characterisation of arbuscular mycorrhizal fungal

107
Characterisation of arbuscular mycorrhizal fungal communities associated with Hieracium lepidulum in Central Otago, New Zealand Max Maurice Leigh Crowe A thesis submitted as partial fulfilment of the requirements for the degree of Master of Science in Ecology University of Otago, New Zealand June 2012

Upload: others

Post on 28-Nov-2021

15 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Characterisation of arbuscular mycorrhizal fungal

Characterisation of arbuscular

mycorrhizal fungal communities

associated with Hieracium lepidulum in

Central Otago, New Zealand

Max Maurice Leigh Crowe

A thesis submitted as partial fulfilment of the requirements for the degree of

Master of Science in Ecology University of Otago,

New Zealand

June 2012

Page 2: Characterisation of arbuscular mycorrhizal fungal
Page 3: Characterisation of arbuscular mycorrhizal fungal

i

Abstract

Alien plants cost the New Zealand economy over $1 billon per annum in lost revenue and

control measures, and can modify native plant communities to the detriment of endemic

biodiversity. Hieracium lepidulum has invaded several regions of New Zealand and is found

in high densities among the hills in Central Otago. The roles of microbes are increasingly

included in theoretical models of plant invasion, and this study investigates the diversity and

spatial structure of a group of ubiquitous organisms, the arbuscular mycorrhizal fungi (AMF),

growing in symbiosis with Hieracium lepidulum.

Three AMF-specific molecular primer sets were tested to determine their relative

sensitivity and specificity for detecting AMF in cultures established from field collected

propagules. The optimal primer set was then used to characterise the AMF community

associated with H. lepidulum in modified subalpine grassland. The fungi from 30 plant

individuals within a 1.8 × 1.8 m plot were characterised using restriction fragment length

polymorphism (RFLP) analysis and cloning. AMF communities colonising individual plants

were found to be diverse, uncorrelated with root biomass, and possess significant

phylogenetic structure. Nine phylogenetically distinct taxa were defined, with no plant

individual possessing more than seven taxa, despite one AMF taxon comprising over 67% of

total abundances. Spatial analysis found evidence of significant positive spatial

autocorrelation in the identities of AMF colonising neighbouring H. lepidulum up to 0.5 m.

Spatial clustering was also detected in the distributions of H. lepidulum individuals at similar

scales, potentially indicating common mechanisms structuring both host and symbiont

distributions.

Phylogenetic analysis of the sequence data found evidence that the detected AMF taxa

were potentially endemic and widespread generalists, indicating that the success of

H.lepidulum as an invader is not likely to be the result of facilitation by coinvasive AMF.

Page 4: Characterisation of arbuscular mycorrhizal fungal

ii

Acknowledgements

I give my thanks to the following individuals and organisations. The Miss E L Hellaby

Indigenous Grassland Research Trust for financial support, my supervisors Dr David

Orlovich and Prof Katherine Dickinson for advice and attention, and my parents Leigh and

Peta Crowe for their love and support. Thanks to the Botany Department, University of

Otago, which is full of good people. Particular thanks to Vicky Tomlinson, Rebecca

MacDonald, and Susan Mackenzie for their help and patience. Thanks to Geoff Brown for

access to Locharburn Station. Thanks also to everyone in the Bannister lab, particularly Mike

Lucas, for help with the details. Special thanks to the following helpful and tolerant friends,

in no particular order: Ella Hayman, Kira Taylor, Matt Thomson, Anna Garden, Ben Myles,

James Mattingley, Kate Ladley, Suus Claessen, and the Ecology and Botany post grads. You

know who you are. Thanks to Brian Niven for statistical advice, Hayley Ridgeway for AMF

cultures and methodological help, and Mike Denham and Alastair Neaves from the

University of Otago department of Surveying for help with the Leica software.

Page 5: Characterisation of arbuscular mycorrhizal fungal

iii

Contents

Chapter 1: Arbuscular mycorrhizal fungi and plant invasions ..................... 1

Introduction ............................................................................................................................ 1

Arbuscular mycorrhizal fungi ............................................................................................ 1

AMF symbiosis .................................................................................................................. 2

AMF evolution and development ...................................................................................... 3

AMF dispersal .................................................................................................................... 4

Global AMF diversity ........................................................................................................ 4

AMF diversity in New Zealand ......................................................................................... 5

AMF and invasion.............................................................................................................. 6

Hieracium lepidulum ......................................................................................................... 7

H. lepidulum distribution ................................................................................................... 7

Mechanisms underlying H. lepidulum invasion................................................................. 8

H. lepidulum and AMF ...................................................................................................... 8

Thesis structure ...................................................................................................................... 9

Chapter Two: The Culture and Detection of AMF ............................................................ 9

Chapter Three: Spatial and genetic diversity of AMF associated with H. lepidulum...... 10

Chapter Four: Conclusions and experimental recommendations .................................... 11

Chapter 2: Establishment and detection of arbuscular mycorrhizal fungi in

trap culture ........................................................................................................ 12

Introduction .......................................................................................................................... 12

Trap culture ...................................................................................................................... 13

Influence of AMF on Hieracium lepidulum growth and competitive ability .................. 14

Page 6: Characterisation of arbuscular mycorrhizal fungal

iv

Research Questions .............................................................................................................. 14

Trap culture establishment ............................................................................................... 14

Molecular detection ......................................................................................................... 15

Methods................................................................................................................................ 16

Site description................................................................................................................. 16

Propagule collection......................................................................................................... 17

Spore isolation ................................................................................................................. 17

Inoculation ....................................................................................................................... 19

Detection .......................................................................................................................... 20

Molecular methods........................................................................................................... 20

Soil DNA extraction ........................................................................................................ 21

Root DNA extraction ....................................................................................................... 21

Spore DNA extraction...................................................................................................... 21

PCR conditions and equipment ........................................................................................ 22

Cloning ............................................................................................................................. 23

RFLP typing ..................................................................................................................... 23

Sequencing ....................................................................................................................... 23

Results .................................................................................................................................. 24

Trap culture ...................................................................................................................... 24

AM1-NS31 primers ......................................................................................................... 25

AmL1-AmL2 primers ...................................................................................................... 26

SSUmAf–LSUmAr, SSUmCf–LSUmBr primers ............................................................ 30

Page 7: Characterisation of arbuscular mycorrhizal fungal

v

Molecular screening of single spore culture .................................................................... 30

Discussion ............................................................................................................................ 31

Trap culture ...................................................................................................................... 31

Molecular methods........................................................................................................... 33

DNA origin ...................................................................................................................... 34

Conclusion ........................................................................................................................... 35

Chapter 3: Spatial and phylogenetic diversity of arbuscular mycorrhizal

fungi associated with Hieracium lepidulum .................................................... 36

Introduction .......................................................................................................................... 36

Study site .......................................................................................................................... 39

Methods................................................................................................................................ 40

Sample collection ............................................................................................................. 40

Sample processing ........................................................................................................... 41

DNA extraction ................................................................................................................ 42

PCR conditions ................................................................................................................ 42

Cloning ............................................................................................................................. 43

RFLP typing ..................................................................................................................... 43

Sequencing ....................................................................................................................... 43

Staining and quantifying AMF infection ......................................................................... 44

Phylogenetic analysis ....................................................................................................... 44

Diversity estimation ......................................................................................................... 45

Community analysis......................................................................................................... 46

Spatial analysis................................................................................................................. 47

Page 8: Characterisation of arbuscular mycorrhizal fungal

vi

Results .................................................................................................................................. 47

Validation of extraction methods ..................................................................................... 47

Validation of cloning methods ......................................................................................... 48

RFLP analysis ...................................................................................................................... 49

Phylogenetic analysis of sequence data ........................................................................... 51

Plant-level variation in AMF communities...................................................................... 55

Rank abundance analysis ................................................................................................. 56

Extrapolation of site diversity .......................................................................................... 57

Spatial analysis................................................................................................................. 58

Spatial structure of H. lepidulum individuals .................................................................. 61

AMF status of conspecifics .............................................................................................. 62

Effect of sampling effort .................................................................................................. 65

Discussion ............................................................................................................................ 66

Key findings ..................................................................................................................... 66

AMF community structure ............................................................................................... 66

Phylogenetic analysis ....................................................................................................... 67

Comparison of diversity measures ................................................................................... 68

Spatial analysis................................................................................................................. 69

Conclusion ........................................................................................................................... 71

Chapter 4: Conclusions and experimental recommendations ...................... 72

Single spore culture screening ............................................................................................. 72

AMF spatial structure .......................................................................................................... 73

Phylogenetic analysis ........................................................................................................... 74

Page 9: Characterisation of arbuscular mycorrhizal fungal

vii

Testing mechanisms of H. lepidulum invasion .................................................................... 75

AMF positive diversity-invasibility model .......................................................................... 76

AMF facilitation/herbivory resistance interaction model .................................................... 76

Limiting factor release ......................................................................................................... 77

Conclusion ........................................................................................................................... 78

References .......................................................................................................... 80

Appendix ............................................................................................................ 96

Page 10: Characterisation of arbuscular mycorrhizal fungal
Page 11: Characterisation of arbuscular mycorrhizal fungal

1

Chapter 1: Arbuscular mycorrhizal fungi and plant invasions

Introduction

The invasion of exotic species is one of the greatest drivers of biodiversity loss worldwide,

posing serious challenges for conservation and economic aspirations (Heywood 1989; Sala et

al. 2000; Pimentel et al. 2005). Plant invaders drive changes in plant community structure

and modify ecosystem processes at several trophic levels (Evans et al. 2001; Mitchell et al.

2006). Understanding the factors that control plant invasions is vital for the formulation of

models to predict and effectively manage invasion.

Arbuscular mycorrhizal fungi (AMF) are a functionally important component of plant

communities that directly influence belowground interactions between plants, including

competition for nutrients and water, through the formation of hyphal networks (van der

Heijden et al. 1998c; Fitter 2005; Finlay 2008). Invasive plants have been shown to interact

with AM fungal networks, affect their structure, and in some cases drive feedback which

promotes further invasion (Reinhart et al. 2003; Callaway et al. 2004; Vogelsang & Bever

2009). Little is known about the distribution, diversity and structures of AMF communities in

New Zealand. This thesis aims to investigate these aspects of an AMF community associated

with an invasive plant species Hieracium lepidulum (Stenstr.) Omang (Asteraceae) in a New

Zealand subalpine grassland ecosystem.

Arbuscular mycorrhizal fungi

Arbuscular mycorrhizal fungi are an ancient group of obligate phytotrophs comprising the

phylum Glomeromycota (Schüβler et al. 2001a). An arbuscule is a “tree-like” fungal

structure within a root cortex cell, and the term mycorrhiza is derived from the Greek

“mycos” and “rhiza”, literally translated as “fungus-root”. The name refers to the symbiotic

relationship where fungal structures form within the roots of plants and hyphae propagate

into the surrounding soil in the form of extraradical mycelium. By colonising plant roots,

AMF obtain photosynthetically derived hexose which they utilise as their primary carbon

source. Evidence of Gomeromycota-like organisms have been found in fossilised plant roots

over 400 million years old (Remy et al. 1994; Redecker et al. 2000; Parniske 2008), and it is

believed that AMF may have been instrumental in the establishment of land plants, although

more recent discoveries also implicate other fungal groups (Bidartondo et al. 2011). AMF are

Page 12: Characterisation of arbuscular mycorrhizal fungal

2

present in almost every terrestrial ecosystem on earth; it is estimated that up to 90% of plant

species are associated with AMF, including members from every major clade within the plant

kingdom. The prevalence of the AMF symbiosis further supports the hypothesis that these

organisms have had a long history of coevolution with land plants (Helgason & Fitter 2005).

AMF symbiosis

Water and vital macronutrients like phosphorus (P) are often present in limiting quantities in

natural environments (i.e., Chapin et al. 1987; Raghothama & Karthikeyan 2005). By

extending several centimetres from the plant root, mycorrhizal hyphae are able to access and

liberate nutrients that are physically or physiologically unavailable to the plant roots (Ollson

et al. 2002; Finlay 2008). The fungal mycelia act as extensions of the host plant’s root

system, effectively expanding the soil surface available for exploitation of resources, with

hyphal length typically in the order of 1500 cm hyphae cm-3 (Olsson et al. 2002). AM hyphae

are also able to obtain nutrients by accelerating decomposition and lysing soil microbes

(Hodge et al. 2001). In many cases plants do not interact directly with the soil environment,

rather they may interact indirectly with soil via their AMF (Smith & Read 2008). Phosphate

crosses the fungal plasma membrane and is converted into polyphosphate, bound within

vesicles (Javot et al. 2007). Polyphosphate is transported back through the fungal hyphae to

the host plant, and through this mechanism AMF are capable of providing up to 90% of their

host plant’s phosphorus requirements (van der Heijden et al. 2008). Uncolonised plants have

been shown to respond to P gradients in soil with the localised stimulation of root growth and

it is believed the AMF exploit this response; AMF colonisation sites may be perceived by the

plant as areas of high P, which stimulates an influx of sugar to the root cell (Helgason &

Fitter 2009). Polyphosphate molecules are also believed act as transport vehicles for other

nutrients, most notably nitrogen in the form of arganine (Govindarajulu et al. 2005). The

AMF symbioses comes at a cost to the plant host, with studies demonstrating that the carbon

allocation to the AMF can be as much as 20% of net host photosynthate (Johnson et al.

2002). Despite the cost, the benefits to the host plant can be significant, including improved

growth (Mosse et al. 1969), protection from pathogens and herbivores (Toussaint 2007;

Koricheva et al. 2009) and enhanced drought tolerance (Auge 2001, 2004). Estimates put the

carbon cost of hyphae at one to two orders of magnitude below the carbon cost of roots

(Helgason & Fitter 2005), which may partially explain the persistence and ubiquity of the

symbioses in terrestrial systems.

Page 13: Characterisation of arbuscular mycorrhizal fungal

3

AMF evolution and development

As obligate symbionts of plants AMF have maintained the ability to form symbioses with

most plant species, leading to the Glomeromycota being characterised as generalist

organisms. It is becoming evident that some combinations of host and fungus are more

compatible than others (van der Heijden et al. 1998c; Klironomos 2003), and that AMF can

be adapted to substrate, with local AMF conferring greater benefits to plants in droughty,

saline and polluted environments (e.g. Juniper & Abbott 1993; Ruiz-Lozano 2003; Christie et

al. 2004). Anthropogenic modification of the environment has been shown to alter the

function of AMF assemblages; soil amendments that alleviate nutrient deficiencies can cause

AMF to parasitise their host (Bethlenfalvay et al. 1982; Koide 1991). The position of the

AMF symbiosis along a parasite-mutualist gradient is therefore contingent on both the biotic

and abiotic environment (Crush 1975; Johnson et al. 1997).

In most environments, the root systems of a mycorrhizal host are colonised by several

AMF strains, which simultaneously associate with several hosts. Laboratory experiments

have indicated that mutual recognition of good “reciprocators” can lead to the formation of

distinct suites of AMF associating with particular plant species (Davison et al. 2011; Kiers et

al. 2011). This phenomenon has also been observed in the field, with several studies

demonstrating that co-occurring species can possess distinct AMF communities (Sykorova et

al. 2007; Davison et al. 2011). Geographical isolation has also been shown to influence the

symbiosis with local AMF-host combinations showing more extreme growth responses (both

positive and negative) than AMF-host combinations of mixed origin (Klironomos 2003).

Evolutionary theory views mutualisms as “reciprocal exploitations”, and predicts that

mutualisms should eventually develop into exploitative or parasitic associations due to the

selection of “cheating” (Foster & Wenseleers 2006). Two species forming a mutualism can

be considered to face the “Prisoner’s Dilemma”, where both members of the symbiosis can

benefit from cooperation, but each benefits most by adapting to cheat their partner (Trivers

1971). AMF are believed to be largely asexual organisms (although see Croll & Sanders

2009; Helgason & Fitter 2009), and as a result should be susceptible to the accumulation of

deleterious mutations within the fungal genome, a process known as Muller’s Ratchet

(Muller 1964; Felsenstein 1974). The asexual mode should result in the AM mutualism being

particularly susceptible to cheating by an exploitative host. It has been demonstrated that both

cheating within symbioses and Muller’s Ratchet have resulted in the decline mutualistic and

asexual groups throughout history (Foster & Wenseleers 2006). The asexual spores of the

Page 14: Characterisation of arbuscular mycorrhizal fungal

4

Glomeromycota differ from typical eukaryote propagules by possessing multiple nuclei and

multiple genomes (Kuhn et al. 2001; Rosendahl & Stukenbrock 2004). These nuclei arrive

within the developing spore via cytoplasmic streaming from the mycelium, and it is

hypothesised that AMF may avoid Muller’s Ratchet by some selective mechanism operating

at the scale of individual hyphae or nuclei (Hijri & Sanders 2005; Jany & Pawlowska 2010).

Other aspects of genetic and spatial organisation are also believed to play a role (Sanders

2002; Pawlowska & Taylor 2004; Yamaura et al. 2004; Croll et al. 2009). It is clear that

knowledge of both genetic and ecological aspects of these organisms is required to

understand the apparent evolutionary stability of the 400 million year AM symbiosis (Kiers

& van der Heijden 2006; Jany & Pawlowska 2010; Kiers et al. 2011).

AMF dispersal

Two forms of infective propagules are recognised in AMF, the hypogeous asexual

glomerospore (sensu Goto & Maia 2006), and the extraradical hyphae. Wind dispersal of

glomerospores has been shown to be important in disturbed arid environments (Warner et al.

1987), and evidence of invertebrate and mammalian dispersal of glomerospores has been

found in several ecosystems, with vectors including rodents and the Australian brushtail

possum (Trichosurus vulpecula) (Cowan 1989; Janos et al. 1995; Mangan & Adler 2002). In

the New Zealand context, the potential for avifauna as dispersal vectors may also be

significant (Johnston 2009).

In many systems, the proliferation of AMF extraradicle hyphae is thought to be the

dominant pathway for infection of new host plants (Klironomos & Hart 2002). Inter- and

intra-specific root systems can become connected via common mycorrhizal networks (CMN),

which have been shown to translocate both plant and fungus-derived nutrition (He et al.

2003; van der Heijden & Horton 2009). These networks are hypothesised to be the pathway

whereby AMF facilitate the maintenance of plant diversity (Grime et al. 1987; Marler et al.

1999); there is evidence that some host plants improve their fitness at the expense of other

plants within the community (Crush 1976; van der Heijden et al. 2008).

Global AMF diversity

To date the Glomeromycota comprises approximately 210 described species (Helgason &

Fitter 2009), in 16 genera (www.amf-phylogeny.com) with broad biogeographic

distributions. The bulk of data concerning the identity and distribution of AMF have been

Page 15: Characterisation of arbuscular mycorrhizal fungal

5

gained through morphotyping spores grown in controlled environments (Öpik et al. 2010).

The adoption of molecular methods has shown that morphotyping provides poor resolution of

species, and is prone to error resulting from environmentally triggered spore phenotype

plasticity (Stockinger et al. 2009). The adoption of molecular methods has also resulted in a

burgeoning database of AMF sequences from environmental samples, unconnected with any

cultures or morphological specimens. As of 1st June 2012, there are nearly 20,000

environmental AMF sequences available on GenBank unattributed to any species (URL:

http://www.ncbi.nlm.nih.gov/genbank). While these data are highly suggestive of as-yet

uncharacterised diversity, it is difficult to determine the significance of this number in terms

of AMF species diversity (Öpik et al. 2006). The most commonly targeted genetic regions

include fragments of the nuclear small ribosomal subunit (18S rSSU), and to a lesser extent,

the internal transcribed spacer region (ITS) and large subunit (25S rLSU) (Öpik et al. 2010).

The multinucleate spores and hyphae of AMF contain multiple genomes, and may express

many copies of the gene regions targeted by molecular primers. Efforts to reconcile

morphological and phylogenetic species concepts are being made, but are limited by the

small number of reference strains that have been successfully isolated and maintained in

culture.

AMF diversity in New Zealand

Compared with other kingdoms of life, there is relatively poor understanding of

fungal diversity and function in New Zealand’s ecosystems; only a third of New Zealand’s

predicted fungal diversity has been discovered, and fungi contain the greatest proportion of

“data deficient” species (Buchanan et al. 2012). Morphological studies from the 1970s

underpin our understanding of AMF species diversity in New Zealand (e.g. Hall 1977;

Johnson 1977; Baylis 1978). There are currently 38 species in the Glomeromycota recognised

within New Zealand, two of which are considered naturalised, with the rest having

“unknown” or “uncertain” status (Buchanan 2012). Of these, five species had unique spore

morphologies found in samples taken from coastal Otago forests (Hall 1977), while the

remainder represent spores extracted from field and culture soils that were similar in

morphology to those described in the treatise on North American AMF spore morphology by

Gerdemann and Trappe (1974). Data concerning AMF molecular diversity from New

Zealand have become available primarily via three published studies (Russell et al. 2002;

Page 16: Characterisation of arbuscular mycorrhizal fungal

6

Russell & Simon 2005; Bidartondo et al. 2011). Each of these studies has revealed

phylogenetically distinct sequences that may represent AMF taxa endemic to New Zealand.

Experimental, descriptive and taxonomic investigations of AMF in New Zealand are

primarily hindered by the lack of AMF culture collections, and up until recently, the relative

high cost of and technical limitations of DNA-based methods. As already discussed, there are

important ecological motivations for improving the state of knowledge of AMF in New

Zealand. There are also important economic motivations, including the roles of AMF in soil

carbon cycling, erosion prevention and as nutrient vectors in agricultural systems (Schwartz

et al. 2006; Bonfante & Genre 2010).

AMF and invasion

It is clear that many factors are involved in determining whether a plant species becomes

invasive, with copious literature investigating the roles of allelopathy, reproductive ability,

novel biotic and abiotic conditions, and the escape from specialist herbivores and pathogens

(e.g. Torchin et al. 2003; Blair & Wolfe 2004; Mitchell et al. 2006). The known influences of

AMF on the relative fitness of individuals within plant communities have led to

investigations into the role of AMF in some plant invasions. These include the positive

reinforcement of invasive species by enhanced nutrient uptake, protection against root

pathogens, and negative feedback on native species through reallocation of soil resources.

Isotope studies have demonstrated that dominant plants may subsidise subdominant plants

within a community through unequal investment in CMNs, and potentially by the

transmission of carbon from one plant to another (Grime et al. 1987).

Glasshouse experiments have shown that the identity of both host and AMF influence

the outcomes of the symbiotic relationship in one-to-one situations (van der Heijden et al.

1998b; Helgason et al. 2002), indicating that relative strength of AMF-conferred growth

enhancement of invasive plants may vary, depending on the associated AMF. However,

studies investigating the AMF communities on invasive plants have found that both

facultative and obligately mycorrhizal invasive plant species gain competitive benefit from a

broad range of local AMF species (Bever et al. 2001; Pringle & Bever 2008). Marler et al.

(1999) found that the exotic spotted knapweed, Centaurea maculosa (Asteraceae), grew

significantly larger in the presence of native species when local AMF species were also

present. Invasive plant species tend to be generalists and it has been hypothesised by Pringle

Page 17: Characterisation of arbuscular mycorrhizal fungal

7

et al. (2009) that generalist plants with broad distributions are likely to associate with

generalist AMF species.

It has been postulated that in some circumstances plant invasion may be facilitated by

coinvasive AMF species(Pringle et al. 2009). AMF are able to be transported in soil, and

countries like New Zealand have probably gained many new species of AMF over the two

centuries of European colonisation. Introduction of new AMF species may still be occurring,

with commercially produced innocula available for importation from several foreign sources.

The disruption of indigenous and agricultural systems by exotic cryptic fungal species has

already been demonstrated in New Zealand. The colonisation of wilding pine trees (Pinus

spp.) by exotic ectomycorrhizal species has facilitated their invasion into agricultural and

indigenous vegetation, with huge implications for indigenous plant communities (Dickie et

al. 2010).

Until recently, the cryptic lifestyle of AMF has prevented the collection of large scale

data on distributions and abundances that are needed to test general hypotheses (Fitter 2005;

Öpik et al. 2006; Öpik et al. 2010), however recent studies using next generation sequencing

methods are beginning to elucidate the identity of AMF associated with invasive plants over

broad geographic regions (Moora et al. 2011).

Hieracium lepidulum

New Zealand’s tussock grasslands, which are managed extensively for high country grazing

and conservation values, face threats from several exotic plant invaders. Of particular concern

in subalpine habitats is Hieracium lepidulum, one of a number related species that are

invasive in New Zealand (Wiser & Allen 2000; Chapman et al. 2004), and elsewhere (Davis

1977; Connor 1992). In its native continental Europe, H. lepidulum generally constitutes a

subdominant portion of the plant communities of subalpine and alpine meadows. In New

Zealand H. lepidulum can grow at very high densities, and forms almost monospecific

patches of several hundred square metres in some locations.

H. lepidulum distribution

H. lepidulum is found in Taranaki in the North Island, and in Nelson, Marlborough,

Canterbury and Otago in the South Island. It is a generalist species, found in many habitats

including grasslands, scrub, Nothofagus and plantation forests. It is most common in the high

country grasslands of Central Otago and Canterbury, particularly in areas that are grazed or

Page 18: Characterisation of arbuscular mycorrhizal fungal

8

which have been retired from grazing (Rose et al. 1995). H. lepidulum is a triploid,

diplosporous apomict (Gadella 1992); all seeds produced are likely to be genetically identical

to the parent plant. A recent analysis of H. lepidulum at several sites in Canterbury and

Central Otago led the authors to conclude that the gene flow among populations was very low

and that each population was likely to have been colonised by a single dispersal event

(Chapman et al. 2004). The analysis also predicted that the Pisa Range in Central Otago was

the most likely founding population, probably as a contaminant of pasture seed in the late 19th

century (Chapman et al. 2004). Invasion by H. lepidulum has negative impacts for both

economic and conservation values in Central Otago, including reductions in the relative

abundance of palatable species (Scott 1993) and indigenous biodiversity (Rose et al. 2004;

Rose & Frampton 2007).

Mechanisms underlying H. lepidulum invasion

Previous studies of the mechanisms underlying H. lepidulum invasion have been unable to

fully elucidate how it becomes such an aggressive invader of indigenous communities. In

glasshouse trials of competitive ability H. lepidulum was found to outcompete native species

only under fertile situations, a result that is inconsistent with the field observations of H.

lepidulum invasion and proliferation in predominantly low nutrient habitats (Radford et al.

2010). In contrast, in field situations, the addition of nutrients results in a significant decrease

in H. lepidulum cover and density (Radford et al. 2010), and management of H. lepidulum

and related species can include the application of fertilisers, which favours the growth of

more desirable, usually pasture species (Walker et al. 2003). A combination of both herbivore

selection and a relatively high resilience to herbivory have been suggested as important

mechanism for the success of H. lepidulum (Radford et al. 2007), a hypothesis consistent

with experimental results and the proliferation of H. lepidulum in areas that have

experienced historical grazing pressure (Rose et al. 1995). However, studies have also

observed H. lepidulum invasion into ungrazed vegetation, suggesting that there are other

mechanisms influencing H. lepidulum success (Rose & Frampton 2007).

H. lepidulum and AMF

Hieracium species have been shown to depend upon AMF symbionts to maintain their

relative abundance in experimental mesocosms (van der Heijden et al. 1998c), and studies

Page 19: Characterisation of arbuscular mycorrhizal fungal

9

have been conducted to investigate the role of AMF in explaining the invasion of

H. lepidulum in New Zealand. Significantly greater establishment, cover and biomass has

been reported in plants inoculated with field soil extracts, and high abundances of AMF

structures associated with the treatment support the hypothesis that the positive growth

responses were mediated by AMF (Downs & Radford 2005). The results from two more

recent studies are less conclusive. Roberts et al. (2009) found no significant change in the

competitive ability of H. lepidulum when inoculated with AMF spores and Spence et al.

(2011) found that H. lepidulum inoculated with field soils from a Nothofagus forest

experienced a depression in growth over a 60 day period. In both of these studies

methodological bias, including the length of the growing period and the lack of specificity in

the AMF treatment, may limit the utility of these results for broader generalisations.

Thesis structure

Chapter Two: The Culture and Detection of AMF

Research Topic One: AMF trap culture

Known intra- and inter-specific variability in the effects of AMF on plant growth indicates

that the use of bulk inocula from field soils may mask complexity within the AMF-host

symbiosis in experimental situations. Little is known about the indigenous AMF of New

Zealand, and no studies have yet attempted to determine the identity of AMF strains

associated with H. lepidulum. Furthermore, a lack of cultured specimens of New Zealand

glomeromycetes prevents the study of these organisms in controlled environments. In order

to investigate the potential of single spore trap cultures as a method of cultivating AMF in a

glasshouse environment, the second chapter of this thesis details the isolation and inoculation

of trap cultures with AMF associated with H. lepidulum in two invaded subalpine plant

communities – an indigenous Chionochloa tussock grassland and a modified short Festuca

grassland pasture system.

Research Topic Two: Molecular detection of AMF

Despite the recognised complexity and as yet unresolved nature of genetic organisation

within AMF species, molecular tools provide an avenue to better understand both the

distribution and diversity of AMF in environmental samples. When applied to single spores,

or cultures derived from single spores, molecular methods can assist with the attribution of

Page 20: Characterisation of arbuscular mycorrhizal fungal

10

ribosomal gene diversity to individuals and species. This information can then assist with the

more authoritative attribution of genetic diversity from environmental samples to taxonomic

diversity. In order to optimise DNA-based detection of AMF from trap culture and

environmental samples, Chapter Two will document the experimental process of PCR primer

selection, optimisation, and sensitivity, and discuss the implications and potential pitfalls of

the chosen methods.

Research Questions

1) Is single spore trap culture a viable method for the isolation of AMF associated with

H. lepidulum?

2) Are the currently available AMF-specific molecular primers equally sensitive in their

ability to amplify AMF DNA from soil, spores and root samples?

Chapter Three: Spatial and genetic diversity of AMF associated with H. lepidulum

Research Topic Three: Molecular diversity of AMF from environmental samples

The development of AMF–specific primers allows AMF diversity from environmental

samples to be studied with better resolution than was previously available using morphology-

based methods. The molecular methods developed in Chapter Two will be applied to a

comprehensive study of the AMF diversity associated with all members of a population of H.

lepidulum from a spatially restricted area. Chapter Three will also discuss the relative merits

of restriction fragment length polymorphisms (RFLPs) to discriminate AMF taxa, the impacts

of sampling intensity on detected diversity, and the interpretation of diversity metrics based

on genes from multigenomic organisms. Phylogenetic methods will be used to compare the

detected sequences with published AMF sequences from New Zealand and abroad.

Research Topic Four: Spatial diversity in AMF communities

It has been shown in several studies that a large sampling effort is required to ensure that all

AMF taxa within a site are detected. It has also been demonstrated that spatial structure in

AMF communities occurs at very small scales (<1 m), if at all (Wolfe et al. 2007; Mummey

& Rillig 2008). In Chapter Three, the distribution of the diversity of the AMF community

Page 21: Characterisation of arbuscular mycorrhizal fungal

11

amongst the plant individuals will be investigated, as will the presence and scales of spatial

autocorrelation in the AMF communities.

Research Questions

1) What is the molecular diversity associated with H. lepidulum at the sample location?

2) Do RFLP-based methods provide sufficient resolution to differentiate AMF taxa?

3) Can phylogenetic methods determine the likely origin of AMF associated with H.

lepidulum?

4) Are all H. lepidulum individuals within the study site associated with the same suite

of AMF?

5) Is there any evidence of spatial autocorrelation in the identities of AMF associated

with H. lepidulum?

6) What are the appropriate levels of replication needed to obtain representative data on

AMF community composition at the study site?

7) Is there any evidence of spatial clustering of H. lepidulum individuals within the site?

Chapter Four: Conclusions and experimental recommendations

The final chapter of this thesis will bring together the findings from the second and third

chapters in the context of the current understanding of AMF diversity, and in terms of their

role in the invasion of H. lepidulum in New Zealand. The chapter will also outline

recommendations for further research into AMF in New Zealand, with a focus on questions

and methodologies that could shed further light on the relationship between the taxonomy of

this cryptic, yet ecologically significant group, and their potential roles in invasion of

indigenous plant communities.

Page 22: Characterisation of arbuscular mycorrhizal fungal

12

Chapter 2: Establishment and detection of arbuscular mycorrhizal

fungi in trap culture

Introduction

The recognition of individuals and species is a relatively straight forward matter for sexual

organisms — the biological species concept applies, and individuals are typically both

physically and genetically distinct. Asexual species can generally be defined using a genetic

species concept, and individuals defined on the basis of physically distinct ramets. However,

in the case of arbuscular mycorrhizal fungi (AMF), both the physical and genetic extents of

individuals are difficult to determine, and as yet no operational species concept exists (Taylor

et al. 2000; Rosendahl 2008).

In terms of physical extent, some estimates put the length of AMF extraradical

mycelia on the order of metres per cubic centimetre of soil (Olsson et al. 2002), and the

biomass of living and dead arbuscular mycorrhizal (AM) hyphae at over 50% of total soil

microbial biomass (Read & Perez-Moreno 2003). The AM mycelia are also capable of

forming anastomoses between hyphae belonging to the same isolate (Giovannetti et al. 2001),

and single isolates are able to colonise many different plant species. In natural systems these

anastomoses could produce an “infinite hyphal network” connecting together different plants

over large areas (Giovannetti et al. 2004). This physiognomy, as well as the microscopic size

and morphologically uninformative nature of both intra- and extra-radical AMF hyphae,

makes the physical differentiation of AMF impossible on the basis of vegetative structures

(Whitcomb & Stutz 2007).

AMF species are known to possess multiple copies of ribosomal (Sanders et al. 1995;

Lloyd-Macgilp et al. 1996; Kuhn et al. 2001), protein coding (Viera & Glenn 1990; Hijri &

Sanders 2005; Hijri et al. 2007) and non-coding regions (Pawlowska & Taylor 2004). Some

studies have indicated that the intra-specific genetic variation may be greater than inter-

specific variation, and that species from separate genera may possess identical copies of some

genes (Clapp et al. 1999). Whether the AM genomic diversity is contained within identical

nuclei with high ploidy (homokaryotic) or whether nuclei contain distinct genomes

(heterokaryotic) remains contentious (Kuhn et al. 2001; Rosendahl & Stukenbrock 2004).

The coenocytic nature of AM hyphae allows nuclei to freely stream throughout the

“individual” (Giovannetti et al. 2001), making it possible for local differentiation in genotype

expression, possibly in response to local conditions (Helgason & Fitter 2009). An operational

Page 23: Characterisation of arbuscular mycorrhizal fungal

13

species concept for AMF requires that the “full range and frequency of sequence variation

[...] be obtained for each discrete taxon” (Clapp et al. 1999). However, such a detailed

analysis may reveal that the premise of “discrete taxa” does not apply to AMF (Bachmann

1998). For example, it has been suggested that ancient asexual lineages may in fact represent

“microspecies”, with each individual “propagating its own genetic heritage” (Coyne & Orr

2004).

Trap culture

Prior to the widespread adoption of molecular methods for diversity studies, total community

diversity of AMF was estimated by differentiating AM spores isolated from field or

glasshouse soils. This led to conclusions that AMF sporulation could be seasonal, and that

some AMF taxa were rarely represented by spores in field soils (Gemma et al. 1989; Sanders

1990). Trap culture is the method whereby individual spores, groups of spores, or other

infective AM propagules are inoculated onto a receptive host plant and grown in a controlled

environment. Trap culture of field soils, particularly successive rounds using the same media,

has been shown to induce sporulation of both recalcitrant and prolific AMF (Smith & Read

2008). Single AMF spores can then be inoculated onto trap plants, allowing the production of

infective propagules from a single AMF strain. Single spore trap culture has advantages over

other methods of producing AMF inoculum, such as bulk soil and multispore culture,

including a reduction in contamination by fungi which parasitise AMF (Brundrett 1991).

Although it is a clear simplification of the ecological processes occurring in functional

ecosystems, cultures initiated from a single multinucleate spore also provide opportunities to

study aspects of intra-“individual” genetic organisation, genomic inheritance to asexual

progeny, and functional roles in symbiosis.

There is a large disparity between number of AMF taxa that have been described from

morphological specimens, and those that have only ever been recorded as DNA fragments

from environmental samples. For every species for which morphological data are available

there are over 50 unique “uncultured” sequences available on GenBank (URL:

http://www.ncbi.nlm.nih.gov/genbank), and there is evidence of many more to be discovered,

particularly in Oceania, where relatively few studies have been carried out (Öpik et al. 2010).

It is difficult to interpret this large body of environmental sequence data in an ecological

context for two major reasons. The first is the lack of phylogenetic resolution afforded by

short (usually ribosomal) DNA fragments, and the second, and perhaps more significant, is

Page 24: Characterisation of arbuscular mycorrhizal fungal

14

the general lack of understanding about the intra- and inter-specific genetic organisation of

the Glomeromycota (Gamper et al. 2009; Öpik et al. 2010). These issues have led several

workers to state that the only way to definitively characterise the genetic diversity within

species of AMF is through their establishment in culture (Redecker et al. 2003; Gamper et al.

2009; Öpik et al. 2010).

Influence of AMF on Hieracium lepidulum growth and competitive ability

By investing photosynthetically derived carbon into AM fungal hyphae, host plants are able

to acquire nutrients at a lower carbon cost than by producing roots (Helgason & Fitter 2005).

Individuals which invest fewer resources into growth can expend more on other structures

and processes, such as reproduction and defence (Berta et al. 1993; Vance et al. 2003;

Bennett et al. 2006), and plant species that are able to do this have been shown to be both

superior competitors and successful invaders (Blossey & Notzold 1995).

It is possible that the invasion by H. lepidulum is facilitated through an association

with AMF and both bulk soil trap culture and multi-spore trap culture have been used to

inoculate seedlings of Hieracium lepidulum in previous studies (Downs & Radford 2005;

Roberts et al. 2009; Spence et al. 2011). H. lepidulum is colonised by AMF in the field and

can be colonised in a glasshouse situation, however the effects of AMF colonisation on

H. lepidulum growth and competitive ability are inconsistent. Host response has been shown

to vary between studies, with some showing enhanced growth, and others showing no effect

or suppressed growth. Inoculation of Pilosella species, which are closely related to

Hieracium, with single spore culture AMF strains has produced significant and positive

growth responses in glasshouse trials (van der Heijden et al. 1998b).

Research Questions

Trap culture establishment

Several sources of bias have been identified to account for the inconsistencies in H.

lepidulum growth response to inoculation, including the roles of other soil organisms, soil

disturbance, AMF selection biased towards sporulating species, and the potential for AMF

related benefits to be restricted to a specific stage in the plant life cycle. The relative benefit

of an AM symbiosis has been shown to depend on the identity of the AMF strain involved. It

may be that interactions between AMF strains and other soil organisms within bulk inoculum

are responsible for the variation observed in host response (e.g. van der Heijden et al. 1998a).

Page 25: Characterisation of arbuscular mycorrhizal fungal

15

In order to test hypotheses of functional diversity within AMF it is necessary to isolate and

compare the relative effects of individual AMF strains on a plant host phenotype.

H. lepidulum has invaded a wide range of vegetation types, and this study seeks to determine

whether the AMF colonising H. lepidulum vary between Chionochloa dominated tussock

grassland and a modified Festuca grassland in subalpine Central Otago. These two vegetation

types were chosen to represent a range of anthropogenic disturbance and plant communities,

both of which could be expected to influence the diversity and composition of the AMF

colonising H. lepidulum. This study uses a single spore culture method to investigate the

isolation of monospecific AMF cultures from H. lepidulum in its invaded range. Verification

of successful trap culture was carried out using a combination of visual inspection of trap

cultured roots for AM structures, and AMF specific primers.

Molecular detection

Several attempts have been made to design DNA primers specific to the Glomeromycota, and

while some have had widespread use, all suffer from major interpretive and methodological

limitations, including lack of specificity and bias. The commonly used gene regions for

ecological studies include the rDNA genes (Clapp et al. 1995; Rosendahl & Stukenbrock

2004), particularly the region of the small subunit (18S SSU rDNA) targeted by the AM1 &

NS31 primers (Helgason et al. 1998). More recently, two new primer sets have been

developed; the ~750 bp 18S SSU rDNA gene fragment primers (Lee et al. 2008), and the

~1800 bp SSU–ITS–LSU fragment nested primer set (Krüger et al. 2009) (Figure 1). The

authors of these primer sets claim that they can be used to amplify all subgroups of AMF

while excluding all other organisms. This study aims to optimise the polymerase chain

reaction (PCR) protocols for the three primer sets mentioned above, and determine which

primer set is best suited for detection of AMF both in trap culture, and for environmental

studies, in terms of sensitivity, specificity and phylogenetic resolution.

Page 26: Characterisation of arbuscular mycorrhizal fungal

16

Figure 1. Primer map showing binding sites of forward and reverse primers (5’

study in sequences from representative members of AMF species

are ~30 bp distant and not differentiated on this primer map.

transcribed spacer (ITS) region and large subunit (LSU) rDNA (5465 bp) of Glomus sp. ‘intraradices’

DAOM197198 (AFTOL-ID48, other culture/voucher identifiers: MUCL43194, DAOM1

numbers AY635831, AY997052, DQ27390)

Methods

Site description

Two mountain ranges were chose

the Old Man Range and Locharbu

because of good road access and in the case of Locharburn station,

previously been granted to researchers from Otago.

elevation, and complementary in terms of representing land use and dominant vegeta

subalpine Central Otago. The Old Man Range site has a diverse range of herbs and grasses

dominated by narrow leaved snow tussock (

lesser extent, exotic grasses

Locharburn site has had a history of disturbance associated with sheep and cattle farming fo

over 100 years, as well as pre

cunninghamii Colenso. remain on rock outcrops, but the dominant plants are

matthewsii (Hack.) Cheeseman,

Agrostis capillaris, with locally abundant

toumatou Raoul. Both sites have been invaded by

populations at Locharburn are much larger in extent and relative dominance than the Old

Man Range populations, which tend to consist of small patches or individuals. It is thought

that the Pisa Range may have been the point of entry for

contaminated pasture seed (Chapman

Figure 1. Primer map showing binding sites of forward and reverse primers (5’–3’ direction) used in this

sentative members of AMF species. The binding sites of AM1 and AML2

nt and not differentiated on this primer map. Small subunit (SSU) rDNA, internal

transcribed spacer (ITS) region and large subunit (LSU) rDNA (5465 bp) of Glomus sp. ‘intraradices’

ID48, other culture/voucher identifiers: MUCL43194, DAOM1

numbers AY635831, AY997052, DQ27390), modified from Krüger et al. (2009).

wo mountain ranges were chosen as field sites in Central Otago: the Obelisk access road on

the Old Man Range and Locharburn Station on the Pisa Range. These sites were

because of good road access and in the case of Locharburn station, and because access had

ted to researchers from Otago. The sites were similar in aspect and

elevation, and complementary in terms of representing land use and dominant vegeta

The Old Man Range site has a diverse range of herbs and grasses

d by narrow leaved snow tussock (Chionochola rigida (Raoul) Zotov

lesser extent, exotic grasses Anthoxanthum odoratum L. and Agrostis

Locharburn site has had a history of disturbance associated with sheep and cattle farming fo

over 100 years, as well as pre-European burning. At this site, refugia of

remain on rock outcrops, but the dominant plants are

(Hack.) Cheeseman, exotic agricultural species Anthoxanthum odoratum

, with locally abundant Aciphylla aurea W. R. B Oliver

. Both sites have been invaded by Hieracium lepidulum

populations at Locharburn are much larger in extent and relative dominance than the Old

an Range populations, which tend to consist of small patches or individuals. It is thought

that the Pisa Range may have been the point of entry for H. lepidulum

(Chapman et al. 2004).

3’ direction) used in this

The binding sites of AM1 and AML2

Small subunit (SSU) rDNA, internal

transcribed spacer (ITS) region and large subunit (LSU) rDNA (5465 bp) of Glomus sp. ‘intraradices’

ID48, other culture/voucher identifiers: MUCL43194, DAOM1816601; accession

: the Obelisk access road on

These sites were chosen

because access had

The sites were similar in aspect and

elevation, and complementary in terms of representing land use and dominant vegetation in

The Old Man Range site has a diverse range of herbs and grasses

(Raoul) Zotov), and, to a

Agrostis capillaris L. The

Locharburn site has had a history of disturbance associated with sheep and cattle farming for

European burning. At this site, refugia of Podocarpus

remain on rock outcrops, but the dominant plants are Festuca

Anthoxanthum odoratum and

W. R. B Oliver and Discaria

Hieracium lepidulum, although the

populations at Locharburn are much larger in extent and relative dominance than the Old

an Range populations, which tend to consist of small patches or individuals. It is thought

. lepidulum, possibly through

Page 27: Characterisation of arbuscular mycorrhizal fungal

17

Propagule collection

Ten kilograms of soil and roots were collected from hapazardly selected H. lepidulum

individuals to a depth of 10 cm at both field sites. The soil was transported and stored in large

sealed low-density polyethylene (LDPE) bags until it reached Dunedin, where it was

processed within 24 hours. The samples from each site were pooled together and passed

through a 2 cm sieve. Five replicate 100 g sub-samples were taken from each pool and used

for spore isolation and seedling inoculation in single spore trap culture as detailed below.

Soil samples collected from the two field sites were also used to establish bulk soil

trap cultures, which were grown in parallel with the single spore culture. The bulk soil trap

culture were composed of bulk soil, containing AMF spores, hyphae and root fragments. The

homogenised soil from each site was combined with an equal volume of autoclaved

horticultural sand and potted up in sterile 1.7 L pots, with 12 replicate pots for each site.

Approximately 50 surface-sterilised Trifolium repens cv. Huia seed were then broadcast sown

in each pot. The pots were grown on in a glasshouse with no added nutrients and watered as

required with tap water. After 3 months of growth, 100 g of mixed soil and roots were taken

from 3 replicates from each site and pooled for extraction of spores. Extracted spores that

were obviously free of damage or parasitism were subsequently used for a second round of

single spore culture.

Spore isolation

AMF spores were isolated from the field soil in 10 batches using a modified wet sieving

method (INVAM, URL: http://invam.caf.wvu.edu) In each batch, 100 g of soil was

suspended in 1.5 L of tap water and left for 2–3 hours to let the heavier fractions settle. Due

to the presence of lipid bodies within the AMF spores they were expected to float in the water

column or be trapped in the meniscus and froth at the water surface (H. Ridgeway, pers.

comm. 2010). The liquid fraction was decanted through two stacked sieves of decreasing

pore size: 500 µm and 32 µm; the first sieve to catch large pieces of floating organic debris

and AMF sporocarps, and the second, smaller sieve to capture AMF spores and fine particles.

The fine particles were then washed with tap water, at low pressure to avoid disrupting the

spores.

A 20/60% sucrose density gradient was set up in 50 mL centrifuge tubes. Half a

teaspoon of the particles from the 32 µm sieve were added to 6 tubes which were centrifuged

for 1 minute at 1000 g. A 50 mL syringe with a 100 mm length of low density polyethylene

Page 28: Characterisation of arbuscular mycorrhizal fungal

18

tube was used to collect the particles suspended in solution and trapped at the sucrose

gradient interface. The particles from the 6 replicate tubes were pooled together in a 32 µm

sieve and rinsed for 1–2 minutes under tap water and then transferred to a sterile 50 mL

centrifuge tube with 20 mL distilled water. Cleaned spores were maintained at 4°C in

distilled water for no more than 7 days prior to inoculation.

Figure 2. a), b): Microscope images of a single spore and multiple AMF spores. c) An AMF spore

inoculated on to germinating T. repens radicle.

Page 29: Characterisation of arbuscular mycorrhizal fungal

19

Inoculation

T. repens seeds were surface-sterilised following the method of Triplett and Barta (1987) and

germinated over two days in a sterile Petri dish. Under a dissecting microscope, spores were

transferred to a watchglass containing distilled water and cleaned of soil debris with a fine

paint brush. Spores were then transferred individually onto the emergent radicle of a surface

sterilised T. repens seed with a 10 µL micropipette (Figure 2 c).

Taking care not to dislodge the spore, the germinating seeds were then carefully planted into

sterile 1 mL plastic pipette tips filled with a 5:4:3 mixture of horticultural sand: unfertilised

potting mix: pumice screened to <0.5 mm and steam-sterilised for 1 hour at 121°C (Ridgway

et al. 2006). The clover seedlings were grown for 2 weeks and then the entire pipette tip was

transplanted into 1.7 L pots containing 1.5 L of an unscreened mixture of the same media as

detailed above, which was also steam-sterilised for 1 hour (Figure 3). Each pot was

inoculated with 50 mL of 1 µm soil filtrate from the original field site in order to introduce

non-AMF soil microbes. The lower 5 mm was cut off each pipette tip to allow roots to

emerge into the pot.

Figure 3. T. repens inoculated with AMF spore in a 1 mL pipette tip transplanted into 1.7 L pot

containing sterile medium.

Additional surface-sterilised clover seeds were broadcast sown into the pots to provide

addition root colonisation habitat for the AMF. For each site, 40 apparently viable spores

were transferred onto individual clover radicles, making a combined total of 80 pots with 5

uninoculated controls. In order to obtain spores from recalcitrant species that may not have

Page 30: Characterisation of arbuscular mycorrhizal fungal

20

been present in the field soils a further 40 single spore culture pots from each site were

inoculated using spores isolated from bulk soil trap culture after 3 months of growth in

culture, for a total of 160 culture attempts. All pots were watered as required with tap water.

Trap culture pots were kept 0.3 m apart throughout the experiment to minimise cross

contamination.

Detection

After a minimum of 4 months growth, root samples were taken from two positions within

randomly selected single spore culture pots: from the circumference, where fine roots were

bound together, and in the centre of the root ball at the point of emergence from the pipette

tip. Ten centimetres of roots, composed of five 2 cm sections of various diameters, were

taken from each location for each pot. Scalpels and forceps were flame-sterilised between

each single spore culture pot to avoid cross contamination. The roots were washed in tap

water to remove soils particles, placed in biopsy cassettes and cleared by immersion in 10%

KOH at 90°C for 15 minutes. The roots were then rinsed 3 times in tap water, and then

immersed in a solution of 5% v/v acetic acid with 0.05% v/v Acme black ink overnight. The

roots were rinsed of excess ink solution and left overnight in fresh water to improve visibility

of fungal structures. The cleared and stained root fragments were systematically screened

under 40× magnification transmission light microscopy. The detection of AMF structures,

including non-septate hyphae, arbuscules and vesicles, were considered evidence of culture

success. Clover roots grown in field soil and the online AM reference database INVAM

(URL: http://invam.caf.wvu.edu) were used as positive controls for the staining methods, and

as a reference for structures associated with AMF colonisation. A total of 40 inoculated root

systems were harvested for staining and microscopic assessment.

Molecular methods

Morphological studies of root fragments are time consuming, and the large number of

inoculated pots used in the present study meant that a faster method for screening culture was

desirable. The extreme sensitivity of molecular techniques and the existence of AMF-specific

primer sets provided an opportunity to test for the presence of AMF simultaneously in several

single spore culture pots by pooling together root samples, and extracting and amplifying the

pooled root DNA. However, prior to this the relative ability of each primer set to detect AM

fungal DNA was tested. Positive controls that were known to contain AM fungal DNA were

Page 31: Characterisation of arbuscular mycorrhizal fungal

21

extracted from environmental samples of H. lepidulum, including some from a previous study

of H. lepidulum AMF community composition (D. Lyttle unpubl.), and bulk soil trap culture

T. repens. The primers were the commonly used AM1–NS31 pair (Helgason et al. 1998), the

“improved” AMF specific primer pair AmL1–AmL2 (Lee et al. 2008), and the nested primer

set designed by Krüger et al. (2009). Upon successful amplification with positive controls,

the primer sets were tested to compare their ability to detect the presence of AMF from single

spore culture. Sample DNA was isolated using three different methods: single spore culture

soil, single spore culture roots and single spores isolated from bulk soil trap culture.

Soil DNA extraction

Ten gram soil samples were taken from the top 50 mm of single spore trap culture substrate,

predominantly under the host root systems. The 20 samples were taken randomly, 10 from

each of the sample sites (Old Man and Locharburn). DNA was extracted from each of the

pots in 250 mg subsamples using the MO BIO PowerSoil® DNA Isolation Kit following the

manufacturer’s “centrifuge” protocol, with a final elution volume of 30 µL.

Root DNA extraction

Root samples were taken using the same root extraction method described in the

morphological study. The roots were sealed in LDPE bags on ice for no more than 4 hours.

The roots were then washed 3 times under tap water, and agitated by hand to remove soil

particles, followed by a further two washes with distilled water under a dissecting microscope

to remove any remaining particles. The roots were then patted dry on paper towels, placed in

paper bags, and dried in a 50 °C oven for 16 hours. Samples were thoroughly disrupted in

liquid nitrogen using an acid washed mortar and pestle. Upon disruption, 20 mg of

homogenous powdered tissue was immediately loaded into a 1.5 mL microcentrifuge tube

and suspended in 400 µL Qiagen lysis buffer AP1. DNA was then extracted following the

Qiagen DNeasy plant mini protocol, with a final elution volume of 50 µL. For the culture

screening experiments, the 20 mg of disrupted root tissue was taken from 10 pooled samples.

Spore DNA extraction

Three AMF spore DNA extraction attempts were made with spores that were isolated from

two bulk soil trap culture pots, both containing soil originating from Locharburn Station. The

spores were isolated using the spore isolation method described above. Single cleaned spores

Page 32: Characterisation of arbuscular mycorrhizal fungal

22

were transferred onto the wall of a 1.5 mL microcentrifuge tube with a fine paintbrush and

crushed with a flame sterilised needle. The disrupted spore was then spun down by

centrifuge, and suspended in 200 µL lysis buffer AP1 (Qiagen). DNA was then extracted

following the Qiagen DNeasy plant mini protocol, with a final elution volume of 30 µL.

PCR conditions and equipment

Unless stated otherwise, all PCR mastermixes were composed of 17 µL Thermoprime®

Reddymix containing 1.25 U Thermoprime Plus DNA polymerase, 75 mM Tri-HC (pH 8.8 at

25°C), 20 mM (NH4)2SO4, 1.5 mM MgCl2, 0.01% (v/v) Tween® 20, and 0.2 mM of each

nucleotide, 10 pmol of both forward and reverse primers, and 1 µL DNA template for a total

PCR reaction volume of 20 µL. All PCR reactions were carried out using a thermocycler

(Eppendorf MastercyclerGradient), and the protocols used are listed below (Table 1). PCR

optimisation led to the inclusion of 0.02 µL 10% (w/v) bovine serum albumin (BSA) in the

AmL1-AmL2 PCR mastermix. The PCR products were loaded on 1% agarose with 1% TAE

buffer with ethidium bromide staining (1 µg mL-1). The gels were visualised by gel

electrophoresis under ultra-violet transillumination using a 1 kb plus ladder (Invitrogen) as a

standard.

Table 1. PCR thermocycling conditions for each primer pair used in this study

AM1-NS31 NS1-NS4 AmL1-AmL2 SSUmAf–

LSUmAr

SSUmCf–

LSUmBr

Initial

denaturation 3 min @ 94°C 3 min @ 94°C 3 min @ 94°C 5 min @ 99°C 5 min @ 99°C

Denaturation 1 min @

94°C

35 ×

30 sec

@ 94°C

30 ×

30 sec

@ 94°C

30 ×

10 sec

@ 99°C

40 ×

10 sec

@ 99°C

30 × Annealing 1 min @

58°C

1 min @

40°C

40 sec

@ 58°C

30 sec

@ 60°C

30 sec

@ 63°C

Extension 1 min @

72°C

1 min @

72°C

30 sec

@ 72°C

1 min @

72°C

1 min @

72°C

Final extension 10 min @ 72°C 10 min @ 72°C 5 min @ 72°C 10 min @ 72°C 10 min @ 72°C

Page 33: Characterisation of arbuscular mycorrhizal fungal

23

Cloning

PCR products were cloned into pCR®2.1-TOPO® TA vector and transformed into

chemically competent Escherichia coli Top10 to the manufacturer’s instructions (Invitrogen).

Transformed cells were screened using blue/white colony screening on 90 mm plates

containing Luria-Bertani (LB) agar, 50 µg mL-1 kanamycin and coated with 40 mg mL-1

X-gal. Thirty positive transformants from each sample were transferred on to patch plates.

Using the same PCR protocol as was used to produce the amplicons, the 30 clones from each

sample were then amplified using 1 µL of diluted transformed cells as template.

Transformants that failed to amplify, or that produced bands of a size other than expected,

were excluded from subsequent analysis.

RFLP typing

Restriction fragment length polymorphism (RFLP) analysis was used to screen the cloned

rDNA fragments. The restriction enzyme HinfI (Roche Applied Science, Mannheim,

Germany) was used in 25 µL reactions comprising 5 µL PCR product, 1 U enzyme, 2.5 µL

10× concentration SuRE/Cut™ buffer A (Roche Applied Science, Mannheim, Germany) in

16.5 µL MilliQ H2O. Reactions were digested for 4 hours at 37 °C. 10 µL of each reaction

was analysed by electrophoresis on 1.5% agarose gels stained with ethidium bromide and

visualised by UV transillumination using a 1 kb plus ladder as standard.

Sequencing

Sequencing reactions were either performed on plasmid purifications of the transformed cells

using a Qiagen minprep plasmid purification kit, and primers directed against the M13f and

M13r promoter regions, or, if direct sequencing was carried out, with the primer set used to

amplify the DNA. Sequencing was carried out by the Genetics Analysis Service, Department

of Anatomy, University of Otago, using dye terminator sequencing chemistry (BigDye

Terminator v. 3.1, Applied Biosystems) on an ABI 3730xl DNA Analyser (Applied

Biosystems). Sequence assemblies and alignments were performed using the software

GENEIOUS v5.1 (Drummond et al. 2010), and sequence identity was verified using BLAST

searches of the NCBI database (URL: http://www.ncbi.nlm.nih.gov).

Page 34: Characterisation of arbuscular mycorrhizal fungal

24

Results

Trap culture

Of the 160 single spore trap cultures initiated, none of the 40 randomly selected samples

screened using microscopic methods contained detectable levels of AMF after 4 months of

growth. No obvious differences in growth or biomass between the inoculated and

uninoculated T. repens occur over this period. However as the plants were not destructively

harvested so empirical comparisons of biomass were not made. This was a precautionary

approach to avoid destroying trap cultures which had successfully been colonised by AMF.

All plants were relatively small and showed signs of nutrient deficiency, particularly

phosphorus as evidenced by chlorosis of the older leaves. Clover roots grown in field soils

were infected with non-septate hyphae, with both arbuscules and vesicles observable (Figure

4). Spore densities were comparable to those present in the field soils, however no attempt

was made to quantify this, or to qualify the differences in spore morphology of spores used to

inoculate the single spore culture. Of the remaining 120 single spore culture samples, 20 were

tested using soil DNA, 15 were tested using root DNA, and the remaining 85 were tested

using pooled root DNA of 10 or fewer samples, as detailed below.

Figure 4. Microscopic images of stained AMF structures within cleared T. repens roots from bulk soil

trap culture. Non-septate hyphae and vesicles as visible as dark regions in the micrographs

Page 35: Characterisation of arbuscular mycorrhizal fungal

25

AM1-NS31 primers

The AM1–NS31 primer pair successfully produced amplicons of the predicted size for three

DNA extractions known to contain AM fungal DNA. Amplification was only successful in

the positive control samples that had been diluted 1:10 or 1:100. DNA from T. reprens single

spore culture roots did successfully amplify, however multiple bands were present, including

amplicons of the predicted size, indicating the presence of non-specific amplification,

possibly in addition to amplification of AM fungal DNA (Figure 5).

Figure 5. Electrophoresis gel showing amplicons of the predicted size from the positive controls (lanes 1-

12, three samples, each at four dilutions), and non-specific amplification from five T. repens single spore

culture root DNA samples.

The positive control OMR2 was amplified across a gradient of primer annealing temperatures

at multiple dilutions in order to increase the stringency of the primer binding and reduce the

amplification of non-target DNA (not shown). Amplification of the positive control ceased at

64°C, and this annealing temperature was used in a high stringency PCR to amplify four

single spore culture root DNA samples. However, multiple bands were still evident in the

single spore culture DNA samples even at this higher annealing temperature, although they

included bands of the expected size (Figure 6).

Page 36: Characterisation of arbuscular mycorrhizal fungal

26

Figure 6. Electrophoresis gel showing a positive control (OMR3, lanes 1-3) and four single spore culture

root DNA samples amplified using the AM1-NS31 primer pair using an annealing temperature of 64°C.

Arrows mark bands excised for purification and sequencing.

Bands representing amplicons of the expected size were excised from the gel, purified and

sequenced to determine their identity. BLAST searches of the excised bands revealed that

they were T.repens 18S rSSU DNA, indicating non-specific amplification, whereas the

positive control was confirmed to be from Glomeromycota (Table 2).

Table 2. Top BLAST search results from direct sequencing of AM1-NS31 amplicons showing sample

DNA origin, and accession number, organism and % similarity of closest match. The “best match” was

based on the highest bit-score result for each sample, and both bit score and percent identity are given for

each match. SSC = single spore culture.

Sample Origin Band Best Match Identities Sequence organism OMR2 + ive control A AF437708 758.701 (98.9%) Glomus sp. Glo3 isolate 20

OMS9r SSC (root) B EF023302 489.99 (96.8%) Phaseoleae environmental sample clone OMS9r SSC (root) C EF023302 489.99 (96.8%) Phaseoleae environmental sample clone OMS4r SSC (root) D EF023302 489.99 (96.8%) Phaseoleae environmental sample clone OMS4r SSC (root) E EF023450 475.57 (96.0%) Phaseoleae environmental sample clone

AmL1-AmL2 primers

Samples which contained AM fungal DNA (positive controls OMR3, OMR2, 34R, 35R)

showed only faint amplification using the AmL–AmL2 primer pair across a range of template

concentrations when the “field protocol” described in Lee et al (2008) was used. As plant and

soil derived compounds can inhibit PCR, it was hypothesised that amplification using the

universal eukaryotic rDNA primers NS1–NS4 would produce a template containing the

sample rDNA, including AM fungal DNA, while excluding any inhibitory compounds, a

Page 37: Characterisation of arbuscular mycorrhizal fungal

27

technique used successfully by Lee et al. (2008) during their design of the primers.

Amplification using NS1–NS4 was successful, and subsequent nested amplification of the

amplicons at 1:100 dilution with the AmL–AmL2 primer pair yielded amplicons of the

expected size (Figure 7, lane 3). The nested approach was also applied to DNA extractions

from AMF culture attempts with unknown AMF status. As with the positive controls, there

was no amplification of any of these samples when the AmL1–AmL2 primers were used

following the “field protocol”, however pre-amplification with NS1–NS4 primers produced a

template which could be successfully amplified using the AMF specific primers (Figure 7,

lanes 4-15).

Figure 7. Electrophoresis gel showing amplicons from AmL1-AmL2 primer pair as nested primers, with

the template DNA being 1 µL amplicons diluted 1:100 from a PCR using universal eukaryotic primers

NS1-NS4. The gel shows amplicons from five samples, with the first three lanes displaying the positive

control (OMR3), and the remaining twelve displaying four individual single spore culture root DNA

extractions (OMS20, OMS13, OMS6 and OMS7). The three lanes for each sample correspond to the

template dilutions in the NS1–NS4 PCR, from left to right: undiluted, 1:10 and 1:100. The final lane is the

negative control.

The nested approach was also used to target AM fungal DNA from single spore culture soil

and bulk soil trap culture spore DNA extractions. No amplicons resulted from the spore DNA

template, but nested PCR using the AmL1-AmL2 primers produced amplicons of the

expected size for the soil DNA samples (Figure 8).

Page 38: Characterisation of arbuscular mycorrhizal fungal

28

Figure 8. Electrophoresis gel showing amplicons from AmL1-AmL2 primer pair as nested primers, with

the template DNA being 1 µL amplicons diluted 1:100 from a PCR using universal eukaryotic primers

NS1-NS4. Lanes 1-3 show the absence of amplification for the AMF spore DNA samples, while lanes 4-15

show the predominantly positive amplification of eleven single spore culture soil samples. The final lane is

the negative control.

Cloning and sequencing was used to determine the identity of the putatively AM fungal

amplicons from root and soil DNA extractions. Representative samples were selected on the

basis of a single band of the expected size from the nested AmL1-AmL2 PCR. Thirty positive

clones from five samples were screened using RFLP analysis with the restriction enzyme

HinfI, and three unique RFLP-types from each sample were selected for sequencing. BLAST

searches of the sequence data revealed that all three of the clones from the positive control

samples belong within Glomeromycota. However the clones from the putatively mycorrhizal

single spore culture roots were revealed to be from the host plant (Trifolium sp.) and

associated basidiomycetes (Sistotrema spp.), while the single spore culture soil clones

included algae and bryophyte rRNA gene fragments (Table 3). Due to the non-specific

amplification resulting from nested PCR of single spore culture samples and the inability to

visually differentiate non-specific amplification using gel electrophoresis, an effort was made

to optimise PCR conditions for the AmL1-AmL2 primer pair so that nested PCR was not

needed. The addition of 0.02 µL 10% (w/v) BSA to each reaction resulted in successful

amplification of positive controls under PCR thermocycling conditions which otherwise

produced only weak amplification (Figure 9). When single spore culture samples which had

been shown to lack AM fungal DNA were amplified in the presence of BSA no non-specific

amplification was detected. Furthermore, the optimised protocol was able to successfully

Page 39: Characterisation of arbuscular mycorrhizal fungal

29

amplify AM fungal DNA from positive controls over a wide range of template dilutions

(Figure 10).

Table 3. Top BLAST search results from cloned AmL1-AmL2 amplicons. Sequences were obtained from

clones (numbered) or direct sequencing of PCR product (ds). The “best match” was based on the highest

bit-score result for each sample, and both bit score and percent identity are given for each match. SSC =

single spore culture.

Sample Clone Origin Best Match Identities Sequence organism OMR3 4 HIElep (root) AJ699061 1337.58 (99.9%) Glomus environmental OMR3 1 HIElep (root) JF414174 1355.62(99.9%) Glomeromycota sp. MIB 8384

OMR3 3 HIElep (root) JF14191 1324.96(99.3%) Glomeromycota sp. MIB 8392

OMS6 8 SSC (root) DQ898712 1332.17(99.9%) Sistotrema brinkmannii

OMS6 7 SSC (root) EF024316 1260.04(98.0%) Phaseoleae environmental

OMS6 5 SSC (root) DQ898718 1323.16 (99.7%) Sistotrema farinaceum

OMS13 10 SSC (root) EF23623 1323.16 (99.7%) Phaseoleae environmental

OMS13 4 SSC (root) EF023244 1243.81(97.2%) Phaseoleae environmental

OMS13 3 SSC (root) DQ898712 1227.58(96.9%) Sistotrema brinkmannii

LBS6 3 SSC (soil) HQ246325 1296.11 (98.9%) Bractracoccus cohaerens

LBS6 2 SSC (soil) FR865603 1319.55 (99.9%) Chlamydomonas rosae

LBS6 1 SSC (soil) EF024801 1076.09 (92.9%) Thaumatomonidida environmental

OMS23 2 SSC (soil) X80980 1341.19(100%) Leptobryum pyriforme

OMS23 4 SSC (soil) EU091862 1043.63(92.5%) Uncultured Banisveld eukaryote clone

OMS23 1 SSC (soil) X80980 1348.4(99.8%) Leptobryum pyriforme

Figure 9. A comparison of amplification success of positive control samples using the AmL1-AmL2

primer pairs with (right) and without BSA, with all other PCR conditions unchanged. Lanes 1-4

correspond to the same samples in each gel, with lane 5 being a negative control.

Page 40: Characterisation of arbuscular mycorrhizal fungal

30

Figure 10. Electrophoresis gel demonstrating the sensitivity of the optimised AmL1–AmL2 PCR protocol

for detecting AM fungal DNA. Lanes 1-6 show serial dilutions of positive control root DNA sample

“OR3” under 6 dilutions: undiluted, 1:10, 1:100, 1:500, 1:1000, 1:5000. Lane 7 is the negative (water)

control.

SSUmAf–LSUmAr, SSUmCf–LSUmBr primers

Positive controls were unable to be amplified using the SSUmAf–LSUmAr primers over a

range of template concentrations 1, 1:10, 1:100, 1:500, and a range of annealing temperatures

(55–60°C). Amplification was not improved by hot start PCR. Despite a lack of detectable

amplification via transillumination, the results of the SSUmAf–LSUmAr reaction were used

as template (undiluted, 1:10, 1:100) for nested PCR using SSUmCf–LSUmBr primers

following the method of Kruger et al. (2009). Similarly, nested PCR failed to produce any

detectable amplification.

Molecular screening of single spore culture

The AmL1-AmL2 primer pair were considered to be the most appropriate for screening

single spore culture for two reasons: they were the only AMF-specific primers that were able

to amplify positive samples while avoiding non-specific amplification in negative samples,

and the primer pair was shown to amplify the target region across a wide range of template

concentrations (Table 4). The root samples from the remaining 85 single spore culture

samples were pooled into 9 samples. DNA was extracted from the pooled root samples and

these extractions were used as template for PCR using the optimised protocol either undiluted

Page 41: Characterisation of arbuscular mycorrhizal fungal

31

and diluted 1:10. None of the 12 reactions showed any amplification at either template

dilution, indicating an absence of AM fungal DNA from all pooled root samples.

Table 4. Summary of PCR amplification success of primers for the detection of AM fungal DNA from

positive control and culture samples. Positive (+) and negative (-) amplification results are tabulated.

SSC= single spore culture, BSC = bulk soil trap culture.

AM1-NS31 NS1-NS4 AmL1-AmL2

(nested)

AmL1–AmL2 SSUmAf–

LSUmAr

SSUmCf–

LSUmBr

+ ive control + + + + (w/ BSA) - -

SSC root + (non-AMF) + + (non-AMF) - - -

SSC soil + (non-AMF) + + (non-AMF) - - -

BSC spore - - - - - -

Discussion

Trap culture

This study has demonstrated that it is possible to grow and detect AMF from both of the field

sites under glasshouse conditions; ten randomly selected bait plants extracted from 24 bulk

soil trap culture found to have mycorrhizal structures when studied by light microscopy.

However, despite high replication in culture attempts, no AMF structures or AM fungal DNA

were detectable in any of the 160 trap cultures inoculated with single spores after 4 months of

growth. Previous studies that have investigated AMF diversity using trap culture methods

have reported successful colonisation rates of 5% (Hall 1977). This statistic may be

illustrative of the large numbers of variables involved in attempting to recreate AMF

symbioses in a glasshouse environment. The large number of replicates undertaken in this

study was an attempt to ensure successful colonisation in at least some cultures despite this

recognised poor success rate. Hall (1977) cites low viability of field collected spores as the

primary cause for low colonisation success, as well as the considerable time required for

some associations to establish (over a year in some cases). The use of glasshouse propagated

bulk soil trap culture spores was an effort to produce spores of a higher quality (i.e., less

parasitised) than field collected spores, however with no successful colonisation there can be

no definitive conclusion regarding the influence of spore origin or parasitism on culture

success. Other potential causes of culture failure may include spore dormancy. Other workers

have suggested that AMF propagules can remain dormant for several months, and that a cool

treatment is required for successful germination (Gemma & Koske 1988; Tommerup 1992).

Page 42: Characterisation of arbuscular mycorrhizal fungal

32

Alternatively, it may be that spore germination was successful but colonisation was too

localised to be detected using either of the methods employed in this study.

Several factors have been shown to influence culture success, including sterilisation

methods, medium type, nutrient availability and host identity. Biocides, particularly

Benomyl, have been used to sterilise soils of AM in experimental studies. Concerns of

residual toxicity prevented their use in this study. Heat based methods have been compared

unfavourably to sterilisation by fumigation or exposure to gamma radiation due to heat

induced release of “nutrient flushes” and heavy metals (Trevors 1996). Despite this, in the

absence of a gamma source, autoclaving was selected over fumigation to avoid the potential

of residual chemical toxicity in the medium. Particle size and bulk density have been shown

to significantly influence the production of spores from trap culture, but the medium

combination used in this study had previously been found to work (Ridgway et al. 2006).

Fertilisation of the trap cultures was avoided, as it is recommended only under signs of

extreme nutrient deficiency (chlorosis of young leaves), and nutrient poor conditions favour

the establishment of AMF (INVAM, URL: http://invam.caf.wvu.edu). It is unlikely that the

choice of T. repens as host plant influenced the outcomes of the inoculation attempts, as other

authors have reported successful use of T. repens (Liu & Wang 2003; Ridgway et al. 2006).

In this study, AMF structures and DNA were successfully detected in the roots of T. repens

from bulk soil trap culture. In studies where single spore cultures have been successfully

initiated (using a different host), seedlings were grown for 10–12 days before inoculation

(Morton et al. 1993), indicating that the developmental stage of the host roots may be a

critical aspect of AMF colonisation not taken into account in this study.

Many methods for extracting spores from soil samples advocate the homogenisation

of the soil samples (i.e., (Klironomos et al. 1993)). It is unlikely, therefore, that the process of

pooling the soil samples used in this study had any significant negative impact on the

viability of the AMF spores. Furthermore, it is considered that the potential benefits from

pooling soil samples outweighed any disadvantages. Pooled soil samples could be expected

to contain a greater diversity of AMF spores, and therefore a greater range of infectivity. The

major disadvantage from pooling soil samples is the loss of spatial resolution, however the

spatial distribution of AMF spores was outside the scope of this study.

Page 43: Characterisation of arbuscular mycorrhizal fungal

33

Molecular methods

The AMF-specific primers tested in this study varied in their ability to amplify positive

controls. Amplification success depended on the PCR protocol, and on the concentration of

AM fungal DNA present within the template. The AM1–NS31 primers can successfully

amplify positive controls, but they were sensitive to template concentration. Undiluted

positive controls could not be amplified, suggesting inhibition either from too much DNA, or

by some other substance within the DNA extractions. Under PCR conditions where AM

fungal DNA is successfully amplified from positive samples, the primer pair was also found

to amplify non-target DNA from negative samples, some fragments of which presented as

amplicons of the expected size (Figures 4-6). This non-specific binding has been reported

from AM1-NS31 in previous studies (Douhan et al. 2005), as has the inability of the AM1–

NS31 primers to detect some basal AM groups (Clapp et al. 1995; Helgason et al. 1999). As

pooled root DNA samples could be expected to amplify even in the absence of AM fungal

DNA the AM1–NS31 primers were considered inappropriate for screening trap cultures.

The AmL1-AmL2 primers were the most successful of the three primer sets tested,

being able to amplify AM fungal DNA from environmental root DNA template, and function

as nested primers with the NS1-NS4 primer pair. However, when the AmL1-AmL2 primers

are used in nested PCR with template lacking AM fungal DNA amplification of vascular

plant, fungal and invertebrate rDNA occurred. Furthermore, the amplicons resulting from this

non-specific binding were the same size as was expected for AMF gene fragments. Nested

PCR using the NS1-NS4 primers was carried out by Lee et al. (2008) to amplify AM fungal

DNA from spores and AMF roots. It was also used to test the specificity of the primers

against non-Glomalean DNA using template from non-AMF roots and leaves, including

T. repens which did not amplify. Lee et al. (2008) reported only very faint amplification of

two plant species (Sorghum bicolour and lilium tigrinum) even under low stringency

conditions (annealing at 50 °C). The experiments in this chapter have shown non-specific

amplification at least six distinct organisms not reported in their publication, including

Basidiomycetes (Sistotrema brinkmannii, Sistotrema farinaceum), algae (Bracteacoccus

cohaerens, Chlamydomonas rosae) and a bryophyte (Leptobryum pyriforme). The authors

stated that the AmL1–AmL2 primer pair were a solution to the issues of non-specificity

detected in the AM1–NS31 primer pair, and that they could be confidently used to

characterise environmental samples using non-sequence based approaches such as terminal

RFLP. However this study has demonstrated that care must be taken when using non-

Page 44: Characterisation of arbuscular mycorrhizal fungal

34

sequence based approaches, even when the “improved” AMF primer set is used, as non-

specific binding is possible.

The addition of BSA to the mastermix resulted in amplification success with the

AmL1-AmL2 primers over a broad range of template concentrations, with no evidence of

non-specific binding. The specific action whereby BSA promotes amplification is poorly

understood, however several studies of AMF from environmental samples have reported its

utility as a PCR additive to improve Taq function (Pivato et al. 2007; Davison et al. 2011). It

is reccommded that future studies investigating AMF using molecular methods include BSA

addition as part of the PCR optimisation process.

The positive control samples could not be amplified using either the SSUmAf-

LSUmAr or the SSUmCf-LSUmBr primer pairs, even when a nested approach was used

under a range of PCR conditions. Other workers have also reported difficulty with this primer

set (H. Ridgway pers. comm. 2011). In their published paper Krüger et al. (2009) state their

use of Phusion High-Fidelity DNA polymerase 2x mastermix (Finnzymes, Espoo, Finland)

for all of their reactions, however it is unlikely that the polymerase type used in this study is

responsible for the lack aplification (D. Orlovich, pers. comm. 2012). BSA was not used as a

PCR additive with these primers, and future studies may find that this will improve their

amplification success.

DNA origin

DNA extraction from single spores is routinely reported in molecular studies of AMF, but

there are several methods used for disrupting and extracting DNA. The methods used in this

study have been successfully used elsewhere (Raab et al. 2005), but others have reported

successful amplification of DNA from single spores crushed directly into the PCR mastermix

(Lee et al. 2008). The lack of amplification by any of the primer pairs used in this study

probably indicates a lack of template DNA, however this was not tested using gel

electrophoresis, and more attempts are required to optimise this methodology.

The failure of soil DNA extractions to produce AM fungal DNA in this study is

attributed to the failure of the single spore culture to produce AMF structures. The presence

of 18S rDNA from other eukaryotic organisms in the sequenced clones shows that DNA was

successfully extracted from the soil samples. In the absence of positive amplification of AM

fungal DNA this study cannot definitively critique the use of soil DNA samples, although the

Page 45: Characterisation of arbuscular mycorrhizal fungal

35

relatively high cost of soil DNA extraction kits makes a less attractive approach as a method

for screening trap cultures.

The most successful DNA extraction method in this study was the use of dried root

tissue. Root DNA based methods are preferable for AMF screening as they are less likely to

contain contaminants which can inhibit PCR success, although this study did find evidence of

PCR inhibition in undiluted root DNA samples in the absence of BSA. This may be the result

of the relatively high proportion of host DNA to AM fungal DNA. Comparisons of the

influence of root tissue preservation methods have found that the drying process had a

negative effect on AM fungal DNA integrity. Other preservation methods which maintain the

hydration of the sample (i.e. freezing) produce DNA of a quality that was comparable to fresh

tissue samples (Bainard et al. 2010). Therefore, freezing should be the preferred method of

preservation in future studies using molecular based methods to screen trap cultures or isolate

AM fungal DNA from environmental samples.

Conclusion

The production of monospecific AMF inoculum is a key step in investigating the functional

roles of AMF, including their potential role in the invasion of indigenous vegetation by exotic

plant pests. In this study, AMF propagules were successfully extracted from fresh field soils,

and from field soil which had been maintained in a glasshouse situation for 3 months, but no

attempts to inoculate seedlings with single spores were successful. In the absence of

successful cultures no comparisons can be made regarding the identities of AMF from the

two field sites. Further studies may investigate culture conditions, including host age at

inoculation, the use of pre-germinated AMF spores, and alternative culture media and may

thereby improve culture success and allow these questions of functional and genetic diversity

to be readdressed. This study has identified a labour saving method for screening multiple

culture attempts simultaneously, without the need for labour intensive microscope-based

observations of root colonisation. However, under certain PCR conditions evidence of non-

specific amplification was found that could not be differentiated from positive amplification,

suggesting that care should be taken when using AMF-specific primers for the verification of

AMF presence in samples with little or no AM-fungal DNA.

Page 46: Characterisation of arbuscular mycorrhizal fungal

36

Chapter 3: Spatial and phylogenetic diversity of arbuscular

mycorrhizal fungi associated with Hieracium lepidulum

Introduction

Both biotic and abiotic processes can be reflected in the spatial structures on biological

communities (Smith & Wilson 2002). The structures themselves are important components of

biodiversity, and a key goal of ecology is the attribution of mechanisms to structures in

nature. The organisms that comprise natural communities differ in scale, with each

experiencing diverse and scale-dependent interactions with the biotic and abiotic

environment. As a result, the processes that produce spatial structures tend also to be scale

dependent (Legendre 1993).

At the largest scales, spatial structures can be attributed to the interactions between

organisms and large scale processes (such as orogenesis, vicariance and climate). At these

scales the abiotic environment drives local similarity by filtering adaptive traits. At smaller

scales, particularly within homogenous environments, the distribution of species and

individuals are more likely to be determined by biotic factors, such as dispersal, competition

and predator-prey interactions (Allen & Starr 1982; Legendre & Fortin 1989). The study of

the spatial structures at small scales can therefore indicate the types and strength of biotic

processes acting within a community. The scales and extents of spatial structures within

communities can be measured by testing the observed distributions against a null model

(Legendre 1993).

One such indicator of spatial structures is spatial autocorrelation (SAc), which is a

measure of whether nearby localities are be more (or less) similar than expected by random

chance. Spatial autocorrelation often occurs because nearby localities are influenced by the

same “generating process” (Legendre & Fortin 1989). Epidemics or natural disasters are

examples of large scale ecological processes that could be described by SAc, however many

smaller scale ecological processes can also produce spatial autocorrelation, generally in the

form of patches or gradients. A priori knowledge of spatial heterogeneity within a given

system is important when testing experimental hypotheses: samples taken from within the

scales at which SAc is operating will not be independent and classical statistical tests will

reject the null hypothesis more often than is warranted due to violation of underlying

assumptions of independence.

Page 47: Characterisation of arbuscular mycorrhizal fungal

37

The spatial structure and diversity of arbuscular mycorrhizal fungi (AMF) in natural

systems have been relatively unstudied (Chaudhary et al. 2008; Dumbrell et al. 2010). Early

studies of AMF distributions found evidence of spatial structure at sub-metre scales

(Klironomos & Moutoglis 1999; Carvalho et al. 2003), however these studies used the proxy

of spore abundance, which is known to vary seasonally and in response to the plant host

identity (Bever et al. 1996). There is also evidence that neither the correlation between spore

abundance and extraradical hyphal biomass, nor the morphological species concept, are

sufficiently resolved to provide meaningful measures of diversity and abundance (Clapp et al.

1995). More recent studies of AMF community structure have utilised deoxyribonucleic acid

(DNA)-based molecular methods. Using AMF-specific primers, typically targeting the 18S

small subunit (SSU) ribosomal DNA (rDNA) genes, AM fungal DNA can be amplified from

environmental DNA samples, typically extracted from either soil or roots. Studies using

DNA-based methods to study AMF spatial structure are few, but indicate that the strength of

spatial autocorrelation within AMF communities depends on the habitat type under

consideration and the scale at which the study is conducted. Significant positive spatial

autocorrelation has been found in an AMF community in a temperate grassland environment

(Mummey & Rillig 2008). Using 18S rDNA-based methods on soil samples AMF

community composition was found spatially structured at scales of < 50 cm, and that

extraradicle hyphal lengths were spatially structured at < 30 cm (Mummey & Rillig 2008).

Glomus 25S large subunit (LSU) rDNA ribotypes isolated from Hieracium pilosella roots

along parallel transects were found to have extents ranging from several centimetres up to ten

metres (Rosendahl & Stukenbrock 2004). Conversely, no evidence of spatial structure was

found in AMF community composition in a calcareous wetland community using 18S rDNA-

based methods (Wolfe et al. 2007). The understanding of how AMF communities are

structured in nature is important for developing theoretical models to explain the distributions

of plants within plant communities, and to understand how AMF can facilitate plant species

coexistence.

AMF facilitation of plant species coexistence is recognised as an important factor

contributing to the maintenance of diversity and productivity in terrestrial ecosystems (Grime

et al. 1987; van der Heijden et al. 1998a; Wagg et al. 2011). However, the relative

importance of AMF facilitation for exotic plant invaders remains unclear. Until recently it

was commonly believed that AMF were “a swarm of interchangeable types” that could

associate equally well with any host plant (Smith & Read 2008). It is increasingly evident

that AMF can form distinct communities in the soil associated with specific plant hosts

Page 48: Characterisation of arbuscular mycorrhizal fungal

38

(Bever et al. 1996; Eom et al. 2000; Davison et al. 2011). Several case studies have found co-

occurring plant species that possess different AMF communities, indicating that some

mechanism for preferential association between plant and fungal taxa is operating within

plant communities (Helgason et al. 2002; Vandenkoornhuyse et al. 2003; Öpik et al. 2008). It

is therefore conceivable that invasive plants that form mycorrhizal associations will modify

AMF communities within an invaded range, with potential impacts on native biodiversity and

implications for more widespread invasion. Invasive plants have previously been implicated

as drivers of feedback within the soil microbial communities influencing the “trajectory of

invasion” (Vitousek & Walker 1989; Bever et al. 1997; Stylinski & Allen 1999; Vogelsang &

Bever 2009).

AMF research in New Zealand is still in its infancy; little is known about the

composition and structure of mycorrhizal communities, the number and distribution of AMF

species, or the relative specificity and identity of AMF associated with particular plant

species. It is unlikely that a single mechanism underlies the success of H. lepidulum as an

invader of subalpine Central Otago, but research into relative growth performance has

identified several avenues for research (Radford et al. 2007), including the potential of AMF

facilitation. The purpose of this study is to determine the composition and sub-metre spatial

structure of arbuscular mycorrhizal fungi growing on the roots of the invasive forb Hieracium

lepidulum, and to investigate the potential of AMF facilitated invasion.

H. lepidulum forms associations with AMF in the field, but neither the identity of the

AMF species involved, nor their functional significance have been determined (Downs &

Radford 2005; Roberts et al. 2009). There are two hypothetical scenarios regarding the

identity of AMF associated with H. lepidulum in its invaded range. It is possible that

European AMF species were introduced by colonial New Zealanders via some vector (e.g.,

bulk soil or contaminated machinery), and that these pre-adapted AMF have facilitated the

invasion of H. lepidulum into new habitats. Alternatively, H. lepidulum may associate with

generalist native AMF that are able to form a symbiosis with the exotic plant. An invader

whose persistence is facilitated by a specific suite of mutualists is expected to have little

variation in AMF community composition between individuals within a homogeneous

environment, whereas a generalist suite of AMF symbionts could vary in identity between

individuals. This study will use 18S rDNA-based methods to characterise the diversity of

AMF taxa associated with H. lepidulum individuals within a field site in Central Otago, New

Zealand. Phylogenetic analysis will be used to distinguish molecular operational taxonomic

Page 49: Characterisation of arbuscular mycorrhizal fungal

39

units (MOTUs) and to determine the relationships of the sequences with described species

and environmental samples from New Zealand and abroad.

It has been demonstrated that comprehensive sampling of AMF diversity requires

intensive sampling effort (Klironomos et al. 1999; Whitcomb & Stutz 2007). In this study,

the sampling effort is focussed on exhaustive sampling of the root systems of H. lepidulum

individuals within a 1.8 × 1.8 m plot, with records made of the physical location of each

individual plant. In addition to providing sufficient replication to ensure that the majority of

AMF taxa associated with H. lepidulum are sampled, this approach also allows the spatial

structure of the AMF community to be measured. Determination of AMF spatial structure

provides a snapshot of the distributions of individual ribotypes and phylogenetically defined

molecular operational taxonomic units (MOTU) at one point in time.

Invasive species can cause positive feedback by modifying the local environment to produce

conditions favourable for establishment of their progeny. Kilometre-scale spatial

autocorrelation in Hieracium spp. cover has previously been observed (Duncan et al. 1997),

and this study seeks determine whether spatial autocorrelation is also present in the

distributions of H. lepidulum individuals at the sub-metre scale. By measuring the positions

of individuals within a 10 × 2 m area, this study will reveal whether there is evidence of

positive feedback (clustering) by comparing the distributions of plant individuals within a

community against a null model of random distribution. The rectangular shape of the

sampling area was chosen to make the mapping of the H. lepidulum individuals with a total

station more systematic, thereby reducing the likelihood of missed individuals, and to reduce

trampling of the vegetation.

Study site

The Pisa Range is thought to be the point of origin of Hieracium lepidulum into the country,

probably via contaminated pasture seed from Europe (Chapman et al. 2003). H. lepidulum is

a prevalent species on the Pisa Range, particularly on Locharburn Station where it has been

present for at least 20 years (Chapman et al. 2004). Locharburn Station comprises 4250 ha of

low-production sheep and beef grazing on the southern slopes of the Pisa Range. The Station

boundaries extend from 300 to 1100 m asl and are bordered by Pisa Station to the west and

Locharburn scenic reserve to the east. Site characteristics include low phosphate retention

and a cool mesic soil temperature regime (Landcare Research 2011). The underling schist is

Page 50: Characterisation of arbuscular mycorrhizal fungal

40

exposed in increasingly larger tors towards the upper reaches of the station, and the pale,

schist-containing soils are of the Typic Argillic Pallic type. Median annual rainfall for the site

is between 700 and 750 mm per annum. The tors appear to have facilitated the persistence of

several indigenous tree species, including Podocarpus cunninghamii Colenso, Kunzea

ericoides (A.Rich.) Joy Thomps., and Phyllocladus alpinus Hook.f.

The proportion of H. lepidulum in the plant communities in the Locharburn station

varies with altitude. At 800 masl, the communities are less heavily grazed, tend to include

higher indigenous diversity than the lower slopes and H. lepidulum is present in relatively

low abundances. The plant communities at this elevation are dominated by matagauri

(Discaria toumatou Raoul), browntop (Agrostis capillaris L.), sweet vernal (Anthoxanthum

odoratum L.) and Bulbinella angustifolia (Cockayne & Laing) L.B.Moore, with patchily

distributed Aciphylla aurea W.R.B.Oliv. and Melicytus alpinus (Kirk) Garn.-Jones. On the

slopes above 800–1000 m H. lepidulum forms low meadows with a varying proportions and

composition of native and exotic herbs and grasses, predominantly Poa colensoi Hook.f.,

Leucopogon fraseri A.Cunn ex DC. and B. angustifolia. H. lepidulum senesces over winter

and can produce several rosettes of leaves from a single, persistent root system. At many of

these sites the density of H. lepidulum is too high to differentiate between individual plants

without disturbing the root systems. The sample site is located -44° 51' 8.26" S, +169° 16'

43.01" E (NZGD 2000) at a height of 800 m asl.

Methods

Sample collection

In order to ensure that replicate samples were not inadvertently taken from the same

individual, samples were taken from a community with obvious spatial separation of

H. lepidulum individuals. This method was also most amenable for determining the spatial

pattern of the plants. The site choice of site was also influenced by the assumption that low

sparse vegetation should have correspondingly shallow roots and therefore be easier to

sample. To reduce the influence of environmental factors the H. lepidulum individuals were

sampled from a site with apparently homogenous environmental conditions, although this

was not empirically tested. Site selection, therefore, was based upon the density of H.

lepidulum individuals and the physiognomy, homogeneity and diversity of site vegetation.

Individual plants were marked with tags and counted until a roughly symmetrical area

of thirty-six H. lepidulum individuals was demarcated. This sampling technique was used to

Page 51: Characterisation of arbuscular mycorrhizal fungal

41

ensure that all individuals in an area were included and that the interpretation of the spatial

structure was not confounded by gaps of unsampled individuals.

A total station (Leica T410) was used to measure the spatial coordinates of each of the

marked H. lepidulum plants within the sample plot, as well as the location of all individual H.

lepidulum rosettes from a 10 × 2 m transect within the same community. The coordinate data

were managed using Leica Geo Office Tools software.

Extraction of the samples was limited by the practical restrictions on representative

sampling in soil environments, namely the fragility and unknown extent of the belowground

root structures. A compromise between representation and practicality was reached by

excavating a symmetrical 15 × 15 × 15 cm mass of soil centred on the plant. For the purposes

of this study, the AMF colonising the roots contained within these soil masses represent the

fungal diversity of each plant sample. Plants with neighbours < 15 cm away necessitated the

excavation of larger clods with overlapping sample extents to accommodate the arbitrary 15

cm spatial grain. Samples were stored in plastic bags and the spades were scrubbed and

rinsed between individuals to avoid soil cross-contamination. The samples were stored at

ambient temperature for 5 hours in transit and then stored at 4°C until processed (no longer

than 48 hours).

Sample processing

H. lepidulum plants and root systems were manually extracted from the soil by manipulation

in tap water, with particular care taken to keep the root system intact. Samples were

processed individually, with the extracted individual plants transferred to plastic bags and

stored on ice. At this stage some neighbouring samples were pooled as it became evident that

they constituted single individuals. This reduced the sample count from 36 to 30, which is the

minimum sample number recommended to detect spatial autocorrelation using the Mantel

test (Legendre & Fortin 1989). Other plant species found within the plant community were

similarly extracted, identified and washed in tap water. Subsamples of these species’ root

systems were preserved in 50 mL centrifuge tubes containing 70% (v/v) ethanol prior to

staining to determine their mycorrhizal status.

Within 24 hours of being removed from soil, the H. lepidulum roots were excised

from the aboveground structures and washed three times in distilled water to remove all

remaining soil particles and unattached root fragments. Any root fragments that were not

attached were discarded. The wet weight of the belowground samples were measured,

Page 52: Characterisation of arbuscular mycorrhizal fungal

42

however the aboveground structures were not weighed for biomass data because of obvious

grazing. The number of culms and rosettes per plant were counted as a proxy for

aboveground biomass. The washed roots from each sample were then patted dry on paper

towels and cut into ~ 2 cm fragments. Root fragments were taken from each sample to from

five replicate 100 mg subsamples. The samples were transferred into sterile 1.5 mL

microcentrifuge tubes and stored at −18 °C. If insufficient root material was available for five

subsamples then fewer subsamples of the same size were taken.

DNA extraction

The sample tubes were immersed in liquid nitrogen and maintained below freezing point

while sterile micropestles were used to grind each sample into a fine powder. While still

frozen, 500 µL of lysis buffer (Genomic DNA Mini Kit (Plant), Genaid) was added and

allowed to thaw while mixing with the micropestle. DNA was extracted from the samples

using the Geneaid Plant Mini Kit following the manufacturer’s instruction, with a final

elution volume of 30 µL.

PCR conditions

An ~800 bp region of AMF 18S rRNA (SSU) gene was selectively amplified using the AMF

specific primer pair AmL1–AmL2 designed by Lee et al., (2008). These primers have been

shown to amplify AMF DNA, including basal groups Ambisporaceae and Paraglomaceae,

making them more representative than the hitherto most commonly used AMF specific

primers AM1 and NS31 (Simon et al. 1992; Helgason et al. 1998; Lee et al. 2008).

Polymerase chain reactions (PCR) were set up in 20 µL volumes comprising: 1 µL

DNA template, 10 pmol of each primer, 0.02 µL of 10% (w/v) bovine serum albumin (BSA),

and 17 µL of 1.1 × concentration Thermoprime® Reddymix containing 1.25 U

Thermoprime® Plus DNA polymerase, 75 mM Tri-HC (pH 8.8 at 25°C), 20 mM (NH4)2SO4,

1.5 mM MgCl2, 0.01% (v/v) Tween® 20, and 0.2 mM of each nucleotide.

PCR conditions were modified from the “field sample” protocol of Lee et al., (2008),

with the initial denaturing step of 94°C for 3 minutes, followed by 30 cycles of 94°C for 30

seconds, 58°C for 40 seconds and 72°C for 30 seconds with a final extension step of 5

minutes. Reactions were run on 1% agarose gels stained with ethidium bromide and

visualised by UV transillumination.

Page 53: Characterisation of arbuscular mycorrhizal fungal

43

Cloning

Replicate PCR reactions were pooled to produce a representative sample of rDNA fragments

from each of the 30 sampled plant root systems. This method of pooling replicate samples has

been shown to detect similar levels of AMF diversity to cloning each replicate individually

(Renker et al. 2006). There was no attempt to control for variable amplification success in the

subsamples, which were pooled in equal volumes. Replicates that failed to amplify were

discarded.

The pooled PCR products were cloned into pCR®2.1-TOPO® TA vector and

transformed into chemically competent Escherichia coli Top10 following the manufacturer’s

instructions (Invitrogen). Transformed cells were screened using blue/white colony screening

on 90 mm plates containing Luria-Bertani (LB) agar, 50 µg mL-1 kanamycin and coated with

40 mg mL-1 X-gal.

Positive transformants were picked from each of the 30 cloning reactions and

transferred to patch plates. Using the same PCR protocol as described above, between 47 and

70 transformants from each sample were then amplified using 1 µL of diluted transformed

cells as template. Transformants that failed to amplify, or that produced bands different to the

expected size were excluded from subsequent analysis.

RFLP typing

Restriction fragment length polymorphism (RFLP) analysis was used to screen the cloned

rDNA fragments. In total, 1418 amplicons were individually digested with restriction

enzymes HinfI, AluI, and Sau3AI (Roche Applied Science, Mannheim, Germany) in 12.5 µL

reactions comprising 2.5 µL PCR product, 0.5 U enzyme, 1.25 µL of 10× concentration

SuRE/Cut™ buffer A (Roche Applied Science, Mannheim, Germany) in 8.25 µL autoclaved

MilliQ H2O. Reactions were digested for 4 hours at 37 °C. A 10 µL aliquot of each reaction

was analysed by electrophoresis on 1.5% agarose gels stained with ethidium bromide and

visualised by UV transillumination using a 1 kb plus ladder as standard. Observed banding

patterns were recorded and categorised using a library of banding-pattern images that were

compiled as the samples were processed.

Sequencing

An attempt was made to sequence every unique RFLP type. Plasmids were purified using a

QIAprep Spin Miniprep Kit (Qiagen). Sequencing reactions were performed by the Genetics

Page 54: Characterisation of arbuscular mycorrhizal fungal

44

Analysis Service, Department of Anatomy, University of Otago, using dye terminator

sequencing chemistry (BigDye Terminator v. 3.1, Applied Biosystems) with vector primers

directed against the M13f and M13r promoter regions. Reactions were sequenced on an ABI

3730xl DNA Analyser (Applied Biosystems). The software GENEIOUS v5.1 (Drummond et al.

2010) was used to simulate the restriction digestion of sequenced DNA fragments, in order to

verify concordance between RFLP-types and sequence data.

Staining and quantifying AMF infection

To determine the AMF status of H. lepidulum and the other plant species found in the study

site, roots from arbitrarily selected individuals were washed and cut into 2 cm fragments,

placed in biopsy cassettes (SimPort) and cleared in 10% w/v KOH at 90°C for 10–20

minutes. Thick-rooted species required longer clearing time than thin rooted-species. The

samples were then rinsed 3 times in tap water and placed in 1 % HCl for 30 minutes to

acidify the roots and improve staining. The samples were stained in 1:1:1

water:glycerine:lactic acid (v/v/v) with 0.05% w/v trypan blue for 12–16 hours. After

staining, the root samples were stored in water at 4 °C for at least 48 hours to allow excess

stain to leach from the plant tissue. AMF infection rates were quantified for each plant using

10 root fragments that were mounted in parallel on slides and scored at 200 × magnification

using the modified line-intersect method (Tennant 1975).

Phylogenetic analysis

The RFLP abundance data and associated sequence data were used to determine the diversity

and relative abundance of AMF ribotypes. In GENEIOUS, the software CLUSTALW2 (Larkin et

al. 2007) was used to align the sequenced 18S SSU rRNA gene fragments with sequence data

from an AMF phylogeny (Schüβler et al. 2001a; Krüger et al. 2012), published sequences

from two New Zealand studies (Russell & Simon 2005; Bidartondo et al. 2011), and the

closest BLAST matches from the NCBI database (URL: http://www.ncbi.nlm.nih.gov).

PAUP* v. 4.0b10 (Swofford 2003) was used to construct a neighbour-joining tree using a

general time-reversible model of nucleotide substitution rates, with equal rates of variation.

Endogone pisiformis was used as the outgroup. PAUP* was also used to construct a maximum

parsimony bootstrap tree. The bootstrap support was calculated using a heuristic search and

random addition of sequences. The analysis was terminated after 10,000 replicates and the

bootstrap replicates were summarised by a 50% majority rule tree. The bootstrap values were

Page 55: Characterisation of arbuscular mycorrhizal fungal

45

then transposed onto the neighbour joining tree. Following the methods employed by most

AMF community studies using molecular methods, the taxonomic units were defined by well

supported clades of ≥ 97% sequence identity.

Functions in the R package PICANTE (Kembel et al. 2011) were used to assess aspects

of community phylogenetic structure. The function pd was used to calculate Faith’s

phylogenetic diversity, for correlation with root biomass and above ground structures (Faith

1992). The function comm.phylo.cor() was used to measure community phylogenetic

structure, using the Jaccard index of co-occurrence, with a null model of phylogeny tip label

shuffling (sample.taxa.labels), on a neighbour joining tree constructed from

representative samples from each MOTU.

Diversity estimation

Non-parametric estimators are useful when it is difficult to determine whether scarcity in the

dataset is the result of natural scarcity, or could be an artefact of the sampling method (Chao

2004). The Chao method (Chao 1987) is a non-parametric estimator that uses the proportion

of rare taxa in the site to predict the number of unsampled taxa:

�� = �� +������

�����

where Sp is the predicted species richness, So is the total number of observed species and a1

and a2 are the number of species detected only once or twice in the collection respectively.

This bias-corrected form of the Chao equation provides richness estimates unaffected by

cases where a2 is zero (Chao 2004). The statistical analyses for this study were performed in

R version 2.13.2 (R Development Core Team 2011b).

A value of Chao and its associated standard deviation was obtained by using the R

function specpool() in the package VEGAN (Oksanen et al. 2011). Because of the

sensitivity of this method to rare species, these extrapolations are highly dependent on the

taxonomic resolution used in this study. Analysis using higher similarity phylogenetic

clusters would produce correspondingly higher extrapolated taxonomic diversity of the site as

more singletons would be represented in the data. Chao diversity was estimated at each

number of samples, using the R function poolaccum(), in the package VEGAN to show

how the estimated diversity was influenced by sampling effort.

,

Page 56: Characterisation of arbuscular mycorrhizal fungal

46

Community analysis

Accumulation curves were constructed in order to assess the rate of taxon discovery at the

site and to visually represent estimated total diversity. The curves were produced by random

permutation of taxon incidence between samples. The analysis was performed using the

function specaccum() in the package VEGAN. In order to determine how the number of

clones screened per sample influenced the detection of diversity within the site, the data were

resampled under progressively more intensive regimes. Clone identities were resampled from

each individual plant to simulate differing levels of sampling effort, with 100 permutations at

each level. The analysis was performed using the function balanced.specaccum()

from the R package BIODIVERSITYR (Kindt & Coe 2005). Self-starting species-area models

for the species accumulation distribution were constructed using their respective functions in

the package VEGAN. Three species-area models were constructed:

Arrhenius: � × ��,

Gleason: √� + � × log �,

and Gitay: � + � × log �� ,

where n = number of H. lepidulum individuals sampled and k and z are estimated model

parameters (k= the expected number of species in a unit and z = steepness of the species-area

curve). These models were tested using nonlinear least-squares regression with the nls()

function in the package STATS (R Development Core Team 2011a) with the relative best fit to

the observed data being ascertained using the Akaike information criterion (AIC) (Akaike

1974). The best-fitting model was then extrapolated over a range of more intensive sampling

regimes.

In order to compare the relative abundances of the MOTUs present in the site, the

clone frequency data were used to construct dominance/diversity plot. The distributions were

fitted against four commonly used models: broken-stick (Null model), geometric

(preemption), general lognormal, Zipf and Zipf-Mandelbrot. The R function radfit(), in

the package VEGAN, was used to fit the models to the data, and find the best fitting model

according to the (AIC).

Page 57: Characterisation of arbuscular mycorrhizal fungal

47

Spatial analysis

The spatial structure of the AMF community was tested using Mantel tests (Mantel 1967)

comparing the AMF taxa (defined by either RFLP or MOTU) between plant root systems at

varying distances. Dissimilarity matrices of AMF data were constructed for taxon incidence

(Jaccard index) and standardised abundance (Bray-Curtis index). Following Sturge’s rule

(Sturges 1926), a matrix of Euclidean distance between samples was binned into equally

sized “distance classes”, which contained a variable numbers of sample pairs. This method

produces correlograms which are easier to draw and interpret (Legendre & Fortin 1989). The

two largest distance classes were excluded from further analysis because they contained too

few pairs to be statistically valid (Legendre & Fortin 1989). A correlation coefficient was

calculated to test for an overall spatial structure in the data using function mantel() in the

R package VEGAN (Oksanen et al. 2011). In order to test for statistically significant

correlation between samples within defined distance classes, correlograms were constructed

with the R function mantel.correlog() in VEGAN. The correlograms indicate the scale

and magnitude of the autocorrelation present in the community data. Progressive Bonferroni

correction was applied to account for multiple significance testing. An underlying assumption

of this analysis is stationarity i.e., that the site has a fixed diversity with no immigration or

emigration (Legendre & Fortin 1989).

Point pattern analysis was applied to test the distribution of H. lepidulum individuals

within the 10×2 m quadrat against a null model of random (Poisson) distribution using

Ripley’s K-function (Ripley 1977) Kest() in SPATSTAT (Baddeley & Turner 2005), with

confidence intervals produced by varblock(). An isotropic correction was applied to

account for bias resulting from the unobserved points outside the sampling area (Ripley

1988).

Results

Validation of extraction methods

AM fungal DNA was successfully extracted and amplified from all but one of the 147

subsamples of H. lepidulum roots. Due to a scarcity of root material from 3 plants, and a

single failed amplification, 4 of the 30 plant samples were represented by 4 (as opposed to 5)

pooled subsamples. Amplification strength was variable both within and between samples,

potentially indicating variation in DNA template concentrations between samples (Fig. 1).

Page 58: Characterisation of arbuscular mycorrhizal fungal

48

This may reflect AMF biomass within the roots, however amplification strength cannot be

quantified using this PCR method.

Figure 1. Electrophoresis gel showing products from the amplification of 4 samples (28, 14, 29 and 34),

with subsamples indicated by the letters a–e. The 1 Kb Plus ladder shows that the amplicons are the

expected size (~750 bp), although they vary in brightness between and within samples.

Validation of cloning methods

Between 45 and 64 positive transformants were successfully reamplified from each sample

using the AML1 and AML2 primer pair. Putative non-AMF DNA fragments of ~500 and

~1000 bp occurred in 1.7% of positive transformants (Fig. 2) and these were excluded from

subsequent analysis on the basis that they were likely to represent non-specific amplification.

Figure 2. Electrophoresis gel showing an example of the variation in amplicon size produced from cloned

DNA fragments from sample 14. Clones 19 and 27 were excluded from the RFLP analysis.

Page 59: Characterisation of arbuscular mycorrhizal fungal

49

RFLP analysis

The positive transformants were differentiated using RFLP analysis. The restriction enzymes

differed in their ability to detect variation between the clones, with HinfI finding the most

diversity (18 unique RFLPs), and AluI (14 unique RFLPs) finding the least (Fig. 3). Both

HinfI and Sau3AI were able to discriminate between several of the most abundant

RFLP-types, whereas AluI predominantly produced only 2 banding patterns (Fig. 4. a–c).

Considered together, the analysis of 1421 clones from 30 plants found 53 unique RFLP-types

(Appendix 1, Supporting data 1-3).

Figure 3. Three RFLP-type accumulation curves, with 95 % confidence intervals, each demonstrating the

total diversity and rate of discovery of new RFLP types by each of the endonucleases used in this study:

HinfI (blue), Sau3AI (red), and AluI (green).

Page 60: Characterisation of arbuscular mycorrhizal fungal

50

Figure 5. Dominance/diversity plot of RFLP abundance (note log scale). The distribution of abundance

data shows extremely high relative abundances of the most common taxa and a large number of

“singleton” RFLP-types, which were only detected once at the site.

Figure 4. An electrophoresis gel

demonstrating the variation in

discriminatory ability shown by the

three endonucleases HinfI (top),

Sau3AI (middle), and AluI (bottom),

on the same 9 sequences (sample 1,

clones 19-27).

Page 61: Characterisation of arbuscular mycorrhizal fungal

51

The RFLP analysis revealed a highly uneven community of AMF within the roots of

H. lepidulum, with the majority of the clones (892/1421) found to be of a single RFLP-type

(type 633), and only one other RFLP-type that was represented by more than 10% of the

clones (type 111). A large proportion (68%) of the total RFLP diversity was comprised of

RFLP-types that were detected only once at the site (Fig. 5). These singleton RFLP-types

were not confined to particular samples, but distributed throughout the site.

Phylogenetic analysis of sequence data

Diversity measured by phylogenetic analysis was lower than diversity detected by the RFLP

analysis due to high sequence similarity between several RFLP types. Attempts to sequence

all unique RFLP types were made, however the majority of the clones representing singleton

RFLP types could not be sequenced. As a result, some putatively rare taxa were not

represented in the later analyses. SSU fragment sequences were produced for 69% (36/52) of

RFLP patterns found, representing 98.9% (1405/1421) of all colonies analysed. Highly

supported terminal clades were used to define phylogenetic clusters, following the method of

Helgason et al. (1999). Retrospective comparison of RFLP-type diversity measures shows a

degree of consistency between the RFLP banding patterns and the clusters defined by

parsimony analysis, however some of the RFLP-types could not be reproduced in silico using

virtual restriction digestions of the sequence DNA, potentially indicating further

uncharacterised diversity. The RFLP-types whose sequences constituted single phylogenetic

clusters generally differed by a single restriction enzyme RFLP, usually HinfI.

In order to determine the phylogenetic relationships between the RFLP-types detected

in this study and elsewhere a neighbour joining tree annotated with bootstrap values from a

maximum parsimony bootstrap analysis was constructed (Figure 6). The tree includes

described AMF taxa sequences from an AMF rDNA phylogeny (Schüβler et al. 2001b), AM

fungal rDNA sequences from two published New Zealand studies, and the most similar

sequences from around world as defined by a BLAST algorithm search (Altschul et al. 1990)

against the NCBI and MAAJAM databases (URL: http://blast.ncbi.nlm.nih.gov;

http://maarjam.botany.ut.ee) were also included (Table 2).

Page 62: Characterisation of arbuscular mycorrhizal fungal

52

Table 2. Top BLAST search results of representative sequences from the MOTUs found in H. lepidulum

root systems, showing NCBI accession number and percent similarity.

MOTU Best Match Identities Host Plant Collection Location

GlomA02 AJ699069 751/753 (99.7%) Marchantia foliacea New Zealand

GlomA04 JF414177 751/751 (100%) Marchantia foliacea New Zealand

GlomA06 JF414176 743/754 (98.5%) Tmesipteris tannensis New Zealand

GlomA05 JF414191 751/754 (99.6%) Symphyogyna subsimplex New Zealand

GlomA03 JF414180 749/753 (99.5%) Phaeoceros carolinianus New Zealand

GlomB01 FN869808 758/760 (99.7%) Psoralea bituminosa Spain

Acau02 FN825900 750/755 (99.3%) Plantago lanceolata Scotland

Acau01 EU332730 726/728 (99.7%) Miscanthus sinensis South Korea

GlomA01 FR848612 745/753 (98.9%) Glycine max United Kingdom

Backbone support of the 50% parsimony bootstrap tree was high at family level, with

Glomeraceae and Diversisporaceae emerging as well-supported sister clades. The genus

Rhizophagus was monophyletic, with 73% bootstrap support, and was nested within a larger,

well-supported (100% bootstrap support) clade containing sequences from environmental

samples. The Rhizophagus clade contained a single sequence (GlomA01) from this study,

which was detected only once at the site, and was most closed related to an environmental

sample from soybean (Glycine max) roots from the United Kingdom (Table 2). The most

abundant phylotype detected in this study (GlomA02) formed a clade with environmental

sequences from New Zealand, Ecuador and the USA. Sister to this clade was a well

supported (91%) monophyletic group containing GlomA03, along with a New Zealand

environmental sample from the bryophyte Phaeoceros carolinianus. The second most

common phylotype from this study (GlomA04) formed a well supported monophyletic group

with environmental samples of New Zealand origin. This was nested within a strongly

supported (98% bootstrap support) clade of environmental samples from China and the USA,

which has an unresolved placement within Glomales, and no closely related, named species.

Sequences of Glomus macrocarpum were monophyletic within a strongly supported (98%

bootstrap support) clade of environmental samples. The placement of the phylotype

GlomA05 was poorly resolved within this clade, although there was weak support (72%

bootstrap support) of a relationship with New Zealand environmental samples. The forth

most common phylotype (GlomA06) formed a well supported monophyletic group (95%

bootstrap support), containing only sequences from this study, nested within environmental

samples from Ecuador. A clade containing sequences of the genus Claroideoglomus was well

supported (98% bootstrap support), however the relationships between sequences from

described species and the environmental samples was poorly resolved. GlomB01 formed a

Page 63: Characterisation of arbuscular mycorrhizal fungal

53

polytomy with environmental samples from New Zealand, Spain and the United Kingdom.

Two sequences were placed within the clade (100% bootstrap support) containing sequences

for Acaulosporaceae. Acau01 was placed within a strongly supported clade (100% bootstrap

support) containingsequences for Acaulospora longula and environmental samples from

South Korea, while the placement of Acau02 within Acaulosporaceae was poorly resolved.

On the basis of this phylogeny, nine AM fungal MOTUs from at least four genera are

proposed: Rhizophagus, Funneliformis, Claroideoglomus (Glomeraceae) and Acaulospora

(Acaulosporaceae) (Schüβler et al. 2001a; Krüger et al. 2012), with evidence of another

strongly supported taxonomic group not characterised by any type specimens (GlomA04), but

which were identical to sequences obtained from the thallus of the liverwort Marchantia

foliacea from a New Zealand podocarp/broadleaf forest.

Page 64: Characterisation of arbuscular mycorrhizal fungal
Page 65: Characterisation of arbuscular mycorrhizal fungal

0.02

Funneliformis fragilistratus W3238

3173

environmental sample, Spain

environmental sample (Alzatea verticillata), Ecuador

663

Glomus macrocarpum W5293

environmental sample (Alzatea verticillata), Ecuador

Diversispora eburnum

environmental sample (Coronilla juncea), Spain

Archeospora trappei W3179

333331

1333

113

Funneliformis constrictum

environmental sample (Alzatea verticillata), Ecuador

6310

Glomus sp. W3347

Endogone pisiformis

12.3

environmental sample (Austrofossombronia australis), NZ12.1

4179

693

Acaulospora laevis W3347

Scutellospora nodosa BEG4

Paraglomus occultum IA702

environmental sample (Marchantia foliacea), NZ

Diversispora trimurales W4124

Scutellospora fulgida W2993

environmental sample (Miscanthus sinensis), South Korea

environmental sample (Psoralea bituminosa), Spain

environmental sample, Spain

Gigaspora sp. W2992

Diversispora aurantia

environmental sample (Prunus africana), Spain

environmental sample (Alzatea verticillata), Ecuador

12141

Scutellospora spinosissima W3009

Acaulospora sp. W3424

443

Claroideoglomus etunicatum

1111

6133

Funneliformis sp. WUM3

Acaulospora longula

Rhizophagus fasciculatus BEG53

Diversispora spurca W3239

411

environmental sample (Marchantia foliacea), NZ

Glomus macrocarpum W5605

Rhizophagus clarus W3776

Rhizophagus clarus W3095

Claroideoglomus lamellosum

551

Geosophon pyriformis

13223

environmental sample (Marchantia foliacea), NZ

111

environmental sample (Alzatea verticillata), Ecuador

Rhizophagus sp.

6313

Acaulospora sp. W4681

Acaulospora brasiliensis W4699

Scutellospora heterogama

Rhizophagus sp.

Funneliformis verruculosus W3295

environmental sample (Glycine max), UK

Acaulospora brasiliensis W4699

environmental sample (soil), Canada

environmental sample (Rhododendron mucronulatum), South Korea

environmental sample ( Podomitrium phyllanthus) , NZ

environmental sample (Symphyogyna hymenophyllum), Australia

613

Scutellospora weresubiae W2988

Rhizophagus clarus W3224

environmental sample (Miscanthus sinensis), South Korea

843

environmental sample (Jensenia connivens), NZ

Scutellospora pellucida W4761

Archeospora schenckii W3571

environmental sample (Glycine max), UK

12.4

638

environmental sample, Spain

973

Scutellospora heterogama

environmental sample (Anthyllis cytisoides), Spain

Acaulospora longula W3302

Acaulospora brasiliensis W4699

Funneliformis caledonius BEG15

environmental sample (Marchantia foliacea), NZ

111

473

Funneliformis caledonius BEG20

2311

Claroideoglomus lamellosum

environmental sample, Spain

Scutellospora aurigloba W3121

Claroideoglomus claroideum

environmental sample (Miscanthus sinensis), South Korea

environmental sample (Asterella tenera), NZ

environmental sample (Anthyllis cytisoides), Spain

environmental sample (Glycine max), UK

environmental sample (Phaeoceros carolinianus), NZ

Rhizophagus irregularis FL208

Claroideoglomus claroideum

environmental sample (Miscanthus sinensis), South Korea

Pacispora franciscana

Acaulospora lacunosa BEG78

Ambispora fennica W3569

Rhizophagus sp.

9173

Diversispora celata

221

Scutellospora gilmorei W3557

environmental sample (Glycine max), UK

693

111

Claroideoglomus sp. W3349

environmental sample (Anthyllis cytisoides), Spain

environmental sample (Marchantia foliacea), NZ

environmental sample (Retama sphaerocarpa), Spain

Acaulospora sp. W2941

environmental sample (Retama sphaerocarpa), Spain

environmental sample (Marchantia foliacea), NZ

environmental sample (Phaeoceros carolinianus), NZ

Acaulospora laevis W3107

environmental sample (Phaeoceros carolinianus), NZ), NZ

433473

Acaulospora cavernata BEG33

environmental sample (Retama sphaerocarpa), Spain

environmental sample (Anthroceros laminiferus), NZ

Pacispora scintillans

environmental sample (Allium �stulosum), South Korea

331

9173

environmental sample (Retama sphaerocarpa), Spain

6123

Archeospora leptoticha NC176

Claroideoglomus sp.

Acaulospora spinosa W3574

environmental sample (Phaeoceros carolinianus), NZ

72

99

98

86

95

78

93

100

87

9894

98

100

97

88

100 90

89

99

9973

91

98

98

96

9162

100

98

67

100

97

91

9894

98

99

100

100

9862

100

84

100

76

100

10084

10069

100

100

100

64

95

92

100

82

63

100

73

environmental sample (Tabuia chrysantha), Ecuadorenvironmental sample (Tabuia chrysantha), Ecuador

environmental sample (Cedrella montana), Ecuador

environmental sample (Soil), USA

environmental sample (Phaeoceros carolinianus), NZ

environmental sample (Marchantia foliacea), NZenvironmental sample (Marchantia foliacea), NZ

environmental sample (Marchantia foliacea), NZenvironmental sample (Marchantia foliacea), NZ

environmental sample (Marchantia foliacea), NZ

environmental sample ( Symphyogyna subsimplex ), NZ

environmental sample (Perilla frutescens), Chinaenvironmental sample (Psilotum nudum), USA

environmental sample (Olea europaea), USA

environmental sample (Podocarpus falcatus), Ethiopiaenvironmental sample (Alzatea verticillata), Ecuador

environmental sample (Alzatea verticillata), Ecuadorenvironmental sample (Alzatea verticillata), Ecuador

environmental sample (Hyeronima sp.), Ecuadorenvironmental sample (Hyeronima sp.), Ecuador

environmental sample (Prunus sp.), Ecuadorenvironmental sample ( Tmespiteris tannesnsis), NZ

environmental sample (Symphyogyna prolifera), NZenvironmental sample ( Podomitrium phyllanthus), NZ

environmental sample (Marchantia foliacea), NZenvironmental sample (Symphyogyna subsimplex), NZ

environmental sample (Sciaphila tosaensis), Japanenvironmental sample (Retama sphaerocarpa), Spain

GlomA01

GlomA02

GlomA03

GlomA04

GlomA05

GlomA06

GlomB01

Acau01

Acau02

Rhizophagus

Funneliformis

Claroideoglomus

Glomus

Acaulospora

Glo

mal

esD

iver

sipo

rale

s

Pacispora

Scutellospora/Gigaspora

Diversispora

Arc

heos

pora

les

Paci

spor

ales

Figure 6. Neighbour joining phylogenetic tree based on concatenated ribosomal small subunit (18S SSU) gene fragments of Glomeromycota with 50% majority rule bootstrap values and outgroup (Endogone pisiforme). Sequence data include AM fungal DNA isolated from Hieracium lepidulum root samples (orange labels), with the closestmatches according to a BLAST algorithm search of the NCBI database, an AMF rDNA phylogeny from Schüβler et al. (2001), and all known fungal rDNA SSU sequences from New Zealand (green labels). Branches with less than 60% bootstrap support were not labelled. Well supported clades of ≥ 97% similar sequences have been labelled to indicate molecular operational taxonomic units (MOTUs) in this study, and the phylogenetic tree has been divided to indicate order and genus level distinctions.

Page 66: Characterisation of arbuscular mycorrhizal fungal

55

Despite repeated efforts to sequence them, 19 unique RFLP patterns remain unidentified. It is

unknown to what extent these represent artifacts introduced via polymerase errors, mutational

hotspots, cloning of heteroduplexes or chimeras (Patel et al. 1996; Speksnijder et al. 2001),

or whether some or all of them represent rare taxa. The 19 unidentified RFLP types account

for 1.3 % of the total clones screened in this study.

Plant-level variation in AMF communities

Individuals plants differed in the number of AM fungal MOTUs detected in their root

systems (Figure 7). While both the richness and evenness of the AMF communities varied

between individual plants, it is not possible to test for variation within each plant as the

replicate root DNA extractions were pooled together. It can be seen that AMF were detected

in the roots of every plant within the community, however no individual possessed all nine

MOTUs detected at the site. Of the 30 plants sampled, the most prevalent AMF taxon,

GlomA02 (67% of site clone abundances), was present 28 of the 30 plant samples. The two

samples which did not posses GlomA02 were the only samples where GlomB01 was

detected. Sequences placed within Glomeraceae were the most prevalent at the site,

colonising between 67–87% of individual plant root systems, apart from GlomA01, which

was detected only once. The members of Acaulospora were comparatively rare at the site:

Acau02 colonised 33% of samples and Acau01 was detected only once.

Figure 7. Histogram showing the frequency of which a given number of AM fungal MOTUs were

detected in individual H. lepidulum root systems

There was a significant negative correlation (p = 0.03) between MOTU co-occurrence and

phylogenetic distance within the site, indicating that the communities were less closely

related then expected at random (phylogenetic overdispersion). No significant correlation was

Page 67: Characterisation of arbuscular mycorrhizal fungal

56

found between AMF Shannon or phylogenetic diversity and the size of the host root system.

Similarly, the proportion of total root system sampled for DNA extraction was not

significantly correlated with AMF diversity.

Rank abundance analysis

MOTUs defined by strongly supported phylogenetic groups may be more representative of

biological species than RFLP-based diversity measures. Under a niche-based model of

species coexistence the relative abundance data represent the relative proportions of resources

utilised by each taxa. In order to determine the relative abundances of the MOTUs within the

site, and to test the distribution against model distributions, a dominance/diversity plot was

constructed (Figure 8).

Fig 8. Dominance/diversity plot of site MOTU relative abundance distribution overlaid with model

distributions. The lognormal distribution (bold green line) was the best model fit of the observed data

(AIC = 100.9).

Ecological interpretation of dominance/diversity plots is underpinned by an assumption that

the relative abundance of each phylotypes is directly proportional to the size of the niche it

occupies. Under the null model, the niche space is sequentially divided into successive

niches, with the probability of division proportional to the size of the niche – the “broken

stick” model of MacArthur (Macarthur 1957). The Akaike information criterion (AIC) found

that of the models tested, the rank abundance data is best fitted by a log-normal distribution.

From an ecological perspective, this distribution can be produced by a conceptually similar

mechanism to the broken stick model, except that the probability of niche division is

Page 68: Characterisation of arbuscular mycorrhizal fungal

independent of niche size (Wilson 1991)

are also recognised to produce log normal distributions.

Extrapolation of site diversity

The MOTU data were used to produce a

a sampled population is static, without immigration or emigration, repeated sampling will

reach an asymptote of total population richness. It can be seen from the shape of the observed

species accumulation curve (S)

were detected at a fairly stable rate

Figure 9. Observed species accumulation curve (S) and Chao estimates of total AMF phylotype richness

for the sampled site, including 95% confidence intervals.

Alongside the taxon accumulation curv

on the Chao estimator. This estimator is a representation of taxon accumulation of the

observed taxa, as well as taxa which were probably present

Unlike the observed taxon acc

the estimated sampled and unsampled diversity at a given sampling effort. D

influence of rare species (singletons and d

diversity, the size of the statistic is highly dependent on

When the taxonomic units comprise unique RFLP

singletons, the site diversity estimate is 379 taxa

diversity is estimated to be 11 taxa

(Wilson 1991). Several other ecological and statistical processes

are also recognised to produce log normal distributions.

Extrapolation of site diversity

OTU data were used to produce a taxon accumulation curve for the site (Fig. 9). When

a sampled population is static, without immigration or emigration, repeated sampling will

reach an asymptote of total population richness. It can be seen from the shape of the observed

species accumulation curve (S) that after 10 host plants were sampled the new phylotypes

were detected at a fairly stable rate.

9. Observed species accumulation curve (S) and Chao estimates of total AMF phylotype richness

for the sampled site, including 95% confidence intervals.

Alongside the taxon accumulation curve is the extrapolated accumulation curve based

on the Chao estimator. This estimator is a representation of taxon accumulation of the

observed taxa, as well as taxa which were probably present within the site but not detected.

Unlike the observed taxon accumulation curve, the Chao diversity estimate is an indication of

the estimated sampled and unsampled diversity at a given sampling effort. D

influence of rare species (singletons and doubletons) on the Chao estimate of

the size of the statistic is highly dependent on how taxonomic units are defined.

When the taxonomic units comprise unique RFLP-types (n = 56), of which the majority were

site diversity estimate is 379 taxa. When MOTUs are used (n= 9) the

diversity is estimated to be 11 taxa (Fig. 9).

57

. Several other ecological and statistical processes

accumulation curve for the site (Fig. 9). When

a sampled population is static, without immigration or emigration, repeated sampling will

reach an asymptote of total population richness. It can be seen from the shape of the observed

the new phylotypes

9. Observed species accumulation curve (S) and Chao estimates of total AMF phylotype richness

e is the extrapolated accumulation curve based

on the Chao estimator. This estimator is a representation of taxon accumulation of the

within the site but not detected.

umulation curve, the Chao diversity estimate is an indication of

the estimated sampled and unsampled diversity at a given sampling effort. Due to the strong

estimate of total site

how taxonomic units are defined.

, of which the majority were

MOTUs are used (n= 9) the site

Page 69: Characterisation of arbuscular mycorrhizal fungal

58

Spatial analysis

In order to detect spatial structure in the AMF communities, and to determine how spatial

structure was influenced by taxonomic resolution, Mantel tests were calculated using RFLP

data, and MOTUs defined by either 97% or 98% sequence similarity (Table 3).

Table 3. A comparison of the size and global significance of the Mantel test of spatial structure in the

AMF community data using either incidence or abundance based similarity indices. * indicates significant

values, ** indicates highly significant values

The distributions of both RFLP types and 98% similar MOTUs showed highly significant

positive spatial autocorrelation when incidence data were used. Where clone frequency was

included, the spatial autocorrelation decreased, and remained significant only for RFLP types

in the absence of singletons. It would be expected that the removal of the RFLP singletons

would have a negligible effect on size or global significance of the Mantel statistic (r), as it

was calculated using rank similarity based indices. The merging of MOTUs at the 97%

similarity level resulted communities defined by six, rather than nine distinct MOTUs due to

the collapse of GlomA01, GlomA02 and GlomA03 into a one MOTU, and GlomA05 and

GlomA06 into another. When the AMF communities were defined by 97% sequence

similarity, no significant spatial autocorrelation was detected.

A correlogram at 98% MOTU similarity was constructed to determine the spatial

extent of spatial autocorrelation in the AMF communities (Figure 10). The first and third

distance classes (< 0.1, 0.27-0.44 m respectively) both contain pairs of samples which are

significantly more similar than expected at random. The significant similarity found in the

smallest distance classes may be expected purely on the basis of the sampling methodology,

as sample pairs within this distance classes were taken within overlapping sampling extents.

Taxonomic unit(s) Mantel statistic (r)

Incidence data Abundance data

RFLP types 0.30** 0.15

RFLP types (ex singletons) 0.28** 0.16*

98 % similar MOTUs 0.26** 0.15

97 % similar MOTUs 0.063 0.09

Page 70: Characterisation of arbuscular mycorrhizal fungal

59

Figure 10. Correlogram showing the correlation over distance of Jaccard similarity in the AMF

phylotypes colonising H. lepidulum individuals, using 98% sequence similarity to define MOTUs. Filled

squares indicate significant deviations from the null model after correction of α for multiple significance

testing.

The third distance class, between 0.27-0.44 m contained 142 sample pairs, and the

AMF communities between pairs of samples in this distance class were significantly more

similar than expected at random. The second distance class spanned from 0.1–0.27 m and

contained 90 pairs of samples, and there was evidence of positive but not significant spatial

autocorrelation.

The point at which the correlogram initially intersects zero can be interpreted as

representative the physical extents of the structures under consideration. Under this

interpretation the patches in this study are estimated to have an extent of ~0.55 m. No

significant SAc was detected in the larger distance classes and it is outside the scope of this

study to describe the influence of distances of more than a metre on AMF community

similarity. The shape of the correlogram is indicative of the spatial structures under

consideration, but as different spatial structures can produce similar correlograms it is

important that it is considered alongside a map showing species distributions (Legendre

1993); this is shown in Figure 11. Similar AMF assemblages can be seen for many

neighbouring samples, e.g. samples 1 and 3 have 100% overlap in AMF taxa detected, and

the only instance of the taxon GlomB01 at the site. While there is evidence that neighbouring

plants to have similar AMF suites, it appears that the scale of the AMF patches depends on

Page 71: Characterisation of arbuscular mycorrhizal fungal

60

the particular taxon under consideration. The two abundant, widespread taxa GlomA02 and

GlomA04 occur in almost all samples, and probably extend beyond the boundaries of this

study. On this basis it could be predicted that these taxa form patches of > 1 m. However,

neither sample 1 nor 3 possess the most common MOTU at the site, indicating that the

distributions of locally abundant species are spatially restricted. Acau02 appears to follow a

less patchy distribution than other MOTUs, having numerous incidences within the site, but

the autocorrelation appears to be anisometric, following a roughly north-south distribution,

rather than an isometric patch. The distribution of the least common species appears to be

patchy at very small scales, with single (i.e., GlomA01, Acau01) or few (GlomB01)

neighbouring incidences.

Stepwise removal of individual AMF MOTUs showed that no single MOTU was

responsible for the observed spatial signal. However, the removal of the second most

abundant MOTU, GlomA04, improved the spatial structure in the AMF community. This

phenomenon may be attributable to the absence of GlomA04 in two samples, 10 and 14,

despite their having several neighbours within close proximity which possess the phylotype.

It is likely that the widespread presence of Acaulopora MOTUs are largely responsible for

the detected signal of phylogenetic overdispersion in the AMF communities.

Page 72: Characterisation of arbuscular mycorrhizal fungal

61

Figure 11. Spatial map of sample locations (numbered) with coloured representations of the AMF MOTU

detected at each sample location.

Spatial structure of H. lepidulum individuals

Significant and positive evidence of clustering was found in the distribution of 379 H.

lepidulum individuals recorded within the 10 × 2 m transect. The observed distribution was

tested against the distribution expected under complete random spatial organisation using

Ripley’s K function. The K function for the null model (Poisson distribution) was calculated

for 99 model runs to determine an upper and lower envelope of K across all distances (r).

Deviations that fall outside these envelopes at any point in the distribution are considered

signficant. The K values of H. lepidulum within the study site were consistently and

significantly larger than expected, indicating the presence of spatial clustering of H.

Page 73: Characterisation of arbuscular mycorrhizal fungal

62

lepidulum at these scales relative to a null model of complete spatial randomness (Figure 12).

Individuals within the transect were 10 cm away from their nearest neighbour on average.

The maximum distance from an individual to its nearest neighbour was 45 cm.

Figure 12. Ripley’s K-function showing the estimated number of additional H. lepidulum (K) within

distance r of a typical random point. The red curve and bounding shaded area shows the distribution

expected under complete spatial randomness with the upper and lower estimates from 100 model runs.

This distribution shows significant clustering for all values of r.

AMF status of conspecifics

Most of the plant species growing within the plant community were found to be colonised by

AMF with the exception of three species (Carex breviculmis, Rumex acetosella, and

Scleranthus uniflorus) that belong to predominantly “non-mycorrhizal” families (Harley &

Harley 1990). The roots of S. uniflorus possessed extensive fine root hairs up to 400 µm in

length, and of a similar diameter to the intraradical hyphae detected in mycorrhizal species

(Fig 13. e). The determination of mycorrhizal status from some species present at the site was

confounded by root pigmentation, and no AMF data were collected for these species (namely

Geranium sessifolium and Leucpogon fraseri). The relative abundance of fungal structures

varied by plant species. The plant species with the highest proportions of hyphal colonisation

(61%) was Pimelea oreophila, but this was represented by only a single individual with 23

observable root sections. The genus Pimelea (Thymeleaceae) is known to form mycorrhizal

Page 74: Characterisation of arbuscular mycorrhizal fungal

63

associations (Brundrett & Abbott 1991), and while no arbuscules were found, several vesicles

with the “irregular” characteristics of Acaulospora species were detected (Fig 11 b). Kellaria

villosa, also a member of Thymeleaceae, possessed the ellipsoid, thin walled vesicles

considered typical of Glomus mycorrhizae, as did Celmisia gracilenta (Asteraceae) (Fig 13.

f) and Wahlenbergia albomarginata (Campanulaceae) (Fig 13. c). All species at the site

belonging to Asteraceae (with the exception of Raoulia subscericea) had relatively high

levels of mycorrhizal colonisation, with fungal structures present in > 50% of root

intersections observed (Table 4).

Table 4. Proportions and types of AM fungal structures detected as quantified by the root intersection

method on stained root samples from the Locharburn study site. N/A values indicate roots which could

not be sufficiently cleared to visualise AM fungal structures.

Host Species observations no structure hyphae arbuscules vesicles

Anisotome brevistylis 43 72% 28% 0% 0%

Anthosachne solandri 35 51% 20% 14% 14%

Anthoxanthum odoratum 56 57% 13% 29% 2%

Bulbinella angustifolia 31 68% 10% 19% 3%

Carex breviculmis 45 100% 0% 0% 0%

Celmisia gracilenta 36 44% 14% 25% 17%

Festuca rubra 57 61% 9% 26% 4%

Geranium sessiliflorum N/A N/A N/A N/A N/A

Hieracium lepidulum 36 47% 31% 19% 3%

Hypochaeris radicata 38 26% 42% 18% 13%

Kellaria villosus 30 73% 23% 3% 0%

Leptinella pectinata subsp. villosa 46 78% 13% 4% 4%

Leucopogon fraseri N/A N/A N/A N/A N/A

Pilosella officinarum 33 36% 30% 27% 6%

Pimelea oreophila 23 30% 61% 0% 9%

Poa colensoi 68 96% 3% 1% 0%

Poa hesperia 128 80% 12% 5% 2%

Raoulia subsericea 33 85% 15% 0% 0%

Rumex acetosella 32 100% 0% 0% 0%

Scleranthus uniflorus 44 100% 0% 0% 0%

Trifolium dubium 30 57% 20% 20% 3%

Trifolium repens 33 79% 15% 3% 3%

Wahlenbergia albomarginata 30 77% 20% 0% 3%

Page 75: Characterisation of arbuscular mycorrhizal fungal

64

Figure 13. Micrographs of plant roots extracted from the study site, which have been cleared and stained

for endogenous AMF structures. Kellaria villosa (a), Pimelia oreophila (b), Wahlenbergia albomarginata

(c), Hieracium lepidulum (d), Scleranthus uniflorus (e), Celmisia gracilienta (f). Scale bars = 100 µm

Page 76: Characterisation of arbuscular mycorrhizal fungal

65

Effect of sampling effort

In order to determine the optimal level of clone screening, clones within samples were

randomly resampled at different levels of sampling effort, and pooled to reveal how this

influenced the total MOTU diversity measured at the site (Fig. 14). The number of taxa

detected at each sampling intensity were fitted to an Arrhenius model which was extrapolated

to show the estimated diversity up to 100 clones per sample. Depending upon the diversity

estimator used, this analysis shows that 66–88% of the described site diversity is likely to be

detected at a sampling intensity of 20 clones per sample — equivalent to 45% of the

screening effort used in this study. The predicted taxon accumulation curve shows that a

much greater sampling effort would be required to ensure representatives of all taxa were

detected using these field sampling and DNA extraction methods.

Figure 14. The effect of clone screening effort per sample on the total site MOTU diversity detected. The

dashed line indicates the extrapolated diversity at sampling sizes >45 clones modelled using the Arrhenius

species-area relationship.

Page 77: Characterisation of arbuscular mycorrhizal fungal

66

Discussion

Key findings

This study has identified nine phylogenetically supported AMF taxa associated with

Hieracium lepidulum in an invaded plant community and shown that the composition,

richness and evenness of AMF suites varies between individuals within a population. Only

one of the detected AMF MOTUs was closely related to a described taxon, Acaulospora

longula. Many of the detected sequences form monophyletic groups with sequences

previously detected in New Zealand, indicating the existence of potentially endemic, but

geographically widespread AMF taxa. Furthermore, this study has demonstrated that AMF

communities associated with H. lepidulum plants are positively autocorrelated at scales below

50 cm.

AMF community structure

When the AMF are considered using clone abundance data, the phylotype GlomA02

is clearly dominant at the site comprising more than 60% of clones screened, indicating a

fungus or group of closely related fungi which are particularly well adapted to both the soil at

the site and colonisation of H. lepidulum. Overdominance, where one or two species account

for more than 40% of the total abundance, generally characterises impoverished or unstable

communities, and is unusual in communities of larger organisms with a lognormal

distribution (Whitaker 1975; Dumbrell et al. 2010). However in a meta-analysis 32 studies

Dumbrell et al. (2010) found overdominance to be a general trait of AMF communities in

plant roots, with the most abundant taxon occupying, on average, 40% of the total abundance.

The meta-analysis also found that the most abundant taxon tended to belong within Glomus

group A, comprising Rhizphagus, Funneliformis and Glomus — a pattern which was also

found in this study. Relative to other members of the Glomeromycota, members of

Glomeraceae generally produce the majority of fungal biomass within the plant root, whereas

Acaulosporaceae has more equal distribution of fungal biomass inside and outside of the host

(Hart & Reader 2002; Powell et al. 2009). It has been proposed that the extensive

colonisation patterns of the Glomeraceae may reflect a functional role in the protection of

roots from infection by soil pathogens (Newsham et al. 1995). This may partially explain the

tendency for members of Glomeraceae to dominate molecular studies based on root samples

(Santos-Gonzalez et al. 2007).

Page 78: Characterisation of arbuscular mycorrhizal fungal

67

The number of AMF taxa detected in the roots of H. lepidulum individuals varied,

ranging from two to seven taxa, with no individual possessing all taxa detected at the site, and

no relationship between host biomass and AMF diversity. This was surprising, considering

that increased root biomass should provide more root surface areas for AMF colonisation.

Phylogenetic analysis revealed that taxa sharing host root systems were significantly less

related than expected. From a niche-based assembly perspective this phylogenetic

overdispersion may represent competitive interactions between AMF taxa, or environmental

filtering of convergent traits (Webb et al. 2002). If phylogenetic overdispersion is considered

in the context of the dominance distribution of the community, it may be that rare AMF are a

functionally important component, rather than a random assemblage. Evidence of

phylogenetic trait conservatism has been found in other AMF studies (Maherali &

Klironomos 2007), and while this study seems to support those findings, further

investigations of AMF in natural communities needs to be conducted to determine whether

phylogenetic overdispersion is a general characteristic of AMF communities.

Phylogenetic analysis

A pattern of monophyly in groups of NZ sequences may represent indicate endemic

taxa, with four of the MOTUs (i.e., GlomA03, GlomA04, GlomA05 and GlomA06) found in

this study being most closely related to sequences of undescribed AMF isolated from New

Zealand bryophytes in podocarp/broadleaf forests (Russell & Simon 2005; Bidartondo et al.

2011). The majority of studies which have investigated molecular diversity of AMF have

found unique taxa and around 50% of taxa detected are found only at one site, even when

those taxa are locally abundant (Öpik et al. 2006). It is unclear whether this statistic is

indicative of high diversity globally, or is a reflection of how few molecular studies have

been carried out thus far (Helgason et al. 2002; Fitter 2005). The “exceptionally” low

diversity of AMF that were found to associate with Marchantia was interpreted as indicating

highly specific symbiotic relationships (Russell & Simon 2005). The low taxonomic

resolution of the AmL1-AmL2 primers may produce monophyletic groups that are not

necessarily indicative of species-level relationships; however the presence of several shared

ribotypes between an exotic asterad in a dry grassland habitat and a thalloid liverwort in a

podocarp/broadleaf forest indicates the presence of geographically widespread generalist

AMF taxa. The ecological implications of such widespread AMF taxa are considerable,

potentially indicating that the podocarp/broadleaf systems are also vulnerable to H. lepidulum

Page 79: Characterisation of arbuscular mycorrhizal fungal

68

invasion and that there may be some mechanism for AMF gene flow between these habitats.

Within the Locharburn study site, further investigation of AMF molecular diversity at other

loci may reveal that some of the MOTUs defined in this study represent several AMF taxa.

This may help to explain the extremely high relative abundance of some MOTUs.

In an experimental study of AMF communities on the palm Trachycarpus fortunei,

introductions of the palm to several sites across Europe resulted in colonisation by a

relatively conserved suite of AMF (Moora et al. 2011). The authors concluded that the

detected AMF taxa represented generalists with global distributions whose life strategy was

particularly amenable to T. fortunei, which is also a generalist species. A broader hypothesis

from the latter authors’ findings is that widely distributed generalist AMF taxa are most likely

to colonise invading plant species, particularly if they also generalists (Pringle et al. 2009;

Moora et al. 2011). Further investigations of AMF associated with H. lepidulum and other

species may reveal whether this hypothesis is supported in the context of plant invasions in

New Zealand.

Evidence against the proposed endemic AMF taxa includes the position of foreign

sequences within the NZ clades, such as the North American environmental sequence within

GlomA02 (Figure 6). A similar situation was found when a Canadian AMF culture sequence

emerged within a phylogeny of Rhizopagus intraradices isolated from a Swiss agricultural

system (Koch et al. 2004). The workers concluded that it represented either gene flow

between the land masses, or that small scale diversity may account for diversity on a much

larger scale. However, more data are required before definitive conclusions are able to be

drawn.

Comparison of diversity measures

Under the current taxonomic delineations, the nine taxa detected in this study represent at

least four distinct genera in two families. However several distinct RFLP-types, which are

likely to represent rare taxa, remain unknown as they could not be sequenced. Furthermore,

there were several RFLP types which could not be reproduced using in silico endonuclease

digestions of the sequence data, potentially indicating preferential amplification or cloning

bias. The probability of there being unsampled taxa within the study site is also supported by

the Chao diversity estimate, based on the proportion of rare taxa detected in the site, which

predicted at least 2 undetected taxa. The MOTU rank abundance data based on clone

frequencies indicates that the AMF community is heavily dominated a few extremely

Page 80: Characterisation of arbuscular mycorrhizal fungal

69

common species, and it is likely that any further species remaining within the community are

very rare and colonise only single plant hosts.

Spatial analysis

All of the attempts to characterise the spatial structure of AMF communities to date

have used systematic sampling of the soil environment, with variable results. This is the first

study that has attempted to characterise the sub-metre spatial structure of a suite of AMF

associated with a particular plant species. This study has shown that the AMF suites of H.

lepidulum individuals < 0.5 m distant are significantly more similar than would be expected if

AMF were randomly assorted among plants. It is likely that the similarity of AMF suites is at

least partially explained by the extents and interactions of the roots systems of plant

individuals. The Mantel correlogram indicates that plants up to 0.2 m apart have significantly

more similar AMF suites than expect at random, consistent with the hypothesis of hyphal

dispersion as the dominant mechanism for colonisation of plants, however it would be

expected that individuals < 0.15 m apart would be similar as they represent overlapping

sample extents. The detection of similarity in these overlapping samples could be considered

indirect support of the accuracy of the sampling methods used in this study, but not explicit

evidence of AMF spatial structure per se. A correlogram of positive spatial autocorrelation is

expected to show a peak of significant autocorrelation at the smallest distance classes, with

the size of the correlation coefficient declining as the distance between pairs of host

individuals increases. It may be that the variation in the magnitude and significance of the

detected spatial structure in the first three distance classes is the result of taxon-specific

variation in patch size (Figure 10). The reduced signal of spatial structure in the Mantel

statistic when clone abundances were used may be due to the variation in amplification

success between samples. It is likely that the relative abundances of clones from each sample

are representative of taxon relative abundance (Suzuki et al. 1998; Dumbrell et al. 2011), but

variation in the absolute abundances of the initial AMF biomass within each root system

requires that the clone abundances be scaled to reflect this.

It can be seen from the map of AMF spatial distributions (Fig. 11), that the AMF

community at the site was dominated by common taxa (predominantly within Glomus group

A) represented in the majority of host plants. It has been shown that AMF species are able to

form anastomoses between individuals possessing the same genotype (Giovannetti et al.

2003). Although there is no direct evidence to support the hypothesis, the detected spatial

Page 81: Characterisation of arbuscular mycorrhizal fungal

70

distribution of the phylotypes, particularly the most dominant phylotypes, is consistent with

what would be expected from common mycorrhizal networks connecting the host root

systems belowground. An alternate and less parsimonious explanation would be that

phylotypes form discrete patches around each host plant, and independently develop AMF

suites from the local species pool. Conversely the rare taxa show stronger evidence of patchy

spatial restriction, having only single instances (i.e. GlomA01 and Acau01) or two nearby

instances (i.e., GlomB01, samples 1 and 3, Fig. 10). These localised fungi, which appear less

able to develop and maintain hyphal networks with H. lepidulum, may represent less

beneficial symbionts that the host plant has discriminated against (Bever et al. 2009; Powell

et al. 2009).

It has been demonstrated that plant hosts are able to discriminate between AMF and

preferentially allocate carbon to the taxa that provide the greatest reciprocal P increase, and

vice versa (Bever et al. 2009; Kiers et al. 2011). Furthermore, it has shown that H. lepidulum

expends a relatively high proportion of energy on investment in belowground structures

under nutrient poor conditions (Radford et al. 2007). This physiological adaptation may give

H. lepidulum a relatively large influence on the composition of the AMF community. The

facilitation may not contingent upon the identity of the AMF taxa per se, rather that it is a

physiological adaptation of H. lepidulum that allows it to associate with local AMF which are

particularly suited to the local environment, and maximise the benefits from these symbiosis,

however this hypothesis would require more work, and is outside the scope of this thesis. In

future studies, temporal and spatial replication could show how the identity of the dominant

taxa fluctuates over space and time, which will help elucidate the roles and scales of positive

feedback within AMF communities. Further investigations are also necessary to determine

other factors that influence the distribution of H. lepidulum and AMF within the study site.

The observed spatial structures are probably partially explained by spatially structured

environmental variables, such as soil physico-chemical properties and the distributions of

mycorrhizal plant species, and by pure spatial effects (Legendre et al. 2005).

Page 82: Characterisation of arbuscular mycorrhizal fungal

71

Conclusion

This study represents the first time that molecular methods have been used to measure AMF

community composition of Hieracium lepidulum, and the first time that molecular methods

have been used to characterise AMF diversity and spatial structure within a NZ grassland

community. H. lepidulum has been shown to associate with at least nine distinct fungal taxa

from four genera, several of which are highly similar to AMF sequences previously amplified

from New Zealand vegetation. The organisation of the fungal taxa within the site has been

shown to be spatially non-random, with nearby plants tending to posses more similar suites

than expected. At the site, the AMF community is dominated by a single taxon, the

abundance and phylogenetic placement of which is consistent with findings from previous

molecular surveys from diverse ecosystems. Further investigations of the AMF associated

with H. lepidulum in its invaded range which include both spatial and temporal replication

will help to determine the distributions and seasonal dynamics of these functionally

significant soil fungi.

Page 83: Characterisation of arbuscular mycorrhizal fungal

72

Chapter 4: Conclusions and experimental recommendations

The preceding chapters of this thesis have explored aspects of the interactions between the

exotic plant Hieracium lepidulum and an ecologically important group of symbiotic fungi, the

Glomeromycota, in subalpine plant communities in Central Otago, New Zealand. In the first

chapter, the state of knowledge concerning these organisms, their interactions, and their

distributions in the context of H. lepidulum invasion in New Zealand was reviewed. In

Chapter Two, the potential of single spore trap culture as a method for isolating individual

AMF strains for experimental manipulation, and the relative sensitivity of various AMF-

specific primers for detecting their presence was investigated. In Chapter Three, the

optimised PCR protocols designed in Chapter Two were used to characterise AMF ribotypes

from H. lepidulum in a spatially explicit manner from the Locharburn field site. In this final

chapter, the literature will be reviewed in the context of the findings from the two

experimental chapters of this thesis, and recommendations will be made for further research

into the relationship between AMF and H. lepidulum in New Zealand. As with many

ecological studies, the results from this research prompt more questions than they provide

definitive answers. In the absence of manipulative experiments we are still no closer to

understanding the functional significance of the AMF symbiosis in the invasion of New

Zealand indigenous vegetation by H. lepidulum. However, this thesis has investigated several

aspects of the H. lepidulum–AMF symbiosis that may prove useful for informing future

experimental work.

Single spore culture screening

The success of other workers (i.e., Koch et al. 2004; Avio et al. 2006) clearly demonstrates

that the production of individual AMF strains through single spore trap culture is possible.

However, this study has reinforced some of the practical limitations of AMF as experimental

units, including the requirement of large numbers of culture attempts and long times frames.

The phylogenetic analysis of sequences from H. lepidulum roots in Chapter Three shows that

over 95% of the ribotypes detected form clades that are distinct in this study from any

cultured species, and form monophyletic groups with “environmental samples” from around

the world. Several workers have speculated that culture collections may consist

predominantly of “weedy” species, which are able to be easily cultured (Helgason et al. 2002;

Fitter 2005), and the position of the sequences from this study in the phylogeny may indicate

that these are species that are difficult to establish and maintain in trap culture. Conversely,

Page 84: Characterisation of arbuscular mycorrhizal fungal

73

the finding of relationships between the sequences growing on H. lepidulum and Marchantia

foliacea may indicated that the AMF species concerned may be generalists, and as such they

should represent species that are relatively easy to cultivate.

Due to the unknown genetic organisation of the AMF species detected in this study

and of the Glomermycota in general, it may be that some or all of the MOTUs identified in

Chapter Three represent more than one AMF species. Koch et al. (2004) found that the rDNA

sequence variation may mask significant, and ecologically relevant genetic variation at other

loci. Future studies investigating the genetic diversity of cultured AMF at different loci will

help to clarify the taxonomic distinctions between AMF species.

This study has identified a method for screening large quantities of single spore

culture attempts that is faster and more practical than traditional root-staining based methods.

This study has shown that the AmL1–AmL2 primer set is sufficiently sensitive for tens of

single spore culture root samples to be pooled and simultaneously screened for the presence

of AM fungal DNA. However it has been shown that the AmL1–AmL2 primers are not as

specific to the Glomeromycota as previous studies have claimed. This has implications for

their efficacy in the situations where AMF status is uncertain, or for studies that do not

employ sequence-based characterisation methods.

AMF spatial structure

The intensive sampling method used in the environmental study has revealed several aspects

of the AMF symbiosis that have hitherto received little attention. Prior to this study the

diversity of AMF assemblages associated with H. lepidulum in New Zealand was unclear. At

least 9 phylogenetically distinct ribotypes comprising at least 4 genera form symbioses with

H. lepidulum at Locharburn station. Furthermore, this study has found evidence for a

lognormal dominance/diversity distribution of ribotypes with one dominant taxa comprising

more than 65% of taxon abundances, a distribution similar to that found in several other

studies (Dumbrell et al. 2010). The ecological significance of this is unclear, however the

similarities found among the dominance/diversity distributions of several independent studies

of AMF communities may indicate the action of a common process structuring AMF

communities (Dumbrell et al. 2010), at least within plant roots.

Future environmental studies and surveys of AMF from natural systems need to

consider spatial structure within AMF communities. If, as postulated in Chapter 3, the spatial

structure of plant hosts and their AM symbionts are related, future studies may show that the

Page 85: Characterisation of arbuscular mycorrhizal fungal

74

scales of spatial structure in AMF communities are correlated with the spatial structure in the

plant hosts, and may vary correspondingly in extent, particularly for plants with high

mycorrhizal dependence.

Phylogenetic analysis

One of the most intriguing discoveries of this study is the phylogenetic relationship between

AMF ribotypes from diverse habitats and hosts within New Zealand. There is evidence of

“New Zealand-only” clades within the AMF phylogeny constructed in Chapter 3, comprising

AMF ribotypes that are common to subalpine grassland and podocarp forests. Geographical

isolation and environmental filtering are two large scale processes that strongly influence the

distribution of plant species, and it would be expected that the associated plant symbionts

would show similar distributions in response to similar environmental drivers. However, not

only has this study shown evidence that AMF strains occur in very different habitat types,

their isolation from widely divergent plant groups indicates that the distributions of the fungi

may also be independent of host identity.

One implication of the detected ribotype distributions may be that the sequences

represent generalist AMF species with widespread distributions across several habitat types

in New Zealand. This hypothesis supported by the evidence of multiple ribotypes common to

both habitats. Alternatively, the low resolution of the SSU rDNA (Gamper et al. 2009) may

mean that the phylogenetic relationships observed between New Zealand sequences may

reflect the distribution of some broader taxonomic grouping, perhaps at genus level.

However, when it is considered that there have been only three published SSU rDNA studies

of AMF in New Zealand, it is surprising that there is such a large overlap of ribotypes found

among the studies. This may indicate that AMF ribotype diversity is nationally limited, which

may in turn reflect a limited AMF species pool. Whereas Russell & Simon (2005) concluded

that their finding of low AMF diversity from several isolates of Marchantia foliacea was

indicative of a group of specialised AMF, this low diversity may in fact be a general

characteristic of New Zealand AMF. Alternatively, as both H. lepidulum and M. foliacea

could be considered generalist species, and a study has been shown that generalist plants

tend to associate with generalist suites of AMF (Moora et al. 2011), it may be that the

similarities found between the AMF colonising the hosts are a coincidence of the host life

strategy. In either case, this study has found evidence that AMF ribotypes isolated from an

invasive species in a New Zealand subalpine grassland are related to ribotypes from a New

Page 86: Characterisation of arbuscular mycorrhizal fungal

75

Zealand podocarp forest. This finding provides evidence that the facilitation of H. lepidulum

invasion is not via a coinvasive exotic AMF species. Conversely, the evidence indicates that

the AMF involved are likely to be endemic to New Zealand.

Testing mechanisms of H. lepidulum invasion

Despite several decades of research, the factors influencing the success of H. lepidulum are

not entirely clear. It is possible that the proximate mechanisms facilitating H. lepidulum

invasion are not the same throughout its invaded range in New Zealand. For example,

kilometre-scale autocorrelation in Hieracium cover was found in Otago but not in Canterbury

(Duncan et al. 1997), a finding that was postulated to reflect the process of invasion being at

an earlier stage in Otago. The relative importance of grazing on H. lepidulum fitness remains

contentious, with contradictory evidence from several studies (e.g. Rose et al. 1995; Norton

et al. 2006). In post-burning situations, grazing can help to suppress the spread of H.

lepidulum, however it will also promote the establishment of other exotic and native

unpalatable species (Mark et al. 2011). The relative herbivory resistance of H. lepidulum

vegetative structures has been identified as a potentially important factor in H. lepidulum

invasion (Radford et al. 2009), as has its ability to establish and persist in low-light situations

(i.e., Nothofagus forest, Chionochloa tussocks). Within Nothofagus forest, H. lepidulum

density is greatest at microsites of high plant diversity (Wiser et al. 1998; Meffin et al. 2010)

and high AMF inoculum potential (Spence et al. 2011). This apparent positive diversity-

invasibility relationship could be attributed to facilitation of exotic invasion by resident

biodiversity, or more specifically, the AMF networks supported by the resident plant

community (Spence et al. 2011).

Hypothetical models of H. lepidulum invasion have been previously published, and

the following paragraphs will elaborate on some of these models in terms of AMF

facilitation. Duncan et al. (1997) ruled out single-factor explanations for H. lepidulum

invasion, and broadly stated that the probability of invasion is a function of two variables: (1)

the suitability of the site for hawkweed establishment (susceptibility of environment) and (2)

the size of the hawkweed propagules rain (strength of invasion). Here we will focus only on

aspects of the susceptibility of the environment to H. lepidulum invasion by considering AMF

facilitation as a variable which may influence community composition via interactive effects

with other environmental parameters. One of the difficulties in testing hypotheses concerning

the roles of AMF in invasion is the myriad functional roles that have been attributed to them,

Page 87: Characterisation of arbuscular mycorrhizal fungal

76

from improved phosphate nutrition, to pathogen resistance, to a kind of plant socialism —

where resources are redirected from dominant plant species to subdominant species (van der

Heijden & Horton 2009). When these factors are added to the already numerous biotic and

abiotic variables, experimental designs quickly become large and unwieldy.

It is likely that a more comprehensive understanding of AMF diversity and

distribution in New Zealand from field sites will develop from molecular studies. Sequencing

costs are dropping rapidly, with current prices at US$0.12 Mb-1, down from US$100 Mb-1 in

2008 and US$1000 Mb-1 in 2004 (NHGRI: genome.gov/sequencingcosts). This continuing

trend will facilitate the expansion of AMF research into more ecosystems and more plant

species, and enable a range of questions similar to those raised in this study to be addressed.

AMF positive diversity-invasibility model

In a longitudinal study over six years, the cover of H. lepidulum was “significantly positively

associated with the richness of basal herb, creeping herb and grass functional groups” at the

scale of the 30 × 30 cm quadrats, but uncorrelated with any of the measured abiotic soil

variables across a range of habitats (Meffin et al. 2010). A similar correlation was found by

Wiser et al. (1998), and Spence et al. (2011) were able to show a correlation between plant

diversity, AMF inoculum potential and H. lepidulum density. Spence et al. (2011)suggest that

this correlation may represent the action of a positive diversity-invasibility relationship for H.

lepidulum within their study site. A large proportion of the AMF diversity which has

previously been detected in New Zealand was of found to be associated with H. lepidulum in

this study. More data are needed to characterise AMF assemblages of H. lepidulum across the

invaded range to determine their identity and distribution. If plant and AMF community

metrics are simultaneously measured in the sampled communities it may be possible to

unravel this hypothesised positive diversity-invasibility relationship and determine whether it

is plant diversity that promotes H. lepidulum invasion, or whether particular AMF ribotypes

are more important predictors of H. lepidulum success.

AMF facilitation/herbivory resistance interaction model

It may be that a combination of AMF facilitation and herbivory underlies the success of

H. lepidulum as an invader: invasion may be facilitated by utilising the existing AMF

networks within a plant community, but it is the relative resistance of H. lepidulum to

herbivory that promotes its maintenance and eventual success within the community. It is

Page 88: Characterisation of arbuscular mycorrhizal fungal

77

believed that the rosette growth form of H. lepidulum makes it less attractive to mammalian

herbivores, particularly stock (Scott 1993). As herbivory pressure is preferentially directed at

other, more easily grazed species within the community, H. lepidulum is able to command a

higher proportion of resources from the AMF network. It has been demonstrated in

glasshouse experiments that the AMF symbionts of an individual of one species can form a

common mycorrhizal network with an individual of another. Isotope studies have shown that

simulated grazing of one individual results in a significant transfer of nutrients to the other,

resulting in an increase of biomass relative to control plants (Jakobsen 2004). It may be that

similar relationships exist within plant communities containing H. lepidulum — common

mycorrhizal networks redirect resources from the belowground structures of grazed

individuals to the ungrazed individuals. Such a feedback mechanism would ensure that

photosynthates remain available to the AMF network under grazing pressure, and would

manifest aboveground as a proliferation of unpalatable species such as H. lepidulum. This

hypothesis could be tested in a field situation by the simultaneous application of fungicide

and grazing to experimental plots containing H. lepidulum. The AMF facilitation model

would predict that in the absence of AMF (fungicide treatment), palatable species will persist

in the community, resulting in overall lower aboveground biomass relative to grazed plots

which contain AMF.

Limiting factor release

Mutalisms tend to be more important biotic processes shaping plant communities within low

nutrient/productivity environments (Bertness & Callaway 1994) and these processes tend to

facilitate the maintenance of diversity and evenness. Conversely, eutrophic conditions tend to

promote competitive interactions, leading to less diverse communities dominated by one or a

few species (Vitousek et al. 1997). It is recognised that in high nutrient soils plants are able to

obtain sufficient nutrients without needing to invest in the AMF network (Collins & Foster

2009). In grassland ecosystems, it has been demonstrated that the dominant plant species

following eutrophication tend to be those that show small growth responses to AMF

colonisation, while the subdominant species remain mycorrhizal (Johnson et al. 2008).

Molecular studies have demonstrated that AMF community composition can also change in

response to eutrophication, potentially driving community shifts via feedback (Egerton-

Warburton et al. 2007). Applications of phosphate fertiliser (50–100 kg ha-1) and oversowing

reduces the proportion of H. lepidulum within plant communities (Scott 1993). The addition

Page 89: Characterisation of arbuscular mycorrhizal fungal

78

of fertilisers may liberate plant species from the AM symbioses, reducing the importance of

the AM network as a source of nutrients for the plant community. In the resulting

competition-dominated system, H. lepidulum would no longer be able to successfully

compete with conspecifics (Radford et al. 2010), and would become subdominant.

Furthermore, in these competition-based systems, the increased prevalence of plant species

with low mycorrhizal dependency, coupled with an overall reduction in plant community

dependence on AMF, may result in the suppression of AMF. As the subdominants are known

to maintain their AM symbioses under eutrophic conditions, upon cessation of nutrient

application there may be a proliferation of those AMF taxa which the subdominant species

have maintained throughout the interim. H. lepidulum is known to possess phenotypic

adaptations that facilitate its persistence in high country plant communities, including

herbivory resistance and a relative tolerance to adverse nutrient conditions (Scott 1993;

Radford et al. 2006). It has also been shown that in low nutrient situations H. lepidulum

diverts a higher proportion of energy towards investment in belowground structures (Radford

et al. 2007). As nutrient levels decline, this increased investment in belowground structures

may represent greater resource availability for AMF in terms of host-derived photosynthate.

As a result the influence of H. lepidulum on the composition and organisation of the AMF

community may increase, potentially driving a switch in the plant community via positive

feedback (Wilson & Agnew 1992). This model of invasion could be tested by applying

different fertiliser treatments to experimental field plots containing H. lepidulum. Using 18S

rDNA based characterisation of root and soil DNA, the AMF community composition could

be compared before, during and after the treatments in both H. lepidulum roots, and in the

soil. This model would predict a reduction in H. lepidulum density and soil AMF diversity

resulting from fertiliser application, and a concomitant increase in H. lepidulum and AMF

taxa associated with H. lepidulum as the soil nutrient levels decline after the treatment has

ended. This model would also predict that the application of fungicide to a eutrophic

grassland would result in the loss of subdominant species such as H. lepidulum. However the

loss of AMF from grassland systems is likely to have significant long term negative impacts

on soil stability and plant productivity.

Conclusion

The diversity and abundance of interactions between organisms is a mechanism thought to

explain the resilience and robustness of ecosystems and ecosystem processes. A full

Page 90: Characterisation of arbuscular mycorrhizal fungal

79

appreciation of the complexity underlying ecosystems may never be possible, yet

technological advances are allowing increasingly detailed analyses at increasingly smaller

scales. The research undertaken in this thesis has demonstrated the scales of spatial structure

in arbuscular mycorrhizae colonising Hieracium lepidulum, and found spatial patterns of

fungi and host that are consistent with facilitated invasion. It is clear that AMF play a

fundamental role in the maintenance of terrestrial diversity, and that the 400 million year

coevolution of AMF and terrestrial plants should inform concepts about vegetation dynamics.

More research is needed to determine whether the identities and spatial structure of AMF are

consistent across the invaded range of H. lepidulum, and to investigate the mechanisms

underlying the observed spatial structures.

Page 91: Characterisation of arbuscular mycorrhizal fungal

80

References

Akaike H 1974. A new look at the statistical model identification. IEEE Transactions on Automatic Control 19: 716-723.

Allen TFH, Starr TB 1982. Hierarchy: Perspectives for ecological complexity. Chicago,

University of Chicago Press. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ 1990. Basic Local Alignment Search

Tool. Journal of Molecular Biology 215: 403-410. Auge RM 2001. Water relations, drought and vesicular-arbuscular mycorrhizal symbiosis.

Mycorrhiza 11: 3-42. Auge RM 2004. Arbuscular mycorrhizae and soil/plant water relations. Canadian Journal of

Soil Science 84: 373-381. Avio L, Pellegrino E, Bonari E, Giovannetti M 2006. Functional diversity of arbuscular

mycorrhizal fungal isolates in relation to extraradical mycelial networks. New Phytologist 172: 347-357.

Bachmann K 1998. Species as units of diversity: an outdated concept. Theory Bioscience

117: 213-230. Baddeley A, Turner R 2005. Spatstat: an R package for analyzing spatial point patterns.

Journal of Statistical Software 12: 1-42. Bainard LD, Klironomos JN, Hart MM 2010. Differential effect of sample preservation

methods on plant and arbuscular mycorrhizal fungal DNA. Journal of Microbiological Methods 82: 124-130.

Baylis GTS 1978. Endomycorrhizas in our native flora. Journal of the Royal New Zealand

Institute of Horticulture 6: 63-69. Bennett AE, Alers-Garcia J, Bever JD 2006. Three-way interactions among mutualistic

mycorrhizal fungi, plants, and plant enemies: Hypotheses and synthesis. American Naturalist 167: 141-152.

Berta G, Fusconi A, Trotta A 1993. Va Mycorrhizal Infection and the Morphology and

Function of Root Systems. Environmental and Experimental Botany 33: 159-173. Bertness MD, Callaway R 1994. Positive interactions in communities. Trends in Ecology &

Evolution 9: 191-193. Bethlenfalvay GJ, Brown MS, Pacovsky RS 1982. Parasitic and Mutualistic Associations

between a Mycorrhizal Fungus and Soybean - Development of the Host Plant. Phytopathology 72: 889-893.

Page 92: Characterisation of arbuscular mycorrhizal fungal

81

Bever JD, Westover KM, Antonovics J 1997. Incorporating the soil community into plant population dynamics: the utility of the feedback approach. Journal of Ecology 85: 561-573.

Bever JD, Morton JB, Antonovics J, Schultz PA 1996. Host-dependent sporulation and

species diversity of arbuscular mycorrhizal fungi in a mown grassland. Journal of Ecology 84: 71-82.

Bever JD, Schultz PA, Pringle A, Morton JB 2001. Arbuscular mycorrhizal fungi: More

diverse than meets the eye, and the ecological tale of why. BioScience 51: 923-931. Bever JD, Richardson SC, Lawrence BM, Holmes J, Watson M 2009. Preferential allocation

to beneficial symbiont with spatial structure maintains mycorrhizal mutualism. Ecology Letters 12: 13-21.

Bidartondo MI, Read DJ, Trappe JM, Merckx V, Ligrone R, Duckett JG 2011. The dawn of

symbiosis between plants and fungi. Biology Letters 7: 574-577. Blair AC, Wolfe LM 2004. The evolution of an invasive plant: an experimental study with

Silene latifolia. Ecology 85: 3035-42. Blossey B, Notzold R 1995. Evolution of Increased Competitive Ability in Invasive

Nonindigenous Plants - a Hypothesis. Journal of Ecology 83: 887-889. Bonfante P, Genre A 2010. Mechanisms underlying beneficial plant-fungus interactions in

mycorrhizal symbiosis. Nature Communications 1: 1-11. Brundrett M, Abbott L 1991. Roots of jarrah forest plants .I. mycorrhizal associations of

shrubs and herbaceous plants. Australian Journal of Botany 39: 445-457. Brundrett MC 1991. Mycorrhizas in natural ecosystems. In: Norris JR, Read DJ, Varma AK

ed. Methods in Microbiology. London, Academic Press. Pp. 1-21. Buchanan PK 2012. Phyla Zygomycota and Glomeromycota. In: Gorden DP ed. New

Zealand Inventory of Biodiversity. Christchurch, New Zealand, Canterbury University Press. Pp. 524-527.

Buchanan PK, Beever RE, Glare TR, Johnston PR, McKenzie EHC, Paulus BC, Pennycook

SR, Ridley GS, Smith JMB 2012. Kingdom Fungi. In: Gorden DP ed. New Zealand Inventory of Biodiversity. Christchurch, Canterbury University Press. Pp. 499-515.

Callaway RM, Thelen GC, Barth S, Ramsey PW, Gannon JE 2004. Soil fungi alter

interactions between the invader Centaurea maculosa and North American natives. Ecology 85: 1062-1071.

Carvalho LM, Correia PM, Ryel RJ, Martins-Loucao MA 2003. Spatial variability of

arbuscular mycorrhizal fungal spores in two natural plant communities. Plant and Soil 251: 227-236.

Page 93: Characterisation of arbuscular mycorrhizal fungal

82

Chao A 1987. Estimating the population-size for capture recapture data with unequal catchability. Biometrics 43: 783-791.

Chao A 2004. Species estimation and applications. Encyclopedia of Statistical Sciences.

Online, John Wiley & Sons, Inc. Chapin FS, Bloom AJ, Field CB, Waring RH 1987. How plants cope: Plant physiological

ecology. BioScience 37: 49-57. Chapman H, Pearson ML, Robson B 2003. Genetic diversity in tussock hawkweed

(Hieracium lepidulum) and use of allele frequencies for identifiying patterns of spread. DoC Science Internal Series.

Chapman H, Robson B, Pearson ML 2004. Population genetic structure of a colonising,

triploid weed, Hieracium lepidulum. Heredity 92: 182-188. Chaudhary VB, Lau MK, Johnson NC 2008. Macroecology of microbes - biogeography of

the Glomeromycota. In: Varma A ed. Mycorrhiza. Berlin, Springer. Pp. 529-562. Christie P, Li XL, Chen BD 2004. Arbuscular mycorrhiza can depress translocation of zinc to

shoots of host plants in soils moderately polluted with zinc. Plant and Soil 261: 209-217.

Clapp JP, Fitter A, Young JPW 1999. Ribosomal small subunit sequence variation within

spores of an arbuscular mycorrhizal fungus, Scutellospora sp. Molecular Ecology 8: 915-921.

Clapp JP, Young JPW, Merryweather JW, Fitter AH 1995. Diversity of fungal symbionts in

arbuscular mycorrhizas from a natural community. New Phytologist 130: 259-265. Collins CD, Foster BL 2009. Community-level consequences of mycorrhizae depend on

phosphorus availability. Ecology 90: 2267-2576. Connor HE 1992. Hawkweeds, Hieracium Spp, in Tussock Grasslands of Canterbury, New-

Zealand, in 1960s. New Zealand Journal of Botany 30: 247-261. Cowan PE 1989. A vesicular-arbuscualr fungus in the diet of brushtail possums, Trichosurus

vulpecula. New Zealand Journal of Botany 27: 129-131. Coyne JA, Orr HA 2004. Speciation. Sunderland, MA, Sinhauer. Croll D, Sanders IR 2009. Recombination in Glomus intraradices, a supposed ancient

asexual arbuscular mycorrhizal fungus. Bmc Evolutionary Biology 9: 1-13. Croll D, Giovannetti M, Koch AM, Sbrana C, Ehinger M, Lammers PJ, Sanders IR 2009.

Nonself vegetative fusion and genetic exchange in the arbuscular mycorrhizal fungus Glomus intraradices. New Phytologist 181: 924-937.

Crush JR 1975. Occurance of endomycorrhizas in soils of the Mackenzie Basin, Canterbury,

New Zealand. New Zealand Journal of Agricultural Research 18: 361-364.

Page 94: Characterisation of arbuscular mycorrhizal fungal

83

Crush JR 1976. Endomycorrhizas and legume growth in some soils of the Mackenzie Basin,

Canterbury, New Zealand. New Zealand Journal of Agricultural Research 19: 473-476.

Davis BNK 1977. The Hieracium flora of chalk and limestone quarries in England. Watsonia

11: 345-351. Davison J, Opik M, Daniell TJ, Moora M, Zobel M 2011. Arbuscular mycorrhizal fungal

communities in plant roots are not random assemblages. FEMS Microbiology Ecology 78: 103-115.

Dickie IA, Bolstridge N, Cooper JA, Peltzer DA 2010. Co-invasion by Pinus and its

mycorrhizal fungi. New Phytologist 187: 475-484. Douhan GW, Petersen C, Bledsoe CS, Rizzo DM 2005. Contrasting root associated fungi of

three common oak-wood land plant species base on molecular identification: host specificity or non-specific amplification? Mycorrhiza 15: 365-372.

Downs TM, Radford IJ 2005. Arbuscular mycorhizal fungal colonisation on Hieracium

lepidulum roots in experimental and field soil inoculated media. New Zealand Journal of Botany 43: 843-85.

Drummond AJ, Ashton B, Cheung M, Cooper A, Kearse M, R. M, Stones-Havas S, Sturrock

S, Thierer T, Wilson A 2010. Geneious v5.1. Dumbrell AJ, Nelson M, Helgason T, Dytham C, Fitter AH 2010. Idiosyncrasy and

overdominance in the structure of natural communities of arbuscular mycorrhizal fungi: is there a role for stochastic processes? Journal of Ecology 98: 419-428.

Dumbrell AJ, Ashton PD, Aziz N, Feng G, Nelson M, Dytham C, Fitter AH, Helgason T

2011. Distinct seasonal assemblages of arbuscular mycorrhizal fungi revealed by massively parallel pyrosequencing. New Phytologist 190: 794-804.

Duncan RP, Colhoun KM, Foran BD 1997. The distribution and abundance of Hieracium

species (Hawkweeds) in the dry grasslands of Canterbury and Otago. New Zealand Journal of Ecology 21: 51-62.

Egerton-Warburton LM, Johnson NC, Allen EB 2007. Mycorrhizal community dynamics

following nitrogen fertilization: a cross-site test in five grasslands. Ecological Monographs 77: 517-577.

Eom AH, Hartnett DC, Wilson GWT 2000. Host plant species effects on arbuscular

mycorrhizal fungal communities in tallgrass prairie. Oecologia 122: 435-444. Evans RD, Rimer R, Sperry L, Belnap J 2001. Exotic plant invasion alters nitrogen dnamics

in an arid grassland. Ecological Applications 11: 1301-1310. Faith DP 1992. Conservation evaluation and phylogenetic diversity. Biological Conservation

61: 1-10.

Page 95: Characterisation of arbuscular mycorrhizal fungal

84

Felsenstein J 1974. The evolutionary advantage of recombination. Genetics 78: 737-756. Finlay RD 2008. Ecological aspects of mycorrhizal symbiosis: with special emphasis on the

functional diversity of interactions involvng the extratradical mycelium. Journal of Experimental Botany 59: 1115-1126.

Fitter AH 2005. Darkness visible: reflections on underground ecology. Journal of Ecology

93: 231-243. Foster KR, Wenseleers T 2006. A general model for the evolution of mutualisms. Journal of

Evolutionary Biology 19: 1283-1293. Gadella TWJ 1992. Notes on some triple and inter-scetional hybrids in Hieracium L.

subgenus pilosella (Hill) S.F. Grey. Proceedings of the Koninklijke Nederlandse Akademie Van Wetenschappen 95: 51-63.

Gamper HA, Walker C, Schüßler A 2009. Diversispora celata sp. nov: molecular ecology and

phylotaxonomy of an inconspicuous arbuscular mycorrhizal fungus. New Phytologist 182: 495-506.

Gemma JN, Koske RE 1988. Seasonal variation in spore abundance and dormancy of

Gigaspora gigantea and in mycorrhial innoculum potential of a dune soil. Mycologia 80: 211-216.

Gemma JN, Koske RE, Carreiro M 1989. Seasonal dynamics of selected species of V-A

mycorrhizal fungi in a sand dune. Mycological Research 92: 317-321. Gerdeman JW, Trappe JM 1974. The Endogonaceae in the Pacific Northwest. Mycological

Memoir (New York Botanical Garden) 5: 1-76. Giovannetti M, Sbrana C, Avio L, Strani P 2004. Patterns of below-ground plant

interconnections established by means of arbuscular mycorrhizal networks. New Phytologist 164: 175-181.

Giovannetti M, Fortuna P, Citernesi AS, Morini S, Nutu MP 2001. The occurence of

anastomosis formation and nuclear exchange in intact arbuscular mycorrhizal networks. New Phytologist 164: 175-181.

Giovannetti M, Sbrana C, Strani P, Agnolucci M, Rinaudo V, Avio L 2003. Genetic diversity

of isolates of Glomus mosseae from different geographic areas detected by vegetative compatibility testing and biochemical and molecular analysis. Applied and Environmental Microbiology 69: 616-624.

Goto BT, Maia LC 2006. Glomerospores: a new denomination for the spores of

Glomeromycota, a group molecularly distinct from the Zygomycota. Mycotaxon 96: 129-132.

Page 96: Characterisation of arbuscular mycorrhizal fungal

85

Govindarajulu M, Pfeffer PE, Jin H, Abubaker J, Douds DD, Allen JW, Bücking H, Lammers PJ, Shachar-Hill Y 2005. Nitrogen transfer in the arbuscular mycorrhial symbiosis. Nature 435: 819-823.

Grime JP, Mackey JML, Hillier SH, Read DJ 1987. Floristic diversity in a model system

using experimental microcosms. Nature 328: 420-422. Hall IR 1977. Species and mycorrhizal infections of New Zealand endogonaceae.

Transactions of the British Mycological Society 68: 341-356. Harley JL, Harley EL 1990. A check-list of mycorrhiza in the British flora - second adenda

and errata New Phytologist 115: 699-711. Hart MM, Reader RJ 2002. Taxonomic basis for variation in the colonization strategy of

arbuscular mycorrhizal fungi. New Phytologist 153: 335-344. He X-H, Critchley C, Bledsoe C 2003. Nitrogen transfer within and between plants through

common mycorrhizal networks (CMNs). Critical Reviews in Plant Sciences 22: 531-567.

Helgason T, Fitter A 2005. The ecology and evolution of the arbuscular mycorrhizal fungi.

Mycologist 19: 96-101. Helgason T, Fitter AH 2009. Natural selection and the evolutionary ecology of the arbuscular

mycorrhizal fungi (Phylum Glomeromycota). Journal of Experimental Botany 60: 2465-2480.

Helgason T, Fitter AH, Young JPW 1999. Molecular diversity of arbuscular mycorrhizal

fungi colonising Hyacinthoides non-scripta (bluebell) in a seminatural woodland. Molecular Ecology 8: 659-666.

Helgason T, Daniell TJ, Husband R, Fitter AH, Young JPW 1998. Ploughing up the wood-

wide web? Nature 394: 431-431. Helgason T, Merryweather JW, Denison J, Wilson P, Young JPW, Fitter AH 2002.

Selectivity and functional diversity in arbuscular mycorrhizas of co-occurring fungi and plants from a temperate deciduous woodland. Journal of Ecology 90: 371-384.

Heywood VH 1989. Patterns, extents and modes of invasions by terrestrial plants. In: Drake

JA, Mooney HA, di Castri F, Groves RH, Kruger FJ, Rejmanek M, Williamson M ed. Biological Invasions: a global perspective. Chichester, UK, John Wiley.

Hijri M, Sanders IR 2005. Low gene copy number shows that arbuscular mycorrhizal fungi

inherit genetically different nuclei. Nature 433: 160-163. Hijri M, Niculita H, Sanders IR 2007. Molecular characterization of chromosome termini of

the arbuscular mycorrhizal fungus Glomus intraradices (Glomeromycota). Fungal Genetics and Biology 44: 1380-1386.

Page 97: Characterisation of arbuscular mycorrhizal fungal

86

Hodge A, Campbell CD, Fitter A 2001. An arbsucular mycorrhizal fungus accelerates decomposition and acquires nitrogen directly from organic material. Nature 41.

Jakobsen I 2004. Hyphal fusion to plant species connections - giant mycelia and community

nutrient flow. New Phytologist 164: 4-7. Janos DP, Sahley CT, Emmons LH 1995. Rodent dispersal of vesicular-arbuscular

mycorrhizal fungi in amazonian peru. Ecology 76: 1852-1858. Jany JL, Pawlowska Teresa E 2010. Multinucleate spores contribute to evolutionary

longevity of asexual Glomeromycota. American Naturalist 175: 424-435. Javot H, Pumplin N, Harrison MJ 2007. Phosphate in the arbuscular mycorrhizal symbiosis:

transport properties and regulatory roles. Plant, Cell & Environment 30: 310-322. Johnson D, Leake JR, Ostle N, Ineson P, Read DJ 2002. In situ 13 CO2 pulse-labelling of

upland grassland demonstrates a rapid pathway of carbon flux from arbuscular mycorrhizal mycelia to the soil. New Phytologist 153: 327-334.

Johnson NC, Graham JH, Smith FA 1997. Functioning of mycorrhizal associations along the

mutualism-parasitism continuum. New Phytologist 135: 575-586. Johnson NC, Rowland DL, Corkidi L, al. e 2008. Characteristics of plant winners and losers

in grassland eutrophication - importance of allocation plasticity and mycorrhiza function. Ecology 89: 2868-2878.

Johnson PN 1977. Mycorrhizal Endogonaceae in a New Zealand Forest. New Phytologist 78:

161-170. Johnston PR 2009. Causes and consequences of changes to New Zealand's fungal biota. New

Zealand Journal of Ecology 34: 175-184. Juniper S, Abbott L 1993. Vesicular-arbuscular mycorrhizas and soil salinity. Mycorrhiza 4:

45-57. Kembel SW, Ackerly DD, Blomberg SP, Cornwell WK, Cowan PD, Helmus MR, Morlon H,

Webb CO 2011. Picante: R tools for integrating phylogenies and ecology. Kiers ET, van der Heijden MGA 2006. Mutualistic Stability in the Arbuscular Mycorrhizal

Symbiosis: Exploring Hypotheses of Evolutionary Cooperation. Ecology 87: 1627-1636.

Kiers ET, Duhamel M, Beesetty Y, Mensah JA, Franken O, Verbruggen E, Fellbaum CR,

Kowalchuk GA, Hart MM, Bago A and others 2011. Reciprocal rewards stabilize cooperation in the mycorrhizal symbiosis. Science 333: 880-882.

Kindt R, Coe R 2005. Tree diversity analysis. A manual and software for common statistical

methods for ecological and biodiverity studies. Nairobi, World Agroforestry Centre (ICRAF).

Page 98: Characterisation of arbuscular mycorrhizal fungal

87

Klironomos J, Moutoglis P, Kendrick B, Widden P 1993. A comparison of spatial heterogeneity of vesicular-arbuscular mycorrhizal fungi in two maple forest soils. Canadian Journal of Botany 71: 1472-1480.

Klironomos JN 2003. Varation in plant response to native and exotic arbuscular mycorrhizal

fungi. Ecology 84: 2292-2301. Klironomos JN, Moutoglis P 1999. Colonization of nonmycorrhizal plants by mycorrhizal

neighbours as influenced by the collembolan, Folsomia candida. Biology and Fertility of Soils 29: 277-281.

Klironomos JN, Hart MM 2002. Colonization of roots by arbuscular mycorrhizal fungi using

different sources of inoculum. Mycorrhiza 12: 181-184. Klironomos JN, Rillig MC, Allen MF 1999. Designing belowground field experiments with

the help of semi-variance and power analyses. Applied Soil Ecology 12: 227-238. Koch AM, Kuhn G, Fontanillas P, Fumagalli L, Goudet J, Sanders IR, Field CB 2004. High

Genetic Variability and Low Local Diversity in a Population of Arbuscular Mycorrhizal Fungi. Proceedings of the National Academy of Sciences of the United States of America 101: 2369-2374.

Koide RT 1991. Nutrient supply, nutrient demand and plant response to mycorrhizal

infection. New Phytologist 117: 365-386. Koricheva J, Gange A, Jones T 2009. Effects of mycorrhizal fungi on insect herbivores: a

meta-analysis. Ecology 90: 2088-2097. Krüger M, Stockinger H, Krüger C, Schüßler A 2009. DNA-based species level detection of

Glomeromycota: one PCR primer set for all arbuscular mycorrhizal fungi. New Phytologist 183: 212-223.

Krüger M, Krüger C, Walker C, Stockinger H, Schüßler A 2012. Phylogenetic reference data

for systematics and phylotaxonomy of arbuscular mycorrhizal fungi from phylum to species level. New Phytologist 193: 970-984.

Kuhn G, Hijri M, Sanders IR 2001. Evidence for the evolution of multiple genomes in

arbuscular mycorrhizal fungi. Nature 414: 745-748. Landcare Research 2011. Geospatial Data Integration Portal

http://gisportal.landcareresearch.co.nz Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, McWilliam H, Valentin

F, Wallace IM, Wilm A, Lopez R and others 2007. ClustalW and ClustalX version 2. Bioinformatics 23: 2947-2948.

Lee J, Lee S, Young JPW 2008. Improved PCR primers for the detection and identification of

arbuscular mycorrhizal fungi. FEMS Microbiology Ecology 65: 339-349. Legendre P 1993. Spatial autocorrelation: trouble or new paradigm? Ecology 74: 1659-1673.

Page 99: Characterisation of arbuscular mycorrhizal fungal

88

Legendre P, Fortin MJ 1989. Spatial pattern and ecological analysis. Plant Ecology 80: 107-

138. Legendre P, Borcard D, Peres-Neto PR 2005. Analyzing beta diversity: partitioning the

spatial variation of community composition data. Ecological Monographs 75: 435-450.

Liu R, Wang F 2003. Selection of appropriate host plants used in trap culture of arbuscular

mycorrhizal fungi. Mycorrhiza 13: 123-127. Lloyd-Macgilp SA, Chambers SM, Dodd JC, Fitter AH, Walker C, Young JPW 1996.

Diversity of the ribosomal internal transcribed spacers within and among isolates of Glomus mosseae and related mycorrhizal fungi. New Phytologist 133: 103-111.

Macarthur RH 1957. On the Relative Abundance of Bird Species. Proceedings of the

National Academy of Sciences of the United States of America 43: 293-295. Maherali H, Klironomos JN 2007. Influence of Phylogeny on Fungal Community Assembly

and Ecosystem Functioning. Science 316: 1746-1748. Mangan S, Adler G 2002. Seasonal dispersal of arbuscular mycorrhizal fungi by spiny rats in

a neotropical forest. Oecologia 131: 587-597. Mantel N 1967. Detection of Disease Clustering and a Generalized Regression Approach.

Cancer Research 27: 209-&. Mark AF, Wilson JB, Scott C 2011. Long-term retierent of New Zealand snow tussock

rangeland: effects on canopy structure, hawkweed (Hieracium spp.) invasion and plant diversity. New Zealand Journal of Botany 49: 243-262.

Marler MJ, Zabinski CA, Callaway RM 1999. Mycorrhizae indirectly enhance competitive

effects of an invasive forb on a native bunchgrass. Ecology 80: 1180-1186. Meffin R, Miller AL, Hulme PE, Duncan RP 2010. Biodiversity research: Experimental

introduction of the alien plant Hieracium lepidulum reveals no significant impact on montane plant communities in New Zealand. Diversity and Distributions 16: 804-815.

Mitchell CE, Agrawal AA, Bever JD, Gilbert GS, Hufbauer RA, Klironomos JN, Maron JL,

Morris WF, Parker IM, Power AG and others 2006. Biotic interactions and plant invasions. Ecology Letters 9: 726-740.

Moora M, Berger S, Davison J, Öpik M, Bommarco R, Bruelheide H, Kühn I, Kunin WE,

Metsis M, Rortais A and others 2011. Alien plants associate with widespread generalist arbuscular mycorrhizal fungal taxa: evidence from a continental-scale study using massively parallel 454 sequencing. Journal of Biogeography 38: 1305-1317.

Morton JB, Bentivenga SP, Wheeler WW 1993. Germ plasm in the International Collection

of Arbuscular and Vesicular-arbuscular Mycorrhizal Fungi (INVAM) and

Page 100: Characterisation of arbuscular mycorrhizal fungal

89

proceedures for culture development, documentation and storage. Mycotaxon 48: 491-528.

Mosse B, Hayman DS, Ide GJ 1969. Growth responses of plants in unsterilized soil to

inoculation with vesicular-arbuscular mycorrhiza. Nature 224: 1031-1032. Muller HJ 1964. The relation of recombination to mutational advance. Mutation Research 1:

2-9. Mummey DL, Rillig MC 2008. Spatial characterization of arbuscular mycorrhizal fungal

molecular diversity at the submetre scale in a temperate grassland. FEMS Microbiology Ecology 64: 260-270.

Newsham KK, Fitter AH, Watkinson AR 1995. Multi-functionality and biodiversity in

arbuscular mycorrhizas. Trends in Ecology & Evolution 10: 407-411. Norton DA, Espie P, Murray W, Murray J 2006. Influece of pastoral management on plant

biodiversity in a depleted short tussock grassland, Mackenzie Basin. New Zealand Journal of Ecology 30: 335-344.

Oksanen J, Blanchet FG, Kindt R, Legendre P, Minchin PR, O'Hara RB, Simpson GL,

Solymos P, Henry M, Stevens H and others 2011. vegan: Community Ecology Package http://cran.r-project.org accessed 10 January 2012

Ollson PA, Jakobsen I, Wallander H 2002. Foraging and resource allocation strategies of

mycorrhizal fungi in a patchy environment. In: van der Heijden MGA, Sanders IR ed. Mycorrhizal Ecology. Berlin, Germany, Springer-Verlag. Pp. 93-115.

Olsson PA, Jakobsen I, Wallander H 2002. Foraging and resource allocation strategies of

mycorrhizal fungi in a patchy environment. In: van der Heijden MGA, Sanders IR ed. Mycorrhizal Ecology. Berlin, Germany, Springer-Verlag. Pp. 93-115.

Öpik M, Moora M, Liira J, Zobel M 2006. Composition of root-colonizing arbuscular

mycorrhizal fungal communities in different ecosystems around the globe. Journal of Ecology 94: 778-790.

Öpik M, Moora M, Zobel M, Saks Ü, Wheatley R, Wright F, Daniell T 2008. High diversity

of arbuscular mycorrhizal fungi in a boreal herb-rich coniferous forest. New Phytologist 179: 867-876.

Öpik M, Vanatoa A, Vanatoa E, Moora M, Davison J, Kalwij JM, Reier Ü, Zobel M 2010.

The online database MaarjAM reveals global and ecosystemic distribution patternsin arbuscular mycorrhizal fungi (Glomeromycota). New Phytologist 188: 223-241.

Parniske M 2008. Arbuscular mycorrhiza: the mother of plant root endosymbioses. Nat Rev

Micro 6: 763-775. Patel R, Lin C, Laney M, Kurn N, Rose S, Ullman EF 1996. Formation of chimeric DNA

primer extension products by template switching onto an annealed downstream

Page 101: Characterisation of arbuscular mycorrhizal fungal

90

oligonucleotide. Proceedings of the National Academy of Sciences of the United States of America 93: 2969-2974.

Pawlowska TE, Taylor JW 2004. Organization of genetic variation in individuals of

arbuscular mycorrhizal fungi. Nature 427: 733-737. Pimentel D, Zuniga R, Morrison D 2005. Update on the environmental and economic costs

associated with alien-invasive species in the United States. Ecological Economics 52: 273-288.

Pivato B, Mazurier S, Lemanceau P, Siblot S, Berta G, Mougel C, Van Tuinen D 2007.

Medicago species affect the community composition of arbuscular mycorrhizal fungi associated with roots. New Phytologist 176: 197-210.

Powell JR, Parrent JL, Hart MM, Klironomos JN, Rillig MC, Maherali H 2009. Phylogenetic

trait conservatism and the evolution of functional trade-offs in arbuscular mycorrhizal fungi. Proceedings of the Royal Society B: Biological Sciences 276: 4237-4245.

Pringle A, Bever JD 2008. Analogous effects of arbuscular mycorrhizal fungi in the

laboratory and a North Carolina field. New Phytologist 180: 162-175. Pringle A, Bever JD, Gardes M, Parrent JL, Rillig MC, Klironomos JN 2009. Mycorrhizal

symbioses and plant invasions. Annual Review of Ecology, Evolution, and Systematics 40: 699-715.

R Development Core Team 2011a. The R Stats Package v. 2.13.2. R Development Core Team 2011b. R: A language and environment for Statistical

Computing. Vienna, Austria, R Foundation for Statistical Computing. Raab PA, Brennwald A, Redecker D 2005. Mitochondrial large ribosomal subunit sequences

are homogeneous within isolates of Glomus (arbuscular mycorrhizal fungi, Glomeromycota). Mycological Research 109: 1315-1322.

Radford I, Dickinson KJM, Lord JM 2009. Does the invader Hieracium lepdiulum have a

comparitive growth advantage over co-occurring plants? High leaf area and low metabolic costs as invasive traits. New Zealand Journal of Botany 47: 395-403.

Radford IJ, Dickinson KJM, Lord JM 2006. Nutrient stress and performance of invasive

Hieracium lepidulum and co-occurring species in New Zealand. Basic and Applied Ecology 7: 320-333.

Radford IJ, Dickinson KJM, Lord JM 2007. Functional and performance comparisons of

invasive Hieracium lepidulum and co-occurring species in New Zealand. Austral Ecology 32: 338-354.

Radford IJ, Dickinson KJM, Lord JM 2010. Does disturbance, competition or resource

limitation underlie Hieracium lepidulum invasion in New Zealand? Mechanisms of establishment and persistence, and functional differentiation among invasive and native species. Austral Ecology 35: 282-293.

Page 102: Characterisation of arbuscular mycorrhizal fungal

91

Raghothama K, Karthikeyan A 2005. Phosphate acquisition root physiology: from Gene to

Function. In: Lambers H, Colmer T ed, Springer Netherlands. Pp. 37-49. Read DJ, Perez-Moreno J 2003. Mycorrhizas and nutirent cycling in ecosystems: a journey

towards relevance? New Phytologist 157: 475-492. Redecker D, Kodner R, Graham LE 2000. Glomalean fungi from the Ordovician. Science

289: 1920-1921. Redecker D, Hijri I, Wiemken A 2003. Molecular identification of arbuscular mycorrhizal

fungi in roots: perspectives and problems. Folia Geobotanica 38: 113-124. Reinhart KO, Packer A, van der Putten WH, Clay K 2003. Plant–soil biota interactions and

spatial distributions of black cherry in its native and invasive ranges. Ecology Letters 6: 1046-1050.

Remy W, Thomas NT, Hass H, Kerp H 1994. Four hundred-million-year-old vesicular

arbuscular mycorrhizae. Proceedings of the National Academy of Sciences of the United States of America 91: 11841-11843.

Renker C, Weißhuhn K, Kellner H, Buscot F 2006. Rationalizing molecular analysis of field-

collected roots for assessing diversity of arbuscular mycorrhizal fungi: to pool, or not to pool, that is the question. Mycorrhiza 16: 525-531.

Ridgway HJ, Kandula J, Stewart A 2006. Optimising the medium for producing arbuscular

mycorrhizal spores and the effect of inoculation on grapevine growth. New Zealand Plant Protection 59: 338-342.

Ripley BD 1977. Modelling spatial patterns (with discussion). Journal of the Royal Statistical

Society, Series B 39: 172-212. Ripley BD 1988. Statistical inference for spatial processes. Cambridge, UK, Cambridge

University Press. Roberts A, Radford IJ, Orlovich DA 2009. Do alterations of arbuscular mycorrhizal fungal

communties change interactions between an invader Hieracium lepidulum and two co-occurring species? A glasshouse study. Australasian Mycologist 28: 29-35.

Rose A, Frampton CM 2007. Rapid short-tussock grassland decline with and without grazing,

Marlborough, New Zealand. New Zealand Journal of Ecology 31: 232-244. Rose AB, Platt KH, Frampton CM 1995. Vegetation change over 25 years in a New Zealand

short-tussock grassland: Effects of sheep grazing and exotic invasions. New Zealand Journal of Ecology 19: 163-174.

Rose AB, Suisted PA, Frampton CM 2004. Recovery, invasion, and decline over 37 years in

a Marlborough short-tussock grassland, New Zealand. New Zealand Journal of Botany 42: 77-87.

Page 103: Characterisation of arbuscular mycorrhizal fungal

92

Rosendahl S 2008. Communities, populations and individuals of arbuscular mycorrhizal fungi. New Phytologist 178: 253-266.

Rosendahl S, Stukenbrock EH 2004. Community structure of arbuscular mycorrhizal fungi in

undisturbed vegetation revealed by analyses of LSU r DNA sequences. Molecular Ecology 13: 3179-3186.

Ruiz-Lozano JM 2003. Arbuscular mycorrhizal symbiosis and alleviation of osmotic stress.

New perspectives for molecular studies. Mycorrhiza 13: 309-317. Russell AJ, Bidartondo MI, Butterfield BG 2002. The root nodules of the Podocarpaceae

harbour arbuscular mycorrhizal fungi. New Phytologist 156: 283-295. Russell J, Simon B 2005. The Liverwort Marchantia foliacea Forms a Specialized Symbiosis

with Arbuscular Mycorrhizal Fungi in the Genus Glomus. New Phytologist 165: 567-579.

Sala OE, Chapin FS, Armesto JJ, Berlow E, Bloomfield J, Dirzo R, Huber-Sanwald E,

Huenneke LF, Jackson RB, Kinzig A and others 2000. Global Biodiversity Scenarios for the Year 2100. Science 287: 1770-1774.

Sanders IR 1990. Seasonal patterns of vesicular-arbuscular mycorrhizal occurence in

grasslands. New Phytologist 120: 517-524. Sanders IR 2002. Ecology and evolution of multigenomic arbuscular mycorrhizal fungi.

American Naturalist 160: S128-S141. Sanders IR, Alt M, Groppe K, Boller T, Wiemken A 1995. Identification of Ribosomal DNA

Polymorphisms among and within Spores of the Glomales - Application to Studies on the Genetic Diversity of Arbuscular Mycorrhizal Fungal Communities. New Phytologist 130: 419-427.

Santos-Gonzalez JC, Finlay RD, Tehler A 2007. Seasonal dynamics of arbuscular

mycorrhizal fungal communities in roots in a seminatural grassland. Applied and Environmental Microbiology 73: 5613-5623.

Schüβler A, Schwarzott D, Walker C 2001a. A new fungal phylum, the Glomeromycota:

phylogeny and evolution. Mycological Research 105: 1413-1421. Schüβler A, Gehrig H, Schwarzott D, Walker C 2001b. Analysis of partial Glomales SSU

rRNA gene sequences: implications for primer design and phylogeny. Mycological Research 105: 5-15.

Schwartz MW, Hoeksema JD, Gehring CA, Johnson NC, Klironomos JN, Abbott LK, Pringle

A 2006. The promise and the potential consequences of the global transport of mycorrhizal fungal inoculum. Ecology Letters, Blackwell Publishing Limited. Pp. 501-515.

Scott D 1993. Response of Hieracium in 2 Long-Term Manipulative Agricultural Trials. New

Zealand Journal of Ecology 17: 41-46.

Page 104: Characterisation of arbuscular mycorrhizal fungal

93

Simon L, Lalonde M, Bruns TD 1992. Specific amplification of 18s fungal ribosomal genes

from vesicular-arbuscular endomycorrhizal fungi colonizing roots. Applied and Environmental Microbiology 58: 291-295.

Smith B, Wilson J 2002. Community convergence: Ecological and evolutionary. Folia

Geobotanica 37: 171-183. Smith S, Read D 2008. Mycorrhizal Symbiosis. London, Academic. Speksnijder AGCL, Kowalchuk GA, De Jong S, Kline E, Stephen JR, Laanbroek HJ 2001.

Microvariation artifacts introduced by PCR and cloning of closely related 16S rRNA gene sequences. Applied and Environmental Microbiology 67: 469-472.

Spence L, Dickie I, Coomes D 2011. Arbuscular mycorrhizal inoculum potential: a

mechanism promoting positive diversity–invasibility relationships in mountain beech forests in New Zealand? Mycorrhiza 21: 309-314.

Stockinger H, Walker C, Schüßler A 2009. ‘Glomus intraradices DAOM197198’, a model

fungus in arbuscular mycorrhiza research, is not Glomus intraradices. New Phytologist 183: 1176-1187.

Sturges HA 1926. The choice of a class interval. Journal of the American Statistical

Association 21: 65-66. Stylinski CD, Allen EB 1999. Lack of native species recovery following severe exotic

disturbance in southern Californian shrublands. Journal of Applied Ecology 36: 544-554.

Suzuki M, Rappe MS, Giovannoni SJ 1998. Kinetic bias in estimates of coastal picoplankton

community structure obtained by measurements of small-subunit rRNA gene PCR amplicon length heterogeneity. Applied and Environmental Microbiology 64: 4522-4529.

Swofford DL 2003. Phylogenetic analysis using parsimony (*and other methods).

Sunderland, Massachusetts, Sinauer Associates Sykorova Z, Wiemken A, Redecker D 2007. Cooccurring Gentiana verna and Gentiana

acaulis and their neighboring plants in two swiss upper montane meadows harbor distinct arbuscular mycorrhizal fungal communities. Appl. Environ. Microbiol. 73: 5426-5434.

Taylor JW, Jacobsen DJ, Kroken S, Kasuga T, Geiser DM, Hibbett DS, Fisher MC 2000.

Phylogenetic species recognition and species concepts in fungi. Fungal Genetics and Biology 31: 21-32.

Tennant D 1975. Test of a modified line intersect method of estimating root length. Journal

of Ecology 63: 995-1001.

Page 105: Characterisation of arbuscular mycorrhizal fungal

94

Tommerup IC 1992. Methods for the study of the population biology of vesicular-arbuscular mycorrhizal fungi. In: Norris JR, Read DJ, Varma AK ed. Methods in Microbiology. London, Academic. Pp. 23-51.

Torchin ME, Lafferty KD, Dobson AP, McKenzie VJ, Kuris AM 2003. Introduced species

and their missing parasites. Nature 421: 628-630. Toussaint JP 2007. Investigating physiological changes in the aerial parts of AMF plants:

what do we know and where should we be heading? Mycorrhiza 17: 349-353. Trevors JT 1996. Sterilization and inhibition of microbial activity in soil. Journal of

Microbiological Methods 26: 53-59. Triplett EW, Barta TM 1987. Trifolitoxin production and nodulation are necessary for the

expression of superior nodulation competitiveness by Rhizobium leguminosarum bv. trifolii Strain T24 on Clover. Plant Physiol. 85: 335-342.

Trivers RL 1971. The evolution of reciprocal altruism. Quarterly Review of Biology 46: 35-

57. van der Heijden M, Klironomos J, Urisic M, Moutoglis P, Strietwolf-Engel R, Boller T,

Wiemken A, Sanders IR 1998a. Mycorrhizal fungal diversity determines plant biodiversity, ecosystem variability and productivity. Nature 396: 69-72.

van der Heijden MGA, Horton TR 2009. Socialism in soil? The importance of mycorrhizal

fungal networks for facilitation in natural ecosystems. Journal of Ecology 97: 1139-1150.

van der Heijden MGA, Bardgett RD, van Straalen NM 2008. The unseen majority: soil

microbes as drivers of plant diversity and productivity in terrestrial ecosystems. Ecology Letters 11: 296-310.

van der Heijden MGA, Boller T, Wiemken A, Sanders IR 1998b. Different arbuscular

mycorrhizal fungal species are potential determinants of plant community structure. Ecology 79: 2082-2091.

van der Heijden MGA, Klironomos JN, Ursic M, Moutoglis P, Streitwolf-Engel R, Boller T,

Wiemken A, Sanders IR 1998c. Mycorrhizal fungal diversity determines plant biodiversity, ecosystem variability and productivity. Nature 396: 69-72.

Vance CP, Uhde-Stone C, Allan DL 2003. Phosphorus acquisition and use: critical

adaptations by plants for securing a nonrenewable resource. New Phytologist 157: 423-447.

Vandenkoornhuyse P, Ridgway KP, Watson IJ, Fitter AH, Young JPW 2003. Co-existing

grass species have distinctive arbuscular mycorrhizal communities. Molecular Ecology 12: 3085-3095.

Viera A, Glenn MG 1990. DNA content of vesicular-arbuscular mycorrhizal fungal spores.

Mycologia 82: 263-267.

Page 106: Characterisation of arbuscular mycorrhizal fungal

95

Vitousek PM, Walker LR 1989. Biological ivasion by Myrica Faya in Hawai'i: plant

demography, nitrogen fixation, ecosystem effects. Ecological Monographs 59: 247-265.

Vitousek PM, Aber JD, Howarth RW, Likens GE, Matson PA, Schindler DW, Schlesinger

WH, Tilman DG 1997. Human alteration of the global nitrogen cycle: sources and conequences. Ecological Applications 7: 737-750.

Vogelsang KM, Bever JD 2009. Mycorrhizal densities decline in association with nonnative

plants and contribute to plant invasion. Ecology 90: 399-407. Wagg C, Jansa J, Schmid B, van der Heijden M 2011. Belowground biodiversity effects of

plant symbionts supports aboveground productivity. Ecology Letters 14: 1001-1009. Walker S, Wilson JB, Lee WG 2003. Recovery of short tussock and woody species guilds in

ungrazed Festuca novae-zealandiae short tussock grassland with fertiliser or irrigation. New Zealand Journal of Ecology 27: 179-189.

Warner NJ, Allen MF, MacMahon JA 1987. Dispersal agents of vesicular-arbuscular

mycorrhizal fungi in a disturbed arid ecosystem. Mycologia 79: 721-730. Webb CO, Ackerly DD, McPeek MA, Donoghue MJ 2002. Phylogenies and community

ecology. Annual Review of Ecology and Systematics 33: 475-505. Whitaker RH 1975. Communities and Ecosystems, 2nd ed. New York, Macmillan. Whitcomb S, Stutz J 2007. Assessing diversity of arbuscular mycorrhizal fungi in a local

community: role of sampling effort and spatial heterogeneity. Mycorrhiza 17: 429-437.

Wilson JB 1991. Methods for fitting dominance/diversity curves. Journal of Vegetation

Science 2: 35-46. Wilson JB, Agnew ADQ 1992. Positive-feedback switches in plant communities. Advances

in Ecological Research 23: 263-336. Wiser SK, Allen RB 2000. Hieracium lepidulum invasion of indigenous ecosystems. In:

Conservation Advisory Science Notes. In: Office DoCH ed. Wellington. Wiser SK, Allen RB, Clinton PW, Platt KH 1998. Community structure and forest invasion

by an exotic herb over 23 years. Ecology 79: 2071-2081. Wolfe B, Mummey D, Rillig M, Klironomos J 2007. Small-scale spatial heterogeneity of

arbuscular mycorrhizal fungal abundance and community composition in a wetland plant community. Mycorrhiza 17: 175-183.

Yamaura N, Higashi M, Behera N, Wakano JY 2004. Evolution of multualism through

spatial effects. Journal of Theoretical Biology 226: 421-428.

Page 107: Characterisation of arbuscular mycorrhizal fungal

96

Appendix

Supporting data contained on the attached computer disc:

S1 – RFLP gel photos

S2 – RFLP-type library

S3 – RFLP-type database