 2009). Relatively little is known about the origin and biogeography of the Orchidaceae. Ramirez et al.

(2007) suggested that the subfamilies had diverged around 65 mya between the late

Cretaceous and early Paleocene when global temperatures were rising (Zachos et al.,

2008). By using molecular dating techniques on orchid fossils of the two major subfamilies and , Gustafsson et al. (2010) suggested that major diversification within these two subfamilies occurred during a cooler period by the end of the Eocene and into the Oligocene.

A number of hypotheses owing to pollination specialization, specialized mycorrhizal interactions, and niche partitioning attempt to explain the diversity within the

Orchidaceae.

1

Texas Tech University, Pablo Antonio Tovar, December 2015

Pollinator specialization hypothesis is one of them (Cozzolino & Widmer, 2005).

The Orchidaceae is marked by a degree of pollinator specialization that probably has not been attained by any other family of angiosperms (Ibish et al., 1996). Certain pollination strategies are argued to involve remarkably specialized relationships between plants and pollinators. Some orchids produce highly specific suites of olfactory and visual stimuli that attract specific pollinator species. For example, in orchids that utilize pseudo- copulation strategies, a male bee or wasp is attracted to these stimuli, visually mistaking the orchid flower for a female conspecific and pollinates through repeated mistaken copulations (Dafni & Bernhardt, 1990). It is also reported that recently derived sub- families show a decrease in pollinator specialization (Tremblay, 1992). Additionally,

Gravendeel et al. (2004) did not record evidence for orchid speciation being driven by pollinator specialization. Instead, the study reports that Epidendroideae does not present a different level of pollinator specialization from other subfamilies. This is especially notable because Epidendroideae represents the most species rich subfamily in

Orchidaceae.

Orchids exemplify diversity in specialized mycorrhizal interactions (van der Pijil &

Dodson, 1966; Dressler, 1981, Rasmussen, 1995; Taylor et al., 2002; Otero & Flannagan,

2006; Barret et al., 2010). Examples for this are the common and widely distributed orchid species like Neottia ovata that associate with a broad spectrum of mycorrhizal fungi from families including Tulasnellaceae, Ceratobasidiaceae, Thelephoraceae, and

Sebacinaceae (Jacquemyn et al. 2015). Also, some species with restricted distributions, such as Piperia yadonii, host a large variation in mycorrhizal communities among

2

Texas Tech University, Pablo Antonio Tovar, December 2015 populations (Pandey et al. 2013). Girlanda et al. (2011) reported specific mycorrhizal associations and partial mycoheterotrophy in four widely distributed photosynthetic orchids. In addition, non-photosynthetic, fully mycoheterotrophic orchid species, reported to have the most narrow associations, usually host mycorrhizal fungi from one fungal family, can range from widely distributed habitats, like Neottia nidus-avis (Selosse et al.,

2002) to restricted distribution like colemanii (Kennedy et al., 2011)

Considering that orchid seeds are rudimentary and lack endosperm, they lack stored energy resources for germination and establishment (Smith and Read, 2008). Thus, they rely upon mycorrhizal symbiosis to provide nutrients and carbohydrates to germinating seeds. Orchid mycorrhizae differ from other types of endomycorrhizae like arbuscular or ericoid mycorrhizae in that hyphae actually penetrate root cell membranes and form hyphal coils (Smith and Read, 2008). Other endomycorrhizae invaginate the cell membrane, but do not penetrate the cell membrane (Peterson et al., 2004). In addition, studies have shown a high level of orchid specificity towards mycorrhizal fungi

(Taylor and Bruns, 1997; Otero et al., 2002; Otero et al., 2004), where most of the fungal associates belong to three families in the basidiomycetes: Ceratobasidiaceae,

Tulasnellaceae and Sebacinaceae (Dearnaley et al., 2012), among others (Cortinariaceae,

Atheliaceae, Auriculariaceae, Cantharellaceae, Telephoraceae). Irregular distributions of orchid mycorrhizal fungi, combined with a diversity of mycorrhizal strategies ranging from broad to highly specific could explain the high diversity of orchid species (Otero and Flannagan, 2006; McCormick and Jacquemyn, 2014).

3

Texas Tech University, Pablo Antonio Tovar, December 2015

Environmental niche partitioning or microhabitat specialization has been demonstrated in tropical epiphytic environments (Gentry and Dodson, 1987; Wolf and

Flamenco, 2003). In tropical forests, is not uncommon to find orchids restricted to different layers across the forest stratification. Biotic and abiotic factors change across the gradient, affecting species distribution (Diez & Pulliam, 2007). Still, this hypothesis has yet to be extensively researched in orchids.

Plant habit

Approximately 70% of the Orchidaceae are photosynthetic epiphytes (Zotz, 2013) that are primarily distributed in the tropical and subtropical regions of the planet.

Epiphytic orchids typically grow numerous roots that increase adhesion to the host

(Gravendeel et al., 2004). Tropical epiphytic species may experience periodic drought due to seasonal rains. Adaptations to an epiphyte habit include modifications in the root system. Roots of epiphytic orchids may be photosynthetic and aid in uptake and storage of water and nutrients (Gravendeel et al., 2004). Succulence, crassulacean acid metabolism (CAM), and sequential production of individual shoots operating as independent physiological units are among the adaptations for nutrient and water uptake in epiphytic orchid taxa (Gravendeel et al., 2004).

The remaining 30% of the orchid taxa occupy terrestrial habitats distributed across all continents. Like epiphytic orchids, terrestrial orchids have specific adaptations according to the environmental pressures they have to face within a habitat. Orchids may exhibit plant organs that are modified for mycorrhizal associations. A protocorm, for 4

Texas Tech University, Pablo Antonio Tovar, December 2015 example, is a tuber-shaped body with rhizoids that make up the entire seedling for a period of time before producing a shoot. Terrestrial orchids also may produce thickened roots, tubers, corms, and / or rhizomes that can serve as belowground storage structures

(Rasmussen, 1995). As stated by Rasmussen (1995), terrestrial orchids are dependent on mycorrhizal association for germination and plant development, especially in the protocorm and seedling stages.

When a seed from terrestrial orchids germinates, it first develops into a protocorm, a structure that bears primordial leaves but no roots. Eventually, the mycorrhizal stage develops when the apical meristem elongates and the first root initials develop

(Rasmussen, 1995). The protocorm meristem produces rhizomes that may vary in shape according to the species (Rasmussen, 1995). This rhizome is a modified subterranean stem that sends out roots and shoots from its nodes. Rhizomes start with mycotrophic functions, but eventually this function is transferred to the roots and the mycotrophic tissue in the rhizome decreases. Storage of water, starch, proteins and other nutrients is another function of the rhizome in orchids (Jang et al., 2006).

Terrestrial orchids typically have thick, fleshy, brittle unbranched roots; root systems are generally simple and small in relation to the plant size. Apart from mycotrophy, root functions include absorption of water and nutrients, storage, and anchorage (Rasmussen, 1995). Several terrestrial orchids produce root modifications that are called tubers for nutrient storage as their principal function. Tubers vary considerably in habit and shape, and usually they are colonized by fungi only in peripheral extensions and in superficial tissue (Rasmussen, 1995). The old tuber senesces during the growing 5

Texas Tech University, Pablo Antonio Tovar, December 2015 season as the new tuber begins to grow (Medina et al., 2009). During the dormant season

(winter, or dry season), the aboveground portions of the plant senesce except for the stem tuber and rhizome that grow a new shoot in the subsequent growing season, producing stems and leaves.

Platanthera praeclara Sheviak and Bowles

Platanthera praeclara (Orchidaceae) (Figure 1.1) is placed in the subfamily

Orchidoideae, tribe Orchideae and subtribe Orchidinae. Plants in this subtribe have a terrestrial habit. Orchidinae are usually non-saprophytic orchids with root-stem tuberoids.

Their roots do not have velamen (usually present in epiphytic orchids) and they have a slender stem (Dressler, 1993). The genus Platanthera includes approximately 200 species that are distributed across temperate and tropical North America, North Africa, Central

America and Eurasia (U.S. Fish and Wildlife Service, 2009). In North America, 24

Platanthera species occur from northern Mexico to the southern province of Manitoba in

Canada (USFWS, 2009). Platanthera is one of the most species-rich genera of temperate orchids in the northern hemisphere (Hapeman and Inoue, 1997), having most of its diversity in North America (Sheviak & Catling, 2002) and in East Asia (Inoue, 1983).

Species of the genus Platanthera occur in a wide variety of habitats and soils, usually acidic, ranging from swamps and meadows to fairly dry woodlands (Correll, 1978;

Williams and Williams 1983; Rasmussen, 1995).

Platanthera praeclara, previously included in the P. leucophaea taxonomic concept, was described by Sheviak and Bowles in 1986 when they encountered plants on 6

Texas Tech University, Pablo Antonio Tovar, December 2015 the Sheyenne in North Dakota that exhibited distinct features. The species is currently listed as 'Threatened' by the U.S. Fish and Wildlife Service under the

Endangered Species Act of 1973 in the United States and as Endangered in Canada

(USFWS, 2009; Environment Canada, 2006). Natural populations of the species are known to occur in 6 mid-western states in the United States (, , Iowa,

Nebraska, Minnesota, and North Dakota) and in the province of Manitoba, Canada

(USFWS, 2009) (Figure 1.2).

Platanthera praeclara may have been historically distributed through a large portion of the western Central Lowlands and eastern Great Plains of the central United

States, and through the Interior Plains in south-central Canada (Brownell, 1984). The species had historically occurred in South Dakota and , but conversion of prairies to agriculture in the latter half of the nineteenth century caused extinction of the species in these states (Bowles and Duxbury, 1986; USFWS, 2009).

Morphology of Platanthera praeclara

Platanthera praeclara, as described by Sheviak and Bowles (1986), is an erect, stout, herbaceous perennial. It grows from 30 to 85 cm tall. Typically, the plant has multiple coarse, fleshy roots. Each plant usually has a single, glabrous, unbranched stem bearing two to five oblong-elliptic to lanceolate, keeled, glabrous leaves. These are usually 7 to 15 cm long and 1 to 4 cm wide. The inflorescence is 5 to 22 cm long and 4 to

7 cm in diameter and is a raceme of 5 to 25 flowers, with lanceolate, acuminate bracts 1.5

7

Texas Tech University, Pablo Antonio Tovar, December 2015 to 4 cm long and 0.4 to 0.7 cm wide. Flowers are white or creamy white and have a perianth which is directed forward and forms a hood over the column. The dorsal sepal is ovate to suborbiculate, concave, 9 to 13 mm long and 5 to 8 mm wide. The lateral sepals are obliquely- obovate, asymmetrical, 7 to 14 mm long and 5 to 10 mm wide. The lateral petals are cuneate to flabelliform, rounded to truncate, 9 to 16 mm long, and 6 to 13 mm wide with lacerate distal margins. The labellum (modified petal) is deeply 3-lobed, usually 2 to 3.2 cm long, 2 to 4 cm wide, fringed, and bears a slender, arcuate, clavate, 4 to 5.5 cm long, sometimes as short as 2 cm long spur. The ellipsoid capsule is 2 to 2.5 cm long and 4 to 6 mm in diameter and releases seeds upon dehiscence.

Habitat of Platanthera praeclara

The western prairie fringed orchid is a perennial herbaceous species occurring in the North American . Tallgrass prairie plant communities are historically fire and grazing adapted and are usually dominated by Andropogon gerardii, A. scoparius, Sorghastrum nutans, Deschampsia caespitosa and Panicum virgatum, all members of the Poaceae (USFWS, 1996; Ghimire et al., 2011). Tallgrass prairie remains in the eastern fringe of the Red River Valley in Manitoba in Canada and in Minnesota in the USA. In eastern North Dakota, it survives in the Sheyenne National Grassland. From

South Dakota, through Minnesota and Iowa tallgrass prairie can be found in the Coteau des Prairies. Through east-central Kansas it can be found in the Flint Hills, and in the south, in can be found in northern Oklahoma. Tallgrass prairie extends further into the east through Illinois, Indiana and a portion of south western Michigan (Jones and

8

Texas Tech University, Pablo Antonio Tovar, December 2015

Cushman, 2004). Platanthera praeclara plants generally occur in unplowed, calcareous prairies and sedge meadows (USFWS, 1996). The largest known population of P. praeclara occurs in southeastern Manitoba, Canada, with 23,530 flowering plants documented in the year 2003 (Environment Canada, 2006).

Platanthera praeclara in Manitoba occur in wet prairies or meadow vegetation developed on Alfisols (poorly drained grey wooded soils) with a thin sandy mantle overlying stony calcareous reworked till (Collicutt, 1992; USFWS, 1996). In North

Dakota, Platanthera praeclara is usually observed in the sedge meadows in the Glacial

Sheyenne Delta in tallgrass prairies (USFWS, 1996). Calcium rich wet prairie soils

(calciaquolls); wet prairie soils with minimum horizon development (haploquolls) and cool prairie soils with minimum horizon development (haploboralls) are the typical soils that support orchid growth on this area (United States Department of Agriculture, 1999;

Wolken, 1995). In Minnesota, populations in the Pembina Trail prairie complex are associated with poorly drained to moderately well–drained soils with gentle slopes

(USFWS, 1996). Populations in Iowa, southwestern Minnesota, northeastern Nebraska, northeastern Kansas and northwestern Missouri occur in wet-mesic to mesic tallgrass prairie (Freeman and Brooks, 1989). Udolls or Udic Ustolls (humid to intermittently dry mollisols, or prairie soils) constitute soils in these areas, on gentle to moderate slopes

(USDA, 1999). Plants in north-central Nebraska typically grow in tall grass prairies or sedge meadows in swales among the Nebraska Sand Hills in poorly developed sandy soils of warm climates (USDA, 1999). Platanthera praeclara growing in eastern

Nebraska occurs in wet mesic prairies and sedge meadows along the floodplain of the

9

Texas Tech University, Pablo Antonio Tovar, December 2015

Platte River (Nagel and Kolstad, 1987). In eastern Kansas P. praeclara is observed in mesic to wet-mesic upland prairies on unglaciated, level to hilly, Pennsylvanian-age sediments covered with a thin, discontinuous mantle of loess residuum (USFWS, 1996).

Platanthera praeclara generally occurs in the wetter facies of such prairies or in associated sedge meadows (USFWS, 1996). Perennial taxa from the family Cyperaceae, with Carex spp. and Eleocharis spp. dominate sedge meadows. In addition, annual and perennial grasses and forbs (like Poa pratensis, Carex lanuginose and Juncus balticus) also occur in this community type (USFWS, 1996). Changes in the plant community might alter orchid presence (Wolken, 2001). Swales in sedge meadows supporting P. praeclara were compared with swales lacking orchids and it has been suggested that plant canopy cover of Juncus balticus, Stachys palustris and Carex spp. was higher in swales supporting the orchids (Wolken, 2001).

Reproduction in Platanthera praeclara

Platanthera praeclara has evolved an obligate outcrossing pollination system, i.e. pollen is delivered to a flower from a different individual (Sheviak and Bowles, 1986).

The flowers lack nectar guides, bear long nectariferous spurs and are fragrant at night; these are features that typify hawkmoth-pollinated plants (Phillips, 2003). In an effort to reach the nectar contained within the flower’s spur, the hawkmoth unintentionally brushes up against the pollinia, and they generally attach to its eyes (USFWS, 1996). The hawkmoth then transfer the pollinia to another P. praeclara flower (Phillips, 2003). Fox et al. (2013) documented six different moth species (including two native species) of 10

Texas Tech University, Pablo Antonio Tovar, December 2015 hawkmoth as pollinaria carrying pollinators of P. praeclara over nine years of observations at the Sheyenne National Grasslands in North Dakota, USA. These include:

Lintneria eremitus, Hyles lineata, Sphinx drupiferarum, , Hyles euphorbiae and Hyles gallii, two of which (S. drupiferarum and H. gallii) had already been reported as pollinators of P. praeclara in Manitoba, Canada (Westwood and

Borkowsky, 2004; Borkowsky and Westwood, 2009).

Plant species with more diverse pollinators are typically at a lower extinction risk than those that rely on a few specialized pollinators (Bond, 1994; Marten-Rodriguez et al., 2010). Abundance of the pollinator species also plays a major role (Tremblay et al.,

2005). Unfortunately, only one (H. euphorbiae) of the known pollinators of P. praeclara is considered abundant. The other pollinator species are not abundant, and two of them,

E. achemon and S. drupiferarum, now appear to be less common (Fox et al., 2013). In addition, it is important to note that the hawkmoths do not depend on the orchid as they can feed from a number of non-orchid nectar sources to support their populations, whereas the orchid depends on the hawkmoths for pollination (USFWS, 1996).

As mentioned above, P. praeclara depends on outcrossing pollination for reproduction (Sheviak and Bowles, 1986). However, low pollination rates and seed production have been observed in most populations (Phillips, 2003). Borkowsky and

Westwood (2009) assessed pollination in the Manitoba Tall Grass Prairie Preserve in

Canada by recording number of removed and remaining pollinaria and number of seed capsules, among other observations, from flowers of P. praeclara. The study reported a

11

Texas Tech University, Pablo Antonio Tovar, December 2015 significant increase in seed capsule production in P. praeclara plants that had ultraviolet lights placed by them. However, the study describes that only 7% of flowers produced seed capsules, a smaller percentage of what has been reported in P. praeclara from

Minnesota and North Dakota in other similar studies (Pleasants, 1993; Pleasants and

Moe, 1993).

Mycorrhizal associations

Mutualistic mycorrhizal symbiosis involves reciprocal transfer of nutrients between a plant and its fungal partner (Kiers et al., 2011). Plants may reward fungal hyphae that are able to allocate greater phosphorus (P) resources. Consequently, fungi respond to this cooperation by supplying increased phosphorus (P) when receiving more carbon (C)

(Kiers et al., 2011).

Members of the Orchidaceae have evolved their own form of mycorrhiza. All orchid seeds are believed to be mycoheterotrophic in nature, or fully dependent on fungi for germination and early protocorm development before plants gain photosynthetic capability (Smith and Read, 2008). Coils of fungal hyphae, called pelotons (Figure 1.4), are digested by the host plant to acquire nitrogen (N), carbon (C), and other nutrients

(Smith, 1967). It is important to add that the fungi involved in the symbiosis do not appear to depend on the host plant to survive (Rasmussen, 1995; Bunch et al., 2013).

Orchids across the planet have been reported to be most commonly associated with members of the fungi previously referred to as ‘rhizoctonia’, which were recorded in older literature as the anamorphic states of heterobasidiomycetes (Currah and Zelmer,

12

Texas Tech University, Pablo Antonio Tovar, December 2015

1992; Sen et al., 1999; Sharma et al., 2003a; Taylor et al., 2003; Zettler et al., 2003;

Zettler and Piskin, 2011). 'Rhizoctonias' were divided into two genera, Ceratorhiza and

Epulorhiza (Roberts, 1999), but this assemblage contains three different Agaromycetes taxa including Sebacinales, Ceratobasidiaceae and Tulasnellaceae. However, as explained by Dearnaley et al. (2012), this classification is not entirely correct since none of these taxa is adequately described by the definition (De Candolle, 1815) of the now obsolete asexual genus Rhizoctonia.

Orchid mycorrhizal fungi are widely diverse outside of their ecological niche inside orchid roots. The fungal family Ceratobasidiaceae includes economically important pathogenic and parasitic species such as Ceratobasidium noxium, which causes black rot of coffee, and Ceratobasidium cereale that causes sharp eyespot of cereals (Roberts,

1999). Many species are saprophytic, occurring in the soil and plant detritus (Smith and

Read 2008). Tulasnellaceae, along with the Ceratobasidiaceae have been detected as taxa in clades forming ectomycorrhizae (Bidartondo et al., 2003; Yagame et al., 2012). Fungi from the Order Sebacinales have been observed as mycorrhizae forming fungi in the

Orchidaceae and Ericaceae, as well as occurring as common endophytes in the roots of many plant species (Selosse et al., 2009; Weiβ et al., 2011). In orchids, molecular taxonomic identification of fungal associates have revealed that Ceratobasidiaceae,

Sebacinaceae and Tulasnellaceae generally associate with orchids worldwide (Dearnaley et al., 2012), with Tulasnellaceae observed as the most frequently fungal family in temperate and tropical regions (Rasmussen, 1995; Yuan et al., 2010).

13

Texas Tech University, Pablo Antonio Tovar, December 2015

Despite the wide diversity and distribution of orchid mycorrhizal forming fungi, different orchid taxa often restrict the associations to specific fungi. The specificity in the association between fungi and orchids can be defined as the phylogenetic breadth in the diversity of the associates (Shefferson et al., 2010). Narrow fungal specificity is more noticeable usually in holomycoheterotrophic (i.e., fully dependend on fungi) orchid species that typically associate with fungi from one fungal family (Selosse et al., 2002;

Kennedy et al., 2011). Conversely, low specificity has been observed in orchids associating with two or three fungal families at the same time (Bonnardeaux et al., 2007;

Pandey et al., 2013).

Mycorrhizal associations in Platanthera praeclara have been studied previously to some extent. However, like majority of the other orchids, these critical associations have not been investigated range-wide or over time. In 1995, Zelmer and Currah first isolated and characterized eight isolates of Ceratorhiza pernacatena from the family

Ceratobasidiaceae by sampling roots from seven plants of P. praeclara from a single population in Manitoba, Canada. From the same orchid population, one isolate belonging to the Tulasnellaceae and fifteen Ceratobasidiaceae isolates were subsequently observed in 11 plants of P. praeclara (Zelmer et al., 1996). Later, Sharma et al. (2003a) isolated, characterized and identified fungi from 21 plants of P. praeclara in its native habitat in

Minnesota (Clay county, Norman county, Polk county and Mower county) and Missouri

(Harrison county). In this study, 78 isolates belonged to the fungal family

Ceratobasidiaceae (e.g. Genbank accession number DQ088771) and nine belonged to the fungal family Tulasnellaceae (e.g. Genbank accession number DQ068772). Results from

14

Texas Tech University, Pablo Antonio Tovar, December 2015 these studies in Manitoba, Minnesota, and Missouri indicated that the orchid may prefer members of the Ceratobasidiaceae when forming mycorrhizal associations. However, a large part of the range was not represented in these studies, which limits the interpretation of range-wide patterns of mycorrhizal diversity and specificity in the species.

Additionally, temporal variation in mycorrhizal associations (phenological stages and / or yearly variation) was not captured in these studies (Zelmer and Currah, 1995; Zelmer et al. 2006; Sharma et al. 2003a). However, studies to test in vitro symbiotic germination of

P. praeclara seeds indicated that germination is more successful when cold-moist stratified seeds are cultured with fungal isolates derived from naturally occurring protocorms of P. praeclara (Sharma et al., 2003b) suggesting that there might be a successional patterns in fungal associations through the various phenological stages of plant growth.

Threats to Platanthera praeclara

The tallgrass prairie ecosystem is one of the most threatened ecosystems in North

America if not on Earth (Jones and Cushman, 2004). Approximately 1% of the original tallgrass prairie remains. Some of the reasons for this loss include the nutrient rich soils formed by prairie ecosystems that are also highly desirable for agricultural use.

Conversion to crop fields physically removes the native prairie ecosystem in addition to causing hydrological changes in any remaining adjacent prairies. Platanthera praeclara prefers late-successional prairies that do not undergo rapid changes, which renders them exceptionally sensitive to habitat destruction and degradation (USFWS, 1996; Goedeke et al., 2008). Anthropogenic factors can induce habitat alteration and degradation while

15

Texas Tech University, Pablo Antonio Tovar, December 2015 inappropriate land management results in ecological changes that disadvantage P. praeclara populations. For example, non-application or misapplication of prescribed fire, grazing or mowing could have a detrimental impact on orchid survival, growth and reproduction (USFWS, 1996). Habitat deterioration and loss could lead to a decline in soil microbiota, local hydrology, and changes in plant and animal community compositions. As a result, exotic or invasive plants that P. praeclara could not out- compete find it easier to establish, displacing the orchid (Kirby et al., 2003). Agricultural production, especially row-cropping has damaged large tracts of mature, late-succession prairie habitat that is considered ideal for the orchids to persist (USFWS, 1996). If, instead, land is under grazing usage, cattle can also damage individual plants by grazing and / or trampling them. While natural low-density grazing may have been viable under historical circumstances when the plant was widely distributed, high-density commercial grazing can be a considerable threat especially for populations expressing low numbers of individuals (Goedeke et al., 2008). Moreover, remnant prairie and grazing lands are decreasing under pressure of urban development (Goedeke et al., 2008).

While the biology and ecology of P. praeclara are somewhat understood, concern increases about the persistence of the species. As populations decline, irregular flowering, dormancy periods, the need for appropriate symbiont and the need for particular pollinators might reduce the likelihood of population survival. Further, with smaller and isolated populations, genetic diversity is reduced and a decline in reproduction is likely to occur (Sharma et al., 2002).

16

Texas Tech University, Pablo Antonio Tovar, December 2015

Other miscellaneous threats to the native Tallgrass Prairie communities and specifically to P. praeclara include damage or destruction by herbicides used against invasive species, roadway maintenance activities, introduction of new species, and / or plant collection by hobbyists (Goedeke et al., 2008).

Conservation status and measures

Platanthera praeclara is currently listed in the USA as 'Threatened' by the U.S.

Fish and Wildlife Service under the Act of 1973 in the United States and as Endangered in Canada (USFWS, 2009; Environment Canada, 2006). The focus of the recovery efforts has been on “maintaining the habitat of known populations on native prairie and providing the highest level of protection appropriate for all populations”

(USFWS, 2009). These goals will be accomplished by warranting that a minimum proportion of P. praeclara occur on lands that are being converted to non-grassland, and by ensuring that appropriate management is administrated where P. praeclara populations are established. However, many orchid populations are in private owned lands, making it challenging to manage plants properly (Goedeke et al., 2008).

Why study this taxon?

The biology and ecology of P. praeclara has been investigated in a little more detail than many other North American taxa native to the prairies due to its relatively wide distribution, its threatened status, and mostly because it is a large-flowered and a charismatic plant, which makes it a more desirable target than diminutive species that might be equally unique and rare (USFWS, 2009). Further, the species is associated with

17

Texas Tech University, Pablo Antonio Tovar, December 2015 a highly unique, endangered, and iconic ecosystem that is vanishing quickly. Despite the body of knowledge available for changes in habitat (Collicut, 1993; Fauske and Rider,

1996; Alexander, 2001; Self and Anthonisen, 2005; USFWS, 2009), reproduction and pollinators (Cuthrell, 1994; Westwood and Borkowsky, 2004; Jordan et al., 2006;

Borkowsky and Westwood, 2009; Fox et al., 2013), germination (Sharma et al., 2003a;

Sharma et al., 2003b), and assessment of mycorrhizae from parts of its range (Zelmer and

Currah, 1995; Zelmer et al., 1996; Sharma et al., 2003a), there is an overall decline in populations, mostly because of habitat loss and / or degradation.

Some of the challenges for conservation of this species include a lack of understanding of its microhabitat requirements and how those conditions shift over space and time. Sieg and King (1995) highlighted changes in soil moisture in the landscape across different land management treatments, showing that soils subjected to long season grazing tend to have lower percentages of soil moisture. Hydrological changes can also influence the availability and community composition of mycorrhizae (Kivlin et al.,

2011). Little or no data are available for range-wide temporal and spatial variation in mycorrhizal associations and / or microhabitat conditions that support the growth of P. praeclara. Considering the critical involvement of fungi in germination, and a life-long association with adult plants, knowledge of distribution, diversity, and shifts in mycorrhizal associations is a necessary component toward understanding the ecological constraints on the persistence of the individuals and populations of any orchid taxon.

Changes in the soil microflora can have implications on the composition and function of the prairie communities and / or on the distribution, abundance, and

18

Texas Tech University, Pablo Antonio Tovar, December 2015 persistence of individual species (Eom et al., 1999). Obligate fungal partners of plants are implicated in influencing vegetative community composition via positive or negative feedback mechanisms (Bertness & Callaway, 1994). These effects are often compounded when the plant-fungus relationships are highly exclusive, as can be the case for many orchid species and their mycorrhizal fungi. Shifts in fungal communities in tall grass prairies have been linked to changes in soil variables due to different types of land management (Murty et al., 2002). As plant community composition is immediately affected by land management, microbial communities are ultimately affected, too (Bossio et al., 1998; Johnson et al., 2003; Wu et al., 2015). Studies have described that shifts in the structure of fungal communities are associated with changes in soil properties like texture (Girvan et al., 2003), soil pH (Fierer and Jackson, 2006) and soil nitrogen and phosphorus availability (Frey et al., 2004; Lauber et al., 2008). Although these variables change depending on the land use, it is not clear if and how they affect biogeographical patterns in soil microbial communities (Lauber et al., 2008) at macro or micro spatial scales.

An assessment of both biotic (fungal associates) and abiotic (physical and chemical soil characteristics) components of the microhabitat of P. praeclara is thus warranted to assist with identifying its ecological niches and for optimizing conservation decisions for the taxon. This study was undertaken to include the objectives listed below:

19

Texas Tech University, Pablo Antonio Tovar, December 2015

Objectives of the study

1) To assess the variation in the distribution and diversity (species richness and

abundance) of mycorrhizal fungal communities within the roots of Platanthera

praeclara in response to space, time, and edaphic characteristics.

It was expected that P. praeclara will exhibit high specificity towards its fungal associates across space, time, and edaphic characteristics despite its wide distribution, given the rarity of the orchid and the previously reported literature on mycorrhizal associations within the taxon.

2) To assess the effect of four experimental land management treatments (Spring burn

every four years, Fall burn every four years, Annual fall haying, and no treatment) on

the temporal variation in the diversity of mycorrhizal fungal communities within the

roots of Platanthera praeclara.

The diversity of fungi associated with P. praeclara was expected to be similar across prairie management treatments and over the two years because of the expectation of overall high mycorrhizal specificity in the species.

Significance of the study

The data from this study will quantify whether and how the distribution and diversity of mycorrhizal fungi associated with Platanthera praeclara shifts spatially and

20

Texas Tech University, Pablo Antonio Tovar, December 2015 temporally. Majority of orchid mycorrhizal assessments are conducted and reported by assuming that relatively few plant roots sampled from few locations at any one time are reliable measures of diversity and specificity in orchid-fungal associations. Range-wide, multi-year investigations by using robust statistical approaches are rare, and only a few other orchid species have been studied in sufficient detail to develop generalized theories about temporal or spatial stasis in orchid-mycorrhizal associations. While spatial shifts are more easily expected, mycorrhizal fungi of orchids are generally assumed to be static over temporal scales. An assessment of the variability in orchid-fungal associations in response to edaphic characteristics are less common still.

The studies described here were designed to reveal both the spatial and temporal variation in orchid mycorrhizal symbiosis by using an emblematic and rare North

American orchid taxon native to an equally threatened ecosystem that is the North

American tallgrass prairie. Along with explaining the dynamics of the orchid mycorrhizal relationships in response to space, time, and soil edaphic variables, the studies described herein will assist conservation biologists with management decisions. Knowledge of the identity and diversity of fungi associated with the orchid under specific management treatments and how they change over time may be utilized to assist in habitat preservation and management programs.

21

Texas Tech University, Pablo Antonio Tovar, December 2015

Literature Cited

Alexander, B.W. (2001). An inventory of the Western Prairie Fringed Orchid (Platanthera praeclara) in the Sheyenne National Grassland, Ransom County, North Dakota in 2001. Report to the North Dakota Parks and Recreation Department and the North Dakota Natural Heritage Inventory program.

Barrett, C.F., Freudenstein, J.V., Taylor, D.L., & Kõljalg, U. (2010). Rangewide analysis of fungal associations in the fully mycoheterotrophic Corallorhiza striata complex (Orchidaceae) reveals extreme specificity on ectomycorrhizal Tomentella (Thelephoraceae) across North America. American Journal of Botany, 97(4), 628-643

Bertness, M.D., & Callaway, R. (1994). Positive interactions in communities. Trends in Ecology & Evolution, 9(5), 191-193.

Bidartondo, M. I., Bruns, T. D., Weiß, M., Sérgio, C., & Read, D. J. (2003). Specialized cheating of the ectomycorrhizal symbiosis by an epiparasitic liverwort. Proceedings of the Royal Society of London B: Biological Sciences, 270(1517), 835-842.

Bonnardeaux, Y., Brundrett, M., Batty, A., Dixon, K., Koch, J., & Sivasithamparam, K. (2007). Diversity of mycorrhizal fungi of terrestrial orchids: compatibility webs, brief encounters, lasting relationships and alien invasions. Mycological research, 111(1), 51-61.

Bond, W.J. (1994). Do mutualisms matter? Assessing the impact of pollinator and disperser disruption on plant extinction. Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences, 344(1307), 83-90.

Borkowsky, C., & Westwood, A.R. (2009). Seed capsule production in the endangered western prairie fringed orchid (Platanthera praeclara) in relation to sphinx moth (Lepidoptera: ) activity. J. Lepid. Soc, 63, 110-117.

Bossio, D.A., Scow, K.M., Gunapala, N., & Graham, K.J. (1998). Determinants of soil microbial communities: effects of agricultural management, season, and soil type on phospholipid fatty acid profiles.Microbial ecology, 36(1), 1-12.

Bowles, M.L., & Duxbury, A. (1986). Report on the status of Platanthera praeclara Sheviak and Bowles in Oklahoma, Kansas, Nebraska, South Dakota, and North Dakota. Submitted to US Fish and Wildlife Service, Denver, Colorado.

22

Texas Tech University, Pablo Antonio Tovar, December 2015

Brownell, V.R. (1984). Status report on the prairie white-fringed orchid : a rare species in Canada. Unpublished report to the Canadian government.

Bunch, W.D., Cowden, C.C., Wurzburger, N., & Shefferson, R.P. (2013). Geography and soil chemistry drive the distribution of fungal associations in lady’s slipper orchid, Cypripedium acaule. Botany, 91(12), 850-856.

Chase, M.W., Cameron, K. M., Barrett, R.L., & Freudenstein, J.V. (2003). DNA data and Orchidaceae systematics: a new phylogenetic classification.Orchid conservation, 69, 89.

Collicutt, D. (1992). Status of the western prairie fringed orchid (Platanthera praeclara) in Canada. Unpublished report to Committee on the Status of Endangered Wildlife in Canada. December 1992. Ottawa

Correll, D.S. (1978). Native orchids of North America north of Mexico. Stanford University Press.

Cozzolino, S., & Widmer, A. (2005). Orchid diversity: an evolutionary consequence of deception? Trends in Ecology & Evolution, 20(9), 487-494.

Currah, R.S., & Zelmer, C. (1992). A key and notes for the genera of fungi mycorrhizal with orchids and a new species in the genus Epulorhiza. Reports of the Tottori Mycological Institute 30, 43-59.

Cuthrell, D.L. (1994). Insects associated with the prairie fringed orchids, Platanthera praeclara Sheviak & Bowles and P. leucophaea (Nuttall) Lindley. MSc Thesis, North Dakota State University, Fargo, ND, USA.

Dafni, A., & Bernhardt, P. (1990). Pollination of terrestrial orchids of Southern Australia and the Mediterranean region: systematic, ecological, and evolutionary implications. Evolutionary biology, Plenum Press 193-252

Davis, S.K. (1995). National recovery plan for the western prairie fringed orchid, Platanthera praeclara. Prepared for the Endangered Plant and Invertebrates of Canada Project, Canadian Nature Federation. Ottawa. 23 pp.

Dearnaley, J.D. (2007). Further advances in orchid mycorrhizal research. Mycorrhiza, 17(6), 475-486.

Dearnaley, J.D. W., Martos, F., & Selosse, M.A. (2012). 12 Orchid Mycorrhizas: Molecular Ecology, Physiology, Evolution and Conservation Aspects. In Fungal associations (pp. 20 207-230). Springer Berlin Heidelberg.

23

Texas Tech University, Pablo Antonio Tovar, December 2015 de Candolle, A. P. (1816). Mémoire sur les Rhizoctones: nouveau genre de champignon qui attaque les racines des plantes, et en particulier celle de la luzerne cultivée. G. Dufour.

Diez, J.M., & Pulliam, H.R. (2007). Hierarchical analysis of species distributions and abundance across environmental gradients. Ecology, 88(12), 3144-3152.

Dressler, R.L. (1981). The orchids: natural history and classification. Harvard University Press, Cambridge, Massachusetts, USA

Dressler, R.L. (1993). Phylogeny and classification of the orchid family. Cambridge University Press.

Environment Canada. (2006). Recovery Strategy for the Western Prairie Fringed Orchid (Platanthera praeclara) in Canada. Species at Risk Act Recovery Strategy Series. Environment Canada, Ottawa.

Eom, A.H., Hartnett, D.C., Wilson, G.W., & Figge, D.A. (1999). The effect of fire, mowing and fertilizer amendment on arbuscular mycorrhizas in tallgrass prairie. The American Midland Naturalist, 142(1), 55-70.

Fauske, G.M. and Rider, D.A. (1996). Pollination, pollinators and herbivores of the Western Prairie Fringed Orchid. Report submitted to North Dakota Natural heritage Program, North Dakota Parks and Recreation Department.

Fierer, N., & Jackson, R.B. (2006). The diversity and biogeography of soil bacterial communities. Proceedings of the National Academy of Sciences of the United States of America, 103(3), 626-631.

Figueroa, C., Salazar, G.A., Zavaleta, H.A., & Engleman, E.M. (2008). Root character evolution and systematics in Cranichidinae, Prescottiinae and Spiranthinae (Orchidaceae, Cranichideae). Annals of Botany, 101(4), 509-520.

Fox, K., Vitt, P., Anderson, K., Fauske, G., Travers, S., Vik, D., & Harris, M.O. (2013). Pollination of a threatened orchid by an introduced hawk moth species in the tallgrass prairie of North America. Biological Conservation, 167, 316-324.

Freeman, C.C., and R.E. Brooks. (1989). Status report on Platanthera praeclara Sheviak and Bowles (western prairie fringed orchid) in Kansas, Nebraska, and South Dakota. Unpublished report to U.S. Fish and Wildlife Service, Denver, Colorado.

Frey, S.D., Knorr, M., Parrent, J.L., & Simpson, R.T. (2004). Chronic nitrogen enrichment affects the structure and function of the soil microbial community in

24

Texas Tech University, Pablo Antonio Tovar, December 2015

temperate hardwood and pine forests. Forest Ecology and Management, 196(1), 159-171.

Gentry, A.H., & Dodson, C.H. (1987). Diversity and biogeography of Neotropical vascular epiphytes. Annals of the Missouri Botanical Garden, 205-233.

Ghimire, S.R., Charlton, N.D., Bell, J.D., Krishnamurthy, Y.L., & Craven, K.D. (2011). Biodiversity of fungal endophyte communities inhabiting switchgrass (Panicum virgatum L.) growing in the native tallgrass prairie of northern Oklahoma. Fungal Diversity, 47(1), 19-27.

Girlanda, M., Segreto, R., Cafasso, D., Liebel, H.T., Rodda, M., Ercole, E., & Perotto, S. (2011). Photosynthetic Mediterranean meadow orchids feature partial mycoheterotrophy and specific mycorrhizal associations. American Journal of Botany, 98(7), 1148-1163.

Girvan, M.S., Bullimore, J., Pretty, J.N., Osborn, A.M., & Ball, A.S. (2003). Soil type is the primary determinant of the composition of the total and active bacterial communities in arable soils. Applied and Environmental Microbiology, 69(3), 1800-1809.

Goedeke, T., Sharma, J., Delphey, P. & Marshall Mattson, K. (2008). Platanthera praeclara. In: IUCN 2013. IUCN Red List of Threatened Species. Version 2013.2. .

Gravendeel, B., Smithson, A., Slik, F.J., & Schuiteman, A. (2004). Epiphytism and pollinator specialization: drivers for orchid diversity?.Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences, 359(1450), 1523- 1535.

Gustafsson, A.L.S., Verola, C.F., & Antonelli, A. (2010). Reassessing the temporal evolution of orchids with new fossils and a Bayesian relaxed clock, with implications for the diversification of the rare South American genus Hoffmannseggella (Orchidaceae: Epidendroideae). BMC evolutionary biology, 10(1), 177.

Hapeman, J.R., & Inoue, K. (1997). Plant-pollinator interactions and floral radiation in Platanthera (Orchidaceae). Molecular evolution and adaptive radiation, 433-54. Cambridge, UK: Cam- bridge Univ. Press Ibisch, P.L., Boegner, A., Nieder, J., & Barthlott, W. (1996). How diverse are Neotropical epiphytes? An analysis based on the ‘Catalogue of the flowering plants and gymnosperms of Peru’. Ecotropica, 2(1), 13-28.

25

Texas Tech University, Pablo Antonio Tovar, December 2015

Inoue, K. (1983). Systematics of the genus Platanthera (Orchidaceae) in Japan and adjacent regions with special reference to pollination. Journal of the faculty of science, University of Tokyo: Section III: Botany 13: 285-374.

Jacquemyn, H., Waud, M., Merckx, V.S., Lievens, B., & Brys, R. (2015). Mycorrhizal diversity, seed germination and long‐term changes in population size across nine populations of the terrestrial orchid Neottia ovata. Molecular ecology, 24, 3269- 3280.

Jang, C.S., Kamps, T.L., Skinner, D. N., Schulze, S.R., Vencill, W.K., & Paterson, A.H. (2006). Functional classification, genomic organization, putatively cis-acting regulatory elements, and relationship to quantitative trait loci, of sorghum genes with rhizome-enriched expression. Plant physiology,142(3), 1148-1159.

Johnson, M.J., Lee, K.Y., & Scow, K.M. (2003). DNA fingerprinting reveals links among agricultural crops, soil properties, and the composition of soil microbial communities. Geoderma, 114(3), 279-303.

Jones, S.R. & R.C. Cushman. (2004). Peterson Field Guides: A Field Guide to the North American Prairie. Houghton Mifflin Company. Boston, U.S.

Jordan, C.R., Fauske, G.M., Harris, M.O., & Lenz, D. (2006). First record of the spurge hawkmoth as a pollen vector for the western prairie fringed orchid.Prairie Naturalist, 38(1), 57.

Kennedy, A.H., Taylor, D.L., & Watson, L.E. (2011). Mycorrhizal specificity in the fully mycoheterotrophic Hexalectris Raf. (Orchidaceae: Epidendroideae).Molecular Ecology, 20(6), 1303-1316.

Kiers, E.T., Duhamel, M., Beesetty, Y., Mensah, J.A., Franken, O., Verbruggen, E., & Bücking, H. (2011). Reciprocal rewards stabilize cooperation in the mycorrhizal symbiosis. Science, 333(6044), 880-882.

Kirby, D.R., Lym, R. G., Sterling, J.J., & Sieg, C.H. (2003). Observation: leafy spurge control in western prairie fringed orchid habitat. Journal of range management, 466-473.

Kivlin, S.N., Hawkes, C.V., & Treseder, K. K. (2011). Global diversity and distribution of arbuscular mycorrhizal fungi. Soil Biology and Biochemistry, 43(11), 2294- 2303.

Lauber, C.L., Strickland, M.S., Bradford, M.A., & Fierer, N. (2008). The influence of soil properties on the structure of bacterial and fungal communities across land-use types. Soil Biology and Biochemistry, 40(9), 2407-2415.

26

Texas Tech University, Pablo Antonio Tovar, December 2015

Mabberley, D.J. (1997). The plant-book. 2nd edn. Cambridge University Press, Cambridge

Martén‐Rodríguez, S., Fenster, C.B., Agnarsson, I., Skog, L.E., & Zimmer, E.A. (2010). Evolutionary breakdown of pollination specialization in a Caribbean plant radiation. New Phytologist, 188(2), 403-417.

McCormick, M. K., & Jacquemyn, H. (2014). What constrains the distribution of orchid populations?. New Phytologist, 202(2), 392-400.

Medina, R. D., Flachsland, E. A., Gonzalez, A. M., Terada, G., Faloci, M. M., & Mroginski, L. A. (2009). In vitro tuberization and plant regeneration from multinodal segment culture of Habenaria bractescens Lindl., an Argentinean wetland orchid. Plant Cell, Tissue and Organ Culture (PCTOC), 97(1), 91-101.

Murty, D., Kirschbaum, M.U., Mcmurtrie, R.E., & Mcgilvray, H. (2002). Does conversion of forest to agricultural land change soil carbon and nitrogen? A review of the literature. Global Change Biology, 8(2), 105-123.

Myers, N., Mittermeier, R. A., Mittermeier, C. G., Da Fonseca, G. A., & Kent, J. (2000). Biodiversity hotspots for conservation priorities. Nature, 403(6772), 853-858.

Nagel, H.G. and O.A. Kolstad. (1987). Comparison of plant species composition of Mormon Island, Crane Meadows, and Lilian Annette Row Sanctuary in central Nebraska. Transactions of the Nebraska Academy of Science 15:37-48.

Otero, J.T., Ackerman, J.D., & Bayman, P. (2002). Diversity and host specificity of endophytic Rhizoctonia-like fungi from tropical orchids. American Journal of Botany, 89(11), 1852-1858.

Otero, J.T., Ackerman, J.D., & Bayman, P. (2004). Differences in mycorrhizal preferences between two tropical orchids. Molecular Ecology, 13(8), 2393-2404.

Otero, J.T, & Flanagan, N.S. (2006). Orchid diversity–beyond deception. Trends in Ecology & Evolution, 21(2), 64-65.

Pandey, M., Sharma, J., Taylor, D., & Yadon, V.L. (2013). A narrowly endemic photosynthetic orchid is non‐specific in its mycorrhizal associations. Molecular ecology, 22(8), 2341-2354.

Peterson, R.L., Massicotte, H. B., & Melville, L. H. (2004). Mycorrhizas: anatomy and cell biology. NRC Research, Canada

27

Texas Tech University, Pablo Antonio Tovar, December 2015

Phillips, L. (2003). Pollination of Western Prairie Fringed Orchid, Platanthera praeclara Sheviak & Bowles: implications for restoration and management. Restoration and Reclamation Review. Student On-Line Journal, Department of Horticultural Science, University of Minnesota.

Pleasants, J.M. (1993). Pollination ecology of the western prairie fringed orchid, Platanthera praeclara. North Dakota Dept. Parks and Tourism, Bismarck, ND.

Pleasants, J.M. & Moe. S. 1993. Floral display size and pollination of the western prairie fringed orchid, Platanthera praeclara (Orchidaceae). Lindleyana 8: 32–38.

Ramírez, S.R., Gravendeel, B., Singer, R.B., Marshall, C.R., & Pierce, N.E. (2007). Dating the origin of the Orchidaceae from a fossil orchid with its pollinator. Nature, 448(7157), 1042-1045.

Rasmussen, H.N. (1995). Terrestrial orchids: from seed to mycotrophic plant. Cambridge University Press.

Roberts, P. (1999). Rhizoctonia-forming fungi. Herbarium, Royal Botanic Gardens.

Self, R. and Anthonisen, E. (2005). Western Prairie Fringed Orchid monitoring study project report for 2005. Nature Conservancy and U.S Forest Service.

Selosse, M.A., WEIß, M., Jany, J.L., & Tillier, A. (2002). Communities and populations of sebacinoid basidiomycetes associated with the achlorophyllous orchid Neottia nidus‐avis (L.) LCM Rich. and neighbouring tree ectomycorrhizae. Molecular Ecology, 11(9), 1831-1844.

Selosse, M. A., Dubois, M. P., & Alvarez, N. (2009). Do Sebacinales commonly associate with plant roots as endophytes?. Mycological research,113(10), 1062- 1069.

Sharma, J., Van Sambeek, J.W. and Starbuck, C.J. (2002). Western Prairie Fringed Orchid: its status, ecology, and in vitro propagation. In: S. Fore (ed.), Proceedings of the Eighteenth Annual North American Prairie Conference: Promoting Prairie: 189-193.

Sharma, J., Zettler, L.W., & Van Sambeek, J.W. (2003a). A Survey of Mycobionts of Federally Threatened Platanthera praeclara (Orchidaceae). Symbiosis, 34, 145- 155

Sharma, J., Zettler, L.W., Van Sambeek, J.W., Ellersieck, M.R., & Starbuck, C.J. (2003b). Symbiotic seed germination and mycorrhizae of federally threatened

28

Texas Tech University, Pablo Antonio Tovar, December 2015

Platanthera praeclara (Orchidaceae). The American midland naturalist, 149(1), 104-120.

Sheviak, C.J., & Bowles, M.L. (1986). The prairie fringed orchids: a pollinator-isolated species pair. Rhodora, 88(854), 267-290.

Sheviak, C.J., and Catling, P.M. (2002). Platanthera Richard. In: Flora of North America Editorial Committee. Eds. Flora of North America North of Mexico 26. Oxford: Oxford University Press, 551 – 571

Sieg, C.H., & King, R.M. (1995). Influence of environmental factors and preliminary demographic analyses of a threatened orchid, Platanthera praeclara. American Midland Naturalist, 134, 307-323.

Smith, S.E. (1967). Carbohydrate translocation in orchid mycorrhizas. New Phytologist, 66(3), 371-378.

Smith, S.E., and Read, D.J. (2008) Mycorrhizal Symbiosis, 3rd edn. Academic Press, London, UK

Swarts, N.D., & Dixon, K.W. (2009). Terrestrial orchid conservation in the age of extinction. Annals of Botany, 104(3), 543-556.

Taylor, D.L., & Bruns, T.D. (1997). Independent, specialized invasions of ectomycorrhizal mutualism by two nonphotosynthetic orchids. Proceedings of the National Academy of Sciences, 94(9), 4510-4515.

Taylor, D.L., T.D. Bruns, J. R. Leake, and D.J. Read. (2002). Mycorrhizal specificity and function in myco-heterotrophic plants. In M. van der Heijden and I. Sanders [eds.], The ecology of mycorrhizas, vol. 157, 375 – 414. Springer, Berlin, Germany.

Taylor, D.L., Bruns, T.D., Szaro, T. M., & Hodges, S.A. (2003). Divergence in mycorrhizal specialization within Hexalectris spicata (Orchidaceae), a nonphotosynthetic desert orchid. American Journal of Botany, 90(8), 1168-1179.

Tremblay, R.L. (1992). Trends in the pollination ecology of the Orchidaceae: evolution and systematics. Canadian Journal of Botany, 70(3), 642-650.

Tremblay, R.L., Ackerman, J.D., Zimmerman, J.K., & Calvo, R.N. (2005). Variation in sexual reproduction in orchids and its evolutionary consequences: a spasmodic journey to diversification. Biological Journal of the Linnean Society, 84(1), 1-54.

29

Texas Tech University, Pablo Antonio Tovar, December 2015

U.S. Department of Agriculture. (1999). Soil : A basic system of soil classification for making and interpreting soil surveys. Second edition. Agriculture Handbook No. 436. U.S. Soil Conservation Service, Washington, D.C.

U.S. Fish and Wildlife Service. (1996). Platanthera praeclara (western prairie fringed orchid) recovery plan. Ft. Snelling, MN.

U.S. Fish and Wildlife Service. (2009). Platanthera praeclara (western prairie fringed orchid) recovery plan. 5 year review: Summary and Evaluation. Ft. Snelling, MN.

Van der Pijl, L. and C.H. Dodson. (1966). Orchid flowers: their pollination and evolution. University of Miami Press, Coral Gables, Florida, USA.

Vásquez, R., Ibisch, P.L., & Gerkmann, B. (2003). Diversity of Bolivian Orchidaceae–a challenge for taxonomic, floristic and conservation research. Organisms Diversity & Evolution, 3(2), 93-102.

Weiβ, M., Sýkorová, Z., Garnica, S., Riess, K., Martos, F., Krause, C., Oberwinkler, F., Bauer, R. & Redecker, D. (2011). Sebacinales everywhere: previously overlooked ubiquitous fungal endophytes. PLoS One 6(2):e16793.

Westwood, A.R., & Borkowsky, C.L. (2004). Sphinx moth pollinators for the endangered western prairie fringed orchid, Platanthera praeclara in Manitoba, Canada. Journal-lepidopterists Society, 58(1), 13-20.

Williams, J.G., & Williams, A.E. (1983). Field guide to orchids of North America. Universe Books, New York, 14(3), 24-53.

Wolf, J.H., & Alejandro, F.S. (2003). Patterns in species richness and distribution of vascular epiphytes in Chiapas, Mexico. Journal of Biogeography, 30(11), 1689- 1707.

Wolken, P.M. (1995). Habitat and life history of the western prairie fringed orchid (Platanthera praeclara). M.S. thesis. University of Wyoming, Laramie.

Wolken, P.M., Sieg, C.H., & Williams, S. E. (2001). Quantifying suitable habitat of the threatened western prairie fringed orchid. Journal of Range Management, 611- 616.

Wu, T., Milner, H., Díaz-Pérez, J.C., & Ji, P. (2015). Effects of soil management practices on soil microbial communities and development of southern blight in vegetable production. Applied Soil Ecology, 91, 58-67.

30

Texas Tech University, Pablo Antonio Tovar, December 2015

Yagame, T., Orihara, T., Selosse, M. A., Yamato, M., & Iwase, K. (2012). Mixotrophy of Platanthera minor, an orchid associated with ectomycorrhiza‐forming Ceratobasidiaceae fungi. New Phytologist, 193(1), 178-187.

Yuan, L., Yang, Z. L., Li, S. Y., Hu, H., & Huang, J. L. (2010). Mycorrhizal specificity, preference, and plasticity of six slipper orchids from South Western China. Mycorrhiza, 20(8), 559-568.

Zachos, J.C., Dickens G.R., Zeebe R.E. (2008). An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature, 451(7176):279-283.

Zelmer, C.D., & Currah, R.S. (1995). Ceratorhiza pernacatena and Epulorhiza calendulina spp. nov.: mycorrhizal fungi of terrestrial orchids. Canadian Journal of Botany, 73(12), 1981-1985.

Zelmer, C.D., Cuthbertson, L., & Currah, R.S. (1996). Fungi associated with terrestrial orchid mycorrhizas, seeds and protocorms. Mycoscience, 37(4), 439-448.

Zettler, L.W., & Piskin, K.A. (2011). Mycorrhizal fungi from protocorms, seedlings and mature plants of the eastern prairie fringed orchid, Platanthera leucophaea (Nutt.) Lindley: A comprehensive list to augment conservation. The American Midland Naturalist, 166(1), 29-39.

Zotz, G. (2013). The systematic distribution of vascular epiphytes–a critical update. Botanical Journal of the Linnean Society, 171(3), 453-481.

31

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 1.1. An inflorescence of Platanthera praeclara. Photo by Jyotsna Sharma

32

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 1.2. A partial map of the US and Canada showing the natural distribution of Platanthera praeclara. The states where the plant is currently known to be present (orange), the counties where the populations occur (gray) and the counties where roots were collected for the studies (yellow) are highlighted.

33

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 1.3. A transverse section of a root of Platanthera praeclara showing colonization by fungal pelotons (examples indicated by arrows). The scale-bar represents 200 microns.

34

Texas Tech University, Pablo Antonio Tovar, December 2015

CHAPTER II

MYCORRHIZAL ASSOCIATIONS OF PLATANTHERA PRAECLARA ACROSS ITS NATURAL DISTRIBUTION

Abstract

Platanthera praeclara is a rare endemic orchid native to parts of the North

American tallgrass prairie. As every other orchid species, P. praeclara survival relies on mycorrhizal associations at early developmental stages. Yet, the associations remain through the rest of the orchid’s life cycle. This study assessed the mycorrhizal diversity from several sites to describe mycorrhizal communities where P. praeclara occurs naturally across its natural distribution. Moreover, mycorrhizal assessments were conducted in two consecutive years to observe specificity and variation through time.

Types of land management that are usually applied to the sites where these orchids naturally occur were taken into account, in addition to edaphic characteristics, in an attempt to explain variation in the orchid mycorrhizal communities that were observed.

Mycorrhizal associations of Platanthera praeclara were studied in plants at 11 sites across its natural distribution in two consecutive years, 2013 and 2014. Across the two years, 39 Operational Taxonomic Units (OTUs) in two fungal families, Ceratobasidiaceae and Tulasnellaceae, were observed in roots of P. praeclara. Ceratobasidiaceae was the dominant fungal family at majority of the sampled sites, and it was represented by 77% of the analyzed nuclear ribosomal internal transcribed spacer sequences. Short pairwise sequence distances (π = 0.039 ± 0.03, N=238) between the sequences of

Ceratobasidiaceae and phylogenetic clustering of the OTUs suggested that P. praeclara 35

Texas Tech University, Pablo Antonio Tovar, December 2015 forms specific associations toward fungi from Ceratobasidiaceae. Fungi from the fungal family Tulasnellaceae (π = 0.135 ± 0.06, N=69) represented a broader phylogenetic breadth, with the orchid being less specific in its associations with the OTUs from this family. Despite the observed specificity, fungal communities in roots tended to vary

(P<0.05) among a large majority of the sites. Changes were observed in mycorrhizal fungal community composition from one year to the next. Clustering of sites with similar land management treatments was not detected. Nonmetric multidimensional scaling analysis did not segregate sites by land management; however, fungal communities appeared to be more similar in northern sites when compared to the fungal communities associated with the species in southern sites. Correlations between the observed fungal communities and edaphic characteristics were not evident.

36

Texas Tech University, Pablo Antonio Tovar, December 2015

Introduction

Mycorrhizal fungi in the plant family Orchidaceae are characterized by intracellular coils of hyphae termed pelotons (Smith and Read, 2008). A large majority of the known orchid mycorrhizae are formed with fungi from the fungal phylum Basidiomycota

(Rasmussen, 1995; Taylor et al., 2002; Taylor and McCormick, 2008; McCormick and

Jacquemyn, 2014). Orchid mycobionts may be identified by isolating the fungi from the roots of the orchids and culturing them (Hadley and Williamson, 1972; Currah et al.,

1987; Currah et al., 1990; Zettler et al., 2001; Sharma et al., 2003b). However, identifying cultured isolates of fungi that associate specifically with orchids presents a number of challenges. Complex fruit bodies are absent or rarely seen, vegetative hyphal morphologies are mostly homogenous within genera (Taylor and McCormick, 2008), and many characters overlap between species, or vary because of the environment or changes during development stages within individuals (Andersen, 1990). Additionally, fungal isolation success may vary depending on the season, capacity of fungi to respond to culture conditions, or disturbance of plant roots previous to isolation (Ramsay et al.,

1986).

A more comprehensive method for characterizing the species richness and abundances of mycorrhizal fungi from plant roots is by amplifying specific DNA barcode regions by performing a polymerase chain reaction (Gardes and Bruns, 1993). Often, fungal DNA needs to be specifically amplified from mix of plant and fungal DNA, which requires targeting barcode regions that are specific to the organism of interest. 37

Texas Tech University, Pablo Antonio Tovar, December 2015

In fungi, the nuclear ribosomal internal transcribed spacer (nrITS) region (between

700 and 800 nucleotides long) contains two variable non-coding regions (ITS1 and ITS2) that are nested within DNA repeats between the highly conserved small subunit (18S),

5.8S, and the large subunit (28S) rRNA genes (Figure 2.2). The internal transcribed spacer is the most widely sequenced target region for molecular identification of fungi because it can be amplified with ‘universal primers’ that are complementary to sequences within the rRNA genes (White et al., 1990). It has a multi-copy nature that makes this region relatively easy to amplify from small, diluted or degraded DNA samples (Gardes and Bruns, 1993). Further, the high variability of the nrITS region among closely related species makes it suitable to differentiate species or operational taxonomic units (OTUs)

(Gardes et al., 1991; Gardes and Bruns, 1996; Anderson and Cairney, 2004; Martin and

Rygiewicz, 2005; Taylor and McCormick, 2008; Seifert K.A., 2009; Schoch et al., 2012;

Oja et al., 2015).

Polymerase chain reaction amplification followed by Sanger sequencing of the nrITS has improved identification and characterization of different types of mycorrhizae

(Gardes et al., 1991; Gardes and Bruns, 1993; Redecker, 2000). In 2008, Taylor and

McCormick published primers that were developed specifically to amplify the nrITS region of fungi known to commonly associate with orchids. While these primers have helped to clarify distributions of orchid mycorrhizae in natural environments, their exclusive use was not implied. Capturing the complete profiles of communities of peloton forming fungi in orchid roots can in fact require a multi-primer approach or use of cloning techniques. With the advent of multiplex sequencing tools such as Next

38

Texas Tech University, Pablo Antonio Tovar, December 2015

Generation Sequencing (NGS), identification of fungi in roots is technically possible to generate large amounts of sequences from individual samples, however separating the multitudes of endophytes from peloton-forming fungi can heavily skew the interpretations with respect to specificity if generic fungal nrITS barcoding is performed.

Platanthera praeclara is an herbaceous, perennial terrestrial orchid endemic to the

North American tallgrass prairie. It is listed as Threatened by the U.S. Fish and Wildlife

Service under the Endangered Species Act of 1973 in the United States and as

Endangered in Canada (USFWS, 2009; Environment Canada, 2006). Natural populations of P. praeclara are known to occur in six mid-western states in the United States

(Kansas, Missouri, Iowa, Nebraska, Minnesota, and North Dakota) and in the province of

Manitoba, Canada (USFWS, 1996) (Figure 2.1). According to the U.S Fish and Wildlife

Service (USFWS, 2009), there were 29,140 individuals of P. praeclara in 2008 in the

United States. Other plants that compose vegetation communities in the North American tallgrass prairies are species historically adapted to fire and grazing like Andropogon gerardii, A. scoparius, Sorghastrum nutans, Deschampsia caespitosa and Panicum virgatum, all members of the Poaceae (USFWS, 1996; Ghimire et al., 2011). Still, there is a vast presence of exotic or invasive plants that individuals of P. praeclara cannot out- compete [e.g., Poa pratensis, Bromus inermis, Trifolium campestre, Festuca arundinaceae and Sorghum halapense (Cully et al., 2003)] and displace the orchid (Kirby et al., 2003; Biederman et al., 2014). As an obligate outcrossed pollinated species, P. praeclara relies on hawkmoth species for reproduction (Fox et al., 2013). However, low

39

Texas Tech University, Pablo Antonio Tovar, December 2015 pollination rates and seed production have been observed in most populations (Phillips,

2003).

Mycorrhizal associations in Platanthera praeclara have been studied previously to some extent. Zelmer and Currah (1995) first isolated and characterized eight isolates of

Ceratorhiza pernacatena from the family Ceratobasidiaceae by sampling roots from seven plants of P. praeclara from a single population in Manitoba, Canada. From the same orchid population, one isolate belonging to the Tulasnellaceae and fifteen

Ceratobasidiaceae isolates were subsequently observed in 11 plants of P. praeclara

(Zelmer et al., 1996). Sharma et al. (2003a) assessed mycorrhizal fungi from P. praeclara from 21 plants in its native habitat in Minnesota (Clay county, Norman county,

Polk county and Mower county) and Missouri (Harrison county). In this study, 78 isolates belonged to the fungal family Ceratobasidiaceae (e.g. Genbank accession number

DQ088771) and nine belonged to the fungal family Tulasnellaceae (e.g. Genbank accession number DQ068772). However, like majority of the other orchids, these critical associations have not been investigated range-wide or over time.

While the ecology and biology of P. praeclara has been investigated in more detail than other North American orchids because of its threatened status and because of its association with a unique and endangered ecosystem that is disappearing rapidly, there is still a lack of understanding of its microhabitat requirements and how those conditions shift over space and time. Because different prairie management treatments are utilized in areas where populations of P. praeclara occur, soil variables known to affect dynamics of soil fungal or mycorrhizal communities likely change in response (Girvan et al., 2003;

40

Texas Tech University, Pablo Antonio Tovar, December 2015

Fierer and Jackson, 2006; Frey et al., 2004; Lauber et al., 2008). Although the edaphic characteristics can vary depending on the ecosystem management, it is not clear if and how they affect biogeographical patterns in soil microbial or mycorrhizal communities

(Lauber et al., 2008) at macro or micro spatial scales.

The objective of the study reported herein was to assess the variation in the distribution and diversity (species richness and abundance) of mycorrhizal fungal communities within the roots of Platanthera praeclara in response to space, time, and edaphic characteristics. Given the rarity of the orchid and the previously reported literature on mycorrhizal associations within the taxon, high specificity towards its mycobionts was expected in the species across time, space, and edaphic characteristics despite its wide distribution.

Materials and methods Collection sites

Root tissues of Platanthera praeclara were sampled from 11 extant sites across four states in the US and one province in Manitoba, Canada. The locations in the United

States included Dinesen Prairie in Shelby County, Iowa; Valentine National Wildlife

Refuge in Cherry County, Nebraska; two different sites separated by a slope at

Bluemound State Park in Rock County, Minnesota, four long-term treatment sites with different management including prescribed fire in spring every 4 years (PS), prescribed fire in fall every 4 years (PF), annual haying in fall (PH), and no management (PC) at

Pembina Trail Preserve) in Polk County, Minnesota; and two sites within the Sheyenne 41

Texas Tech University, Pablo Antonio Tovar, December 2015

National Grasslands in Ransom County, North Dakota. In Canada, plants were sampled at the Manitoba Tallgrass Preserve in the Province of Manitoba. The largest distance between any two sites (Manitoba and Iowa) was 838 km whereas the shortest distance between any two sites (Blumound uphill and Bluemound downhill) was 3 km.

All locations and sites from where plants were sampled for this study represented some form of prairie management practice including no management, prescribed fire, grazing, or mowing, etc. While the historic information on frequency and intensity of these practices was not always precise or complete, information on recent applications during at least the past 5 years was obtained and recorded. The site in Manitoba has typically received prescribed burns every 5 to 7 years, unless there are severe droughts preventing such events or wildfires occur undesirably. For example, an unprescribed fire occurred at the study site in MB in 2012. In ND, high intensity grazing is applied to the

AA unit from May to November, however variable sections are fenced off each year to protect the individuals of P. praeclara. The Viking unit is not subjected to any kind of land management, and for that reason it was categorized as a control/no management treatment although a wildfire occurred in 2012 at this site. The two sites separated by a slope in Rock County, MN are subjected to prescribed burn periodically. In NE, prescribed grazing with light pressure is administrated during the growing season once every 5 to 6 days with rest periods of 2 to 3 months. Annual late summer haying is also administrated at this site while individually marking and protecting the reproductive plants of P. praeclara so that the capsules can mature and release seeds naturally.

Prescribed burns are administrated once every three years at Dinesen Prairie in IA.

42

Texas Tech University, Pablo Antonio Tovar, December 2015

Sample collection

In 2013 and 2014, roots from individual plants were sampled between mid-June and early July. By this time in the growing season, differentiating between seedlings, juveniles or vegetative (individuals that will not produce an inflorescence in the growing season), and reproductive individuals was possible. Roots from two seedlings, two vegetative and two reproductive plants were sampled at each of the 11 sites. Roots from one or two extra plants were sampled with permission. Seventy-three and 66 plants were sampled across the 11 sites during 2013 and 2014, respectively (Table 2.1), which is a remarkably significant sampling effort considering the status of the species. Whole plants were carefully excavated to examine the complete root system (Figure 2.3). To keep plants alive after sampling, just a few roots were extracted from the plant, and the sampled plant was replanted. Sampled plants were watered thoroughly immediately after replanting to maximize reestablishment. Because seedlings have very small root systems, the whole seedling was collected and all roots were processed for fungal identification

(Figure 2.4). Roots were stored at 4°C until they were shipped overnight to the laboratory at Texas Tech University for further processing.

In each of the two years, one aggregate soil sample was collected at each site where orchids were sampled by adding together at least three subsamples. Approximately 500g of soil were extracted from sites where the plants were present. The soil samples were stored at room temperature until processed further. Physicochemical analyses were carried out by A&L Plains Labs, Inc., Lubbock, Texas. Organic matter (OM), pH, macro 43

Texas Tech University, Pablo Antonio Tovar, December 2015

nutrients (K, Mg, Ca, Na), cation exchange capacity (CEC), NO3-N, other nutrients (S,

Zn, Mn, Fe, Cu, B) and textural components (% sand, silt and clay) were measured in each sample.

Fungal identification from root sections

Roots were rinsed under tap water to wash them free of soil and debris before photo-documentation (Figure 2.5). Roots from each individual plant were subsequently placed in individual 50 mL plastic tubes. Inside a laminar flow hood and under sterile conditions, the surface sterilization was carried out by following these steps: (1) roots were first rinsed in a 70% EtOH (ethanol) solution by shaking thoroughly for 50 seconds;

(2) roots were then rinsed with a 3% NaOCl (sodium hypochlorite) solution by shaking thoroughly for 30 seconds; (3) a 70% ethanol rinse for 50 seconds was carried out; (4) roots were finally rinsed with ultrapure, sterile water several times to remove residues of

EtOH and NaOCl. Surface sterilization treatment was modified as needed based on the thickness and the number of the roots (Pandey et al., 2013).

The epidermis of each root was gently shaved off using a sterile scalpel to remove additional microbes that might have survived the surface sterilization treatment. Peloton inspection was conducted by cutting transversal thin slices of root tissue and examining under a compound microscope (Figure 2.6). If pelotons were observed, 2.5 to 3 cm long segments adjacent to the inspected slice were collected, sliced thinly and stored

44

Texas Tech University, Pablo Antonio Tovar, December 2015 individually in 2 mL plastic tubes at -80°C until DNA was extracted (Pandey et al.,

2013).

DNA extraction was conducted using Qiagen DNeasy Plant Mini Kit (Qiagen,

Valencia, CA, USA) by following the manufacturer's protocol with a slight modification.

All samples were incubated in a 3.3% solution of polyvinylpyrrolidone in AP1 lysis buffer at 65°C for 2 hours while mixing once every 30 minutes before the RNase digestion step. Subsequently, 4 μL of RNase (supplied with the kit) was added to each sample and samples were incubated at 65°C for 15 minutes, mixing once every 4 minutes. Total DNA was eluted in 50 μL of elution buffer and stored at -80°C until further processing.

PCR and sequencing

Polymerase chain reaction was performed using total DNA extracted from individual root segments. Each reaction was prepared to a final volume of 25 μL by using

Promega GoTaq Flexi DNA Polymerase reagent kit (Promega, Madison, Wisconsin,

USA). Concentrations for each reagent were: 5x Green GoTaq Flexi Buffer, 10 mM/μL dNTPs (100 mM of each dNTP), 25 mM of MgCl2, 10 μM/μL of each primer, 10 μg/μL of BSA and 5 u/μL GoTaq DNA polymerase (Pandey et al., 2013). PCR consisted of 35 cycles in an epGradients Master Cycler (Eppendorf, Hamburg, Germany) and included a

2 minute initial denaturation at 94°C before thermocycling with a 45 seconds denaturation at 94°C followed by a 45-second annealing at different temperatures

45

Texas Tech University, Pablo Antonio Tovar, December 2015 depending on the primer pair (ITS1-OF and ITS4-OF; ITS4-TUL and ITS1) used (Table

2.4 in Chapter II) and 72°C elongation for 1 minute. Finally, the last cycle was followed by an extension at 72°C for 5 minutes.

Two primer pairs ITS1-OF/ITS4-OF and ITS1/ITS4-TUL (Taylor & McCormick,

2008) were used. A total of 498 root segments were assayed for molecular analysis. Each

DNA sample was first subjected to PCR using the ITS-OF primer pair at 58°C annealing temperature. Samples that did not amplify were then run with the same primer pair at

52°C annealing temperature. If amplification was still not obtained, ITS1/ITS4-TUL was used with 54°C annealing temperature on the samples that did not amplify previously. A

2% agarose gel electrophoresis was run to verify amplification. Samples showing a single band of the expected size range (600 – 800 bp) were cleaned using DNA Clean and

Concentrator 5 kit (Zymo Research, Irvine, CA, USA). DNA quantity and quality were measured using a NanoDrop 2000c spectrophotometer (Thermo Scientific, Wilmington,

DE, USA). Samples showing multiple bands were processed using GenElute Gel

Extraction Kit (Sigma-Aldrich, Poole, United Kingdom) by isolating the band closer to the expected size range. Sanger sequencing of all cleaned PCR products was carried out by the DNA Analysis Facility on Science Hill at Yale University. Samples were sequenced in one direction using the reverse primer (ITS4-OF or ITS4-TUL) (Pandey et al., 2013).

46

Texas Tech University, Pablo Antonio Tovar, December 2015

Cloning

Polymerase chain reaction products from samples with sequencing difficulties were purified either using PureLink PCR Purification kit (Invitrogen Life Technologies,

Carlsbad, California) or GenElute Gel Extraction Kit (Sigma-Aldrich, Poole, United

Kingdom). Once a clean PCR product was obtained, it was used to carry out a cloning reaction and cloning transformation using the TOPO TA Cloning kit for Sequencing along with One Shot TOP10 and DH5α-T1 competent cells (Invitrogen Life technologies,

Carlsbad, California), following the manufacturer protocol.

To analyze the transformants, 3-5 colonies were selected and cultured overnight in

1 mL of LB (Luria-Bertani) (ThermoFisher Scientific, Waltham, MA, USA) medium containing 50 μg/mL of ampicillin. Plasmid DNA was isolated using the PureLink Quick

Plasmid Miniprep Kit, (Invitrogen Life Technologies, Carlsbad, CA) following the protocol included in the kit.

Polymerase chain reaction was performed on the extracted plasmid DNA following the amplification protocol described earlier using ITS1-OF/ITS4-OF primer pair.

Polymerase chain reaction products were purified using GenElute Gel Extraction Kit

(Sigma-Aldrich, Poole, United Kingdom) and DNA concentrations were measured using a NanoDrop 2000c spectrophotometer (Thermo Scientific, Wilmington, DE, USA).

Sanger sequencing of all cleaned PCR products was carried out as described above.

Data analyses 47

Texas Tech University, Pablo Antonio Tovar, December 2015

DNA Analysis Facility on Science Hill at Yale University performed a basic trimming on the sequences. However, all sequences were double-checked and a more thorough trim was implemented using Geneious 4.8.5. Trimmed sequences were then identified through BLAST searches using the Megablast option for finding highly similar

DNA sequences.

OTU clustering

Sequences were grouped into operational taxonomic units (OTUs) at 97% sequence similarity criterion using the OTU pipeline from the University of Alaska at Fairbanks

Life Science Informatics Portal (http://www.borealfungi.uaf.edu; Taylor & Houston

2011). The longest and highest quality representative sequence from each OTU was used for further analyses. Operational taxonomic Units were obtained separately for sequences obtained from root fragments collected in year 2013 and 2014. Additionally, for general diversity estimation, sequences were also clustered into OTUs by combining the data sets for both years.

Mean pairwise sequence distance (pi, π)

To estimate sequence divergence within the identified fungal families, separate multiple alignments of sequences were constructed for years 2013 and 2014 data sets, and for the combined data set, using MAFFT version 7 with L-INS-I model (Katoh and

Standley, 2013). Mean pairwise sequence distance (pi, π; Nei and Kumar, 2000) among

48

Texas Tech University, Pablo Antonio Tovar, December 2015 all individual sequences within a fungal family was then estimated using MEGA 6

(Tamura et al., 2013). Mean pairwise sequence distances obtained using the combined 2- year data set from P. praeclara mycorrhizal fungi were compared with the previously reported pi distances among mycorrhizal fungi from other orchids (Pandey et al., 2013).

Fisher’s exact test

The statistical significance of the variation in fungal diversity between sites was tested using the Fisher’s exact test based on abundances of individual OTUs across the sampled sites per year of collection. The test was performed in R (R Development Core

Team 2010) using all the OTUs detected per sampled population for 2013 and 2014 separately and also for the combined data set.

Two-way hierarchical ordination

Two-way hierarchical ordination analysis (Bray and Curtis., 1957) using OTU abundances was performed to test the relationships between the mycorrhizal fungal communities of the sites based on dissimilarities. PC-ORD (McCune et al., 2002) was used to perform the two-way hierarchical ordination using the Sorensen (Bray-Curtis) distance measure, and setting the group linkage criterion as ‘average’. This test was conducted for each sampling year separately and also for the combined two-year data set.

The resulting dendrogram clustered sites according to similarity in mycorrhizal fungal communities held by the sampled plants across the sampled sites, and served as an aid to

49

Texas Tech University, Pablo Antonio Tovar, December 2015 observe whether sites sharing land management types (prescribed burn, haying or grazing, no management) clustered together.

Multiple response permutation procedure

We divided all sampled sites into groups by land management and by latitude to test whether fungal community composition abundance varies among the groups. Sites including MB, PF, PS, BU, BD and IA represented prescribed fire, NDG, NE, and PH represented hayed or grazed prairies, and hayed or grazed prairies, and no active management / control was represented by NDVK and PC. When grouped by latitude,

MB, NDG, NDVK, PC, PH, PF and PS were grouped as northern sites while BD, BU, IA and NE were grouped as southern sites. A multiple response permutation procedure

(MRPP) using PC-ORD (McCune et al., 2002) was carried out by using these groups for data sets from year 2013 and 2014 separately as well as for both data sets combined.

Diversity curves

Cumulative OTU diversity curves across populations were estimated using the sample-based Mao Tao and Chao 1 methods to examine whether the abundance of OTUs increases with increasing sample size using EstimateS 9.0 (Colwell, 2005). This analysis was conducted by using out OTU abundance data obtained from the dataset that represented both sampling years together.

50

Texas Tech University, Pablo Antonio Tovar, December 2015

Nonmetric multidimensional scaling

Nonmetric multidimensional scaling (NMDS) in PC-ORD (McCune et al., 2002) was used to ordinate OTU abundance in roots of P. praeclara and to visualize the correlation between edaphic characteristics [OM, organic matter; ENR, estimated nitrogen release; P1, phosphorus 1 (readily available to plants); P2, phosphorus (readily available to plants plus part of the active reserve in the soil) 2; K, potassium; Mg, magnesium; Ca, calcium; Na, sodium; CEC, cation exchange capability; NO3, nitrate; S,

Sulphur; Zn, zinc; Mn, manganese; Fe, iron; Cu, copper and B, boron) and OTUs observed at each sampled site separately for each sampling year. The abundance-based version of the Sorensen index (Bray-Curtis) was used to calculate the distances.

Dimensionality of the ordination was determined to choose the lowest dimensionality that captured most of the variation. Data sets from 2013 and 2014 were both best described by

3-dimensional solutions with instabilities below 0.00001.

Phylogenetic analyses

To examine the phylogenetic relationships among the fungi associated with

Platanthera praeclara and other known orchid fungi, known orchid mycorrhizal sequences were added to the data used to generate the trees to place the fungi identified in this study in the context of previously known orchid mycorrhizae. Separate multiple alignments were constructed using MAFFT version 7 with L-INS-I model (Katoh and

51

Texas Tech University, Pablo Antonio Tovar, December 2015

Standley, 2013) for each of the fungal families, that is, Ceratobasidiaceae and

Tulasnellaceae (Taylor and McCormick, 2008; Pandey et al., 2013) by using the sequences generated in this study and the respective reference sequences. Reference sequences were selected from previously published studies on orchid mycorrhizae. The resulting alignments were improved manually in AliView Version 1.17.1 (Larsson,

2014).

Phylogenetic relationships within each fungal family were estimated by determining the best-fit substitution model in ModelTest (Posada and Crandall, 1998) to construct maximum-likelihood trees using MEGA 6 (Tamura et al., 2013). Support values were estimated via 10000 bootstrap replicates in the same program. Because of the accelerated diversification of Tulasnella nuclear ribosomal ITS regions, a family level alignment using only the conserved 5.8S region extracted from the full length ITS sequences was first constructed (Shefferson et al., 2007; Taylor & McCormick, 2008) to build a maximum-likelihood tree in MEGA 6. The resulting tree was used to identify subclades within Tulasnellaceae containing the Platanthera praeclara observed OTUs, and full-length ITS alignments from these subclades were then constructed to generate the trees presented herein. Sistotrema sp. (Genbank accession number EU218893) was used for rooting Ceratobasidiaceae tree. Tulasnellaceae tree was midpoint rooted and subclade trees were rooted based on relationships shown in the broader 5.8S tree.

Additionally, posterior probability trees were built in MrBayes version 3.2.5 (Ronquist and Huelsenbeck, 2003) using the substitution models T92+G for the Ceratobasidiaceae tree, and K2+G for the Tulasnellaceae trees. The priors of the substitution models were

52

Texas Tech University, Pablo Antonio Tovar, December 2015 set to default values. More than one million generations were carried out for analysis and continued until the standard deviation value of <0.01 was obtained. Cold chains were sampled after 100 generations and cross diagnosis of two runs was carried out after 1000 generations with the 25% burn in fraction and 0.2 temperature. All phylogenetic trees were visualized and edited in FigTree software (http://tree.bio.ed.ac.uk/).

Results

Operational taxonomic unit (OTU) diversity

Of the 498 root fragments from 139 sampled plants across 2013 and 2014, 306 root fragments from 109 plants produced workable DNA sequences after processing sequentially with several primer pairs and subject to cloning as needed (Table 2.1).

Ceratobasidiaceae and Tulasnellaceae were the two fungal families associated with

Platanthera praeclara across all 11 sites from where plants were sampled (Figures 2.7 and 2.8).

2013

Nineteen OTUs representing both mycorrhizal fungal families were identified from samples collected in the year 2013 (Table 2.5). Tulasnellaceae was represented by 10

OTUs, and Ceratobasidiaceae by 9 OTUs. Across all plants and sites, infection of a single root system by 2 to 4 fungal OTUs was observed in 46% of all sampled plants (Figure

53

Texas Tech University, Pablo Antonio Tovar, December 2015

2.9a). The remaining 54% of plants contained one OTU each. A majority of the plants

(48 of 54 plants) possessed fungi belonging to a single family, while only six plants had fungi belonging to both families (Figure 2.9b).

2014

When additional plants were sampled in 2014, Ceratobasidiaceae and

Tulasnellaceae were represented by 14 and 5 OTUs, respectively (Table 2.6). Across all plants and sites, infection of a single root system by 2 to 4 fungal OTUs was observed in

27% of all sampled plants (Figure 2.10a). The remaining 73% of plants had only one

OTU each. A majority of the plants (51 of 55 plants) possessed fungi belonging to a single family, while only four plants had fungi belonging to the two families (Figure

2.10b).

2013 and 2014 combined

Thirty-nine OTUs representing both fungal families were identified when data from

2013 and 2014 were pooled. The Fungal family Ceratobasidiaceae was represented by 27

OTUs and Tulasnellaceae was represented by 12 OTUs (Table 2.7). Across all sampled plants and sites, infection of a single root system by 2 to 4 fungal OTUs was observed in

50% of all sampled plants (Figure 2.11a). The remaining 50% of sampled individuals yielded only one OTU. A majority of the plants (100 of 109 plants) possessed fungi 54

Texas Tech University, Pablo Antonio Tovar, December 2015 belonging to a single family, while roots of nine plants yielded fungi belonging to both families (Figure 2.11b).

Mean pairwise sequence distance (π)

Mean pairwise distance of sequences from Tulasnellaceae was consistently larger than for Ceratobasidiaceae regardless of the year of sampling, and in the combined data set (Table 2.8). By combining the data sets for both years (2013 and 2014), the mean distance observed among Ceratobasidiaceae was 0.039 ± 0.003 and the mean distance for

Tulasnellaceae was 0.135 ± 0.006. Additionally, mycorrhizal fungal associates of P. praeclara from the family Ceratobasidiaceae had the lowest mean pairwise sequence distance among orchid species with similar wide distribution except for Serapias vomeraceae. Mean pairwise distances among the Tulasnellaceae documented in the roots of P. praeclara were larger than most of the orchids with similar distribution patterns

(Table 2.9).

Fisher’s exact test

The family Ceratobasidiaceae was represented in roots of P. praeclara from all populations sampled in the year 2013 and 2014. The fungal family Tulasnellaceae was absent in roots sampled from four of the sampled sites (PC, PH, NDVK and NE) in 2013

(Figure 2.7). Significant differences were observed in 84% of the sites when the

55

Texas Tech University, Pablo Antonio Tovar, December 2015 mycorrhizal fungal community abundances were compared to each other by using the

Fisher’s exact test (Table 2.10). In the year 2014, Tulasnellaceae was not documented in the roots sampled from three sites (PH, NDVK and BD) (Figure 2.8). The Fisher’s exact test indicated that for 2014, 66% of the sites had significant differences in the composition of their mycorrhizal fungal abundance (Table 2.11). In the combined data set representing both years, the Fisher’s exact test including all OTUs per population revealed that P. praeclara roots host similar fungal communities in 13% of the sampled sites (Table 2.12).

Two-way hierarchical ordination

The site-wise two-way hierarchical ordination based on the OTU abundances indicated that sites did not cluster according to the type of land management (prescribed burn, haying/grazing, and no management) (Figure 2.12 and 2.13). When the two-way hierarchical ordination analysis was conducted for the combined data set (years 2013 and

2014), no clear clustering of sites based on management or latitude was detected (Figure

2.14).

Multiple response permutation procedure

The multiple response permutation procedure did not explain the variation in the mycorrhizal fungal communities to be related to type of land management (P > 0.05 in

2013, 2014 and combined data sets; Table 2.13). The variation was neither explained by

56

Texas Tech University, Pablo Antonio Tovar, December 2015 geographical positioning of the sites in 2013 or 2014. However, variation in the mycorrhizal fungal communities was observed in the combined data set when separating most northern and most southern sites (Table 2.13).

Diversity curves

Sample-based cumulative OTU diversity curves built from the combined data set

(2013 and 2014) showed that estimates of Chao 1 (55 OTUs) are higher than the estimates of Mao Tau (39 OTUs), indicating that OTU diversity was not saturated with the current sampling and that more OTUs would be detected with additional sampling

(Figure 2.15).

Nonmetric multidimensional scaling

Nonmetric multidimensional scaling (NMS) ordination analysis of OTU abundances in orchid roots from samples collected in 2013 showed that distinction in orchid-associated fungi was correlated to Mn and Zn in NE; ENR and Ca in PC; OM,

CEC, S, and Salts in PH; and Na and pH in BD (Figure 2.16). For the 2014 data set, distinction in orchid-associated fungi was correlated to Silt, Clay and H in BU; S in

NDVK; pH and P2 in PF; and B in BD (Figure 2.17).

57

Texas Tech University, Pablo Antonio Tovar, December 2015

Phylogenetic analyses

Maximum likelihood trees and posterior probability Bayesian trees showed similar relationships among the OTUs detected in this study and reference sequences. The maximum likelihood tree built with the OTUs from the fungal family Ceratobasidiaceae observed in P. praeclara along with reference nrITS sequences from published research showed four major clades (Figure 2.18). Ninety percent of the OTUs clustered together in one clade, along with one representative Ceratobasidium sp. sequence obtained from previously sampled roots of P. praeclara (GenBank accession DQ068771, J. Sharma unpublished data). The sister clade, without any OTUs from this study, was represented by mycorrhizal fungi belonging to the genus Ceratobasidium reported from roots of epiphytic neotropical orchids Ionopsis utricularoides (Otero et al., 2007; Borges et al.,

2013) and Tolumnia variegate (Otero et al., 2002). The same clade was also represented by a Ceratobasidium sp. obtained from the roots of tropical orchids in Singapore (Ma et al., 2003). Two of the OTUs from this study were separated from these two clades. One

OTU (OTU 24) was located in the third clade and was closely related to Ceratobasidium sp. observed in decayed roots of conifer seedlings in Lithuania (Menkis et al., 2006) and in Picea mariana forest soils in Alaska in the United States (Taylor et al., 2014). In the fourth clade, OTU 23 was closely related to Thanatephorus sp. observed in roots of

Vanilla aphylla in Cuba (Porras-Alfaro and Bayman, 2007).

The first sub-tree of the Tulasnellaceae (Figure 2.19a) contained five of the 12

OTUs observed for this mycorrhizal fungal family in P. praeclara. Operational

Taxonomic Units 1, 10 and 11 were not closely related to any previously accessioned 58

Texas Tech University, Pablo Antonio Tovar, December 2015 sequences. However, the sister clade of the clade described before contained OTU 3, which is closely related to a reported Tulasnella sp. from P. praeclara roots (GenBank accession DQ068773, J. Sharma unpublished data) and to a Tulasnella calospora obtained from roots of the orchid Paphiopedilum charlesworthii in Thailand

(Nontachaiyapoom et al., 2010). A third clade with OTU 12 was related to Tulasnella calospora obtained from roots of Acianthus exsertus, an orchid from the Andean cloud forest (Suarez et al., 2006) and is also related to Tulasnella sp. observed in the New

Zealand orchid Nematocera iridescens (Watkins, 2012). The second sub-tree of the

Tulasnellaceae (Figure 2.19b) also included five OTUs. Two of the five OTUs (OTU4 and OTU6) were closely related to Tulasnella bifrons, which have been related to mycorrhizal fungi from the North American native orchid Goodyera pubescens

(McCormik et al., 2004) and to Tulasnella sp. obtained from roots of Dactylorhiza sp.

(Jacquemyn et al., 2012). Tulasnella sp. associated with the orchid Tipularia discolor

(McCormik et al., 2004) were closely related to OTU2 and OTU 8. Operational

Taxonomic Unit 7 was closely related to Tulasnella irregularis observed in the roots of an orchid from Thailand (Nontachaiyapoom et al., 2010). In the third sub tree (Figure

2.19c), OTU 5 and OTU 9 were observed closely related to an uncultured Tulasnellaceae isolate from the roots of the orchid Cypripedium parviflorum (Shefferson et al., 2007) from Illinois in the United States.

59

Texas Tech University, Pablo Antonio Tovar, December 2015

Discussion

Studies quantifying range-wide mycorrhizal communities of grassland orchids are severely limited, and are rarer still, or non-existent, within North America. Moreover, majority of the studies that report mycorrhizal association in orchids include samples collected only once from a few sites, thus restricting the interpretations with respect to the species-wide patterns of specificity over space and time (Rasmussen, 1995;

McCormick et al., 2006; Smith and Read, 2008; Dearnaley et al., 2012; Ercole et al.,

2015). We report both spatial and temporal variation patterns in the fungal communities within the roots of a rare terrestrial orchid native to North American tallgrass prairie.

Roots of Platanthera praeclara associated with two of the most common fungal families, Ceratobasidiaceae and Tulasnellaceae, that orchids across the world typically associate with. The fungal family Ceratobasidiaceae was represented at all of the 11 sampled sites in each of the two sampling years. Within the roots of the orchid, fungi from this family were more abundant than those of the fungal family Tulasnellaceae; however, an exception to this pattern was noted at the site BU in Minnesota across both sampling years. These patterns suggest that the orchid is specific in its choice of mycorrhizal fungi, and prefers the members of Ceratobasidiaceae (Figures 2.7 and 2.8).

Majority of the Ceratobasidiaceae OTUs identified in this study clustered with a previously known isolate from P. praeclara (Sharma et al. 2003) suggesting that the orchid has retained the associations with this narrow group of fungi for over a decade.

Rarity in orchids has been related to their specific associations with mycorrhizal fungi (Selosse et al., 2004; Shefferson et al., 2008; McCormik and Jacquemyn, 2014). In 60

Texas Tech University, Pablo Antonio Tovar, December 2015 addition to orchid rarity, geographic range of a species should be considered when assessing specificity in the association between orchids and mycorrhizal fungi (Otero et al., 2007). Data obtained over two years from >100 plants yielded 39 fungal OTUs across

11 sampled sites. Although 39 OTUs might appear to be a large number of different taxonomic units to have relationships with, it is important to note that specificity is defined by the phylogenetic breadth of the symbionts (Molina et al., 1992; Taylor et al.,

2002; McCormick et al., 2004). While Ceratobasidiaceae was observed to be the dominant fungal family across the natural range of the species when data from 2013 and

2014 were combined (represented by 238 nrITS sequences and 27 OTUs), their sequence distance (π = 0.039 ± 0.003) is lower than that of Tulasnellaceae (π = 0.135 ± 0.006). Its sequence distance is also lower than the distances from the same fungal family in other widely distributed orchids (Table 2.9) except for Serapias vomaracea. Interestingly, when comparing pi-distances from sequences of the fungal family Ceratobasidiaceae observed in P. praeclara with Ceratobasidiaceae known from orchids with limited distributions, only Hexalectris grandiflora, a nonphotosynthetic orchid, exhibited a smaller sequence distance. Pairwise sequence distances were supported by the constructed phylogenetic trees, displaying a narrow phylogenetic breadth in the OTUs observed in the family Ceratobasidiaceae, and a broader phylogenetic breadth in the

OTUs of the Tulasnellaceae. Among the Ceratobasidiaceae, 25 of 27 OTUs clustered together in a separate clade. Operational taxonomic units 3 and 12 were the most closely related to a previously published nrITS sequence of Ceratobasidiaceae from P. praeclara

(Genbank accession code DQ068771) (Figure 2.18). This particular genotype was

61

Texas Tech University, Pablo Antonio Tovar, December 2015 detected in 2003, suggesting that the same isolate can remain associated with the species for over a decade. In previous studies (Sharma et al. 2003b), this isolate induced the highest germination in vitro when seeds of P. praeclara were cultured symbiotically.

Tulasnellaceae phylograms on the other hand are an example on a wider phylogenetic breadth, as the OTUs are distributed through the constructed sub trees (Figures 2.19a, b and c). Differences between the fungal communities associated with P. praeclara were noted when analyzing the data from each year separately. In the year 2013,

Ceratobasidiaceae was composed of 9 OTUs (from 122 nrITS sequences) versus 10

OTUs (from 48 nrITS sequences) in Tulasnellaceae. Despite having more OTUs in

Tulasnellaceae, sequence distances were smaller for Ceratobasidiaceae (π = 0.017 ±

0.003) in comparison to Tulasnellaceae (π = 0.109 ± 0.006). In the year 2014, the mycorrhizal fungal community of P. praeclara shifted to 14 OTUs in Ceratobasidiaceae

(116 nrITS sequences) and 5 OTUs in Tulasnellaceae (21 nrITS sequences). Still, sequence distances remained smaller for Ceratobasidiaceae (Table 2.8). For P. praeclara, the number of OTUs observed coupled with the sequence distances provides strong quantitative evidence that this orchid has high specificity compared to other orchids, majority of which have been studied with a much smaller sample size. Still, it is important to notice that rarefaction curves (Mao Tao and Chao 1 in Figure 2.15) showed that higher number of OTUs might have been detected if more plants were investigated although our results suggest that the increased diversity will most likely be attributed to variants within a narrow clade of Ceratobasidiaceae.

62

Texas Tech University, Pablo Antonio Tovar, December 2015

Operational Taxonomic Unit data for the years 2013 (Table 2.5) and 2014 (table

2.6), and for the combined data set (Table 2.7) showed the community composition in terms of OTU for each sampled site. Although there are a few OTUs representing more than 5 sites in each data set (OTUs c1, c4 and c8 in 2013; OTUs c1 and c11 in 2014;

OTUs c1, c12 and c20 in the combined data set), most of the less common OTUs were observed to be restricted to one or two different sites. Moreover, composition of the mycorrhizal fungal communities at each sampled site varied from one year to the other.

Nine OTUs from the fugal family Ceratobasidiaceae were observed in 2013. The next year, 14 OTUs were detected for the same fungal family. For Tulasnellaceae, 10 OTUs were identified in 2013, and 5 were observed for 2014. Shifts in the community of fungal mycorrhizae are not unusual. Seasonal changes might alter the communities through time

(Ercole et al., 2015). Research in other, non-orchid mycorrhizal associations have shown temporal dynamics of fungal assemblages (Twieg et al., 2007; Dumbrell et al., 2011;

Bennett et al., 2013) to change within a year (Cowden and Peterson, 2013). Others have described no change in the community across seasons (Richard et al., 2011). For Orchid mycorrhizae, McCormick et al., (2006) observed that individuals of the orchid Goodyera pubescens maintained associations with single fungal genotypes for periods longer than four years. The reasons that explain shifts in fungal mycorrhizal communities are not clearly elucidated yet. Using high throughput sequencing, studies have shown stable communities of arbuscular mycorrhizae fungi in soil through time (Davison et al., 2012), as well as shifts in ECM fungi in soils of Quercus plants (Jumpponen et al., 2010).

63

Texas Tech University, Pablo Antonio Tovar, December 2015

Multiple fungal taxa were detected in root systems of individual plants of P. praeclara. Several studies have reported this situation in other photosynthetic orchids:

Platanthera spp. (Zelmer et al., 1996); Tipularia discolor (McCormick et al. 2004);

Cephalantera damasonium, Epipactis atrorubens and Platanthera chlorantha

(Bidartondo et al., 2004); Aphyllorchis spp. (Roy et al., 2009); Orchis spp. (Jacquemyn et al., 2010; Lievens et al., 2010); Piperia yadonii (Sharma et al., 2007; Pandey et al., 2013)

Nervilia nipponica (Nomura et al., 2013); Neottia ovata (Jacquemyn et al., 2015).

However, photosynthetic orchid Goodyera pubescens (McCormick et al., 2006) roots were observed to be colonized by a single fungal taxon. Individual root systems associating with several fungal taxa are usually documented in non-orchid species that typically display low fungal specificity (Lodge & Wentworth, 1990; Perotto et al., 1994 and 1996; Monreal et al., 1999). Mycorrhizae from nonphotosynthetic orchids, which usually have high specificity towards mycorrhizal fungi, are colonized only by a single fungal taxon. Examples of nonphotosynthetic orchids presenting these pattern include

Corallorhiza maculata (Taylor and Bruns, 1999), Hexalectris revoluta (Taylor et al.,

2003) and Neottia nidus-avis (Selosse et al., 2002).

Variation in the dominance of orchid mycorrhizal fungi across sites was observed in P. praeclara. The Fisher’s test for the combined data set (Table 2.12), as well as for the data sets for years 2013 and 2014 independently (Table 2.10 and 2.11), indicated significant differences between the mycorrhizal fungal communities in most of the sampled sites. Since communities are different from one year to the other, similarities or differences between populations vary from one year to the next. The two-way

64

Texas Tech University, Pablo Antonio Tovar, December 2015 hierarchical clustering analyses based on Bray-Curtis dissimilarities elucidated the relationships between sampled sites based on their mycorrhizal fungal communities by exhibiting different relationships in both years. Sites with similar mycorrhizal fungal communities were not subjected to the same type of land management in any of the data sets (2013, Figure 2.12; 2014, Figure 2.13; combined data set, Figure 2.14). The multiple response permutation procedure with all data sets ((P > 0.05) (Table 2.13) further confirmed that the variation of the mycorrhizal fungal communities across sites is not influenced by the overall prairie management practices. Also, nonmetric multidimensional scaling ordination analysis did not segregate sites according to their type of land management in either of the yearly data sets (Figures 2.16 and 2.17).

However, studies have shown how changes in habitat management can have significant effects on microbial communities (Carlile et al., 2001; Steenwerth et al., 2002; Johnson et al., 2003). Research has also shown that changes in the structure of fungal communities can be associated with alterations in soil properties like soil nitrogen availability (Frey et al., 2004), soil texture (Girvan et al., 2003), or soil pH (Fierer and Jackson, 2006).

Moreover, soil fungi might be sensitive to vegetation shifts that come along with different types of land management; this may be especially true for mycorrhizal fungi that form associations with particular plants (Heinemeyer et al., 2004). Lauber et al. 2008 observed significant differences in soil texture, carbon, nitrogen, and phosphorus measured under different land management practices. In the same study, phosphorus concentrations were correlated with fungal communities in that similar concentrations of the element fostered similar fungal communities, but no correlation was observed between land management

65

Texas Tech University, Pablo Antonio Tovar, December 2015 and fungal communities in roots of P. praeclara. This suggests that phosphorus might be an important factor that regulates the biogeographical patterns in fungal communities in both soil and plant roots. Rousk et al. (2010) observed that pH and fungal community composition can be related, but the relationships are often weak in part because of the wide pH tolerance that fungal species usually have (Wheeler et al., 1991; Nevarez et al.,

2009). However, few soil variables with strong vectors appeared to correlate with the mycorrhizal fungal communities in orchid roots observed at different sites. In 2013, pH and sodium, exhibited a degree of correlation with the fungal community within orchid roots at PC (Pembina Control plot). Additional strong vectors like Mn and P1 did not appear to correlate with any of the remaining sites. In 2014, clay and H appeared to correlate with the mycorrhizal community within orchid roots at BU. Sites PH and

NDVK seemed to be correlated with S, and site BD may have a correlation with B.

Although the strength of the observed correlations can be inferred from the length of the vector, correlation analysis remain to be performed between these variables and the observed mycorrhizal fungal communities observed at each site to further infer the relationship between edaphic variables and fungal communities within the orchid roots.

Sharma et al. (2003a) isolated and identified mycorrhizae from roots of 21 P. praeclara plants sampled form four sites in Minnesota and one site in Missouri. They reported a higher proportion of the fungal family Ceratobasidiaceae (78 isolates) and a lower proportion for Tulasnellaceae (9 isolates). The current study analyzed 109 plants from 11 sites spanning a much wider geographic area to represent the entire natural range of the species and confirmed that the fungal family Ceratobasidiaceae was dominant and

66

Texas Tech University, Pablo Antonio Tovar, December 2015 more abundant throughout the entire range. Our expectation with respect to the mycorrhizal specificity exhibited by the species was largely corroborated given the narrow phylogenetic breadth observed in the nrITS sequences of the fungal family

Ceratobasidiaceae. However, while the associations are specific, mycorrhizal communities of P. praeclara were not static, as shifts over time (from year 2013 to 2014) and space (from site to site) were evident.

67

Texas Tech University, Pablo Antonio Tovar, December 2015

Literature Cited

Andersen, T.F. (1990). A study of hyphal morphology in the form genus Rhizoctonia. Mycotaxon, 37, 25-46.

Anderson, I.C., & Cairney, J.W. (2004). Diversity and ecology of soil fungal communities: increased understanding through the application of molecular techniques. Environmental Microbiology, 6(8), 769-779.

Bidartondo, M.I., Burghardt, B., Gebauer, G., Bruns, T.D., & Read, D.J. (2004). Changing partners in the dark: isotopic and molecular evidence of ectomycorrhizal liaisons between forest orchids and trees. Proceedings of the Royal Society of London B: Biological Sciences, 271(1550), 1799-1806.

Biederman, L. A., Beckman, J., Prekker, J., Anderson, D., Sather, N. P., & Dahle, R. (2014). Phenological monitoring aids habitat management of threatened plant. Natural Areas Journal, 34(1), 105-110.

Borges, R., Valadares, S., & Cardoso, E.J.B.N. (2013). Ceratobasidium lineages associated with two epiphytic orchids in Colombia. In Fourth Scientific Conference on Andean Orchids (p. 147).

Bray, J. R., & Curtis, J. T. (1957). An ordination of the upland forest communities of southern Wisconsin. Ecological monographs, 27(4), 325-349.

Carlile, M.J., Watkinson, S.C., & Gooday, G.W. (2001). The fungal diversity. The Fungi.

Colwell, R.K. (2005). EstimateS: Statistical estimation of species richness and shared species from samples.

Cowden, C.C., & Peterson, C.J. (2013). Annual and seasonal dynamics of ectomycorrhizal fungi colonizing white pine (Pinus strobus) seedlings following catastrophic windthrow in northern Georgia, USA. Canadian Journal of Forest Research, 43(3), 215-223.

Cully, A. C., Cully, J. F., & Hiebert, R. D. (2003). Invasion of exotic plant species in tallgrass prairie fragments. Conservation Biology, 17(4), 990-998.

Davison, J., Öpik, M., Zobel, M., Vasar, M., Metsis, M., & Moora, M. (2012). Communities of arbuscular mycorrhizal fungi detected in forest soil are spatially heterogeneous but do not vary throughout the growing season. PLoS One, 7(8), e41938. 68

Texas Tech University, Pablo Antonio Tovar, December 2015

Dearnaley, J.D. W., Martos, F., & Selosse, M.A. (2012). 12 Orchid Mycorrhizas: Molecular Ecology, Physiology, Evolution and Conservation Aspects. In Fungal associations (pp. 20 207-230). Springer Berlin Heidelberg.

Dumbrell, A.J., Ashton, P. D., Aziz, N., Feng, G., Nelson, M., Dytham, C., & Helgason, T. (2011). Distinct seasonal assemblages of arbuscular mycorrhizal fungi revealed by massively parallel pyrosequencing. New Phytologist, 190(3), 794-804.

Environment Canada. (2006). Recovery Strategy for the Western Prairie Fringed Orchid (Platanthera praeclara) in Canada. Species at Risk Act Recovery Strategy Series. Environment Canada, Ottawa.

Ercole, E., Adamo, M., Rodda, M., Gebauer, G., Girlanda, M., & Perotto, S. (2015). Temporal variation in mycorrhizal diversity and carbon and nitrogen stable isotope abundance in the wintergreen meadow orchid Anacamptis morio.New Phytologist, 205(3), 1308-1319.

Fierer, N., & Jackson, R. B. (2006). The diversity and biogeography of soil bacterial communities. Proceedings of the National Academy of Sciences of the United States of America, 103(3), 626-631.

Fox, K., Vitt, P., Anderson, K., Fauske, G., Travers, S., Vik, D., & Harris, M.O. (2013). Pollination of a threatened orchid by an introduced hawk moth species in the tallgrass prairie of North America. Biological Conservation, 167, 316-324.

Frey, S.D., Knorr, M., Parrent, J.L., & Simpson, R.T. (2004). Chronic nitrogen enrichment affects the structure and function of the soil microbial community in temperate hardwood and pine forests. Forest Ecology and Management,196(1), 159-171.

Gardes, M., White, T.J., Fortin, J.A., Bruns, T.D., & Taylor, J.W. (1991). Identification of indigenous and introduced symbiotic fungi in ectomycorrhizae by amplification of nuclear and mitochondrial ribosomal DNA. Canadian Journal of Botany, 69(1), 180-190.

Gardes, M., & Bruns, T.D. (1993). ITS primers with enhanced specificity for basidiomycetes‐application to the identification of mycorrhizae and rusts. Molecular ecology, 2(2), 113-118.

Gardes, M., & Bruns, T.D. (1996). ITS-RFLP matching for identification of fungi. In Species Diagnostics Protocols (pp. 177-186). Humana Press.

69

Texas Tech University, Pablo Antonio Tovar, December 2015

Ghimire, S.R., Charlton, N.D., Bell, J.D., Krishnamurthy, Y.L., & Craven, K.D. (2011). Biodiversity of fungal endophyte communities inhabiting switchgrass (Panicum virgatum L.) growing in the native tallgrass prairie of northern Oklahoma. Fungal Diversity, 47(1), 19-27.

Girvan, M.S., Bullimore, J., Pretty, J.N., Osborn, A.M., & Ball, A.S. (2003). Soil type is the primary determinant of the composition of the total and active bacterial communities in arable soils. Applied and Environmental Microbiology,69(3), 1800-1809.

Goedeke, T., Sharma, J., Delphey, P. & Marshall Mattson, K. (2008). Platanthera praeclara. In: IUCN 2013. IUCN Red List of Threatened Species. Version 2013.2. .

Heinemeyer, A., Ridgway, K.P., Edwards, E.J., Benham, D.G., Young, J.P.W., & Fitter, A.H. (2004). Impact of soil warming and shading on colonization and community structure of arbuscular mycorrhizal fungi in roots of a native grassland community. Global Change Biology, 10(1), 52-64.

Jacquemyn, H., Honnay, O., Cammue, B., Brys, R., & Lievens, B. (2010). Low specificity and nested subset structure characterize mycorrhizal associations in five closely related species of the genus Orchis. Molecular Ecology, 19(18), 4086- 4095.

Jacquemyn, H., Deja, A., De Hert, K., Cachapa Bailarote, B., & Lievens, B. (2012). Variation in mycorrhizal associations with tulasnelloid fungi among populations of five Dactylorhiza species. PloS one, 7(8), e42212.

Jacquemyn, H., Waud, M., Merckx, V.S., Lievens, B., & Brys, R. (2015). Mycorrhizal diversity, seed germination and long‐ term changes in population size across nine populations of the terrestrial orchid Neottia ovata. Molecular ecology.

Johnson, M.J., Lee, K. Y., & Scow, K.M. (2003). DNA fingerprinting reveals links among agricultural crops, soil properties, and the composition of soil microbial communities. Geoderma, 114(3), 279-303.

Jumpponen, A.R.I., Jones, K.L., David Mattox, J., & Yaege, C. (2010). Massively parallel 454‐sequencing of fungal communities in Quercus spp. ectomycorrhizas indicates seasonal dynamics in urban and rural sites.Molecular Ecology, 19(s1), 41-53.

Katoh, K., & Standley, D. M. (2013). MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Molecular biology and evolution, 30(4), 772-780.

70

Texas Tech University, Pablo Antonio Tovar, December 2015

Kirby, D.R., Lym, R. G., Sterling, J. J., & Sieg, C. H. (2003). Observation: leafy spurge control in western prairie fringed orchid habitat. Journal of range management, 466-473.

Kivlin, S. N., Hawkes, C. V., & Treseder, K. K. (2011). Global diversity and distribution of arbuscular mycorrhizal fungi. Soil Biology and Biochemistry,43(11), 2294- 2303.

Larsson, A. (2014). AliView: a fast and lightweight alignment viewer and editor for large datasets. Bioinformatics, 30(22), 3276-3278.

Lauber, C.L., Strickland, M.S., Bradford, M.A., & Fierer, N. (2008). The influence of soil properties on the structure of bacterial and fungal communities across land-use types. Soil Biology and Biochemistry, 40(9), 2407-2415.

Lievens, B., Van Kerckhove, S., Justé, A., Cammue, B. P., Honnay, O., & Jacquemyn, H. (2010). From extensive clone libraries to comprehensive DNA arrays for the efficient and simultaneous detection and identification of orchid mycorrhizal fungi. Journal of Microbiological Methods, 80(1), 76-85.

Lodge, D.J., & Wentworth, T.R. (1990). Negative associations among VA-mycorrhizal fungi and some ectomycorrhizal fungi inhabiting the same root system. Oikos, 347-356.

Ma, M., Tan, T.K., & Wong, S.M. (2003). Identification and molecular phylogeny of Epulorhiza isolates from tropical orchids. Mycological Research,107(09), 1041- 1049.

Martin, K.J., & Rygiewicz, P.T. (2005). Fungal-specific PCR primers developed for analysis of the ITS region of environmental DNA extracts. BMC microbiology, 5(1), 28.

McCormick, M.K., Whigham, D.F., & O'Neill, J. (2004). Mycorrhizal diversity in photosynthetic terrestrial orchids. New Phytologist, 163(2), 425-438.

McCormick, M.K., Whigham, D.F., Sloan, D., O'Malley, K., & Hodkinson, B. (2006). Orchid-fungus fidelity: a marriage meant to last?. Ecology, 87(4), 903-911.

McCormick, M.K., & Jacquemyn, H. (2014). What constrains the distribution of orchid populations?. New Phytologist, 202(2), 392-400.

McCune, B., Grace, J. B., & Urban, D. L. (2002). Analysis of ecological communities (Vol. 28). Gleneden Beach, OR: MjM software design.

71

Texas Tech University, Pablo Antonio Tovar, December 2015

Menkis, A., Vasiliauskas, R., Taylor, A.F.S., Stenström, E., Stenlid, J., & Finlay, R. (2006). Fungi in decayed roots of conifer seedlings in forest nurseries, afforested clear‐cuts and abandoned farmland. Plant pathology,55(1), 117-129.

Molina, R., Massicotte, H., & Trappe, J.M. (1992). Specificity phenomena in mycorrhizal symbioses: community-ecological consequences and practical implications. Mycorrhizal functioning: an integrative plant-fungal process, 357, e423.

Monreal, M., Berch, S.M., & Berbee, M. (1999). Molecular diversity of ericoid mycorrhizal fungi. Canadian Journal of Botany, 77(11), 1580-1594.

Nei, M., & Kumar, S. (2000). Molecular evolution and phylogenetics. Oxford University Press.

Nevarez, L., Vasseur, V., Le Madec, A., Le Bras, M.A., Coroller, L., Leguérinel, I., & Barbier, G. (2009). Physiological traits of Penicillium glabrum strain LCP 08.5568, a filamentous fungus isolated from bottled aromatised mineral water. International journal of food microbiology, 130(3), 166-171.

Nomura, N., Ogura-Tsujita, Y., Gale, S. W., Maeda, A., Umata, H., Hosaka, K., & Yukawa, T. (2013). The rare terrestrial orchid Nervilia nipponica consistently associates with a single group of novel mycobionts. Journal of plant research,126(5), 613-623.

Nontachaiyapoom, S., Sasirat, S., & Manoch, L. (2010). Isolation and identification of Rhizoctonia-like fungi from roots of three orchid genera, Paphiopedilum, Dendrobium, and Cymbidium, collected in Chiang Rai and Chiang Mai provinces of Thailand. Mycorrhiza, 20(7), 459-471.

Oja, J., Kohout, P., Tedersoo, L., Kull, T., & Kõljalg, U. (2015). Temporal patterns of orchid mycorrhizal fungi in meadows and forests as revealed by 454 pyrosequencing. New Phytologist, 205(4), 1608-1618.

Otero, J.T., Ackerman, J.D., & Bayman, P. (2002). Diversity and host specificity of endophytic Rhizoctonia-like fungi from tropical orchids. American Journal of Botany, 89(11), 1852-1858.

Otero, J.T., Flanagan, N.S., Herre, E.A., Ackerman, J.D., & Bayman, P. (2007). Widespread mycorrhizal specificity correlates to mycorrhizal function in the neotropical, epiphytic orchid Ionopsis utricularioides (Orchidaceae).American Journal of Botany, 94(12), 1944-1950.

72

Texas Tech University, Pablo Antonio Tovar, December 2015

Pandey, M., Sharma, J., Taylor, D., & Yadon, V. L. (2013). A narrowly endemic photosynthetic orchid is non‐ specific in its mycorrhizal associations. Molecular ecology, 22(8), 2341-2354.

Perotto, S., Actis-Perino, E., & Bonfante, P. (1994). Fungal Biodiversity within Ericoid Mycorrhizal Roots Analysed with Pcr-Rapd Techniques. In Advances in Molecular Genetics of Plant-Microbe Interactions (pp. 187-190). Springer Netherlands.

Phillips, L. (2003). Pollination of Western Prairie Fringed Orchid, Platanthera praeclara Sheviak & Bowles: implications for restoration and management. Restoration and Reclamation Review. Student On-Line Journal, Department of Horticultural Science, University of Minnesota.

Porras-Alfaro, A., & Bayman, P. (2007). Mycorrhizal fungi of Vanilla: diversity, specificity and effects on seed germination and plant growth. Mycologia, 99(4), 510-525.

Posada, D., & Crandall, K. A. (1998). Modeltest: testing the model of DNA substitution. Bioinformatics, 14(9), 817-818.

Ramsay, R.R., Dixon, K.W., & Sivasithamparam, K. (1986). Patterns of infection and endophytes associated with Western Australian orchids.Lindleyana, 1(3), 203- 214.

Rasmussen, H.N. (1995). Terrestrial orchids: from seed to mycotrophic plant. Cambridge University Press.

Redecker, D. (2000). Specific PCR primers to identify arbuscular mycorrhizal fungi within colonized roots. Mycorrhiza, 10(2), 73-80.

Richard, F., Roy, M., Shahin, O., Sthultz, C., Duchemin, M., Joffre, R., & Selosse, M.A. (2011). Ectomycorrhizal communities in a Mediterranean forest ecosystem dominated by Quercus ilex: seasonal dynamics and response to drought in the surface organic horizon. Annals of Forest Science, 68(1), 57-68.

Ronquist, F., & Huelsenbeck, J.P. (2003). MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics, 19(12), 1572-1574.

Roy, M., Watthana, S., Stier, A., Richard, F., Vessabutr, S., & Selosse, M.A. (2009). Two mycoheterotrophic orchids from Thailand tropical dipterocarpacean forests associate with a broad diversity of ectomycorrhizal fungi. BMC biology,7(1), 51.

73

Texas Tech University, Pablo Antonio Tovar, December 2015

Schoch, C.L., Seifert, K.A., Huhndorf, S., Robert, V., Spouge, J.L., Levesque, C.A., & Griffith, G.W. (2012). Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi.Proceedings of the National Academy of Sciences, 109(16), 6241-6246.

Sharma, J., Zettler, L.W., & Van Sambeek, J.W. (2003a). A Survey of Mycobionts of Federally Threatened Platanthera praeclara (Orchidaceae). Symbiosis, 34, 145- 155

Sharma, J., Zettler, L.W., Van Sambeek, J.W., Ellersieck, M.R., & Starbuck, C.J. (2003b). Symbiotic seed germination and mycorrhizae of federally threatened Platanthera praeclara (Orchidaceae). The American midland naturalist, 149(1), 104-120.

Sharma, J., Ishida, M.L., & Yadon, V.L. (2007). Mycorrhizal diversity of an endemic terrestrial orchid. Lankesteriana, 7, 215-218.

Shefferson, R.P., Taylor, D.L., Weiß, M., Garnica, S., McCormick, M.K., Adams, S., & Lee, Y.I. (2007). The evolutionary history of mycorrhizal specificity among lady's slipper orchids. Evolution, 61(6), 1380-1390.

Shefferson, R.P., Kull, T., & Tali, K. (2008). Mycorrhizal interactions of orchids colonizing Estonian mine tailings hills. American Journal of Botany, 95(2), 156- 164.

Seifert, K.A. (2009). Progress towards DNA barcoding of fungi. Molecular Ecology Resources, 9(s1), 83-89.

Selosse, M.A., WEIß, M., Jany, J. L., & Tillier, A. (2002). Communities and populations of sebacinoid basidiomycetes associated with the achlorophyllous orchid Neottia nidus‐ avis (L.) LCM Rich. and neighbouring tree ectomycorrhizae. Molecular Ecology, 11(9), 1831-1844.

Selosse, M. A., Faccio, A., Scappaticci, G., & Bonfante, P. (2004). Chlorophyllous and achlorophyllous specimens of Epipactis microphylla (Neottieae, Orchidaceae) are associated with ectomycorrhizal septomycetes, including truffles. Microbial Ecology, 47(4), 416-426.

Smith, S.E., and Read, D.J. (2008) Mycorrhizal Symbiosis, 3rd edn. Academic Press, London, UK

Steenwerth, K. L., Jackson, L. E., Calderón, F.J., Stromberg, M.R., & Scow, K.M. (2002). Soil microbial community composition and land use history in cultivated

74

Texas Tech University, Pablo Antonio Tovar, December 2015

and grassland ecosystems of coastal California. Soil Biology and Biochemistry, 34(11), 1599-1611.

Suárez, J.P., Weiß, M., Abele, A., Garnica, S., Oberwinkler, F., & Kottke, I. (2006). Diverse tulasnelloid fungi form mycorrhizas with epiphytic orchids in an Andean cloud forest. Mycological research, 110(11), 1257-1270.

Tamura, K., Stecher, G., Peterson, D., Filipski, A., & Kumar, S. (2013). MEGA6: molecular evolutionary genetics analysis version 6.0. Molecular biology and evolution, 30(12), 2725-2729.

Taylor, D.L, & Bruns, T.D. (1999). Population, habitat and genetic correlates of mycorrhizal specialization in the ‘cheating’orchids Corallorhiza maculata and C. mertensiana. Molecular Ecology, 8(10), 1719-1732.

Taylor, D.L., T.D. Bruns, J.R. Leake, and D.J. Read, (2002). Mycorrhizal specificity and function in myco-heterotrophic plants. In M. van der Heijden and I. Sanders [eds.], the ecology of mycorrhizas, vol. 157, 375 – 414. Springer, Berlin, Germany.

Taylor, D.L., Bruns, T.D., Szaro, T. M., & Hodges, S.A. (2003). Divergence in mycorrhizal specialization within Hexalectris spicata (Orchidaceae), a nonphotosynthetic desert orchid. American Journal of Botany, 90(8), 1168-1179.

Taylor, D.L., & McCormick, M.K. (2008). Internal transcribed spacer primers and sequences for improved characterization of basidiomycetous orchid mycorrhizas. New Phytologist, 177(4), 1020-1033.

Taylor, D.L., Hollingsworth, T.N., McFarland, J.W., Lennon, N.J., Nusbaum, C., & Ruess, R.W. (2014). A first comprehensive census of fungi in soil reveals both hyperdiversity and fine-scale niche partitioning. Ecological Monographs, 84(1), 3-20.

Twieg, B.D., Durall, D.M., & Simard, S.W. (2007). Ectomycorrhizal fungal succession in mixed temperate forests. New Phytologist, 176(2), 437-447.

U.S. Fish and Wildlife Service. (1996). Platanthera praeclara (western prairie fringed orchid) recovery plan. Ft. Snelling, MN: US Fish and Wildlife Service.

U.S. Fish and Wildlife Service. (2009). Platanthera praeclara (western prairie fringed orchid) recovery plan. 5 year review: Summary and Evaluation. Ft. Snelling, MN: US Fish and Wildlife Service.

75

Texas Tech University, Pablo Antonio Tovar, December 2015

Watkins, R. L. (2012). The biogeography, ecology and endophyte mycorrhiza of the New Zealand alliance (Orchidaceae): specifically, Nematoceras iridescens (Irwin et Molloy) Molloy, DL Jones & MA Clem.(species): a thesis presented in partial fulfilment of the requirements for the degree of Doctor of Philosophy in Plant Biology at Massey University, Palmerston North, New Zealand (Doctoral dissertation).

Wheeler, K. A., Hurdman, B. F., & Pitt, J. I. (1991). Influence of pH on the growth of some toxigenic species of Aspergillus, Penicillium and Fusarium.International journal of food microbiology, 12(2), 141-149.

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.

Zelmer, C. D., Cuthbertson, L., & Currah, R. S. (1996). Fungi associated with terrestrial orchid mycorrhizas, seeds and protocorms. Mycoscience, 37(4), 439-448.

76

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.1. Number of plants sampled and number of samples that yielded useable sequences from roots of Platanthera praeclara in 2013 and 2014. Plant roots were sampled for mycorrhizal diversity studies across 11 sites representing the natural distribution of the species.

Samples collected Samples yielding useable sequences

2013 2014 Total 2013 2014 Total

Number of plants 73 66 139 54 55 109

Number of root fragments 267 231 498 168 138 306

77

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.2 Soil chemical characteristics of 11 sites from where Platanthera praeclara roots were sampled in 2013 for mycorrhizal analyses. OM, organic matter; ENR, estimated nitrogen release; P1, phosphorus 1 (readily available to plants); P2, phosphorus 2 (readily available to plants plus part of the active reserve in the soil); K, potassium; Mg, magnesium; Ca, calcium; Na, sodium and CEC, cation exchange capability. Please see next page for the remaining Table.

Soil characteristics

OM ENR P1 P2 K Mg Ca Na CEC pH Site (%) (g/ha) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (meq/100g) PC 2.5 89668 27 39 8.2 167 607 3155 53 21 PH 3.1 103118 29 46 8.5 150 852 2582 54 20 PF 1.6 70613 40 45 8.3 137 516 1578 52 13 PS 2.7 94151 33 39 7.8 168 673 1605 57 14 NDG 0.9 53800 29 40 7.2 156 117 749 33 5 NDVK 1.2 60526 30 37 7.0 124 176 923 34 6 BU 3.1 101997 28 31 6.8 314 503 1923 45 15 BD 2.9 99755 28 41 7.8 262 552 4198 42 26 NE 3.1 104239 27 44 7.1 85 86 2145 43 12 IA 1.7 72855 36 39 6.7 413 320 1701 36 13

78

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.2 Continued Additional soil chemical characteristics at 11 sites from where Platanthera praeclara roots were sampled in 2013 for mycorrhizal analyses. NO3, nitrate; S, Sulphur; Zn, zinc; Mn, manganese; Fe, iron; Cu, copper; B, boron; Soluble salts and soil textural components (sand, silt and clay).

Soil characteristics

NO3 S Zn Mn Fe Cu B Soluble salts Sand Silt Clay Site (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (mmhos/cm) (%) (%) (%)

PC 4 9.6 1.2 11.9 60 1.0 0.9 0.4 74 10 16 PH 5 7.8 1.6 10.4 53 1.2 0.7 0.3 72 10 18 PF 3 6.9 0.8 6.4 44 0.8 0.9 0.3 74 12 14 PS 4 6.8 1.5 13.0 89 0.7 0.9 0.4 66 20 14 NDG 6 6.7 0.7 9.4 35 0.5 1.1 0.2 84 6 10 NDVK 6 7.7 1.4 14.6 105 0.5 0.8 0.1 82 6 12 BU 4 7.1 2.5 28.9 241 1.2 1.3 0.1 38 42 20 BD 3 7.7 1.5 17.2 52 1.6 1.0 0.2 32 40 28 NE 3 10.7 2.1 30.1 166 0.7 1.2 0.3 72 14 14 IA 2 6.2 2.1 41.9 76 1.3 0.9 0.2 28 48 24

79

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.3 Soil chemical characteristics of 11 sites from where Platanthera praeclara roots were sampled in 2014 for mycorrhizal analyses. OM, organic matter; ENR, estimated nitrogen release; P1, phosphorus 1; P2, phosphorus 2; K, potassium; Mg, magnesium; Ca, calcium; Na, sodium; H, hydrogen and CEC, cation exchange capability. Please see next page for the remaining Table.

Soil characteristics

OM ENR P1 P2 K Mg Ca Na H CEC pH Site (%) (g/ha) (uL/L) (uL/L) (uL/L) (uL/L ) (uL/L) (uL/L) (uL/L) (meq/100g)

MB 3.4 110964 7.9 28 7.7 130 574 4497 50 0.0 28 PC 3.8 118810 14.3 38 6.9 141 665 2702 54 0.3 20 PH 4.0 123294 14.3 75 7.4 151 913 4407 72 0.0 30 PF 4.0 124411 15.3 36 7.0 161 698 2653 57 0.0 20 PS 4.4 133381 17.1 25 6.8 170 732 2300 65 0.6 19 NDG 1.5 67251 10.9 21 6.2 94 127 1152 36 1.0 8.0 NDVK 1.2 59405 21.5 33 6.2 260 192 841 31 0.9 7.0 BU 3.3 107602 14.4 15 5.8 326 632 3281 36 5.3 28 BD 3.2 105360 11.2 26 6.6 270 641 4720 40 1.9 32 NE 3.3 108723 10.0 14 5.8 112 84 2366 47 3.0 16 IA 3.3 106481 30.5 41 6.1 553 447 2950 37 3.3 23

80

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.3 Continued Additional soil chemical characteristics of 11 sites from where Platanthera praeclara roots were sampled in 2014 for mycorrhizal analyses. NO3, nitrate; S, Sulphur; Zn, zinc; Mn, manganese; Fe, iron; Cu, copper and B, boron; Soluble salts and soil textural components (sand, silt and clay).

Soil characteristics

NO3 S Zn Mn Fe Cu B Soluble salts Sand Silt Clay Site (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (mmhos/cm) (%) (%) (%)

MB 3 4.7 0.9 18 118 1.2 0.3 0.2 72 12 16 PC 1 4.1 1.3 30 144 1.8 0.8 0.3 74 14 12 PH 1 10.6 1.6 36 126 2.4 0.5 0.4 72 12 16 PF 1 11.0 1.5 19 93 1.8 0.7 0.3 74 12 14 PS 1 3.7 1.4 30 128 1.4 0.7 0.4 76 12 12 NDG 3 5.8 0.7 14 60 0.6 0.5 0.1 82 4 14 NDVK 4 5.4 1.6 5 32 0.3 0.4 0.1 82 8 10 BU 3 4.4 3.4 39 153 1.5 0.5 0.2 38 38 24 BD 3 3.3 1.7 33 70 2.0 0.5 0.1 36 38 26 NE 2 3.1 0.6 11 173 0.6 0.5 0.1 76 10 14 IA 4 3.9 3.4 43 78 1.5 0.4 0.1 36 40 24

81

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.4 Primer pairs utilized, oligo sequences, and annealing temperature to amplify the fungal ITS region of the nuclear ribosomal DNA via polymerase chain reaction from the DNA extracted from root fragments of Platanthera praeclara.

Annealing Paired Temperature Primer Sequence primer (°C)

ITS1 -OF Forward AACTCGGCCATTTAGAGGAAGT ITS4 -OF 58 or 52 (Mix these two primers) AACTTGGTCATTTAGAGGAAGT ITS4 -OF Forward GTTACTAGGGGAATCCTTGTT ITS1 -OF

ITS4-Tul Reverse CCGCCAGATTCACACATTGA ITS1 ITS1 Forward TCCGTAGGTGAACCTGCGG ITS4-Tul 54

82

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.5 Number of root sections (i.e. sequences) representing each of the 19 operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi at 11 sites in the year 2013. Values inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed.

Site

OTU MB PC PH PF PS NDG NDVK BU BD NE IA

Ceratobasidiaceae C1 2(2) 8(4) 6(3) 6(1) 7(3) 13(4) 8(4) 7(6)

C2 1(1) 1(1)

C3 1(1)

C4 1(1) 4(2) 1(1) 4(2) 5(2) 1(1) 1(1) 12(5) 9(3)

C5 1(1)

C6 2(1)

C7 1(1) 2(2)

C8 1(1) 3(1) 3(2) 1(1) 5(3) 1(1) 2(1)

C9 1(1)

Tulasnellaceae

T1 1(1)

T2 1(1)

T3 3(2) 13(3) 4(2)

T4 5(3) 5(2) 3(2) 1(1)

T5 4(2)

T6 2(1)

T7 2(2)

T8 1(1)

T9 1(1)

T10 1(1)

Total 10(6) 12(5) 9(4) 21(6) 12(5) 23(4) 5(2) 22(5) 16(4) 25(8) 13(5)

83

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.6 Number of root sections (i.e. sequences) representing each of the 19 operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi at 11 sites in the year 2014. Values inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed.

Site

OTU MB PC PH PF PS NDG NDVK BU BD NE IA

Ceratobasidiaceae C1 6(4) 2(2) 10(4) 14(2) 5(1) 15(6) 4(4) 5(4)

C2 2(1) 1(1)

C3 2(1)

C4 3(2)

C5 1(1)

C6 1(1)

C7 1(1)

C8 1(1) 1(1)

C9 2(1)

C10 2(1)

C11 7(4) 1(1) 3(3) 5(3) 1(1) 2(2) 4(2) 5(2)

C12 2(1)

C13 4(1)

C14 4(1)

Tulasnellaceae T1 2(1)

T2 1(1) 6(4) 1(1)

T3 2(1) 4(2) 1(1)

T4 1(1)

T5 3(1)

Total 16(8) 9(4) 18(4) 22(5) 17(7) 10(4) 22(6) 6(3) 4(4) 11(4) 6(3)

84

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.7 Number of root sections (i.e. sequences) representing each of the 39 operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi at 11 sites in the combined data set (years 2013 and 2014). Values inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed.

Site OTU MB PC PH PF PS NDG NDVK BU BD NE IA Ceratobasidiaceae C1 3(3) 8(3) 11(7) 15(3) 7(3) 17(5) 7(5) 12(8) 12(10) C2 3(2) C3 2(1) C4 2(1) C5 4(1) C6 1(1) 1(1) 8(2) 1(1) C7 1(1) 2(2) C8 2(1) 1(1) C9 4(4) C10 4(1) C11 1(1) 1(1) C12 2(2) 5(3) 6(5) 1(1) 5(4) 1(1) C13 1(1) C14 1(1) C15 1(1) C16 1(1) C17 1(1) C18 1(1) C19 1(1) C20 9(6) 2(2) 3(3) 7(3) 5(4) 5(3) 7(4) 1(1) 1(1) 16(7) 13(5) C21 1(1) C22 1(1) 1(1) C23 2(1) C24 2(1) C25 1(1) C26 1(1) C27 1(1) Tulasnellaceae T1 1(1) T2 6(5) 6(4) 5(2) 3(3) 1(1) T3 2(1) 2(1) 14(4) 4(2) 1(1) T4 5(1) T5 2(1) T6 1(1) T7 2(2) T8 2(1) 1(1) 1(1) T9 3(2) T10 1(1) T11 1(1) T12 1(1)

Total 27(13) 21(9) 25(8) 45(11) 29(11) 33(9) 23(8) 24(8) 21(8) 34(12) 19(8) 85

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.8 Mean pairwise sequence distances and standard errors (SE) for all observed fungal ITS sequences from roots of Platanthera praeclara sampled across 11 sites in years 2013 and 2014, and for the combined data set (years 2013 and 2014), separated by fungal family. The values presented within the parenthesis are the numbers of individual ITS sequences (n) composing each group.

Data set Fungal Family (n) Pi - Distance SE

2013 and 2014 Ceratobasidiaceae (238) 0.039 0.003 combined Tulasnellaceae (69) 0.135 0.006

Ceratobasidiaceae 2013 (122) 0.017 0.003 2013 Tulasnellaceae 2013 (48) 0.109 0.006

Ceratobasidiaceae 2014 (116) 0.072 0.005 2014 Tulasnellaceae 2014 (21) 0.073 0.006

86

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.9 Mean pairwise fungal ITS sequence distances (pi-distance) estimated based on combined transitions and transversions substitution model, within fungal families identified in roots of Platanthera praeclara across 11 sites in North America. Mean pairwise sequence distances from publically available fungal ITS sequences obtained from orchids other than Platanthera praeclara and Nervilia nipponica were calculated by Pandey et al., 2013.

Taxon Tulasnellaceae Ceratobasidiaceae Natural distribution n pi-distance n pi-distance

Platanthera praeclara 69 0.135 ± 0.006 238 0.039 ± 0.003 Wide Piperia yadonii 58 0.231 ± 0.026 71 0.077 ± 0.006 Restricted Anacampis laxiflora 12 0.186 ± 0.009 12 0.071 ± 0.006 Wide Ophrys fuciflora 12 0.225 ± 0.010 7 0.063 ± 0.006 Wide Orchis purpurea 9 0.089 ± 0.008 Wide Serapias vomeracea 27 0.097 ± 0.007 8 0.038 ± 0.004 Wide Chiloglottis aff. jeanesii 14 0.014 ± 0.003 Restricted Chiloglottis trapeziformis 12 0.006 ± 0.002 Wide Goodyera foliosa 7 0.062 ± 0.009 Wide Goodyera hachijoensis 5 0.103 ± 0.009 Moderate Goodyera tesselata 5 0.044 ± 0.006 Wide Cypripedium calceolus 12 0.005 ± 0.002 Moderate Cypripedium candidum 7 0.003 ± 0.002 Wide Cypripedium japonicum 18 0.033 ± 0.006 Wide Nervilia nipponica 9 0.157 ± 0.007 Wide Hexalectris grandiflora 6 0.023 ± 0.003 Moderate

87

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.10 Statistical significance (P values) from the Fisher’s exact test comparing the fungal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) at each site for 2013 below the diagonal. Pairwise geographic distances (kilometers) between population pairs are presented above the diagonal. MB, Manitoba; PC, Pembina Control; PH, Pembina Hay; PF, Pembina Fall fire; PS, Pembina Spring fire; NDG, North Dakota Graze; NDVK, North Dakota Viking; BU, Bluemound Uphill; BD, Bluemound Downhill; NE, Nebraska and IA, Iowa.

MB PC PH PF PS NDG NDVK BU BD NE IA

MB - 167 167 167 167 309 298 610 609 805 838

PC < 0.05 - 0 0 0 158 142 444 443 667 670

PH < 0.05 1.00 - 0 0 158 142 444 443 667 670

PF 0.111 < 0.05 < 0.05 - 0 158 142 444 443 667 670

PS 0.185 0.270 0.249 0.371 - 158 142 444 443 667 670

NDG < 0.05 0.922 0.671 < 0.05 0.308 - 22 319 319 509 553

NDVK < 0.05 < 0.05 < 0.05 0.050 < 0.05 < 0.05 - 321 321 527 554

BU < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 - 3 378 234

BD < 0.05 0.212 0.119 0.068 0.127 < 0.05 < 0.05 < 0.05 - 382 234

NE < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 0.649 < 0.05 < 0.05 - 443

IA < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 1.000 < 0.05 < 0.05 < 0.05 -

88

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.11 Statistical significance (P values) from the Fisher’s exact test comparing the fungal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) at each site for 2014 below the diagonal. Pairwise geographic distances (kilometers) between population pairs are presented above the diagonal. MB, Manitoba; PC, Pembina Control; PH, Pembina Hay; PF, Pembina Fall fire; PS, Pembina Spring fire; NDG, North Dakota Graze; NDVK, North Dakota Viking; BU, Bluemound Uphill; BD, Bluemound Downhill; NE, Nebraska and IA, Iowa.

MB PC PH PF PS NDG NDVK BU BD NE IA

MB - 167 167 167 167 309 298 610 609 805 838 PC < 0.05 - 0 0 0 158 142 444 443 667 670 PH 0.056 < 0.05 - 0 0 158 142 444 443 667 670 PF < 0.05 < 0.05 < 0.05 - 0 158 142 444 443 667 670 PS < 0.05 < 0.05 < 0.05 < 0.05 - 158 142 444 443 667 670 NDG 0.052 < 0.05 0.194 0.072 < 0.05 - 22 319 319 509 553 NDVK < 0.05 < 0.05 0.179 < 0.05 < 0.05 0.066 - 321 321 527 554 BU < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 - 3 378 234 BD 0.228 < 0.05 0.850 1.00 < 0.05 0.832 1.000 < 0.05 - 382 234 NE 0.648 < 0.05 0.340 < 0.05 < 0.05 0.517 0.102 < 0.05 0.384 - 443 IA 0.117 < 0.05 < 0.05 < 0.05 0.140 < 0.05 < 0.05 < 0.05 < 0.05 0.117 -

89

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.12 Statistical significance (P values) from the Fisher’s exact test comparing the fungal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) at each site using the combined data set (years 2013 and 2014) below the diagonal. Pairwise geographic distances (kilometers) between population pairs are presented above the diagonal. MB, Manitoba; PC, Pembina Control; PH, Pembina Hay; PF, Pembina Fall fire; PS, Pembina Spring fire; NDG, North Dakota Graze; NDVK, North Dakota Viking; BU, Bluemound Uphill; BD, Bluemound Downhill; NE, Nebraska and IA, Iowa.

MB PC PH PF PS NDG NDVK BU BD NE IA MB - 167 167 167 167 309 298 610 609 805 838 PC < 0.05 - 0 0 0 158 142 444 443 667 670 PH < 0.05 0.064 - 0 0 158 142 444 443 667 670 PF < 0.05 0.066 < 0.05 - 0 158 142 444 443 667 670 PS < 0.05 < 0.05 < 0.05 < 0.05 - 158 142 444 443 667 670 NDG < 0.05 0.114 0.381 0.05 < 0.05 - 22 319 319 509 553 NDVK < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 - 321 321 527 554 BU < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 - 3 378 234 BD < 0.05 < 0.05 < 0.05 0.06 < 0.05 0.281 < 0.05 < 0.05 - 382 234 NE < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 - 443 IA < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 < 0.05 -

90

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 2.13 Multiple response permutation procedure (MRPP) showing the significance of the variation of the mycorrhizal fungal communities by using abundance data observed at 11 sites of Platanthera praeclara by grouping the sites according to the type of land management (prescribed fire, haying or grazing, or no management). The MRPP was performed with the combined data set (years 2013 and 2014) as well as years 2013 and 2014 separately.

Combined data set (2013 and 2014) 2013 2014 p-value p-value p-value

Management 0.380 0.305 0.198 Latitude 0.034 0.070 0.192

91

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.1 A partial map of the US and Canada showing the natural distribution of Platanthera praeclara. The states where the plant is currently known to be present (orange), the counties where the populations occur (gray) and the counties where roots were collected for the studies (yellow) are highlighted.

92

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.2 A map of the nuclear fungal ribosomal internal transcribed spacer (ITS) region. The two ITS regions are between the SSU18S and LSU28S ribosomal RNA genes and are separated by the 5.8S rRNA gene. Image taken from http://www.gatc- biotech.com

93

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.3 Plants of Platanthera praeclara excavated for sample collection while keeping the root system intact. Photo on the left was taken by Nancy Sather.

94

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.4 A whole seedling of Platanthera praeclara along with severed root and tuber

95

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.5 Root fragments collected from plants of Platanthera praeclara.

96

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.6 A transversal section of a root of Platanthera praeclara showing colonization by fungal pelotons in cortical cells (examples indicated by arrows). The scale-bar represents 400 microns.

97

Texas Tech University, Pablo Antonio Tovar, December 2015

MB PC PH 0 0

2

5 12 9

PF PS NDG 1

3 10 11 9 22

NDVK BU BD 0 2 6 10 5 20

NE IA 0 Ceratobasidiaceae 2

Tulasnellaceae 25 11

Figure 2.7 Pie charts showing the distribution of the fungal families Ceratobasidiaceae and Tulasnellaceae observed in roots of Platanthera praeclara at 11 sites in the year 2013. Numeric values represent the individual ITS sequences belonging to each fungal family.

98

Texas Tech University, Pablo Antonio Tovar, December 2015

MB PC PH 0 3 3

6 13 18

PF PS NDG 1

3 2

19 16 8

NDVK BU BD 0 0

2

4 18 4

NE IA

Ceratobasidiaceae 1 1

Tulasnellaceae 10 5

Figure 2.8 Pie charts showing the distribution of the fungal families Ceratobasidiaceae and Tulasnellaceae observed in roots of Platanthera praeclara at 11 sites in the year 2013. Numeric values represent the individual ITS sequences belonging to each fungal family.

99

Texas Tech University, Pablo Antonio Tovar, December 2015

(a) 35 30

25 20 15

# of plants of # 10 5 0 1 2 3 4 # of OTUs

(b) 60 50

40 30

# of plants of # 20 10 0 1 2 # of fungal families

Figure 2.9 One to four operational taxonomic units in individual plants of Platanthera praeclara (a), and number of individual plants hosting one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b), across 11 sites in 2013.

100

Texas Tech University, Pablo Antonio Tovar, December 2015

(a) 50

40

30

20 # of plants of # 10

0 1 2 3 4 # of OTUs

(b) 60 50

40 30

# of plants of # 20 10 0 1 2 # of fungal families

Figure 2.10 One to four operational taxonomic units in individual plants of Platanthera praeclara (a), and number of individual plants that hosted one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b), across 11 sites in 2014.

101

Texas Tech University, Pablo Antonio Tovar, December 2015

(a) 50

40

30

20 # of plants of # 10

0 1 2 3 4 # of OTUs

(b) 60 50

40 30

# of plants of # 20 10 0 1 2 # of fungal families

Figure 2.11 One to four operational taxonomic units in individual plants of Platanthera praeclara (a), and number of individual plants that hosted one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b), across 11 sites for the combined data set (years 2013 and 2014).

102

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.12 Two-way hierarchical cluster tree and matrix coding based on abundance of 19 fungal operational taxonomic units (OTUs) observed at 11 sites where roots of Platanthera praeclara were sampled in the year 2013.

103

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.13 Two-way hierarchical cluster tree and matrix coding based on abundance of 19 fungal operational taxonomic units (OTUs) observed at 11 sites where roots of Platanthera praeclara were sampled in the year 2014.

104

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.14 Two-way hierarchical cluster tree and matrix coding based on abundance of 39 fungal operational taxonomic units (OTUs) observed at 11 sites where roots of Platanthera praeclara were sampled in years 2013 and 2014.

105

Texas Tech University, Pablo Antonio Tovar, December 2015

60

50

40

30

Mao Tau # ofOTUs # Chao 1 20

10

0 27 55 82 109 137 164 192 219 246 274 301 # of sequences

Figure 2.15 Sample based observed (Mao Tau) and rarefaction (Chao 1) cumulative fungal operational taxonomic unit (OTU) abundance curves for Platanthera praeclara. Fungal communities within the roots of plants were determined by sampling plants from 11 sites that host the orchid species.

106

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.16 Nonmetric multidimensional scaling (NMS) ordination of the mycorrhizal fungal communities within the roots of Platanthera praeclara sampled from 11 sites with correlations to soil variables [Na, Mn, Zn, Ca, OM (organic matter), CEC (cation exchange capability), soluble salts and pH] for the year 2013.

107

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.17 Nonmetric multidimensional scaling (NMS) ordination of the fungal communities of Platanthera praeclara sampled from 11 sites with correlations to soil variables (S, B, H, clay and silt) for the year 2014.

108

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.18 Partial Maximum likelihood tree of the fungal family Ceratobasidiaceae constructed with operational taxonomic unit (OTU) sequences observed in Platanthera praeclara roots collected from 11 sites across its natural habitat. Sequences of known orchid mycorrhizal fungi were added to the alignments to place our sequences in a broader context. Scale bar represents estimated number of DNA substitutions per site. Only branch support values over fifty percent are displayed (maximum likelihood / posterior probability). Please see the next page for the remaining tree. 109

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.18. Continued Maximum likelihood tree of the fungal family Ceratobasidiaceae constructed with operational taxonomic unit (OTU) sequences observed in Platanthera praeclara roots collected from 11 sites across its natural habitat. Only branch support values over fifty percent are displayed (maximum likelihood / posterior probability).

110

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.19a Maximum likelihood tree of the fungal family Tulasnellaceae constructed with operational taxonomic unit (OTU) sequences observed in Platanthera praeclara roots collected from eleven populations across its natural habitat. Sequences of known orchid mycorrhizal fungi were added to the alignments to place our sequences in a broader context. Scale bar represents estimated number of DNA substitutions per site. Only branch support values over fifty percent are displayed (maximum likelihood / posterior probability). 111

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.19b Maximum likelihood tree of the fungal family Tulasnellaceae constructed with operational taxonomic unit (OTU) sequences observed in Platanthera praeclara roots collected from eleven populations across its natural habitat. Sequences of known orchid mycorrhizal fungi were added to the alignments to place our sequences in a broader context. Scale bar represents estimated number of DNA substitutions per site. Only branch support values over fifty percent are displayed (maximum likelihood / posterior probability).

112

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 2.19c Maximum likelihood tree of the fungal family Tulasnellaceae constructed with operational taxonomic unit (OTU) sequences observed in Platanthera praeclara roots collected from eleven populations across its natural habitat. Sequences of known orchid mycorrhizal fungi were added to the alignments to place our sequences in a broader context. Scale bar represents estimated number of DNA substitutions per site. Only branch support values over fifty percent are displayed (maximum likelihood / posterior probability).

113

Texas Tech University, Pablo Antonio Tovar, December 2015

CHAPTER III

MYCORRHIZAL ASSOCIATIONS OF PLATANTHERA PRAECLARA IN RESPONSE TO PRESCRIBED FIRE AND HAYING

Abstract

Grassland plant communities are affected by vegetation management practices including fire, grazing, haying, etc. Consequently, edaphic characteristics within the system can also change in response. Changes in soil can lead to shifts in functioning and composition of microbial communities, including mycorrhizal fungi, which are critical in germination and survival of orchid species. We tested the effect of prairie management treatments on mycorrhizal fungal communities within the roots of

Platanthera praeclara which is a rare terrestrial orchid native to the tallgrass prairies in North America. Root sampling, along with characterizing the edaphic variables, was conducted over two years at a long-term experimental site in Minnesota, USA.

Across the two years, Ceratobasidiaceae and Tulasnellaceae were observed as the fungal families composing mycorrhizal fungal communities in the roots of P. praeclara. Twenty operational taxonomic units (OTUs) were observed through both years of sampling. Fifteen OTUs were identified for Ceratobasidiaceae and five for

Tulasnellaceae. Variation in the mycorrhizal fungal communities was observed from

2013 (four OTUs) to 2014 (15 OTUs). Significant differences (P<0.05) were observed only between the composition of fungal communities of PC (no management) and PF (fall burn treatment) in 2013. In 2014, the composition of fungal communities between all four treatments was significantly different (P<0.05).

114

Texas Tech University, Pablo Antonio Tovar, December 2015

Composition of mycorrhizal fungal communities did not appear to be associated with the type of prairie management treatment. However, temporal shifts appeared to be most noticeable under annual haying and both prescribed fire treatments by yielding an increase in the richness of OTUs. Further, fungal communities within orchid roots appeared to vary over time in response to edaphic variation across the four prairie management treatments. In 2013, sand and pH were correlated with the fungal community observed in treatment PC (no management); the fungal community in PH

(annual haying) was correlated with clay; silt and sodium (Na) were correlated to the fungal community observed in the treatment PS (spring fire); and the fungal community in PF (fall fire) was correlated with boron (B) and readily available phosphorus (P1). In 2014, the fungal community observed in treatment PC was correlated with volumetric water content in June (VWC J) and silt; sand, organic matter (OM) and P1 were correlated with the fungal community observed in treatment PS; clay and electrical conductivity in May (EC M) were correlated with the fungal community in PH; and sulfur was correlated with the fungal community observed in PF. We conclude that despite the relative specificity of the orchid toward its fungi (Ceratobasidiaceae π = 0.049 ± 0.04, N=97; Tulasnellaceae π = 0.203 ±

0.09, N=23), OTUs within the entire suite of its fungal associates change in response to time and management combinations.

115

Texas Tech University, Pablo Antonio Tovar, December 2015

Introduction

Anthropogenic factors continue to induce habitat destruction and degradation across the tallgrass prairies of the Great Plains in North America, which is one of most biodiverse ecosystems on earth. This ecosystem hosts many rare and endemic taxa including the federally threatened terrestrial orchid Platanthera praeclara (USFWS,

1996). It is known that management practices such as prescribed fire, grazing, haying, and/or fertilization can lead to changes in the ecosystem components above and below ground (Bentivenga et al., 1992; Dhilhon and Anderson, 1993; Bossio et al., 1998; Bossio et al., 2005). Habitat degradation and loss can also lead to loss of pollinators.

Additionally, local hydrology can change, and plant communities can shift over time. As the plant communities shift, soil microbial function and communities often shift concurrently. Soil respiration, metabolic profiles (Bendig et al., 2000) and enzyme activities (Bandick and Dick, 1999) are some of the measurements that have shown response to different types of land use because they rapidly respond to disturbances in the soil environment (Johnson et al., 2003). Land use changes may have significant and long- lasting effects on soil properties like carbon and nutrient content, soil texture and pH

(Murty et al., 2002). These effects are usually a consequence of changes in plant species composition. Changes in land use or management ultimately affect microbial communities (Bossio et al., 1998; Johnson et al., 2003; Wu et al., 2015). Studies have shown that shifts in the structure of fungal communities are associated with changes in soil properties such as texture (Girvan et al., 2003), soil pH (Fierer and Jackson, 2006) and soil nitrogen and phosphorus availability (Frey et al., 2004; Lauber et al., 2008).

Although these variables change depending on land use or management, it is not clear 116

Texas Tech University, Pablo Antonio Tovar, December 2015 how they affect biogeographical patterns in soil microbial communities (Lauber et al.,

2008).

Density of fungal propagules in soil is known to be affected by land management practices (Wu et al., 2007). Increased fungal hyphal biomass is usually observed in organically managed crop systems when compared to soil from conventionally managed systems (Sivapalan et al 1993; Shannon et al., 2002; Wu et al., 2007). Animal grazing, soil tillage and crop rotation had been observed to affect fungal densities in soil (Frey et al., 1999; Hedlund, 2002; Singh and Rai, 2004).

Mycorrhizal fungal communities may be susceptible to shifts in vegetation as they usually associate with specific plant types (Heinemeyer et al., 2004). For example, fungi from the phylum Basidiomycota that are usually involved in decomposing lignified plant detritus might be more affected by turning forest soils into agricultural soils (Bardgett and McAlister, 1999). Bentivenga and Hetrick (1992) examined arbuscular mycorrhizal

(AM) fungal species composition in tallgrass prairie subjected to annual fire, mowing and fertilization treatments, but no significant effects of these management types were observed on the AM fungal composition. On the other hand, Dhillion and Anderson

(1993) have shown that AM fungal communities in grasslands are influenced by application of fire, mowing, grazing and fertilization practices. Moreover, mycorrhizal root colonization has been reported to increase (Bentivenga and Hetrick, 1991), decrease

(Rashid et al., 1997) and be unaffected (Anderson and Menges, 1997) in response to prescribed fire. These conflicting results may suggest that fungal communities likely respond to a combination of factors including management and edaphic characters (Eom 117

Texas Tech University, Pablo Antonio Tovar, December 2015 et al., 1999). Also, the response time to allow detection of the shifts in communities may vary depending on the vegetation and / or fungal community.

Although an increasing number of mycorrhizal fungal communities are being described and characterized from roots of orchids, there is no evidence for how these communities are affected by vegetation management practices. We utilized a long-term

(over 20 years) experimental set-up at Pembina Trail Preserve in Polk County,

Minnesota, USA, where one of the largest populations of P. praeclara occurs to test the hypothesis that the diversity of fungi associated with P. praeclara will be similar across four management treatments (i.e., prescribed fire in spring every 4 years, prescribed fire in fall every 4 years, annual haying, and no active vegetation management) over the two sampling years because of the expectation of overall high mycorrhizal specificity in the species.

Materials and methods Collection sites

Root tissue from individual plants of P. praeclara were sampled at the Pembina

Trail Preserve in Polk County, Minnesota, USA. Four experimental treatment plots with different land management [prescribed fire in spring every 4 years (PS), prescribed fire in fall every 4 years (PF), annual haying (PH), and no vegetation management (control;

PC)] were sampled to allow a comparison of fungal associations under each prairie

118

Texas Tech University, Pablo Antonio Tovar, December 2015 management treatment (Figure 3.1). The most recent fall burn was applied in November of 2011, whereas the last spring burn was applied in April, 2012.

Sample collection

In 2013 and 2014, roots from individual plants were sampled between mid-June and early July. By this time in the growing season, differentiating between seedlings, juveniles or vegetative (individuals that will not produce an inflorescence in the growing season), and reproductive individuals was possible. Roots from two seedlings, two vegetative and two reproductive plants were sampled within each prairie management treatment plot. Roots from one or two extra plants were sampled with permission.

Twenty-four and 20 plants were sampled across the 11 sites during 2013 and 2014, respectively (Table 3.3). Whole plants were carefully excavated to examine the complete root system (see Figure 2.3 in Chapter II). To keep plants alive after sampling, just a few roots were extracted from the plant, and the sampled plant was replanted. Sampled plants were watered thoroughly immediately after replanting to maximize reestablishment.

Because seedlings have very small root systems, the whole seedling was collected and all roots were processed for fungal identification (see Figure 2.4 in Chapter II). Roots were stored at 4°C until they were shipped overnight to the laboratory at Texas Tech

University for further processing.

In each of the two years, one aggregate soil sample was collected at each treatment where orchids were sampled by adding together at least three subsamples. Approximately

119

Texas Tech University, Pablo Antonio Tovar, December 2015

500g of soil were extracted from sites where the plants were present. The soil was stored at room temperature until it was processed further. Physicochemical analyses were carried out by A&L Plains Labs, Inc., Lubbock, Texas. Organic matter (OM), pH, macro nutrients (K, Mg, Ca, Na), cation exchange capacity (CEC), NO3-N, other nutrients (S,

Zn, Mn, Fe, Cu, B) and textural components (% sand, silt and clay) were measured in each sample.

Environmental data

A cr800 data logger (Campbell Scientific, Salt Lake City, Utah) was installed with sensors recording continuous data (one reading per hour) on soil temperature, soil volumetric water content, and soil electrical conductivity. The data logger was set up at a location from where probes could simultaneously gather information in each of the four different treatment plots. Two sensors were utilized as replicates in each treatment plot.

Averages of the volumetric water content (VWC), electrical conductivity (EC) and soil temperature at orchid root depth were computed over the two replicate sensors for April,

May and June. We selected these months because the soil temperatures were below freezing prior to April and presumed to induce minimal or no mycorrhizal (plant or fungus) activity. Similarly, data beyond June were not used because root sampling occurred in the month of June (Table 3.1).

120

Texas Tech University, Pablo Antonio Tovar, December 2015

Fungal identification from root sections

Roots were rinsed under tap water to wash them free of soil and debris before photo-documentation (see Figure 2.5 in Chapter II). Roots from each individual plant were subsequently placed in individual 50 mL plastic tubes. Inside a laminar flow hood and under sterile conditions, the surface sterilization was carried out by following these steps: (1) roots were first rinsed in a 70% EtOH (ethanol) solution by shaking thoroughly for 50 seconds; (2) roots were then rinsed with a 3% NaOCl (sodium hypochlorite) solution by shaking thoroughly for 30 seconds; (3) a 70% ethanol rinse for 50 seconds was carried out; (4) roots were finally rinsed with ultrapure, sterile water several times to remove residues of EtOH and NaOCl. Surface sterilization treatment was modified as needed based on the thickness and the number of the roots (Pandey et al., 2013).

The epidermis of each root was gently shaved off using a sterile scalpel to remove additional microbes that might have survived the surface sterilization treatment. Peloton inspection was conducted by cutting transversal thin slices of root tissue and examining under a compound microscope (see Figure 2.6 in Chapter II). If pelotons were observed,

2.5 to 3 cm long segments adjacent to the inspected slice were collected, sliced thinly and stored individually in 2 mL plastic tubes at -80°C until DNA was extracted (Pandey et al.,

2013).

DNA extraction was conducted using Qiagen DNeasy Plant Mini Kit (Qiagen,

Valencia, CA, USA) by following the manufacturer's protocol with a slight modification.

All samples were incubated in a 3.3% solution of polyvinylpyrrolidone in AP1 lysis buffer at 65°C for 2 hours while mixing once every 30 minutes before the RNase 121

Texas Tech University, Pablo Antonio Tovar, December 2015 digestion step. Subsequently, 4 μl of RNase (supplied with the kit) was added to each sample and samples were incubated at 65°C for 15 minutes, mixing once every 4 minutes. Total DNA was eluted in 50 μl of elution buffer and stored at -80°C until further processing.

PCR and sequencing

Polymerase chain reaction was performed using total DNA extracted from individual root segments. Each reaction was prepared to a final volume of 25 μL by using

Promega GoTaq Flexi DNA Polymerase reagent kit (Promega, Madison, Wisconsin,

USA). Concentrations for each reagent were: 5x Green GoTaq Flexi Buffer, 10 mM/μL dNTPs (100 mM of each dNTP), 25 mM of MgCl2, 10 μM/μL of each primer, 10 μg/μL of BSA and 5 u/μL GoTaq DNA polymerase (Pandey et al., 2013). PCR consisted of 35 cycles in an epGradients Master Cycler (Eppendorf, Hamburg, Germany) and included a

2 minute initial denaturation at 94°C before thermocycling with a 45 seconds denaturation at 94°C followed by a 45-second annealing at different temperatures depending on the primer pair (ITS1-OF and ITS4-OF; ITS4-TUL and ITS1) used (Table

2.4 in Chapter II) and 72°C elongation for 1 minute. Finally, the last cycle was followed by an extension at 72°C for 5 minutes.

Two primer pairs ITS1-OF/ITS4-OF and ITS1/ITS4-TUL (Taylor & McCormick,

2008) were used. A total of 498 root segments were assayed for molecular analysis. Each

DNA sample was first subjected to PCR using the ITS-OF primer pair at 58°C annealing

122

Texas Tech University, Pablo Antonio Tovar, December 2015 temperature. Samples that did not amplify were then run with the same primer pair at

52°C annealing temperature. If amplification was still not obtained, ITS1/ITS4-TUL was used with 54°C annealing temperature on the samples that did not amplify previously. A

2% agarose gel electrophoresis was run to verify amplification. Samples showing a single band of the expected size range (600 – 800 bp) were cleaned using DNA Clean and

Concentrator 5 kit (Zymo Research, Irvine, CA, USA). DNA quantity and quality were measured using a NanoDrop 2000c spectrophotometer (Thermo Scientific, Wilmington,

DE, USA). Samples showing multiple bands were processed using GenElute Gel

Extraction Kit (Sigma-Aldrich, Poole, United Kingdom) by isolating the band closer to the expected size range. Sanger sequencing of all cleaned PCR products was carried out by the DNA Analysis Facility on Science Hill at Yale University. Samples were sequenced in one direction using the reverse primer (ITS4-OF or ITS4-TUL) (Pandey et al., 2013).

Cloning

Polymerase chain reaction products from samples with sequencing difficulties were purified either using PureLink PCR Purification kit (Invitrogen Life Technologies,

Carlsbad, California) or GenElute Gel Extraction Kit (Sigma-Aldrich, Poole, United

Kingdom). Once a clean PCR product was obtained, it was used to carry out a cloning reaction and cloning transformation using the TOPO TA Cloning kit for Sequencing along with One Shot TOP10 and DH5α-T1 competent cells (Invitrogen Life technologies,

Carlsbad, California), following the manufacturer protocol. 123

Texas Tech University, Pablo Antonio Tovar, December 2015

To analyze the transformants, 3-5 colonies were selected and cultured overnight in

1 mL of LB (Luria-Bertani) (ThermoFisher Scientific, Waltham, MA, USA) medium containing 50 μg/mL of ampicillin. Plasmid DNA was isolated using the PureLink Quick

Plasmid Miniprep Kit, (Invitrogen Life Technologies, Carlsbad, CA) following the protocol included in the kit.

Polymerase chain reaction was performed on the extracted plasmid DNA following the amplification protocol described earlier using ITS1-OF/ITS4-OF primer pair.

Polymerase chain reaction products were purified using GenElute Gel Extraction Kit

(Sigma-Aldrich, Poole, United Kingdom) and DNA concentrations were measured using a NanoDrop 2000c spectrophotometer (Thermo Scientific, Wilmington, DE, USA).

Sanger sequencing of all cleaned PCR products was carried out as described above.

Data analyses

DNA Analysis Facility on Science Hill at Yale University performed a basic trimming on the sequences. However, the obtained sequences were double-checked and a more thorough trim was implemented using Geneious 4.8.5. Trimmed sequences were then identified through BLAST searches using the Megablast option for finding highly similar

DNA sequences.

OTU clustering

Sequences were grouped into operational taxonomic units (OTUs) at 97% sequence similarity criterion using the OTU pipeline from the University of Alaska at

Fairbanks Life Science Informatics Portal (http://www.borealfungi.uaf.edu; Taylor &

124

Texas Tech University, Pablo Antonio Tovar, December 2015

Houston 2011). The longest and highest quality representative sequence from each OTU was used for further analyses. Operational taxonomic Units were obtained separately for sequences obtained from root fragments collected in year 2013 and 2014. Additionally, for general diversity estimation, sequences were also clustered into OTUs by combining the data sets for both years.

Mean pairwise sequence distance (pi, π)

To estimate sequence divergence within the identified fungal families, separate multiple alignments of sequences were constructed for years 2013 and 2014 data sets, and for the combined data set, using MAFFT version 7 with L-INS-I model (Katoh and

Standley, 2013). Mean pairwise sequence distance (pi, π; Nei and Kumar, 2000) among all individual sequences within a fungal family was then estimated using MEGA 6

(Tamura et al., 2013). Mean pairwise sequence distances obtained using the combined 2- year data set from P. praeclara mycorrhizal fungi were compared with the previously reported pi distances among mycorrhizal fungi from other orchids (Pandey et al., 2013).

Sequence divergence for the observed sequences groups by treatment

(independently of the fungal family they belonged to) were also computed.

Fisher’s exact test

The statistical significance of the variation in fungal diversity between treatments was tested using the Fisher’s exact test based on abundances of individual OTUs across 125

Texas Tech University, Pablo Antonio Tovar, December 2015 the sampled treatments per year of collection. The test was performed in R (R

Development Core Team 2010) using all the OTUs detected in each of the sampled treatment for 2013 and 2014 separately and also for the combined data set.

Two-way hierarchical ordination

Two-way hierarchical ordination analysis (Bray and Curtis., 1957) using OTU abundances was performed to test the relationships between the mycorrhizal fungal communities of the sites based on dissimilarities. PC-ORD (McCune et al., 2002) was used to perform the two-way hierarchical ordination using the Sorensen (Bray-Curtis) distance measure, and setting the group linkage criterion as ‘average’. This test was conducted for each sampling year separately and also for the combined 2-year data set.

The resulting dendrogram clustered sites according to similarity in mycorrhizal fungal communities held by the sampled plants across the four treatments, and served as an aid to observe whether sites sharing land management types (prescribed burn, haying or grazing, no management) clustered together.

Nonmetric multidimensional scaling

Nonmetric multidimensional scaling (NMDS) in PC-ORD (McCune et al., 2002) was used to ordinate OTU abundance in roots of P. praeclara and to visualize the correlation between edaphic characteristics [OM, organic matter; ENR, estimated nitrogen release; P1, phosphorus 1 (readily available to plants); P2, phosphorus (readily

126

Texas Tech University, Pablo Antonio Tovar, December 2015 available to plants plus part of the active reserve in the soil) 2; K, potassium; Mg, magnesium; Ca, calcium; Na, sodium; CEC, cation exchange capability; NO3, nitrate; S,

Sulphur; Zn, zinc; Mn, manganese; Fe, iron; Cu, copper and B, boron) and OTUs observed at each sampled site separately for each sampling year. The abundance-based version of the Sorensen index (Bray-Curtis) was used to calculate the distances.

Dimensionality of the ordination was determined to choose the lowest dimensionality that captured most of the variation. Data sets from 2013 and 2014 were both best described by

3-dimensional solutions with instabilities below 0.00001.

Results Operational taxonomic Units diversity

Out of 161 root fragments in 44 plants obtained from across all four treatments, 120 root fragments in 39 plants produced workable DNA sequences (Table 3.3).

Ceratobasidiaceae and Tulasnellaceae were identified as the mycorrhizal fungal families associated to Platanthera praeclara across the sampled land management treatments in both sampling years (Figures 3.2 and 3.3).

2013

Four OTUs representing both mycorrhizal fungal families were identified for samples collected in the year 2013 (Table 3.4). Tulasnellaceae was represented by 3

OTUs, and Ceratobasidiaceae by 1 OTU. Across all plants and treatments, infection of a single root system by 2 to 3 fungal OTUs was observed in 33% of all sampled plants 127

Texas Tech University, Pablo Antonio Tovar, December 2015

(Figure 3.4a). The remaining 67% of plants contained only one OTU. A majority of the plants (18 of 19 plants) possessed fungi belonging to a single family, while only one plant had fungi belonging to the two families (Figure 3.4b).

2014

Ceratobasidiaceae and Tulasnellaceae were represented by 13 and 2 OTUs respectively (Table 3.5). Across all plants and land management types, infection of a single root system by 2 to 4 fungal OTUs was observed in 35% of all sampled plants. The remaining 65% of plants contained only one OTU. A majority of the plants (18 of 20 plants) possessed fungi belonging to a single family, while only two plants had fungi belonging to the two families (Figure 3.5a and 3.5b).

2013 and 2014 combined

Twenty OTUs representing both fungal families were identified when data sets for

2013 and 2014 where pooled (Table 3.6). Fungal family Ceratobasidiaceae was represented by 15 OTUs and Tulasnellaceae was represented by 5 OTUs. Across all plants and populations, infection of a single root system by 2 to 4 fungal OTUs was observed in 48% of all sampled plants. The remaining 52% of plants contained only one

OTU. A majority of the plants (36 of 39 plants) possessed fungi belonging to a single family, while only three plants had fungi belonging to the two families (Figure 3.6a and

3.6b).

Mean pairwise sequence distance (pi, π)

128

Texas Tech University, Pablo Antonio Tovar, December 2015

Mean pairwise distance of sequences from Tulasnellaceae were always larger than those for Ceratobasidiaceae, regardless of the year of sampling or by combining the data sets for both years (Table 3.7). Pairwise sequence distance was also measured by treatment per year (Table 3.8). Both fire treatments (PS and PF) had a larger sequence distance in 2013 than the annual hay treatment or the control treatment (πPS = 0.201 ±

0.008 and πPF = 0.295 ± 0.01). This suggests a wider mycorrhizal fungal diversity in the roots of P. praeclara collected in the plots with prescribed fires. On the other hand, the treatments with a large sequence distance for the year 2014 were prescribed spring fire, prescribed fall fire and the control treatment.

Fisher’s exact test

The family Ceratobasidiaceae was represented in roots of P. praeclara from all populations sampled in the year 2013 and 2014. The fungal family Tulasnellaceae was not represented in two of the sampled treatments (PC, PH) in 2013. In the year 2014,

Tulasnellaceae was not observed in roots of one of the sampled treatments (PH). The

Fisher’s exact test revealed that mycorrhizal fungal communities were mostly similar in the year 2013 between treatments, except for the fungal communities in PF and PC (P <

0.05) (Table 3.9). Conversely, the communities observed in 2014 sampling are all significantly different (P < 0.05) (Table 3.10). Using the combined data set, the Fisher’s exact test for count data showed significant differences in all the treatments except between PH and PS (P > 0.05) (Table 3.11).

129

Texas Tech University, Pablo Antonio Tovar, December 2015

Two-way hierarchical ordination

The treatment-wise two-way hierarchical ordination analysis based on the OTUs observed per year of sampling showed clustering between the treatments that were not significantly different according to the Fisher’s exact test for count data (Figures 3.7 and

3.8).

The two-way hierarchical ordination analysis was also run for the combined data set (Figure 3.9). Plots where prescribed fires were applied clustered together along with the annual haying treatment. This suggests that fungal communities were more similar between the fire treatments and the annual haying, and were more different from the control land management treatment.

Nonmetric multidimensional scaling

Nonmetric multidimensional scaling ordination analysis of fungal communities in samples collected in 2013 from the different land management treatments showed that distinction in orchid-associated fungi is most strongly related to sand and pH at PC (no management plot). Clay appeared to have a strong correlation with the mycorrhizal fungal community observed at PH (annual haying treatment). Silt and Na presented correlation with the fungal community observed at PF (fall burn treatment) (Figure 3.10).

In 2014, mycorrhizal fungal community observed in PH was correlated with clay and electrical conductivity measured in May. Mycorrhizal fungal community observed in PF appeared to have a strong correlation with S. Volumetric water content measured in June and silt presented correlation with the fungal community observed in PC. Sand and P presented correlation with the fungal community in PS. Other edaphic variables and 130

Texas Tech University, Pablo Antonio Tovar, December 2015 environmental data presented strong variables but no evident correlation with any of the sampled sites (Figure 3.11).

Discussion

Across the two years of sampling, Platanthera praeclara associated with

Ceratobasidiaceae in each of the four land management treatments. The fungal family

Tulasnellaceae was not present in P. praeclara roots in PC (no management plot) in 2013 and was absent from PH (annual haying treatment) in both sampling years. Mean pairwise sequence distance separated by family (Table 3.7) suggested that due to the narrow distances within nrITS sequences in Ceratobasidiaceae, P. praeclara plants were specific towards its associations with this fungal family. Given the sequence distances within Tulasnellaceae, associations with the Tulasnellaceae were broader. Pi – distances were also calculated between sequences within each of the fungal families represented in each of the four management treatment (Table 3.8). In 2013, distances were much larger for both fire treatments (i.e., spring and fall burn) than distances from the haying and control treatments; the latter two were only composed by sequences belonging to the

Ceratobasidiaceae. In the year 2014, fungi from Tulasnellaceae were not detected in the annual haying treatment (PH), therefore presenting the lowest sequence distance between treatments for the year. These findings suggest that burn treatments might favor characteristics in the vegetation community or in the soil characteristics that allow for a greater microbial diversity within plant roots. Bleho et al. (2015) reported that P. praeclara plants might benefit from burn treatments applied once every 2 to 3 years. 131

Texas Tech University, Pablo Antonio Tovar, December 2015

When comparing the fungal communities among treatments, the pairwise Fisher’s test revealed that in the year 2013, only PF presented a significantly different mycorrhizal fungal community in orchid roots as compared to PC, while the rest of the pairwise comparisons were not different (Table 3.9). Conversely, the four land management types presented significant differences in their mycorrhizal fungal communities in the year

2014 (Table 3.10) indicating temporal variation in mycorrhizal preferences of the orchid.

Although there is still high specificity according to the mean pairwise sequence distances, fungal community assemblies within the orchid roots from each treatment showed large variation. In 2013, four OTUs were observed across the four treatments. In the same year, the fungal family Tulasnellaceae presented three OTUs and Ceratobasidiaceae presented on. However, the single Ceratobasidiaceae OTU was observed in each treatment, whereas

Tulasnellaceae occurred only in the two burn treatments. In 2014, 15 OTUs (13 from

Ceratobasidiaceae) were observed across all treatments. These temporal differences suggest that fungal community assemblies in plant roots can shift drastically from one growing season to another, as has been previously recorded in other studies (Twieg et al.,

2007; Drumbell et al., 2011; Cowden and Peterson, 2013; Ercole et al., 2015).

The hierarchical clustering analysis suggested segregation of the mycorrhizal fungal communities by clustering together the communities observed in the prescribed fire treatments, and separating the fungal communities observed in the annual haying and control treatments when the total diversity from the combined data sets was analyzed

(Figure 3.9). However, the Fisher’s test for the combined data set suggested that mycorrhizal fungal community composition of PS and PH did not have significant

132

Texas Tech University, Pablo Antonio Tovar, December 2015 differences. These inconclusive patterns implied that perhaps a stronger sampling effort adding more replicates of each treatment should be carried out to make the conclusions more robust.

Land management can affect soil moisture, temperature, carbon, nitrogen, and pH, etc., which are variables that are known to affect fungal communities in grasslands

(Bever et al., 1996). Annual spring fires in tallgrass prairie significantly increase soil temperature and alter soil moisture during the growing season (Knapp and Seastedt,

1986). However, fire treatments are applied to the experimental plots at Pembina Trail plots once every four years, and over one year had already passed by the time plants were sampled for this study in 2013. Regardless, management treatments can have long-lasting effects on edaphic characteristics (Post and Mann, 1990; Murty et al., 2002). A multi- year, long-term study of soil physicochemical profiles across multiple replicates of management treatments can likely improve our understanding of the effects management practices may have on the edaphic characteristics over space and time.

Components of soil texture, which have been associated with alterations in fungal communities previously (Girvan et al., 2003) appeared to be strongly correlated with the mycorrhizal fungal communities in both years 2013 and 2014 according to the nonmetric multidimensional scaling analysis. Phosphorus, pH, nitrates and other variables that have been associated to fungal communities are represented as strong vectors but given that only one replicate plot was sampled to represent each management treatment, strong conclusions cannot be drawn. With the use of a larger number of replicates, soil variables could be correlated to type of land managements although vegetation and soil 133

Texas Tech University, Pablo Antonio Tovar, December 2015 composition also varies in relation to other ecosystem features such as slope, water table, and proximity to edges, etc. Regardless, while the experimental prairie management treatments have been in place for over 20 years at Pembina Trail Preserve, we report the first comparison of the mycorrhizal fungal communities associated with P. praeclara under individual management treatments and over time.

134

Texas Tech University, Pablo Antonio Tovar, December 2015

Literature Cited

Anderson, R., & Menges, E. (1997). Effects of fire on sandhill herbs: nutrients, mycorrhizae, and biomass allocation. American Journal of Botany, 84(7), 938- 938.

Bandick, A.K., & Dick, R.P. (1999). Field management effects on soil enzyme activities. Soil Biology and Biochemistry, 31(11), 1471-1479.

Bending, G.D., Putland, C., & Rayns, F. (2000). Changes in microbial community metabolism and labile organic matter fractions as early indicators of the impact of management on soil biological quality. Biology and Fertility of Soils, 31(1), 78- 84.

Bardgett, R.D., & McAlister, E. (1999). The measurement of soil fungal: bacterial biomass ratios as an indicator of ecosystem self-regulation in temperate meadow grasslands. Biology and Fertility of Soils, 29(3), 282-290

Bentivenga, S.P., & Hetrick, B.A.D. (1991). Relationship between mycorrhizal activity, burning, and plant productivity in tallgrass prairie.Canadian Journal of Botany, 69(12), 2597-2602.

Bever, J.D., Morton, J.B., Antonovics, J., & Schultz, P.A. (1996). Host-dependent sporulation and species diversity of arbuscular mycorrhizal fungi in a mown grassland. Journal of ecology, 71-82.

Bissett, A., Richardson, A. E., Baker, G., & Thrall, P. H. (2011). Long-term land use effects on soil microbial community structure and function. Applied soil ecology, 51, 66-78.

Bleho, B. I., Koper, N., Borkowsky, C. L., & Hamel, C. D. (2015). Effects of Weather and Land Management on the Western Prairie Fringed-orchid (Platanthera praeclara) at the Northern Limit of its Range in Manitoba, Canada.The American Midland Naturalist, 174(2), 191-203.

Bossio, D.A., Scow, K.M., Gunapala, N., & Graham, K.J. (1998). Determinants of soil microbial communities: effects of agricultural management, season, and soil type on phospholipid fatty acid profiles.Microbial ecology, 36(1), 1-12.

Bossio, D.A., Girvan, M.S., Verchot, L., Bullimore, J., Borelli, T., Albrecht, A., Scow, K.M., Ball, A.S., Pretty, J.N. & Osborn, A. M. (2005). Soil microbial community response to land use change in an agricultural landscape of western Kenya. Microbial ecology, 49(1), 50-62. 135

Texas Tech University, Pablo Antonio Tovar, December 2015

Bray, J. R., & Curtis, J. T. (1957). An ordination of the upland forest communities of southern Wisconsin. Ecological monographs, 27(4), 325-349.

Certini, G. (2005). Effects of fire on properties of forest soils: a review.Oecologia, 143(1), 1-10.

Cowden, C. C., & Peterson, C. J. (2013). Annual and seasonal dynamics of ectomycorrhizal fungi colonizing white pine (Pinus strobus) seedlings following catastrophic windthrow in northern Georgia, USA. Canadian Journal of Forest Research, 43(3), 215-223.

Dhilhon, S.S., & Anderson, R.C. (1993). Growth dynamics and associated mycorrhizal fungi of little bluestem grass [Schizachyrium scoparium (Michx.) Nash] on burned and unburned sand prairies. New phytologist, 123(1), 77-91.

Docherty, K. M., Balser, T. C., Bohannan, B. J., & Gutknecht, J. L. (2012). Soil microbial responses to fire and interacting global change factors in a California annual grassland. Biogeochemistry, 109(1-3), 63-83.

Dumbrell, A.J., Ashton, P.D., Aziz, N., Feng, G., Nelson, M., Dytham, C., & Helgason, T. (2011). Distinct seasonal assemblages of arbuscular mycorrhizal fungi revealed by massively parallel pyrosequencing. New Phytologist, 190(3), 794-804.

Eom, A.H., Hartnett, D.C., Wilson, G.W., & Figge, D.A. (1999). The effect of fire, mowing and fertilizer amendment on arbuscular mycorrhizas in tallgrass prairie. The American Midland Naturalist, 142(1), 55-70.

Ercole, E., Adamo, M., Rodda, M., Gebauer, G., Girlanda, M., & Perotto, S. (2015). Temporal variation in mycorrhizal diversity and carbon and nitrogen stable isotope abundance in the wintergreen meadow orchid Anacamptis morio.New Phytologist, 205(3), 1308-1319.

Fierer, N., & Jackson, R.B. (2006). The diversity and biogeography of soil bacterial communities. Proceedings of the National Academy of Sciences of the United States of America, 103(3), 626-631.

Frey, S.D., Elliott, E.T., & Paustian, K. (1999). Bacterial and fungal abundance and biomass in conventional and no-tillage agroecosystems along two climatic gradients. Soil Biology and Biochemistry, 31(4), 573-585.

Frey, S.D., Knorr, M., Parrent, J.L., & Simpson, R.T. (2004). Chronic nitrogen enrichment affects the structure and function of the soil microbial community in

136

Texas Tech University, Pablo Antonio Tovar, December 2015

temperate hardwood and pine forests. Forest Ecology and Management,196(1), 159-171.

Girvan, M.S., Bullimore, J., Pretty, J.N., Osborn, A.M., & Ball, A.S. (2003). Soil type is the primary determinant of the composition of the total and active bacterial communities in arable soils. Applied and Environmental Microbiology,69(3), 1800-1809.

Heinemeyer, A., Ridgway, K. P., Edwards, E. J., Benham, D.G., Young, J.P.W., & Fitter, A.H. (2004). Impact of soil warming and shading on colonization and community structure of arbuscular mycorrhizal fungi in roots of a native grassland community. Global Change Biology, 10(1), 52-64.

Hedlund, K. (2002). Soil microbial community structure in relation to vegetation management on former agricultural land. Soil Biology and Biochemistry, 34(9), 1299-1307.

Johnson, M.J., Lee, K.Y., & Scow, K.M. (2003). DNA fingerprinting reveals links among agricultural crops, soil properties, and the composition of soil microbial communities. Geoderma, 114(3), 279-303.

Katoh, K., & Standley, D. M. (2013). MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Molecular biology and evolution, 30(4), 772-780.

Knapp, A.K., & Seastedt, T.R. (1986). Detritus accumulation limits productivity of tallgrass prairie. BioScience, 662-668.

Larsson, A. (2014). AliView: a fast and lightweight alignment viewer and editor for large datasets. Bioinformatics, 30(22), 3276-3278.

Lauber, C.L., Strickland, M.S., Bradford, M.A., & Fierer, N. (2008). The influence of soil properties on the structure of bacterial and fungal communities across land-use types. Soil Biology and Biochemistry, 40(9), 2407-2415.

McCune, B., Grace, J. B., & Urban, D. L. (2002). Analysis of ecological communities (Vol. 28). Gleneden Beach, OR: MjM software design.

Murty, D., Kirschbaum, M.U., Mcmurtrie, R.E., & Mcgilvray, H. (2002). Does conversion of forest to agricultural land change soil carbon and nitrogen? A review of the literature. Global Change Biology, 8(2), 105-123.

Post, W.M., & Mann, L.K. (1990). Changes in soil organic carbon and nitrogen as a result of cultivation. Soils and the greenhouse effect, 401-406.

137

Texas Tech University, Pablo Antonio Tovar, December 2015

Rashid, A., Ahmed, T., Ayub, N., & Khan, A.G. (1997). Effect of forest fire on number, viability and post-fire re-establishment of arbuscular mycorrhizae.Mycorrhiza, 7(4), 217-220.

Shannon, D., Sen, A.M., & Johnson, D.B. (2002). A comparative study of the microbiology of soils managed under organic and conventional regimes. Soil Use and Management, 18(s1), 274-283

Singh, S.K., & Rai, J.P. (2004). Soil microbial population and enzyme activity related to grazing pressure in alpine meadows of Nanda Devi Biosphere Reserve. Journal of environmental biology/Academy of Environmental Biology, India, 25(1), 103-107.

Sivapalan, A., Morgan, W.C., & Franz, P.R. (1993). Monitoring populations of soil microorganisms during a conversion from a conventional to an organic system of vegetable growing. Biological Agriculture & Horticulture, 10(1), 9-27.

Tamura, K., Stecher, G., Peterson, D., Filipski, A., & Kumar, S. (2013). MEGA6: molecular evolutionary genetics analysis version 6.0. Molecular biology and evolution, 30(12), 2725-2729.

Thomson, B. C., Tisserant, E., Plassart, P., Uroz, S., Griffiths, R. I., Hannula, S. E., Buee, M., Mougel, C., Ranjard, L., Van Veen, J.A., Martin, F., Bailey, M. & Lemanceau, P. (2015). Soil conditions and land use intensification effects on soil microbial communities across a range of European field sites.Soil Biology and Biochemistry, 88, 403-413.

Twieg, B.D., Durall, D.M., & Simard, S.W. (2007). Ectomycorrhizal fungal succession in mixed temperate forests. New Phytologist, 176(2), 437-447.

U.S. Fish and Wildlife Service. (1996). Platanthera praeclara (western prairie fringed orchid) recovery plan. Ft. Snelling, MN: US Fish and Wildlife Service.

Wu, T., Chellemi, D.O., Martin, K.J., Graham, J.H., & Rosskopf, E.N. (2007). Discriminating the effects of agricultural land management practices on soil fungal communities. Soil Biology and Biochemistry, 39(5), 1139-1155.

Wu, T., Milner, H., Díaz-Pérez, J.C., & Ji, P. (2015). Effects of soil management practices on soil microbial communities and development of southern blight in vegetable production. Applied Soil Ecology, 91, 58-67.

138

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.1 Soil chemical characteristics in each prairie management treatment (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) from where Platanthera praeclara roots were sampled in 2013 for mycorrhizal analyses. OM, organic matter; ENR, estimated nitrogen release; P1, phosphorus 1 (readily available to plants); P2, phosphorus 2 (readily available to plants plus part of the active reserve in the soil); K, potassium; Mg, magnesium; Ca, calcium; Na, sodium; CEC, cation exchange capability; and NO3, nitrate. Please see next page for the remaining table.

Soil characteristics

Prairie OM ENR P1 P2 pH K Mg Ca Na CEC NO3 management (%) (g/ha) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (meq/100g) (uL/L)

PC 2.5 89668 27 39 8.2 167 607 3155 53 21 4 PH 3.1 103118 29 46 8.5 150 852 2582 54 20 5 PF 1.6 70613 40 45 8.3 137 516 1578 52 13 3 PS 2.7 94151 33 39 7.8 168 673 1605 57 14 4

139

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.1 Continued Soil chemical characteristics in each prairie management treatment (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) from where Platanthera praeclara roots were sampled in 2013 for mycorrhizal analyses. S, Sulphur; Zn, zinc; Mn, manganese; Fe, iron; Cu, copper; B, boron and soil textural components (sand, silt and clay).

Soil characteristics

Prairie S Zn Mn Fe Cu B Soluble salts Sand Silt Clay management (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (mmhos/cm) (%) (%) (%)

PC 9.6 1.2 11.9 60.7 1.0 0.9 0.4 74.4 10 15.6 PH 7.8 1.6 10.4 53.5 1.2 0.7 0.3 72.4 10 17.6 PF 6.9 0.8 6.4 43.6 0.8 0.9 0.3 74.4 12 13.6 PS 6.8 1.5 13.0 88.8 0.7 0.9 0.4 66.4 20 13.6

140

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.2 Soil environmental data and chemical characteristics in each prairie management treatment (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) from where Platanthera praeclara roots were sampled in 2014 for mycorrhizal analyses. VMC, volumetric water content; EC, electrical conductivity; T, temperature, OM, organic matter; ENR, estimated nitrogen release; P1, phosphorus 1 (readily available to plants); P2, phosphorus 2 (readily available to plants plus part of the active reserve in the soil) and K, potassium.

Soil Characteristics

VWC VWC VWC EC EC EC T T T OM ENR P1 P2 K April May June April May June April May June pH Prairie (%) (g/ha) (uL/L) (uL/L) (uL/L) Management (m³/m³) (m³/m³) (m³/m³) (dS/m) (dS/m) (dS/m) (°C) (°C) (°C)

PC 0.49 0.50 0.51 0.23 0.35 0.50 4.0 10.3 17.3 3.8 118810 14 38 6.9 141 PH 0.43 0.48 0.49 0.22 0.39 0.60 1.7 8.8 18.0 4.0 123294 14 75 7.4 151 PF 0.51 0.50 0.50 0.28 0.39 0.54 1.8 10.1 17.4 4.0 124414 15 36 7.0 161 PS 0.44 0.48 0.49 0.20 0.36 0.51 2.6 11.0 17.9 4.4 133381 17 25 6.8 170

141

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.2 Soil chemical characteristics in each prairie management treatment (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) from where Platanthera praeclara roots were sampled in 2014 for mycorrhizal analyses. Mg, magnesium; Ca, calcium; Na, sodium; CEC, cation exchange capability; NO3, nitrate; S, Sulphur; Zn, zinc; Mn, manganese; Fe, iron; Cu, copper and B, boron and soil textural components (sand, silt and clay).

Soil characteristics

Prairie Mg Ca Na CEC NO3 S Zn Mn Fe Cu B Soluble salts Sand Silt Clay Management (uL/L) (uL/L) (uL/L) (meq/100g) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (uL/L) (mmhos/cm) (%) (%) (%)

PC 665 2702 54 20 1 4.1 1.3 30 144 1.8 0.8 0.3 74 14 12 PH 913 4407 72 30 1 11 1.6 36 125 2.4 0.5 0.4 72 12 16 PF 698 2653 57 20 1 11 1.5 19 93 1.8 0.7 0.3 74 12 14 PS 732 2300 65 19 1 3.7 1.4 30 128 1.4 0.7 0.4 76 12 12

142

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.3. Number of plants sampled and number of samples that yielded useable sequences from roots of Platanthera praeclara in 2013 and 2014. Plant roots were sampled for mycorrhizal diversity studies across four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) at Pembina Trail Preserve in Minnesota, USA.

Samples collected Samples yielding useable sequences

2013 2014 Total 2013 2014 Total

Number of plants 24 20 44 19 20 39 Number of root fragments 81 80 161 54 66 120

143

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.4 Number of root sections (i.e. sequences) representing each of the four operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi in each of the four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in 2013. Values inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed.

Prairie management treatments OTU type PC PH PF PS Ceratobasidiaceae C1 12(5) 9(4) 11(3) 9(3)

Tulasnellaceae T1 6(2) 3(2) T2 2(1) T3 2(1)

Total 12(5) 9(4) 21(7) 12(4)

144

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.5 Number of root sections (i.e. sequences) representing each of the 20 operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi in each of the four prairie management treatment plots (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in 2014. Values inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed.

Prairie management treatments OTU type PC PH PF PS Ceratobasidiaceae C1 3(2) 12(4) 13(2) C2 1(1) C3 3(1) C4 1(1) C5 1(1) C6 1(1) 4(3) C7 3(2) C8 2(1) C9 4(1) C10 4(1) C11 2(1) C12 1(1) C13 1(1)

Tulasnellaceae T1 6(4) 1(1) T2 3(1)

Total 9(5) 16(4) 24(5) 17(7)

145

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.6 Number of root sections (i.e. sequences) representing each of the 20 operational taxonomic units (OTUs) of Platanthera praeclara mycorrhizal fungi in each of the four prairie management treatment plots (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) using the combined data set (years 2013 and 2014). Values inside the parenthesis denote the number of sequences representing an OTU, while the values outside are the number of plants in which an OTU was observed.

Prairie management treatments OTU type PC PH PF PS Ceratobasidiaceae C1 9(3) 19(3) 20(8) 15(7) C2 4(3) 1(1) 1(1) C3 2(1) C4 2(1) C5 2(2) C6 1(1) C7 1(1) C8 1(1) C9 1(1) C10 3(2) C11 4(1) C12 4(1) C13 1(1) 4(2) C14 1(1) C15 1(1)

Tulasnellaceae T1 1(1) 3(3) T2 3(2) 6(2) 3(2) T3 2(1) T4 3(2) T5 2(1)

Total 29(11) 45(11) 25(8) 21(9)

146

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.7 Mean pairwise sequence distances and standard errors (SE) for all observed fungal ITS sequences from roots of Platanthera praeclara sampled from four experimental prairie management treatments in years 2013 and 2014, and for the combined data set (years 2013 and 2014), separated by fungal family Ceratobasidiaceae and Tulasnellaceae. The values presented within the parenthesis are the numbers of individual ITS sequences (n) composing each group.

Data set Fungal Family Pi - Distance SE Tulasnellaceae (23) 0.203 0.009 2013+2014 Ceratobasidiaceae (97) 0.049 0.004

Ceratobasidiaceae 2013 (41) 0.070 0.005 2013 Tulasnellaceae 2013 (13) 0.168 0.007 Ceratobasidiaceae 2014 (56) 0.009 0.002 2014 Tulasnellaceae 2014 (10) 0.257 0.010

147

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.8 Mean pairwise sequence distances and standard errors (SE) for all observed fungal ITS sequences from roots of Platanthera praeclara in each of four prairie management treatment plots (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) for years 2013 and 2014 separated by land management treatment. The values presented within the parenthesis are the numbers of individual ITS sequences (n) composing each group.

Data set Prairie management Pi - Distance SE PS (12) 0.201 0.008 PF (21) 0.295 0.010 2013 PH (9) 0.006 0.002 PC (12) 0.007 0.002

PS (17) 0.133 0.006 PF (24) 0.130 0.005 2014 PH (16) 0.038 0.004 PC (9) 0.259 0.010

148

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.9 Statistical significance (P values) from the Fisher’s exact test comparing the fungal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) observed in each of the four prairie management treatment plots (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) for 2013.

Prairie management treatments PC PH PF PS PC -

PH 1.00 -

PF < 0.05 0.111 -

PS 0.095 0.228 0.546 -

149

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.10 Statistical significance (P values) from the Fisher’s exact test comparing the mycorrhizal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) observed in each of the four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) for 2014.

Prairie management treatments

PC PH PF PS

PC -

PH < 0.05 -

PF < 0.05 < 0.05 -

PS < 0.05 < 0.05 < 0.05 -

150

Texas Tech University, Pablo Antonio Tovar, December 2015

Table 3.11 Statistical significance (P values) from the Fisher’s exact test comparing the fungal communities of Platanthera praeclara based on the abundances of operational taxonomic units (OTUs) observed in each of the four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) for the combined data set (years 2013 and 2014).

Prairie management treatments PC PH PF PS

PC -

PH < 0.05 -

PF < 0.05 < 0.05 -

PS < 0.05 0.142 < 0.05 -

151

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 3.1 Satellite image of the arrangement of the four prairie management treatment plots (PC, no management; PS, spring burn; PF, fall burn, and PH, annual haying) where mycorrhizal fungi of Platanthera praeclara were sampled in years 2013 and 2014. Experimental plots were located at Pembina Trail Preserve, Polk County, Minnesota, USA. Image taken from Google Earth.

152

Texas Tech University, Pablo Antonio Tovar, December 2015

PC PH 0 0

12 9

PF PS

3

10

11

9

Ceratobasidiaceae

Tulasnellaceae

Figure 3.2 Pie charts showing the distribution of the fungal families Ceratobasidiaceae and Tulasnellaceae observed in roots of Platanthera praeclara sampled from four experimental prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in the year 2013. Numeric values represent the individual ITS sequences belonging to each fungal family. See below please.

153

Texas Tech University, Pablo Antonio Tovar, December 2015

PC PH 0

3

6 18

PF PS

3 1

19 16

Ceratobasidiaceae

Tulasnellaceae

Figure 3.3 Pie charts showing the distribution of the fungal families Ceratobasidiaceae and Tulasnellaceae observed in roots of Platanthera praeclara sampled from four experimental prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in the year 2014. Numeric values represent the number of individual ITS sequences belonging to each fungal family.

154

Texas Tech University, Pablo Antonio Tovar, December 2015

(a) 14 12

10 8 6

# of plants of # 4 2 0 1 2 3 # of OTUs

(b) 20 18 16

14 12 10 8 # of plants of # 6 4 2 0 1 2 # of fungal families

Figure 3.4 One to three operational taxonomic units in individual plants of Platanthera praeclara (a), and number of individual plants hosting one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b), across four experimental prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in 2013.

155

Texas Tech University, Pablo Antonio Tovar, December 2015

(a) 14 13 12

10 8 6 5

# of plants of # 4 2 1 1 0 1 2 3 4 # of OTUs

(b) 20

15

10 # of plants of # 5

0 1 2 # of fungal families

Figure 3.5 One to four operational taxonomic units in individual plants of Platanthera praeclara (a), and number of individual plants hosting one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b), across four experimental prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) in 2014.

156

Texas Tech University, Pablo Antonio Tovar, December 2015

(a) 25

20

15

10 # of plants of # 5

0 1 2 3 4 # of OTUs

(b) 40 35

30

25 20

15 # of plants of # 10 5 0 1 2 # of fungal families

Figure 3.6 One to four operational taxonomic units in individual plants of Platanthera praeclara (a), and number of individual plants hosting one or both fungal families (i.e., Ceratobasidiaceae and Tulasnellaceae) (b), across four experimental prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment). Data represent samples collected from each experimental treatment over two years (2013 and 2014).

157

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 3.7 Two-way hierarchical cluster tree and matrix coding based on abundance of four fungal operational taxonomic units (OTUs) within the roots of Platanthera praeclara representing four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment). Data represent samples collected from each experimental treatment in 2013.

158

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 3.8 Two-way hierarchical cluster tree and matrix coding based on abundance of 15 fungal operational taxonomic units (OTUs) within the roots of Platanthera praeclara representing four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment). Data represent samples collected from each experimental treatment in 2014.

159

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 3.9 Two-way hierarchical cluster tree and matrix coding based on abundance of 20 fungal operational taxonomic units (OTUs) within the roots of Platanthera praeclara representing four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment). Data represent samples collected from each experimental treatment over two years (2013 and 2014).

160

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 3.10 Nonmetric multidimensional scaling (NMS) ordination of the fungal communities of Platanthera praeclara in 4 land management treatment plots (PS, spring burn; PF, fall burn; PH, annual haying and PC, no treatment) with correlations to soil variables (clay, silt, sand, S, P1, Na, B and pH) for the year 2013.

161

Texas Tech University, Pablo Antonio Tovar, December 2015

Figure 3.11 Nonmetric multidimensional scaling (NMS) ordination of the fungal communities of Platanthera praeclara representing four prairie management treatments (PS, spring burn; PF, fall burn; PH, annual haying, and PC, no treatment) with correlations to soil physicochemical variables (clay, silt, sand, S, P1, OM and pH) and soil environmental variables (electrical conductivity in June, EC J; electrical conductivity in May, EC M, and volumetric water content in June, VWC J) for the year 2014.

162

Texas Tech University, Pablo Antonio Tovar, December 2015

CHAPTER IV

CONCLUSIONS

The federally threatened orchid Platanthera praeclara was observed to form mycorrhizal associations with fungi from the families Ceratobasidiaceae and

Tulasnellaceae when sampled in two consecutive years (2013 and 2014) across eleven sites representing its natural distribution. Although ITS sequences from

Ceratobasidiaceae were detected more frequently, they also exhibited smaller pairwise sequence distances than ITS sequences from Tulasnellaceae. Short sequence distances suggest that P. praeclara makes more specific associations with very closely related fungi within Ceratobasidiaceae, while making associations with genetically diverse fungi from the Tulasnellaceae. In addition to the pairwise sequence distances, phylogenetic analyses consistently supported the observation that mycorrhizal fungal associates belonging to Ceratobasidiaceae have a rather narrow phylogenetic breadth. Moreover, fungi from Ceratobasidiaceae were present in roots of the sampled plants from each site, while fungi from Tulasnellaceae were absent or weakly represented at some of the sampled sites, suggesting that it is more important for P. praeclara to have more fungal associates belonging to Ceratobasidiaceae. Furthermore, shifts in the mycorrhizal fungal diversity in P. praeclara roots over time showed that although fungal community in the roots can be dynamic, the orchid restricted itself to associating with specific taxa from the fungal family Ceratobasidiaceae.

163

Texas Tech University, Pablo Antonio Tovar, December 2015

Although the composition in the orchid mycorrhizal fungal communities within the roots was not linked to the type of land management, there were correlations with soil variables that have been reported to affect soil fungal community dynamics and biogeographical patterns. The overall mycorrhizal fungal diversity observed within the

Pembina Trail Preserve experimental prairie management treatments presented inconclusive patterns that could explain the observed mycorrhizal fungal communities in

P. praeclara in relation to the land management treatment that was administrated.

However, the two prescribed burn treatments hosted more diversity in terms of OTU richness, while orchid roots from the annual haying treatment lacked Tulasnellaceae symbionts at this site.

It is important to take into account the wide natural distribution of P. praeclara. It is usually expected that orchids with a wider distribution present more general associations towards its mycorrhizae. However, the orchid mycorrhizal fungal diversity observed in this study suggests that orchid taxa with wide distributions can also be specific towards their fungal associates. It is clear that few generalizations about orchid fungal specificity, diversity, or dynamics can be made for a plant family that holds between 20,000 and 35,000 species, and caution is warranted when extrapolating the findings to other orchid species with similar distributions or habits. At the same time, it is possible that terrestrial orchids native to grasslands exhibit similar patterns of mycorrhizal associations, and investigation of such patterns, along with an investigation of ecological drivers that influence the shifts in these associations, is warranted.

164