The diversity of fungi in four Irish forest types By ... - CiteSeerX

Post on 18-Jan-2023

4 views 0 download

Transcript of The diversity of fungi in four Irish forest types By ... - CiteSeerX

i

The diversity of fungi in four Irish forest types

By Richard O’Hanlon B.Sc. (Ed)

A thesis submitted for the degree of Doctor of Philosophy, At the Faculty of Science and Engineering,

University of Limerick, Ireland.

Supervisor:

Dr Thomas Harrington, Department of Life Sciences,

University of Limerick.

Submitted to the University of Limerick: May 2011

ii

iii

“The task of an ecologist”

There is an old story about a man who, returning home one night found his

neighbour searching the ground beneath a street lamp. “Can I help you find

something?” he asked. “I lost my key” replied the neighbour. “Do you know

about where you dropped it?”, “Yes” replied the neighbour “over there” pointing

to a dark corner of the street. “If you dropped it over there then why are you

looking here” asked the man. “Because this is where the light is” replied the

neighbour.

The task of the ecologist is not to bring the search to where the light is, but to bring the light to where the search is.

Perry et al. (2008)

iv

v

Abstract

Sampling of the macrofungal sporocarps, ectomycorrhizal morphotypes and vascular plants was carried out in 28 plots from four forest types (ash, oak, Scot’s pine, Sitka spruce) between the years 2007 and 2009. A total of 409 macrofungal species, 51 ectomycorrhizal morphotypes and 68 vascular plant species were recorded over the three years. It was found that at equal sampling intensities, there were no significant differences in total macrofungal species or ectomycorrhizal morphotype richness between the oak, Scot’s pine and Sitka spruce forest types. Species richness estimation revealed that between 45 and 77% of the above-ground macrofungal species richness and between 57 and 100% of the below-ground ectomycorrhizal morphotype richness were recorded. Cortinarius,

Mycena, Russula, Lactarius and Inocybe were the most species-rich genera recorded. Forty-eight macrofungal species are new records to Ireland and five of the species recorded are on the British Red-Data List.

Based on sporocarp frequency over the three year’s sampling, distinctive macrofungal communities were distinguished using nonmetric multi-dimension scaling, which corresponded to the dominant tree type of the forest. Ash forests lacked the ectomycorrhizal species component, oak forests had many wood- and litter-decay species present, Scot’s pine forests had some specific Lactarius species present (e.g. L. rufus, L. hepaticus) and Sitka spruce forests had a large diversity of Mycena species.

The below-ground ectomycorrhizal communities were surveyed in soil cores taken from four plots from each of three of the forest types. The ectomycorrhizal communities of the forest types (oak, Scot’s pine and Sitka spruce) were distinctly different according to the dominant tree species of the plot. The use of mantle descriptions, RFLP and sequencing methods allowed for the identification of 36 ectomycorrhizal species. The morphology and anatomy of remaining 15 morphotypes is described.

The ability of plantation forests to provide a supplementary habitat for native fungal species richness and diversity is discussed.

vi

vii

Declaration

I hereby declare that I am the sole author of this thesis and that it has not been submitted for any other academic award. References and acknowledgements have been made, where necessary, to the work of others. . Richard O’Hanlon Department of Life Sciences University of Limerick Ireland

Date:

viii

ix

Acknowledgements

I wish to thank my supervisor Dr Tom Harrington for all his help and guidance throughout the work involved in this thesis. I am very grateful for the many hours spent in the lab passing on his taxonomic knowledge of fungi and also for the help he gave me to develop this thesis and also my writing skills. I would like to thank all the staff of the Department of Life Sciences for their help, support and friendship during my years as both an undergraduate and post-graduate. Prof. John Breen, University of Limerick and Dr Gareth Griffith, Aberystwyth University, provided valuable comments and corrections to an earlier draft of this thesis. I am very grateful to the Council for Forest Research and Development (COFORD) for funding this PhD through the FUNCTIONALBIO project. I would like to express my gratitude to Prof. Tom Bolger, School of Biology and Environmental Science, UCD, for his organisation and management of the FUNCTIONALBIO project. I would like to thank my sponsors whom accepted me as a Visiting Scholar during this research. Dr Sue Grayston, Department of Forest Sciences, University of British Columbia, for all her help in organising my trip to Canada. Dr Dan Luoma and Dr Joyce Eberhart for their kindness, support and explanation of the methods for describing ectomycorrhizal communities during my Visiting Scholar trip to Oregon. Both Visiting Scholar exchanges greatly broadened my views on ecology and mycology. COFORD is gratefully acknowledged for funding both Visiting Scholar exchanges. I am grateful to my family for their support and encouragement during my time at University. I would also like to thank all of my friends and fellow postgrads for talking about topics other than my research with me. I would like to thank Dr Chris Quine, Forest Research, and Dr Jonathan Humphrey for allowing access to the “Biodiversity in Britain’s planted forests” data.

x

xi

List of Abbreviations

Abbreviation Full term

ACE Abundance-based Coverage Estimator ANOVA Analysis of Variance CHAO2 Chao 2 richness estimator CWD Coarse Woody Debris DBH Diameter at Breast Height DNA Deoxyribonucleic Acid ECM Ectomycorrhizal FAO Food and Agriculture Organisation FSC Forest Stewardship Council ICE Incidence-based Coverage Estimator ITS Internal Transcribed Spacer IV Indicator Value JI Jaccard Index MRPP Multi-Response Permutation Procedure NMS Nonmetric Multi-dimensional Scaling PAR Photosynthetic Active Radiation PCR Polymerase Chain Reaction PRS Plant Root Simulator RFLP Restriction Fragment Length Polymorphism SD Standard Deviation SFM Sustainable Forest Management SP Scot’s pine SS Sitka spruce U.K. United Kingdom

xii

xiii

Usage of Terms

Shown below is a word cloud created from the most frequently used words in the main body of the thesis. The size of the word relates to the frequency of its use. The usage frequencies range from 60 to 1700 uses, depending on font size.

Species

Soil Ireland Coniferous Plot Vegetation

Sitka spruce Lactarius Scot’s pine Variables Forest

Diversity Function Ash Rarefaction Fungi CWD Tree Russula

Ectomycorrhizal Native Cortinarius Indicator Richness Significant EstimationGenus Plantation Analysis

Biodiversity Habitat

Community

Oak Irish

xiv

xv

Contents

Abstract…………………………………………….…….…………...……...…..iii Declaration…………………………………………….………...……………….iv Acknowledgements………………………………….……………………………v Listof Abbreviations………………………………...……………………………xi Usage of terms………………..…………………………………………………xiii Chapter 1: Introduction 1.1 The changing face of Irish forests ........................................................................... 3

1.2 Biodiversity and Irish forests .................................................................................. 4

1.3 The Functionalbio project ....................................................................................... 6

1.4 Macrofungi in forest ecosystems ............................................................................ 7

1.5 Layout of the thesis ................................................................................................. 8

Chapter 2: Literature review 2.1 Sustainable forest management and certification .................................................. 13

2.2 COFORD and recent forest research in Ireland .................................................... 14

2.3 Studies of fungi in forest ecosystems. ................................................................... 16

2.4 Forest management practices and their effects on forest fungi ............................. 17

2.5 Functional groupings for fungi ............................................................................. 21

2.6 Below-ground ectomycorrhizal diversity in forests .............................................. 26

2.7 Soil and site variables and their relationship with fungal diversity ...................... 27

Chapter 3: The sites: vegetation and site variables measured 3.1 Introduction ........................................................................................................... 41

3.1.1 Forests of Ireland: Extent and current trends ..................................................... 41

3.1.2 Classification of Irish forest habitats ................................................................. 43

3.1.3 Rational for site selection ................................................................................... 44

3.2 Aims of this chapter .............................................................................................. 47

3.3 Materials and methods .......................................................................................... 48

3.3.1 The sites ............................................................................................................. 48

3.3.2 Data collection methods ..................................................................................... 52

3.3.3 Statistical analysis .............................................................................................. 54

3.4 Results ................................................................................................................... 58

3.4.1 Soils physical and chemical attributes. .............................................................. 58

3.4.2 Relationship between plots based on soil variables measured ........................... 67

3.4.3 The species list ................................................................................................... 69

3.4.4 The vegetation and classification of the individual sites ................................... 75

3.4.5 Structural descriptions of the sites ..................................................................... 79

3.4.6 Meteorological data ............................................................................................ 83

3.4.7 Community structure of the vegetation in the plots ........................................... 85

3.4.8 Forest management related to plant species richness ....................................... 105

3.5 Discussion ........................................................................................................... 107

3.5.1 Significantly different plant communities across the four forest types ........... 107

3.5.2 The plant communities of the four forest types ............................................... 108

3.5.3 Relationship of these findings to possible fungal species richness .................. 114

3.6 Conclusions ......................................................................................................... 120

Chapter 4: Macrofungal species richness and diversity in the different forest types 4.1 Introduction ......................................................................................................... 123

4.1.1 Fungal diversity in temperate forest ecosystems ............................................. 123

xvi

4.1.2 Fungal diversity in Irish forests and forests similar in tree species composition and climate ............................................................................................ 124

4.1.3 Factors affecting macrofungal diversity in forests ........................................... 125

4.1.4 Difficulties in estimating fungal diversity in forests ........................................ 127

4.2 Aims of this chapter ............................................................................................. 129

The aims of this chapter are:...................................................................................... 129

4.3 Materials and Methods ....................................................................................... 130

4.3.1 Sites .................................................................................................................. 130

4.3.2 Macrofungal assessment ................................................................................... 130

4.3.3 Statistical analysis ............................................................................................ 131

4.3.4 Species richness comparisons and the estimation of fungal diversity in the forest types and sites .................................................................................................. 132

4.3.5 Species diversity and evenness analysis of the sites......................................... 133

4.3.6 Species richness of functional groups in the different forest types .................. 134

4.3.7 Functional group frequencies across the forest types ....................................... 135

4.3.8 Influence of site variables on species richness ................................................. 135

4.3.9 Seasonal effects on fungal phenology .............................................................. 136

4.4 Results ................................................................................................................. 137

4.4.1 Macrofungal species in the forest plots ............................................................ 137

4.4.2 Species richness estimation of fungal diversity in the forest types .................. 153

4.4.3 Species richness estimation: variation within forest types ............................... 161

4.4.4 Species diversity and evenness ......................................................................... 165

4.4.5 Functional groups of macrofungi ..................................................................... 167

4.4.6 Influence of site variables on species richness ................................................. 171

4.4.7 Effect of weather variables on fungal fruiting .................................................. 176

4.5 Discussion ............................................................................................................ 182

4.5.1 Fungal species richness and the effect of forest type ....................................... 182

4.5.2 Rare and common fungi in Irish forests ........................................................... 188

4.5.3 Functional groups of fungi in Irish forests ....................................................... 191

4.5.4 Species and fruitbody abundance correlated to weather parameters ................ 193

4.5.5 Estimating fungal diversity in Irish forest habitats ........................................... 194

4.6 Conclusions ......................................................................................................... 199

Chapter 5: Macrofungal communities of the forest types 5.1 Introduction ......................................................................................................... 203

5.1.1 Mycocoenological study methods .................................................................... 203

5.1.2 Macrofungal communities of forest ecosystems .............................................. 204

5.1.3 Past studies of macrofungal communities in temperate oak, ash, Scot’s pine and Sitka spruce forest habitats ................................................................................. 206

5.2 Aims of this chapter ............................................................................................. 209

5.3 Materials and methods ......................................................................................... 210

5.3.1. Site list ............................................................................................................. 210

5.3.2 Macrofungal assessment ................................................................................... 210

5.3.3 Abiotic variables ............................................................................................... 210

5.3.4 Stand structural attributes ................................................................................. 210

5.3.5 Statistical and multivariate analysis ................................................................. 210

5.3.6 Multivariate analysis ........................................................................................ 211

5.3.7 Year to year variation ....................................................................................... 213

5.4 Results ................................................................................................................. 215

5.4.1 Species abundance patterns .............................................................................. 215

5.4.2 Analysis of macrofungal community similarity between the sites ................... 224

xvii

5.4.3 Ordination of the forest plots. .......................................................................... 230

5.4.4 Indicator species analysis ................................................................................. 237

5.4.5 Change in fungal community over sample years ............................................. 245

5.5 Discussion ........................................................................................................... 247

5.5.1 Evidence for the existence of distinctive communities of macrofungi in Irish forests ........................................................................................................................ 247

5.5.2 The macrofungal communities of the forest types ........................................... 249

5.5.3 Host preference in macrofungi ......................................................................... 258

5.6 Conclusions ......................................................................................................... 260

Chapter 6: Ectomycorrhizal morphotype richness and community analysis of the forest types 6.1 Introduction ......................................................................................................... 263

6.1.1 Ectomycorrhizas in forest ecosystems ............................................................. 263

6.1.2 Quantifying and recording ectomycorrhizal fungi in forests ........................... 264

6.1.3 Ectomycorrhizal research in Ireland and other temperate countries ................ 265

6.2 Aims of this chapter ............................................................................................ 268

6.2 Methods ............................................................................................................... 269

6.3.1 The plots ........................................................................................................... 269

6.3.2 Sampling and enumeration ............................................................................... 269

6.3.3 Molecular identification of ECM types............................................................ 270

6.3.4 Statistical analysis ............................................................................................ 272

6.3.5 Community similarity analysis ........................................................................ 272

6.3.6 Multivariate community analysis ..................................................................... 273

6.4 Results ................................................................................................................. 275

6.4.1 ECM morphotypes ........................................................................................... 275

6.4.2 ECM richness and abundance in the forest types............................................. 286

6.4.3 Similarity of below-ground ECM assemblage and above-ground ECM sporocarp assemblage................................................................................................ 300

6.4.4 ECM community analysis ................................................................................ 304

6.5 Discussion ........................................................................................................... 316

6.5.1 Ectomycorrhizal diversity ................................................................................ 316

6.5.2 The ECM communities of the forest types ...................................................... 318

6.5.3 Relationship between above-ground ECM sporocarps and below-ground ECM morphotypes .................................................................................................... 322

6.5 Conclusions ......................................................................................................... 328

Chapter 7: General discussion and conclusions 7.1 Macrofungal diversity in Irish forest sites .......................................................... 331

7.2 Macrofungal functional group diversity.............................................................. 333

7.3 Below-ground ectomycorrhizal diversity ............................................................ 334

7.4 The macrofungal communities of Irish forests ................................................... 335

7.5 The value of plantation forests as habitats for macrofungal diversity ................ 340

7.6 Unanswered questions and future research directions ........................................ 342

7.7 Final conclusions: Plantations as a habitat for native macrofungi ...................... 345

7.8 Key findings of this project ................................................................................. 346

References ................................................................................................................. 348

Colour plates ............................................................................................................. 388

Appendix 1 ................................................................................................................ 395

Appendix 2 ................................................................................................................ 396

xviii

1

Chapter 1: Introduction

2

3

Introduction

1.1 The changing face of Irish forests

The Food and Agriculture Organization of the United Nations states that over two

thirds of the worlds land based species live in forests (Anon. 2005). Up until

2005, very little research has been carried out on the biodiversity of forest

plantations in Ireland and how it changes through different stages of the forest

cycle. Investigation into the ecological impacts of the exotic conifer Sitka spruce

(Picea sitchensis) plantations, which will account for at least 60% of the forest

cover in Ireland up to 2030 (Anon. 1996), is necessary. Currently, native tree

species comprise less than 25% of the forested area (National Forest Inventory

2007), but generous grants are available to encourage the planting of more mixed

crops and native broadleaf trees.

Ireland was once covered in dense forests of oak, elm, pine, ash and hazel,

but anthropogenic influence (Cole and Mitchell 2003) has reduced the area of

native woodland to less than 1% of the land area, one of the lowest levels in

Europe (FRA 2010). Native woodlands in Ireland belong to a number of

phytosociological categories, but the principal types are acidophilus oak woods

(Blechno-Quercetum petraeae) dominated by sessile oak Quercus petraea, and

ash-elm-hazel woodland (Corylo-Fraxinetum) on more fertile soils (White and

Doyle 1982). What remains is highly fragmented and frequently modified by

naturalized non-indigenous species such as beech (Fagus sylvatica) and sycamore

(Acer pseudoplatanus) (Perrin et al. 2008).

The trend in Europe is very similar to that in Ireland with major

afforestation being carried out on former agricultural land (UNECE 2003). The

increase of forested area in Europe is 770,000 ha per year (FRA 2010) with over

50% of this increase due to afforestation in Spain, Italy, Norway and Russia. Of

the 50 countries examined in Europe, only Albania and Estonia reported a net loss

of forest area (total loss 8000ha; FRA 2010). In countries such as Belgium,

Denmark, Netherlands, Portugal, United Kingdom and Ireland, exotic species

dominate a large proportion (>30%) of the forest area (UNECE 2000). The

majority of afforestation in Ireland includes the non-native conifer Sitka spruce,

often as the dominant species in mixed plantations (Joyce and O’Carroll 2002).

4

Sitka spruce is particularly well suited to growth conditions in Ireland and if

properly managed on fertile sites, can reach a maximum annual mean increment

of 30m3/ha/annum (Horgan et al. 2004). The fast growth of Sitka spruce in Ireland

makes it an ideal candidate for carbon sequestration. Sitka spruce forests are

second only to unthinned poplar forests in their carbon sequestration ability in

Ireland (Kilbride et al. 1999).

Coniferous tree species such as Sitka spruce and Douglas fir become

commercially mature much faster than deciduous trees such as oak and beech. By

way of planting restrictions and generous grants, the Irish government is

attempting to promote more planting of native and deciduous tree species. Current

planting restrictions state that a plantation must contain at least 10% broadleaf

species and that these trees should be planted “in swathes and not as single stems

within the canopy” (Anon. 2000). As the majority of current afforestation involves

conifers on improved/enclosed land, non-intimate mixes of conifers and

broadleaves are, therefore, likely to become the dominant configuration in future

afforestation (Smith et al. 2005).

1.2 Biodiversity and Irish forests

The word “biodiversity” has been defined in a variety of ways (more than 80

definitions in De Long 1996). In this study the definition used is that of the

Convention on Biological Diversity CBD:

Biological diversity means the variability among living organisms from all sources including, inter-alia, terrestrial, marine and other aquatic ecosystems and the ecological complexes of which they are part; this includes diversity within species, between species and of ecosystems.

(Anon. 1992)

Biodiversity in this form is extremely difficult to measure and therefore

only organism diversity at the level of species is examined because it

encompasses different hierarchies of biological and ecological diversity. Thus,

this project examines the biodiversity of fungal species in Irish forest ecosystems.

Biodiversity is known to have large beneficial effects on ecosystem productivity

5

and stability (Gaston and Spicer 2007) and therefore its conservation is extremely

important.

In this study, functional group biodiversity or functional biodiversity is

also investigated. This is the diversity of species which fit into a specific

functional group and are capable of carrying out that ecosystem function. The

meaning of functional biodiversity and its value to ecosystems was examined by

Bolger (2001), who puts forwards a number of hypotheses as to how we should

look at biodiversity and its value to an ecosystem. The common theme in the

article is that each living organism in an ecosystem is limited in the tasks it can do

and therefore an ecological “division of labor” occurs among the organisms in the

ecosystem (Hector 2011). If one organism were to disappear then there is a chance

that another organism can fill its functional role. However there is also the

possibility that some organisms are keystone participants to the ecosystem and

once lost, their function will be lost as well. This would lead to changes in the

ecosystem structure, chemistry and biological characteristics. Functional

redundancy has been identified in soil decomposer fungi in the U.K., where it was

found that when abundant species were removed from the ecosystem, occasional

species could fill their functional position (Deacon et al. 2006).

It is generally accepted that “biodiversity per se is a good thing; that its

loss is bad and that something should be done to preserve it” (Gaston 1996).

Giller and O’Donovan (2002) provide a good review of the question ‘Does

biodiversity matter’ and list a number of reasons why biodiversity should be

conserved and protected. One of the reasons they list is that biodiversity and its

effect on ecosystems has a monetary value. Constanze et al. (1997) has estimated

the value of the ecological services provided by ecosystems at US$33 trillion. In a

more recent assessment, Bullock et al. (2008) explained that although it is

difficult to put an exact monetary figure on biodiversity in Irish forests, if it is

taken that biodiversity has positive effects on wood production, wood loss

prevention, increased carbon sequestration as well as an increase in aesthetic

value and forest visits, then the total worth of biodiversity to Irish forests could be

more than €700 million per annum. When dealing with such a large source of

income, it is wise to plan and research how possible actions will affect future

growth predictions. As the famous American ecologist Aldo Leopold stated, the

6

first precaution of intelligent tinkering is to "keep every wheel and cog" (Leopold

1949).

A significant start has been made in examining and recording the

biodiversity in Irish forests through the BIOFOREST programme

(http://bioforest.ucc.ie/). The BIOFOREST research group revealed the

biodiversity of vegetation, bryophytes, hoverflies, spiders and birds in both native

and plantation forests as well as making recommendations of how to increase the

biodiversity in the selected forests (Smith et al. 2005; Iremonger et al. 2006;

Smith et al. 2006).

1.3 The FUNCTIONALBIO project

Following on from the large amount of work completed by the BIOFOREST

project, the functional biodiversity of Irish native and plantation forests was

identified as a key research area. The FUNCTIONALBIO project

(http://www.coford.ie/) was set up in 2007 and funded by the Council for Forest

Research and Development (COFORD). The overall aim of the

FUNCTIONALBIO project is to investigate the “Functional biodiversity in

forests, including the diversity of soil decomposers and predatory and parasitic

arthropods. The project is composed of two work packages:

(1) Fungal diversity (macrofungi and ectomycorrhizal diversity) package.

Investigated by Mr Richard O’Hanlon and Dr Thomas Harrington, Department of

Life Sciences, University of Limerick.

(2) Predatory and parasitic arthropods package.

Investigated by Dr Julio Arroyo and Prof. Thomas Bolger, School of Biology and

Environmental Science, University College Dublin.

Both packages are examining the same forest sites over the same time

period (2007-2009). This thesis is concerned with package 1 of the

FUNCTIONALBIO project, which investigates the fungal diversity of the forest

types. A final report on the project will draw together the key findings of both

packages and identify the factors that affect functional biodiversity in these forest

types.

7

1.4 Macrofungi in forest ecosystems

In terms of biomass, diversity and their ecological functions, fungi are the most

important group of organisms in forest ecosystems after the trees themselves (Paul

2007). It is estimated that the fungal biomass in soils exceeds the biomass of all

other soil organisms combined, except plant roots (Paul 2007). A study by Hunt

and Wall (2002) which modelled the removal of different biotic components of

the soil found that the removal of only two component groups, bacteria and

saprophytic (decay) fungi caused drastic changes in net primary productivity. This

suggests that as a functional group (decomposers in the latter case), fungi are

essential for soil processes.

With an estimated 1.5 - 4.5 million species worldwide (Hawksworth 1991;

O’Brien et al. 2005) fungi are one of the most diverse groups of organisms. They

vary from microscopic yeasts to gigantic mushrooms which can grow to weigh up

to 316 kg (Burdsall et al. 1996). A fungus also holds the record for the biggest

living organism ever recorded (Armillaria bulbosa which was found in forest soil

in Oregon which spread for over 40 hectares; Ferguson et al. 2003). The

biodiversity of fungi in an ecosystem has been shown to affect plant diversity (van

der Heijden et al. 1998) and thus primary productivity in terrestrial ecosystems.

Fungi hold key roles in the maintenance of forest ecosystems in particular. Moore

et al. (2001) lists these as:

• Nutrient cycling, retention and formation of soil structure.

• Food in detritivore food webs in forests and forest streams.

• Micro-habitat creation in forests by fungal pathogens.

• Mycorrhizal mutualisms.

Many countries have recognised the important roles fungi play in

ecosystems and have created laws and regulations which protect endangered

species. Slovakia has listed 52 species of fungi as having a ‘special legal status’

which enables the prevention of damage to habitats where these fungi are found

(Lizon 1999). The United Kingdom (English Nature 1999) and Switzerland (Egli

et al. 1995) have also taken steps to halt the loss of fungal diversity by creating

codes of practice and recommendations to achieve fungal conservation. In

contrast, Ireland has no fungi listed as protected in the most recent Checklist of

Protected Species in Ireland (Anon. 2003). On a broader scale, the exclusion of

8

fungi from red lists continues. There are 45,000 species on the International Union

for the Conservation of Nature (IUCN) international red-list of which only one,

Pleurotus nebrodensis, is a fungus (IUCN 2009). Despite the recognition of fungi

as a key part of most ecosystems, studies on fungi are still relatively limited when

compared to the number of studies on vegetation and arthropods in ecosystems.

This may be due to the limited number of researchers working in mycology. Atlas

et al. (1992) pointed out the decline of scientists studying mycology when he

stated “the number of microbial taxonomists has diminished to a level that

qualifies those now working in this speciality as themselves being considered

members of an endangered species”.

1.5 Layout of the thesis

This thesis is composed of seven chapters and two appendices. Chapter 1 is an

introduction to biodiversity and fungal diversity in forests. It also gives the

background to the FUNCTIONALBIO project. Chapter 2 is a literature review

examining the theory and practical findings of previous research into macrofungal

and ectomycorrhizal diversity in forests. Chapter 3 examines the vascular plant

diversity of the plots and gives a rationale for the selection of the forest types. The

vascular diversity is used to predict likely macrofungal diversity. Chapter 4 deals

with the species richness and functional group richness of macrofungi in the forest

types. The completeness of the macrofungal survey is statistically estimated using

species richness estimators. Chapter 5 describes the macrofungal community

composition of the different forest types and examines the relationship between

site variables and community composition. Chapter 6 focuses on the below-

ground ectomycorrhizal diversity of the forest types, describing the

ectomycorrhizal morphotypes using both morphological and molecular methods.

Finally, Chapter 7 provides a general discussion and links the findings of Chapters

3-6 into an overall description of fungal biodiversity in Irish forests and its

relationship to the biodiversity of other organisms in Irish forests. Appendix 1 is

supplied as an attached CD containing images related to this project. Appendix 2

contains an uncorrected proof of the article “Diversity and distribution of

9

mushroom forming fungi (Agaricomycetes) in Ireland” (O’Hanlon and Harrington

in press).

10

11

Chapter 2: Literature review

12

13

2.1 Sustainable forest management and certification

With the percentage of land use devoted to forestry increasing in Ireland, it is of

the utmost importance that more research is conducted to analyze the suitability of

the Forest Biodiversity Guidelines (Anon. 2000) for conserving and protecting

biodiversity and improving our understanding of the effects these plantation

forests will have on the environment and the biodiversity of the Irish countryside.

It is through Sustainable Forest Management (SFM) that both the ecological and

economical aspects of forests can be protected. As forest-related enterprises

employ directly and indirectly over 3.35 million people in Europe, it is important

that forest health is protected, so that Europe can continue to compete with other

nations in terms of forest products created and exported (Anon. 2006b).

The term Sustainable Forest Management covers many aspects of forest

health and sustainability but broadly speaking it can be taken to mean the use and

conservation of forests for the benefit of present and future generations. It is a

term which originated in the 1990s and has been added to and improved over the

years to suit current forest requirements and recommendations. In the most recent

Food and Agriculture (FAO) forestry publication title “State of the World’s

Forests 2007” (FAO 2007) seven thematic elements have been identified through

research by groups such as the United Nations Forum on Forests. These elements

are:

• Extent of forest resources.

• Biological diversity.

• Forest health and vitality.

• Productive functions of forest resources.

• Protective functions of forest resources.

• Socio-economic functions.

• Legal, policy and institutional framework.

In Ireland, sustainable forest certification is carried out by the Forest

Stewardship Council (FSC) (www.fsc.org). The FSC is a not-for-profit body

which is one of two forest certification bodies in Europe, the other being the

14

Programme for the Endorsement of Forest Certification Schemes

(http://www.pefc.org). The FSC has 10 principles which must be adhered to if a

forest is to be FSC certified. The certification is issued for 5 years and Coillte

successfully retained its certification in 2006.

The main principles of the FSC (www.coillte.ie) that deal with biological

diversity (and therefore fungal diversity) on a site are:

• Environmental impact: management will work to conserve biological

diversity so as to protect the ecological functions and integrity of the

forest.

• Benefits from the forest: management will encourage the use of the forests

for production of many products and services to ensure economic viability.

Edible or medicinal fungal fruitbodies would be an example of non-wood

forest products which have economic values.

• Maintenance of high conservation value forests: management decisions

which affect forests of high conservation status will maintain or enhance

the attributes which define such forests.

In the most recent FSC draft report, it was highlighted that a decrease in

the planting of Sitka spruce, increase in the planting of broadleaf species and an

increase in open space/biodiversity enhancement area are necessary in order to

meet FSC criteria (Anon. 2009). This project will produce data that will contribute

to the fulfillment of the three FSC principles outlined above and therefore help

Irish forests to receive the FSC standard.

2.2 COFORD and recent forest research in Ireland

With many consumers becoming more conscious of environmental issues, it is

through certification and eco-labeling that independent reassurance and Forest

Certification is assuring consumers that the timber or timber products they buy

come from a sustainable and well managed forest. Without this eco-friendly

certification, exports of Irish timber would suffer and revenue from timber exports

would drop. To research Irish forest practices and their effects on biodiversity, the

forest service division of the Department of Agriculture, Fisheries and Food set up

15

the National Council for Forest Research and Development (COFORD).

COFORD invests a large sum of money into the research and development of

forest policies and priorities. Some of the projects funded in the past which relate

to forest biodiversity and ways to increase it are the BIOFOREST projects

1. Assessment of biodiversity at different forest stages (Smith et al. 2005)

2. Experimental methods to enhance biodiveristy in plantation forests

(Iremonger et al. 2006)

3. Biodiveristy assessment of afforested sites (Smith et al. 2006)

The first project, undertaken by Smith et al. (2005), identified a number of

indicators of good forest health and high diversity in both ash and Sitka spruce

plantation forests. Indicators such as the type of vegetation present in a site or the

presence of a specific bird species were identified and related to forest

biodiversity. Overall the project made a number of recommendations on how

forest procedures and planting regimes might be changed to enhance forest

biodiversity of vegetation, bryophytes, hoverflies, spiders and birds in forests

(Smith et al. 2008).

The second project, by Iremonger et al. (2006) pointed out that there is no

“one size fits all” with regard to increasing forest biodiversity. Priority must be

given to retaining areas which are of high conservation value. Another

recommendation relates to the importance of open spaces in Irish plantation

forests (Smith et al. 2007; French et al. 2008). Guidelines were given for the

minimum area which should be retained as open spaces and also for the minimum

sizes for forest paths and roads.

The third project, investigated by Smith et al. (2006), assessed the changes

which occurred in afforested lands in the species assemblages of vegetation,

spiders, hoverflies and birds. It was found that the changes that took place in the

first five years of afforestation were not as marked as the changes that took place

as forests age from pre-thicket to thicket, and mature to over-mature (Oxbrough et

al. 2005, 2006, 2006b). However, Oxbrough et al. (2006) also pointed out a

limitation of their study, in that it is not very meaningful to compare the diversity

of different habitat types such as a wetland, which may have a low biodiversity

when surrounded by other wetlands can have a very high biodiversity if it was

16

situated in an area of improved grassland. With this in mind they accord caution to

comparisons of different habitat types.

Of the listed recommendations given in the three projects above, the ones

most related to this project are:

1. The use of surrogate measures of biodiversity as an indicator of fungal

species richness, e.g. quantity of Coarse Woody Debris (CWD) quantity

(Ferris et al. 2000a).

2. The effect of mixed tree crops on biodiversity of plantation forests e.g.

birch spruce mixtures should be investigated for its effect on biodiversity.

3. The retention of high conservation value areas within forests e.g. variable

retention, where an area which has a high level of biodiversity is retained

and a new forest cycle is planted around it thus conserving and possibly

promoting the spread of the plants and animals back into the newly

forming forest.

The National Forest Inventory was completed by the Forest Service of the

Department of Agriculture and Food in 2007 and has also collected much data on

forest composition, health and biodiversity (National Forest Inventory 2007). It

has produced a large body of data on the diversity of plants and lichens in Irish

woodlands but no such work on fungi has taken place.

2.3 Studies of fungi in forest ecosystems.

There have been many studies carried out in the United Kingdom (U.K.)

and Europe looking at the diversity of fungi in different forested habitats (Table

2.1). Biodiversity or ecological diversity is scale-dependent and has a hierarchy in

nature (Whittaker 1977). Diversity can be broken up into levels, alpha, beta,

gamma and epsilon. Alpha diversity, which is also known as within habitat

diversity, is the diversity in a patch or in the case of this project within-site

diversity. Beta diversity describes the contribution of multiple habitats to the

overall diversity of a site (Zak and Willig 2004). In the case of this project, beta

diversity is the between-site diversity. Gamma diversity indicates the number of

species within a region or landscape type. For this project, gamma diversity is

taken to mean between-forest diversity. Epsilon diversity, which is the highest

17

level of diversity put forward by Whittaker (1977), is a measure of the diversity of

a large biogeographic region. Epsilon diversity was not examined in this project

because of the small area of forest landscape and fragmented nature of the

remaining native forests.

Macrofungal diversity is gaining more research interest in Europe in recent

years (Table 2.1). The projects listed that are most related to the

FUNCTIONALBIO projects are the two projects in the U.K., by Ferris et al.

(2000a) and by Humphrey et al. (2000). These studies found, contrary to the

existing opinion, that fungal diversity is usually lower in non-native forests, exotic

conifers such as Sitka spruce and Norway spruce can support a large and varied

collection of fungi, lower than, but nonetheless comparable to that of native

forests. Many of the other studies (Table 2.1) found that fungal diversity and

species abundance is affected by many different factors including soil physical

and chemical variables along with forest structural variables.

Table 2.1 List of previous fungal diversity studies from forests in Europe

Country Scale Reference

England Gamma Ferris et al. 2000a

England Gamma Humphrey et al. 2000

England Alpha Tofts and Orton 1998

England Beta Orton 1987

Estonia Beta Tedersoo et al. 2006

Ireland Beta Heslin et al. 1992

Norway Beta Hoiland and Bendickson 1996

Portugal Alpha Baptista et al. 2010

Scotland Gamma Alexander and Watling 1987

Scotland Beta Newton et al. 2002

Spain Beta Fernandez et al. 2006

Spain Alpha Sarrionandia et al. 2009

Spain Beta Oria de rueda et al. 2010

Sweden Beta Ruhling and Tyler 1990

2.4 Forest management practices and their effects on forest fungi

It has been suggested that even moderate forest management practices can have a

negative effect on macrofungal species diversity in forested sites (Bader et al.

1995). Studies by Penttila (1995), Kotiranta and Niemela (1996) and Lindbald

(1998) in Eastern Europe, Finland and Norway show that species richness of

18

wood-inhabiting fungi is lower in managed forests than in pristine or near-pristine

forests. Forest management is nevertheless important because forest managers and

ecologists strive to strike a balance in the best interest of both disciplines. This is

one of the key principles of SFM, that forests are used in a sustainable and

ecologically safe way for the benefit of current and future generations. There are

many practices, some old and some new, that are used to generate and harvest

high quality wood from forests. These practices can have both positive and

negative effects on forest biodiversity.

The retention of biodiversity in forests has been recognised as a key factor

in the forestry policy of countries such as U.K. (Forestry Commission 1998). The

Heritage Council of Ireland has created a framework for the conservation of

biodiversity in Irish habitats at a local level, and has listed the conservation of

biodiversity, and the sustainable use of the components of biodiversity, as its main

goals (Anon. 2003). Some of the forestry practices currently in place or being

investigated are: thinning of the understory of coniferous forests, planting of

mixed species of trees, retention of coarse woody debris (CWD) on site, the

retention of trees beyond financial maturity and the creation of open spaces within

the forest. A recent and beneficial technique to be developed as a spin off from the

COFORD project ForestScan (COFORD 2009), is the use of terrestrial laser

scanning to accurately measure total wood volume. By estimating how much

wood will be harvested, forest managers can reduce the amounts of waste timber

harvested and also leave areas forested until the timber is needed for processing.

The planting of mixed species of trees as an alternative to monocultures

has been investigated in Ireland for many years. Despite the fact that Sitka spruce

and the other imported conifers are capable of growing on sites of poor fertility

that cannot be used for farming, it has been realized that having single dominant

species has economic, biological and environmental risks. An example of a

biological risk caused by planting of monocultures was the recent outbreak of the

fungal pathogen Phytophthora ramorum, which has caused widespread damage to

larch plantations in the U.K. (Webber et al. 2010), indicating an apparent host

shift from broadleaf to coniferous tree species. To protect against these risks,

planting rules were created (Anon. 1996), that (1) reduced the annual planting of

Sitka spruce to below 60% of the total trees planted and (2) set out guidelines to

ensure that all sites had a measure of heterogeneity. The benefits of mixed crops

19

of trees have been investigated in Ireland and elsewhere, and show that a mixture

of the primary tree species with a nursing tree species produces a more vigorous

growth rate (O’Carroll 1978; Scherer-Lorenzen 2005), less need for fertilizer

application (Carey et al. 1988), resistance to pests and diseases (Su et al. 1996;

Muller 1998; Lugo 1997) and a more substantial monetary return at the end of the

cycle (Joyce and O’Carroll 2002). In studies in Europe and the United States,

mixed forests of two or more tree species have been found to have higher

diversity of vascular plants (Felton et al. 2010), bryophytes (Coote et al. 2008),

lichens (Humphrey et al. 1998) and fungi (Bills et al. 1986; Ferris et al. 2000a)

compared to monocultures. Massicotte et al. (1999) suggest that by having a

mixed crop of trees in a site, there are more opportunities for fungi (especially

ectomycorrhizal fungi) to colonize one or more tree species, due to many fungi

being host generalists and spreading vegetatively to different tree species.

The retention of trees beyond financial maturity is listed in the Forest

Biodiversity Guidelines (Anon. 2000) as a key factor in increasing biodiversity in

Irish forests. This practice is supported by the findings of many researchers

(Peterken et al. 1992; Ferris et al. 2000a; Jones et al. 2003; Humphrey 2005), who

hypothesised that retaining some old growth trees will help to re-establish the

biota that were present in the forest before felling. The general consensus is that

retaining a patch containing a number of trees is much more biologically efficient

than retaining single trees scattered around the site (Luoma et al. 2004). It has

been shown that during primary establishment, young trees depend on

mycorrhizal inoculum from neighbouring trees (Jumpponen et al. 2002;

Ashkannejhad and Horton 2006; Cline et al. 2005). Jones et al. (2003), Dickie and

Reich (2005) and Outerbridge and Trofymow (2004) have shown that the extent

of colonization of roots of immature trees by ectomycorrhizas of older trees

decreases with distance from the older trees. Luoma et al. (2004) found that any

removal of mature trees causes a loss in biomass of the next season’s mushroom

and truffle crop, and that retention of 15% of the stems is not sufficient to prevent

wholesale loss of the mushroom crop.

Just as biodiversity has been shown to be positively related to retention of

trees beyond financial maturity, clear-cutting has been shown to have negative

effects on forest biodiversity. The negative effects of clear-cutting have been

investigated by many researchers and reviewed by Rosenvald and Lohmus (2008).

20

These negative effects include drastic changes in abundance, biomass and

community structure of soil fauna (Huhta 1971, 1976; Huhta et al. 1967, 1969;

Grayston et al. 2006), woodland plant species (Hill 1979), and woodland fungi

(Baath 1981; Jones et al. 2003). In fact, it has been found that many ECM fungi

cannot survive clear-cutting (Amaranthus 1992; Borchers and Perry 1990) and

therefore depend on pockets of surviving inoculum to re-colonise the site. It is not

just the removal of the trees which negatively effects fungal diversity; the

processes related to the removal of the timber also damage soil fungal networks.

Machines used for tree removal damage the soil layer and cause major soil

disturbances to the soil’s moisture and temperature regime, stand structure and

soil micro-site conditions (Franklin and DeBell 1973; Rogerson 1976; Stark

1982). The effect of forest litter-layer disruption on fungi has been investigated by

Luoma et al. (2006a) and depending on the depth of the soil disturbance, a large

reduction (up to 75%) was found in the fruiting of the matsutake fungus during

the ensuing nine years of the study.

Pruning and thinning are very important for the creation of a high quantity

and quality timber in Sitka spruce forests (Joyce and O’Carroll 2002). The

removal of low branches both allows more light to permeate to the ground layers

and also significantly adds to the CWD present of the forest floor. Pruning also

promotes plant diversity in conifer forests (Iremonger 1999; Ferris et al. 2000b)

by allowing shade-limited species to re-colonize from the seed bank or invade

from surrounding habitats. Thinned Sitka spruce stands also have greater aesthetic

appeal. It has been found by Vesterdal et al. (1995) that in un-thinned forests,

carbon, nitrogen and phosphorus accumulated in unavailable forms when litter

production rates exceeded decomposition rates of the ground litter. Thinning, in

addition to wind-throw, is an important contributor to the volume of deadwood in

a stand, in the form of stumps and felled trees (Smith et al. 2005).

Although thinning can have positive effects on biodiversity in forests,

heavy-intensity thinning has been shown to reduce fungal sporocarp numbers by

studies in the Pacific Northwest forests (Pilz and Perry 1984; Waters et al. 1994;

Colgan et al. 1999) and in oak forests in Sweden (Norden et al. 2008). It has been

shown that recently thinned forests experience an overall reduction of fungal

sporocarp production. This finding is in agreement with Edmonds and Lebo

(1998), which found that where bryophyte cover increased (possibly due to

21

increased light penetration to the forest floor) sporocarp production was reduced

or absent. A contributory factor may be increased soil desiccation due to the

reduction in canopy cover in thinned forests (Siitonen et al. 2005). The report by

Colgan et al. (1999) found that, in variably-thinned forests, a reduction in total

hypogeous fungi fruitbodies was found, but that total species richness was higher

in lightly-thinned stands. The authors believe that the process of thinning induced

the fruiting of these extra species, possibly through a survival response, similar to

when plants allocate much of their resources to creating reproductive structures to

ensure the survival of future generations of the plant.

The occurrence of various types of CWD has been listed as a key factor

for increasing biodiversity in forest ecosystems (Franklin et al. 1987; Esseen et al.

1992; Kohm and Franklin 1997). Current recommendations promote the retention

of high quality and a high quantity of CWD in forest sites (Bader et al. 1995;

Christensen and Emborg 1996; Smith et al. 2005; Sippola and Renvall 1999). The

forest biodiversity guidelines (Anon. 2000) state that CWD should be left on site

as standing stems, fallen logs and wind-thrown areas after thinning and final

harvesting. This provides a habitat or refuge for some forest species and can then

promote re-colonization of the forest from the left-over CWD (Renvall 1995).

CWD left on site is one of the main substrates for many fungal fruitbodies,

including corticoid basidiomycetes which have been described as hatcheries for

diverse groups of insects and arthropods (Norden et al. 2008; O’Connell and

Bolger 1997; Okland et al. 2005, 2008). In fact, CWD has been given such a

positive link with enhancing biodiversity that it is being used as a surrogate to

measure relative forest health by forest managers (Kohl 1996; Tomppo 1996;

Gunnar-Jonsson and Jonsell 1999).

2.5 Functional groupings for fungi

The term functional diversity is used in this project to quantify diversity and

richness in groups of fungal species that have identical methods of acquiring

nutrition. The examination of functional groups instead of the number of species

is useful as some relationships between diversity and the environment only

emerge when a functional approach is employed (Smith et al. 2005; Williams and

Hero 2001; Keil et al. 2008). As Waksman (1916) explained “the question is not

22

how many numbers and types of fungi can be found in the soil, but what

organisms lead an active life in soil”. Within a single taxon such as genus or

family, the diversity of functional groups is likely to be determined by completely

different factors than is species richness (Williams & Hero 2001; Schweiger et al.

2007). Schweiger et al. (2007) found that had an analysis of species diversity only

been carried out on their hoverfly data, then the environmental variables would

not have explained the diversity or abundance of hoverflies found on their sites.

However, when the species were grouped based on an array of traits covering a

broad spectrum of a species’ ecology, it was found that environmental variables

had a significant effect on these functional groups. This suggests that a functional

approach may provide a better base for conservation and applied research

(Schweiger et al. 2007).

Functional group analysis is important because in most cases species

extinction is not a random occurrence, and is driven by species dependence on a

resource or on the limited dispersal of that species. Functional groups which have

a large dependence on a resource and also have limited dispersal mechanisms are

even more at risk of extinction due to these non-random events. The functional

groupings being used in this project are the same as those used in Ferris et al.

(2000a). These are ectomycorrhizal fungi (ECM), wood-decay (WD), litter-decay

(LD) and parasitic (P) fungi. Functional groups that are not of prime consideration

in this study include entomogenous fungi, phylloplane fungi, lichens, endophytic

fungi, yeasts, fungal parasites of insects and vertebrates, and aquatic

hyphomycetes.

Wood- and litter-decay fungi (Saprobic fungi)

Saprobic fungi have important recycling functions in terrestrial ecosystems. The

fungal hyphae release enzymes that break down remains of plants and animals,

and release these nutrients back into the soil when the plant and animal material

decomposes. With up to 70% of the above-ground biomass of forests in the form

of perennial plants (Rodin and Basilevic 1968), it is important that the nutrients

stored in this biomass are recycled into the forest soil. The saprobic fungi

examined in this study are litter- and wood-decay fungi. Litter-decay fungi break

down leaves and coniferous needles and so release nutrients from these into the

ecosystem in available forms. The diversity and community composition of these

23

fungi are tightly linked to litter layer variables, and litter variables can vary

significantly in different forest types.

Wood-decay fungi break down branches and CWD in the forest

environment. It has been shown that the diversity of wood-decay fungi is closely

related to the amount of CWD left in situ (Ferris et al. 2000a). Factors such as

CWD size (Odor et al. 2006; Heilmann-Clausen and Christensen 2005), CWD

quality (Sippola et al. 2005) and a range of CWD at different decay stages (Boddy

et al. 1987; Hoiland and Bendickson 1996) are most important in the promotion of

a high diversity of wood-decay fungi in forests.

It has been suggested that the amount of CWD can be used as an indicator

for biodiversity of forest ecosystems (Humphrey et al. 2003). Forest sites with a

high amount of CWD are often rich in red-list species. Studies by Heilmann-

Clausen and Christensen (2003, 2005) have shown that many red-listed polypores

are often restricted to unmanaged forests that characteristically have a high

amount of CWD present. Crites and Dale (1998) found that species richness of

lichens and bryophytes on logs in contact with the ground increased as the logs

progressed through decay stages.

Ectomycorrhizal (ECM) fungi

Ectomycorrhizas are but one of a number of different groups of mycorrhizal fungi

found on roots of forest trees and some forest shrubs. The others include

arbuscular, ericoid and ectendomycorrhizal types, which are not a primary

consideration in this study, because the principal tree species investigated in this

study (oak, Scot’s pine and Sitka spruce) are ECM hosts, and because ECM

diversity is reflected to some extent in above-ground sporocarps. ECM fungi form

symbiotic relationships with the majority of angiosperm and coniferous forest

trees (Molina et al. 1992) in temperate and boreal regions, primarily with oaks

(Quercus), beeches (Fagus), birches (Betula), poplars (Populus), pines (Pinus),

spruces (Picea), firs (Abies) and larches (Larix). In forest ecosystems where soil

fertility is relatively low, trees rely heavily on mycorrhizal associations for

nutrient uptake (Allen 1991; Leake et al. 2004; Smith and Read 2008).

Mycorrhizal relationships are one of the oldest and most important

relationships in terrestrial ecosystems (Brundrett 2002; Stubblefield and Taylor

1988). The fungus forms a mantle around the roots of the plant and aids it in

24

absorbing macro- and micro-nutrients from the soil (Cumming 1996), and thus

confers better growth and survival on the tree (Amaranthus and Perry 1987; Perry

et al. 1990). It has been shown that the fungal part of the mutualism can absorb

almost any limiting nutrient depending on the host (Allen 1991; Smith and Read

2008). Andersson et al. (1996) found increased concentration of phosphorus,

calcium, magnesium and potassium in plants with the ECM fungus Paxillus

involutus present on its roots. Trees with ECM fungi present on their roots also

have increased ability to acquire water, a factor which increases the tree’s ability

to survive in drought-affected areas (Lehto and Zwiazek 2011). In exchange for

this acquisition of nutrients, the plant provides the fungus with up to 20% of its

photosynthate (Graham 2000; Hobbie 2006). In temperate and boreal forest

ecosystems, it has been shown that ECM fungi colonize more than 90% of the

fine roots of trees (Markkola et al. 1996).

Work by Baxter and Dighton (2001) highlighted the importance of ECM

diversity to the host trees, showing that plants with increased ECM diversity on

their roots had higher concentrations of phosphorus in their tissues. The fungal

part of the ECM association (mycobiont) also has the ability to acquire nutrients

directly from decomposing litter (Leake and Read 1997), from live animals such

as springtails (Klironomos and Hart 2001), dead nematodes (Perez-Moreno and

Read 2001) and from stone particles in the soil (Jongmans et al. 1997; Landeweert

et al. 2001) so directly linking the above-ground and below-ground ecosystems.

Plants with mycorrhizae present on their roots have also been shown to have an

increased tolerance to harsh conditions, such as heavy metal toxicity and drought

(Hartley et al. 1997) and also a higher resistance to plant pathogens (Duchesne et

al. 1988).

Although a number of ECM fungi are host specialist in that they only

occur with a specific tree type (Harley and Harley 1987; Molina et al. 1992),

laboratory studies have shown that the majority of ECM fungi are host generalist

and often colonise more than one tree species (Molina and Trappe 1982; Smith et

al. 1995; Simard 1997b). Of the four tree species being investigated in this

project, ash is the only species which does not form ECM structures (Table 2.2)

while oak supports the greatest diversity of ECM fungi in the U.K. with similar

results expected for Ireland.

25

Table 2.2 Tree species distribution in Ireland (National Forest Inventory 2006) and associated ECM species for the tree genera in the U.K. (Newton and Haigh 1998).

Host genus Number of

associated ECM

fungi

Number of

specific ECM

fungi

Area occupied by

host genus in

Ireland (ha)

Ash (Fraxinus) 0 0 20,000

Oak (Quercus) 233 30 15,000

Pine (Pinus) 201 14 70,000

Spruce (Picea) 151 3 350,000

Another important function of ectomycorrhizas is that trees connected by

common mycorrhizal fungi can share carbon, nitrogen and phosphorus through

inter connecting mycelia in the soil (Simard et al. 1997a). Although this transfer

of carbon has been shown experimentally, Simard and Durall (2004) are hesitant

to identify net carbon transfer between trees by ectomycorrhizas as a source of

amelioration to younger or struggling trees, because they have not found the

labelled carbon in the high concentrations of the plant tissues to warrant this

statement. However, it is true that the transfer of carbon between the

ectomycorrhizas of different trees is a subsidy to the nutrient-gathering activities

of a plant, and one which can potentially affect plant community dynamics (Bever

2003).

Parasitic fungi

Parasitic fungi cause much damage to trees and crops all over the world. The

pathogenic fungi Heterobasidium annosum and Armillaria species cause

considerable damage to tree crops in Ireland each year (Joyce and O’Carroll

2002). Although many forest guidelines list these fungi as unwanted in forests,

they serve important functions such as nutrient recycling and also as population

control agents of trees in forests (Moore et al. 2001). Often, disease outbreaks are

indicative of the existence of less than optimum underlying forest conditions, and

the activities of pathogenic fungi alone are not necessarily the cause of the

perceived reduced forest health status (Ostry and Laflamme 2009). Pathogenic

fungi create habitats for many bird and small mammal species through the process

of heart-rot in large trees (Witt 2010), with around 40% of the bird species of

North American forests nesting in tree cavities (McComb and Lindenmayer

26

1999). Trees weakened by parasites are more susceptible to wind-throw, which

creates large canopy gaps and promotes regeneration of other tree species. Other

species of fungal pathogens attack specific tree species only, and so promote a

level of tree species diversity in forests. Annosus root rot caused by

Heterobasidion annosum attacks conifer species, but species such as ash

(Fraxinus spp.) and oak (Quercus spp.) are left undamaged, thus promoting

colonization by these and other broadleaves (Korhonen et al. 1998).

2.6 Below-ground ectomycorrhizal diversity in forests

The study of ECM populations presents particular problems for fungal ecologists.

There are two principal difficulties, identification and quantification of ECM

populations. In practice, most studies use a combination of morphological and

molecular methods for identifying ECM fungi. This method has been used by

Harrington and Mitchell (2002), Sakakibara et al. (2002), Horton and Bruns

(2001) and Kaldorf et al. (2004), where mycorrhizas were identified firstly by

morphological examination of the outer surface and also the mantle pattern

according to Agerer (1987-2002), and then by molecular methods to identify to

species level. It is important to examine the ECM structures on the roots as well as

above-ground sporocarps, as previous studies have shown sporocarps alone to be

bad indicators of the ECM fungi present in the soil (Clapp et al. 1995; Cripps

2004; Danielson 1984; Luoma et al. 2004). Quantification of population sizes is

difficult, because of the near impossibility of defining a fungal “individual”, and

also because of the fact that ectomycorrhizas are non-randomly distributed in the

vertical and horizontal dimensions of the soil layer (Horton and Bruns 2001;

Taylor 2002; Tedersoo et al. 2003). As population estimation is not an aim of the

present study, it will not be considered further here.

Past studies have shown that many ectomycorrhizas are host-specific

(Molina et al. 1992; Newton and Haigh 1998; Bills et al. 1986; Kranabetter et al.

1999; Ishida et al. 2007) and so it would be expected that differences in ECM

communities between the different forest types and also between chronological

stages of the forests (Last et al. 1987; Mason et al. 1987; Smith et al. 2002; Wang

et al. 2005) would be found. ECM communities are not only influenced by tree

species, but also by their position on the roots. It has been shown that ECM

27

communities vary in composition, according to their depth in the soil profile

(Dickie et al. 2002; Rosling et al. 2003; Tedersoo et al. 2003) and also their

distance from the tree (Bruns 1995; Deacon and Fleming 1992; Luoma et al.

2006a). Differing levels of carbon allocation between roots, near and far from the

tree, were once thought to be one of the main drivers of high ECM diversity in

forests (Bruns 1995), but recent findings such as niche specificity (Tedersoo et al.

2003; Buee et al. 2007) and functional ECM exploration types (Peay et al. 2011)

are now thought to have a stronger influence on ECM diversity at the horizontal

level in soil.

2.7 Soil and site variables and their relationship with fungal diversity

Soil biotic and abiotic variables have been suggested to have an influence on

fungal diversity (Enttema and Wardle 2002; Villeneuve et al. 1989). At the

smallest scale, fungi respond to differences in soil pore space, soil aggregates,

organic matter position and the distribution of fine roots in the soil horizon. At a

larger scale, soil type, previous land use and topography, influence the fungi that

can survive in the soil. Soil moisture was found to be positively correlated with

species richness of vascular plants in boreal forests (Zinko et al. 2005). A study

by Krebs et al. (2008) showed that by artificially increasing summer rainfall, the

biomass of fungal sporocarps increased two fold. Soil pH has been regarded as

one of the best indicators of fertility in terrestrial ecosystems such as deciduous

and coniferous forests in northern Europe (Nihlgard and Lindgreen 1977;

Diekmann 1994, 1999; Engelmark and Hytteborn 1999). The fact that pH can

have a large influence on the diversity of fungi in an ecosystem has been shown in

previous studies (Arnolds 1982; Bohus 1984; Harrington and Mitchell 2005a).

Humphrey et al. (2000) used principal components analysis to identify the main

soil factors affecting fungal diversity and found that pH, K, Mg and Ca were the

most influential soil factors in forests. It is well known that pH is related to many

other soil variables, and so pH is sometimes used as a proxy variable for other soil

chemical factors. For instance, at a low pH, soil nitrifying bacteria are limited in

their ability to convert soil ammonium into nitrate (Allison and Prosser 1991;

Jiang and Bakken 1999). Tyler (1985) analysed the base saturation or the

combined percent saturation of the four major cations that have a basic or alkaline

28

reaction (K+, Ca++, Mn++ and Mg++), and found that the fungal diversity was found

to be closely related to the base saturation of the soil.

The effect of, and response to soil abiotic variables is not a prime issue of

this study, as this study is mainly concerned with an examination of fungal

diversity in Irish forests.

The selection of variables known to affect forest fungal diversity

It has been found that the diversity of fungi in forest ecosystems is affected by

physical, chemical, biological, geographical and meteorological factors (Dickie

and Reich 2005; Villeneuve et al. 1989; Ferris et al. 2000a; Cripps 2004; Tyler

1985; Baar 1996; Schmit et al. 2005; Smith et al. 2003; Baar and Kuyper 1998;

Jones et al. 2003; Lee and Lee 2004; Krebs et al. 2008). The most used methods

of measuring these variables are dealt with in the following sections.

The variables examined here are:

1) Edge effect.

2) Soil physical structure.

3) Litter depth.

4) Canopy gaps.

5) Coarse woody debris (CWD).

6) Soil chemical characteristics (pH, Soil P, K, N).

7) Organic matter quantity.

8) Moisture content.

9) Trees species.

10) Tree age.

11) Vegetation layer.

12) Forest structural diversity.

13) Past land use.

14) Altitude.

15) Aspect.

16) Weather.

29

The edge effect and the measurement of forest canopy gaps

Forest edges can be hotspots of diversity due to the numerous biotic and abiotic

gradients that occur where the forest ecosystem meets the adjoining ecosystem.

Due to their significant effect on migration, evolutionary changes of organisms

and species diversity and abundance, edges have become one of the most

researched areas in ecology (Ries et al. 2004). A significant edge effect has been

found in studies on the diversity of fungi (Allen 1987; Kranabetter and Wylie

1998; Kernaghan and Harper 2001; Heilmann and Christenson 2003; Cripps 2004;

Dickie and Reich 2005; Siitonen et al. 2005), plants, birds, mammals and

invertebrates (for review see Ries et al. 2004). Areas within a forest stand with

breaks in the canopy that allow more light to penetrate to the ground floor, are

known as forest gaps and also constitute edges.

Forest gaps have also been identified as areas with high diversity of

vascular plants and bryophytes (Smith et al. 2007a; French et al. 2008). Both

forest edges and forest gaps in the canopy can experience: (a) an increase in

ground floor vegetation due to the higher availability of solar radiation, (b) an

increase in soil moisture from the increase of rainfall reaching the soil, and (c) an

increase in wind turbulence at the forest floor when compared to an area within

the forest interior. All of these effects can influence fungal diversity in the

resulting microsites created. Certain ECM fungal species have been shown to be

more common near areas close to gaps in the forest canopy (Cripps 2004). This

could be due to the preference of some ECM species for areas of low root density

(Peay et al. 2011). Also, fungal species that rely on air turbulence to disperse

spores often occur near the edge of forests and near forest gaps, where their spores

have a greater chance of being dispersed by wind currents (Allen et al. 1993).

The methods of identifying and quantifying canopy gaps have been

reviewed, and their advantages and disadvantages identified (Table 2.3).

Measurements of light or Photsynthetically Active Radiation (PAR) reaching the

ground are time-consuming. Simpler methods to evaluate gaps in the forest

canopy involve measurement of canopy cover and/or canopy closure. The

definition and measurement of these two variables can be found in review papers

by Jennings et al. (1999) and Paletto and Tosi (2009).

30

Table 2.3 Techniques available for the measurement of canopy gaps.

Technique Reference Pros Cons

Photo-reactive paper

Friend 1961; Jennings et

al. 1999

Cheap and easy to set up.

Only approximate estimation of PAR given. Need to be calibrated at each site using quantum sensors. Takes a large amount of time to record.

Leaf area index (LAI) using commercial canopy analyser

Breda 2003 Simple method, results calculated quickly.

Sky conditions need to be similar at each site when measurement is made. Expensive equipment needed.

Hemispherical photography

Breda 2003; Brown et al. 2000

High repeatability, easy to use, robust comparisons between sites can be made.

Expensive equipment needed. Equipment very fragile, slow procedure

Canopy closure analysis by canopy scope

Brown et al. 2000

Cheap, easy to carry, High repeatability between different observers

Cannot be translated into PAR value

Crown position index using the crown illumination estimation

Clarke and Clarke 1992; Jennings et

al. 1999

No equipment necessary. Quick method of assessment

Poor repeatability of results between observers.

Canopy closure using the Moosehorn

Garrison 1949; Brown et al. 2000

Simple method to use. Has a long history of use in forestry research.

Equipment is cumbersome and fragile. Field of view restricted to directly above the observer

Soil physical characteristics

The physical, chemical and biotic characteristics of soil have profound influences

on fungal diversity (Enttema and Wardle 2002; Villeneuve et al. 1989). At the

smallest scale, fungi respond to differences in soil pore space, soil aggregates,

amount and location of organic matter and the distribution of fine roots in the soil

horizon. A soil with tightly packed soil particles restricts root growth (Brady and

Weil 2008) and so can also affect the diversity and distribution of ectomycorrhizal

fungi on root systems. Physical variables which best represent the physical

structure of the soil are bulk density and soil texture. Methods for measuring both

are found in Watson (2007) and Anon. (2006b).

31

Litter depth has also been found to be related to fungal diversity. Ferris et

al. (2000a) found that as litter depth increased, so did the fruiting of wood

saprotrophs and ECM fungi in British forests. The authors attribute this to

increased water-holding capacity of the soil due to the covering layer of leaf litter.

However, de Vries et al. (1995) found the opposite in nitrogen-enriched (due to

atmospheric pollution) forests in the Netherlands; it was found that sod cutting

(manual removal of the litter layer) increased the fruiting of saprobic and ECM

sporocarps significantly. Baar and Kuyper (1998) found that removal of the

Deschampsia flexuosa sward from their Scot’s pine plot in the Netherlands

significantly increased ECM fruiting. The tight mat of roots formed by D.

flexuosa is thought to inhibit ECM fruiting (Arnolds 1991). The current

dominance of another forest grass species, Luzula sylvatica, in Irish acidophilous

oak forests (Anon. 2005a) may have similar effects on the ECM fungi in these

forests. Soil depth and the horizon of the soil are also of great importance,

especially when dealing with ECM root sampling. Many studies have shown that

ECM species have certain preferences for either the organic or mineral soil layers

(Stendell et al. 1999; Taylor and Bruns 1999).

Soil organic matter (SOM) has also been shown to be related to the

diversity of fungi in forests. Tyler (1985) found that diversity and abundance of

sporocarps was highly correlated with SOM in Swedish forests. Harrington (2003)

and Harrington and Mitchell (2005a) found that the distribution and abundance of

several mycorrhizal and saprophytic basidiomycetes species in the Burren Dryas

heaths varied with SOM content. Humphrey et al. (2000) found that SOM was a

key soil variable used in ordinating their sites based on their fungal communities.

Erland and Taylor (2003) reviewed the effect of SOM on fungi, and concluded

that at least in the case of ECM fungi, certain species show a greater affinity for

high and low organic matter concentrations.

Coarse woody debris CWD

Dead wood is closely associated with the life-histories of a variety of flora and

fauna, including fungi (Norden et al. 2004b), bryophytes (Odor et al. 2006) and

invertebrates (Kappes 2006; Topp et al. 2006). The amount of CWD on site can

be measured in a number of ways. The total volume of each piece of wood in a

32

site may be found if a restricted number of samples are being focused on, as in

Hoiland and Bendiksen (1996). In studies which involve a number of different

sites at a number of different chronological stages, other methods of estimating

CWD are necessary. Newton (2007) provides a good introduction to these

methods. The line intercept method is the quickest method, and gives an estimate

of the volume of CWD in a site; calculated by counting the pieces of wood

crossing a transect of known length. The longer the transect length and the more

transects can be taken per site, the more accurate the estimate of CWD volume.

The line intercept method was originally devised by Warren and Olsen (1964),

and has been modified by Kirby et al. (1998) to give final estimates in m3/ha

according to the relationship:

V= nd2 π

2 10

4/8t

Where V= volume of deadwood in m3 per ha n = number of times diameter class crosses transect line d = diameter class of deadwood at point of intersection in meters t= length of transect line in meters

Similar methods have been used in fungal diversity studies by Ferris et al.

(2000a) and Humphrey et al. (2000), and have found that the diversity of wood-

decay fungi was significantly correlated to the amount of CWD present on site.

The decomposition stage of the CWD is also an important variable, and Newton

(2007) devised a system of rating CWD on a scale of 1-5 with 1 being freshly

fallen timber and 5 being well rotten timber. Previous studies have shown that

there is a definite chronosequence of fungal colonizers of CWD (Boddy et al.

1987; Boddy and Watkinson 1995). In studies on fungi in forests, it has been

found that many rare and threatened fungi such as Sclerophora coniophaea and

Skeletocutis odora in Sweden, have a distinct preference for CWD in the

advanced stages of decay (Kruys et al. 1999).

Soil chemical variables

Fungal diversity has been linked to soil chemical variables in many studies

(Termorshuizen 1991; Goodman and Trofymow 1998; Villeneuve et al. 1989;

Humphrey et al. 2000). The main variables measured include: pH, available and

total soil phosphorus and nitrogen, soil potassium, soil organic matter, and

moisture content. Soil pH has been regarded as one of the best indicators of

33

fertility in terrestrial ecosystems, such as deciduous and coniferous forests in

northern Europe (Nihlgard and Lindgreen 1977; Diekmann 1999; Engelmark and

Hytteborn 1999). pH can have a large influence on the species richness and

diversity of fungi in an ecosystem (Arnolds 1982; Humphrey et al. 2000).

The concentrations of different ions and anions in the soil has also been

found to affect fungal populations, with edaphic factors such as cation exchange

capacity (CEC), % base saturation (Hansen 1988; Tyler 1985; Ruhling and Tyler

1990) and richness value (standardised measure of calcium, magnesium,

potassium, phosphorus and nitrogen in soil; Nantel and Newman 1992), having

been found to be related to fungal diversity in previous studies.

Soil nitrogen has been shown to affect ECM species composition and

abundance (Newton and Piggot 1991). In repeated fertilization trials in the

coniferous temperate forests of British Columbia, the percentage colonization of

fine roots was found to be negatively correlated with the addition of fertilizer to

forests (Berch et al. 2006). Further still, some fungal genera have been found to be

either nitrophobic or nitrophilic, such as species from the genera Cortinarius and

Piloderma being nitrophobic and species of Tomentella and Thelephora being

nitrophilic (Lilleskov et al. 2002).

The total measurement of soil nutrients through a one-off collection of a

sample is valid, but tells very little about the flux of nutrients available to the plant

roots over time. A novel approach is to use an in situ method of measurement for

soil nutrient availability. Unlike conventional soil extractions, in situ burials of

ion exchange membranes (IEM) integrate all of the principal edaphic factors

affecting nutrient uptake by plants (soil moisture and temperature, mineralization

and immobilization, buffer power, dissolution, ion diffusion from greater

distances and free ion activities), regardless of soil type (Qian and Schoenau

2002). They have been used to great effect in the areas of forestry (Coll et al.

2007; Hangs et al. 2004), agriculture (Jowkin and Schoenau 1998) and

environmental research (Liang and Schoenau 1995). The results found using this

method for soil N in spruce and pine plots have been found to be similar

throughout a number of studies (Huang and Schoenau 1997; Johnson et al. 2001;

Hangs et al. 2004). The PRSTM probes system (Plant Root Simulator probes,

Western Ag Innovations Inc., Saskatoon, Sask.) differs from the traditional

34

methods of soil analysis as destructive sampling and disturbance of the site is

minimal.

Tree species and tree age

The main tree species and total diversity of tree species present in a site has a

large effect on the diversity of fungi found (Schmit et al. 2005; Ferris et al. 2000a;

Humphrey et al. 2000; Lee and Lee 2004; Ishida et al. 2007). Decay fungi are

known to be specific in the organic compounds they can break down in leaves,

needles and woody debris (Osono et al. 2003; Boddy and Rayner 1983;

Unterseher and Tal 2006; Ludley et al. 2008), and some ECM fungi have been

shown to be highly specific in the symbiotic relationships which they will form

with forest trees (Molina et al. 1992; Newton and Haigh 1998). Stand age in itself

is not normally a strong indicator of fungal species richness in a site; many studies

have found no significant differences between the total numbers of species found

in forests of different ages, but have found that the species profile of the sites did

differ between the age classes (Dighton et al. 1986; Humphrey et al. 2000;

Fernandez et al. 2006; Grebenc et al. 2009; Kranabetter et al. 2005). It has been

proposed that these differences in species composition between different aged

forests are more as a result of changes in the forest soil (Blasius & Oberwinkler

1989; Tyler 1989) and not because of the increase of tree age.

Forest floor vegetation

The vegetation composition of the forest floor is a variable which is often

recorded in studies on fungal diversity. The majority of the forest floor vegetation

in deciduous forests is composed of woodland herbs, the most diverse group of

vascular plants in these habitats (McCarthy and Bailey 1996; Hermy et al. 1999).

There can be large differences in the diversity and structure of vegetation in

broadleaved forests of the same association, due to differences in gazing pressure

and management history. Many woodland herbs and bryophytes of broadleaved

forest can colonize conifer plantations, but generally these plantations are much

poorer in ground vegetation, at least when young (Ferris et al. 2000b). The

National Forest Inventory (National Forest Inventory 2007) found that there were

large differences in the richness of the ground flora between coniferous and

deciduous forests in Ireland. Less than 10% of the deciduous forests surveyed had

35

less than five vascular plant species, while in the coniferous plantations surveyed,

over 20% of the survey sites had less than five plant species in the ground layer.

Forest floor vegetation diversity has been shown to be a good indicator of

fungal diversity in some studies (Perini et al. 1995; Gable and Gable 2007), but

others have found only weak correlations between plant diversity of the forest

floor vegetation (including vascular plants and bryophytes) and fungal diversity

(Humphrey et al. 2000, 2003; Ruhling and Tyler 1990; Sastad 1995). Despite

this, the diversity vegetation has been used in total fungal species richness

estimates (Hawksworth 1991). Watling (1995) has estimated that total fungal and

plant species richness is often at a 6:1 ratio in temperate countries, while

Villeneuve et al. (1989) found the macrofungal : plant ratio of temperate forests at

2:1 in Canada.

Vegetation can often serve as a proxy for soil chemical composition and

soil fertility. Barr (1996) conducted an interesting study which looked at the

decline of ectomycorrhizal fungi in forests, due to the perceived increase in

atmospheric nitrogen deposition. It was known that increased nitrogen deposition

also increased the ground cover of the grass Deschampsia flexuosa, which was

found to have adverse effects on the fruiting of some ECM fungi (Baar and Ter

Braak 1996). The thick mats made by the roots of Deschampsia species probably

inhibit the formation of many fungal sporocarps. The effect of a ground flora

dominated by woodrush Luzula sylvatica, may have similar effects on the

formation of fungal sporocarps, although studies investigating the effect of matt

forming herbs on fungal sporulation are lacking.

Forest structural composition

Managed commercial forests typically consist of one or two canopy layers, as

competition from “wild trees” is minimised to allow for full growth potential of

the cultivated tree crop. In natural forests, there can be multiple canopy layers due

to heterogeneity in the ages of the dominant canopy trees, and also the presence of

understory canopy trees and tall shrubs such as holly (Ilex aquifolium) and hazel

(Corylus avellana). Studies have shown that forests with a larger percentage cover

in multiple canopy layers have higher values for biodiversity (MacArthur 1964;

Wilson 1974), bird diversity (MacArthur and MacArthur 1961), plant diversity

(Ferris et al. 2000b) and fungal diversity (Ferris et al. 2000a; Humphrey et al.

36

2000) than those with a limited number of canopy layers. One method of

recording the species composition and structural composition of vegetation in a

site is the vertical cover index (Ferris-Kaan et al. 1998). Humphrey et al. (2000),

using this method, found significant correlations between the vertical cover index

and fungal diversity in native oak and plantation Sitka spruce and Scot’s pine

forests in the U.K. The method for determining the vertical cover index in

managed forests and semi-natural forests is based on visual assessments of plant

cover in four defined strata. The cover index of the site can then be calculated

using the formula from Humphrey et al. (2003):

CI = 1.9s1 + 3s2 + 10s3 + 5s4 CI = cover index s1 (field) 10 cm –1.9 m in height s2 (shrub) 2–5 m s3 (lower canopy) 5.1–15 m s4 (upper canopy) 15.1–20 m

The numbers before the strata layer in the formula refer to the height of the

layer being measured e.g. the s2 layer spreads from 2-5meters in height and

therefore the percentage cover in this layer is multiplied by 3. This index is not

meant to replace traditional quadrat sampling techniques used in ecological

inventories. It was devised as a method of describing the bottom-to-top

configuration of above-ground vegetation in a forest (Brokaw and Lent 1999).

Site history and past land use

Due to anthropogenic pressures, the land use of much of Europe and North

America has changed over the years. Large-scale deforestation by humans has led

to a situation where in 1996 it was estimated that only 53% of the world’s original

forest cover remained (World Resources Institute 2004 as cited in Perry et al.

2008). Land use prior to planting is of utmost importance when fungal diversity of

a forest is being examined. If land has been used for agricultural purposes in the

past, then elevated nutrient levels can affect the fungi that may colonise that area

(Arnolds 1988; Tyler 1985; Griffith et al. 2004). If the land had previously been

planted with forests, then surviving fungal propagules in the form of vegetative

mycelium and spores could increase the rate of fungal colonization and the

diversity of colonizers (Jones et al. 2003). Site histories can be key to explaining

the fungal diversity of a particular site (Heilmann-Clausen and Christensen 2005).

37

A study by Harrington and Mitchell (2002) described populations of ECM fungi

on Dryas shrub vegetation in Ireland, and suggested that populations of the ECM

host tree Scot’s pine (Pinus sylvestris) in the habitat that became locally extinct

(ca. 500 AD), had provided the necessary inoculation material for the currently

thriving ECM communities there. This finding from the Burren region of Ireland

is similar to that found for ECM species in Oregon, with some ECM species

forming arbutoid mycorrhizas with Arctostaphylos spp. and Arbutus menziesii

after the ECM hosts (Douglas fir) have been removed from the area (Molina and

Trappe 1982).

Site geographic characteristics

Altitude has been shown to be a determining factor in the survival of mycorrhizal

fungi, as many mycorrhizal fungi decrease in numbers as elevation increases

(Lagana et al. 1999; Kernagahan and Harper 2001). The aspect of the site may

also have an impact on the fungi appearing, as south-facing slopes will generally

get more sunlight than north-facing slopes, so a south-facing slope can expect

higher soil and site temperatures and lower moisture contents of the soil. Lee and

Lee (2004) found that fungal diversity was increased in their plots which faced

north, due to higher relative moisture levels in the soil.

Weather patterns and their effect on fungal diversity assessment

Fungal fruiting patterns have been shown to be highly influenced by weather

conditions (Ohenoja 1993; Gulden et al. 1992; Carrier 2003; Krebs et al. 2008)

making a comprehensive study of fungi on a site very difficult. Straatsma and

Krisai-Greilhuber (2003) found no clear relationship between the variation in

weather conditions and fungal diversity over the seven years of their study in

deciduous and coniferous forests in Switzerland. However, there was a positive

relationship between fungal productivity and the amount of rainfall in the July

prior to fruiting in autumn. They also found that relatively high temperatures in

November allowed for a prolonged fruiting season. Mild temperatures suit the

fruiting of macrofungi best, as Humphrey et al. (2000) found that fungal species

richness was positively correlated with accumulated temperature (AT; the number

of degree-days above 5ºC). Moisture deficit (MD) was also found to be correlated

(positively) to fungal species richness in their plots. MD is a measure of the

dryness of a site and is calculated as the maximum accumulated amount that

38

potential evapotranspiration exceeds precipitation. The finding that both AT and

MD are related to fungal diversity and fruiting makes them stand out as possible

standards for quantifying rainfall and ambient temperature. Other studies

(Heilmann-Clausen and Christensen 2005) have used another ratio of precipitation

to temperature called Lang’s index (annual precipitation/annual temperature; Lang

1915) and found it to be significantly related to fungal diversity.

39

Chapter 3: The sites: vegetation and site

variables measured

40

41

3.1 Introduction

3.1.1 Forests of Ireland: Extent and current trends

Forests currently cover over 650,000 ha (10%) of land in Ireland, with the figure

set to increase to 1.2 million ha (17%) by 2030 (Anon. 1996). Currently native

tree species comprise less than 25% of the forested area (National Forest

Inventory 2007), but plans have been put into place to decrease the planting of

exotic species and increase the planting of native broad leaves (Anon. 2009).

Compared to other European countries, Ireland is a deforested country. Just over

81,000 ha of Irish woodlands are classed as semi-natural (FRA 2010). This

equates to 1% of the total land area, the third lowest in Europe after Netherlands

and the Czech Republic but this is increasing at the eleventh fastest rate in the

European Union (FRA 2010).

Up to Neolithic times, Ireland was covered in dense forests of oak

(Quercus species), elm (Ulmus glabra), ash (Fraxinus excelsior), Scot’s pine

(Pinus sylvestris) and alder (Alnus glutinosa) (Mitchell 1995). Anthropogenic

processes associated with the expansion of farming, and climate change in the

Sub-Boreal Period have, over the past 4,500 years, brought about dramatic

changes in the abundance and extent of native forest types. Palynological records

have shown that elm underwent extreme reduction commencing 5000 years ago

and Scot’s pine had become extinct in Ireland by 2000 before present (B.P.)

(Cross 1998). Were it not for the influence of humans, the vegetation of Ireland

would be much different from the current status. Cross (2006) estimated that left

undisturbed, the majority of land area in Ireland would be covered in dense

pedunculate and sessile oak forests with a lesser component of ash, hazel and

alder woods. Present day land cover is vastly different from this hypothetical

situation. The four most abundantly planted tree species in Ireland are exotic

species (Table 3.1).

The majority of the forests planted for commercial gain in Ireland are of

the exotic conifer Sitka spruce (Picea sitchensis Bong Carr.). Sitka spruce is

particularly well suited to growth conditions in Ireland, and if properly managed

42

on fertile sites can reach a maximum annual mean increment of 30m3/ha/annum

(Horgan et al. 2004). This impressive growth rate of Sitka spruce in Ireland makes

it the most planted tree species, either in single species stands or as the

predominant species in a mixed stand (Joyce and O’Carroll 2002). This is

followed by lodgepole pine (Pinus contorta), which is also well suited to the wet

climate encountered in Ireland. Norway spruce (Picea abies) is tolerant to sites of

low nutrient status and grows well on moist sites, which makes it suitable for use

as the main tree species in a forest as well as in mixed stands with Sitka spruce.

Japanese larch (Larix kaempferi) is often planted as a nurse tree species with Sitka

spruce to help the spruce avoid the condition known as “check” (Horgan et al.

2004), which manifests itself as a yellowing of foliage and reduction in growth.

Ash is a fast-growing tree, the timber of which is valuable and in demand.

The trend in Ireland is very similar to that in Europe, with major

afforestation being carried out on former agricultural land (UNECE 2003). The

European temperate forests in general are dominated by deciduous broadleaf trees.

However, due to mass afforestation, this is changing to a more coniferous/mixed

forest composition due to plantation forestry (Millennium Ecosystem Assessment

2005).

Table 3.1 Approximate extent of native and plantation forest types in Ireland. Area (%) refers to the proportion of the Irish forests which are composed of this tree type (National Forest Inventory 2006).

Species Area (ha) Area (%)

Sitka spruce 327830 52.3

Lodgepole pine 46410 7.4

Norway spruce 25960 4.1

Japanese larch 21510 3.4

Ash 19200 3.1

Oak 14600 2.6

Scot’s pine 7400 1.2

The area of forested land in Europe increased by over 800,000 ha per year

between the 20th and 21st century (UNECE 2003). These plantation forests are

seen as one of the key drivers of fungal species loss in Europe, as many native

broadleaf forests (oak and beech) are being replaced by monocultures of spruce

and fir (Senn-Irlet et al. 2007; Dahlberg et al. 2009). In total, around 200,000 ha

43

(20%) of newly afforested land in Europe is planted with introduced species every

year, but this figure varies very much depending on the country in question

(UNECE 2003). Much of this afforestation is taking place on land previously used

for agriculture. In Ireland, 90% of current private afforestation is being carried out

by agricultural land owners (Teagasc 2005). In recent years the relationship

between biodiversity and forest productivity, health and other positive benefits of

biodiversity on ecosystem functioning have been identified (for review see Loreau

et al. 2007). This has brought about changes in forestry practices to promote

biodiversity in forests, and protect habitats which are rich in native biodiversity,

such as native Irish oak forests and semi-native beech forests (Anon. 2009).

3.1.2 Classification of Irish forest habitats

In order to meet the large scale changes in the composition of Irish forests from

native to non-native tree species, parallel changes in the classification systems

used for the vegetation of Irish forests are being researched. In the first major

classification system for vegetation in Irish habitats, all plantation coniferous

forests were grouped together under the classification Vaccinio –Piceetea (Br. –

Bl. Et Vlieger 1939; Kelly and Kirby 1982). The later classification systems for

Irish habitats by Fossitt (2000), again grouped the plantation forests under one

heading “conifer plantation WD4”. Recent work on the vegetation of Irish

plantation forests has significantly expanded this classification system, by

devising 8 distinct groups of vegetation composition in Irish plantation forests

(French et al. 2008).

The current classification system in use for Irish habitats is “A guide to

habitats in Ireland” by Fossitt (2000). A criticism of this system is that it groups

all coniferous woodlands into 4 sub-groups, regardless of the ground flora

vegetation of the sites (Roche et al. 2009; French et al. 2008). Earlier systems in

use in Ireland similarly grouped the coniferous forest into a small number of

groups (White and Doyle 1982), and others did not designate coniferous forests as

a habitat, because coniferous species are not native to Ireland (Perrin et al. 2008).

In recent years, there have been more focussed keys to Irish forest habitats

developed to complement the already established keys. In oak woods in Ireland,

Kelly and Kirby (1982), Kelly and Iremonger (1997) and Kelly (2005) have all

44

designed classification systems for their plant communities. Ash woodlands have

been the focus of a study by Kelly and Kirby (1982), with plantation ash forests

being examined by French et al. (2008). The study by French et al. (2008) also

examined the ground flora classification of Irish plantation Sitka spruce forests,

while Roche et al. (2009) examined Scot’s pine forests in Ireland with a view to

classifying their plant communities.

Previous research in forests of Scotland (Hawksworth 1991), Canada

(Villeneuve et al. 1989) and the U.S.A. (Gable and Gable 2007) have found a

significant relationship between the vascular plant communities and macrofungal

communities of temperate forests. With this in mind, it may be possible to

estimate macrofungal richness and diversity based on the vascular plant richness

and diversity in the forests.

3.1.3 Rational for site selection

To compare the different vegetational composition of Irish unmanaged and

plantation forests, four forest types were selected for study: ash woodlands (ash),

oak woodlands including both pedunculate and sessile oak (oak), Scot’s pine

woodlands (SP) and Sitka spruce woodlands (SS). In the interest of replication,

geographical spread, and the limitations of man-hours, it was decided to limit the

forest types to these four dominant tree species. Representative study sites were

chosen both from an existing list used in the BIOFOREST program (Smith et al.

2005, 2006; Iremonger et al. 2006), and also from known sites which fit the forest

types chosen for the study.

The rationale for choosing these four forest types was:

1. Broadleaved forest types comprising ectomycorrhizal tree species (such as

sessile and pedunculate oak) would be expected to support the greatest species

richness of native macrofungi. Estimates of extent and patterns of fungal diversity

from such forest types would provide a yardstick against which macrofungal

diversity estimates from plantation forests can be compared.

2. Ash is a native sub-climax tree that does not support ectomycorrhizal fungi.

The ash forest type was included because this forest type is found widely,

45

particularly in the midlands, and planting of ash has expanded greatly in recent

years.

3. Sitka spruce and Scot’s pine are non-native conifers, both of which support a

wide range of macrofungi, including ectomycorrhizal species, in their native

ranges. Sitka spruce is the most widely-planted plantation conifer in Ireland, and

assessing fungal diversity of these plantations was a principal objective of the

study. Scot’s pine plantations were included in the study because, while they are

of little commercial extent nowadays, they are anecdotally considered to support a

wider fungal diversity than other faster-maturing plantation forest types. If this is

true, it may be not least be because Scot’s pine plantations are considerably older

than other plantation types.

46

47

3.2 Aims of this chapter

The aims of this chapter are:

• To give the rationale for choosing the forest types to be examined in

further chapters.

• To give the rational for choosing which variables would be measured in

the selected forests.

• To describe the vegetation, structural and chemical characteristics of the

selected plots.

• To group the plots according to the measured variables, as a precursor to

the groupings based on fungal communities used in later chapters.

• To estimate macrofungal species richness using the recorded relationships

between vegetation species richness and macrofungal species richness.

48

3.3 Materials and methods

3.3.1 The sites

The locations of the sites and the corresponding site codes along with geographic

data and past land use data are presented in Fig. 3.1 and Table 3.2. One plot per

site was deemed sufficient, as each site was also represented by a replicate site

based on similarity of dominant tree species, vegetation cover and stand age as

suggested by Coll and Bolger (2007).

49

Fig 3.1 Distribution of study sites around Ireland. SS= Sitka spruce, SP= Scot’s pine. Shaded circles are oak sites, empty circles are ash sites, Squares are Sitka spruce sites and cross is Scot’s pine sites.

50

Table 3.2 Location and descriptions of the sites surveyed. Age abbreviations; Y = young, M= Mature. Vegetation classification follows the habitat descriptions by Fossitt (2000); WD1=highly modified broadleaf woodland, WN2=oak-ash-hazel woodland, WN1=oak-birch-holly woodland, WN7=Semi natural bog woodland, GS3= acid grasslands and WD4=conifer plantation. Past land use refers to the land use of the site previous to planting the current crop. CF= Cultivated Farmland, DW= Deciduous Woodland, BU= Bog or Uncultivated, MW= Mixed Woodland and CW= Coniferous Woodland. a =Tomies wood was used for two plots. Numbers in column 1 refer to the image of the site (Folder 1 Appendix 1).

Site name and

image no.

Site code Tree type GPS

coordinates

Age Vegetation

classification

Soil type Parent material Past land

use

Ballykillcavan (3.1)

Bkil Ash 53o01’N 07o07’W

Y WD1 Grey brown podzol

Limestone gravely till CF

Donadea park (3.2)

Donad Ash 53o20’ N 06o45’W

Y WN2 Gleys Limestone glacial till DW

Killough (3.3)

Killo Ash 52o36’N 07o50’W

M WN2 Gleys Alluvium DW

Ross island (3.4) Ross Ash 52o03’N 09o32’W

M WN2 Brown podzolic Sandstone DW

St Johns wood (3.5)

StJon Ash 53o33’N 08o00’W

M WN2 Brown earth and redzinas

Limestone till DW

Abbeyleix (3.6,3.7)

Abbey Oak (Quercus robur)

52o53’N 07o22’W

M WN2 Grey brown podzol

Stony Limestone glacial till

DW

Kilmacrea (3.8) Kilmac Oak (Q. petraea)

52o54’N 06o10’W

M WN1 Acid brown earths

Ordovician-Cambrian shale

DW

Raheen (3.9,3.10) Raheen Oak (Q. petraea)

52o53’N 08o31’W

M WN1 Grey brown podzol

Limestone glacial till DW

Tomies wooda

(3.11,3.12,3.13) Toomi Oak

(Q. petraea) 52o02’N 09o35’W

M WN1 Brown podzolic Sandstone DW

Union (3.14,3.15) Union Oak (Q. petraea)

54o12’N 08o29’W

M WN1 Grey brown podzol

Limestone glacial till DW

Annagh (3.16,3.17)

Anna Scot’s pine 53o06’N 07o58’W

Y WN7 Basin peat Basin peat BU

Ballygawley (3.18)

BgawSP Scot’s pine 54o12’N 08o27’W

Y WN7 Lithosol Granite and sandstone glacial till

BU

Ballylug (3.19,3.20)

Blug Scot’s pine 52o58’N 06o15’W

M GS3 Brown podzol

Ordovician-Cambrian shale

BU

Bansha (3.21) Bnsha Scot’s pine 52o27’N 08o06’W

M WN1 Podzols Sandstone, granite, mica

MW

51

Brittas (3.22,3.23) Britt Scot’s pine 53o09’N 07o32’W

M WN1 Peaty gley Mica, quartzite, sandstone

MW

Derryhogan (3.24,3.25)

Derry Scot’s pine 52o38’N 7o41’W

Y WN7 Basin peat Basin peat BU

Gortnagowna (3.26,3.27)

Gort Scot’s pine 52o52’N 07o47’W

Y WN7 Brown podzolic Sandstone BU

Torc (3.28,3.29) Torc Scot’s pine 52o00’N 09o31’W

M WN1 Peaty podzol

Granite and sandstone MW

Ballygawley (3.30,3.31)

BgawSS Sitka spruce 54o12’N 08o28’W

M WD4 Lithosol

Granite and sandstone glacial till

DW

Bohatch (3.32,3.33)

Bohat Sitka spruce 52o57’N 08o27’W

M WD4 Blanket peat Basin peat CW

Chevy mature (3.34,3.35)

ChevM Sitka spruce 53o02’N 08o41’W

M WD4 Gleys Sandstone MW

Chevy young (3.36,3.37)

ChevY Sitka spruce 53o02’N 08o41’W

Y WD4 Gleys Sandstone CF

Cloonagh (3.38) Cloon Sitka spruce 54o10’N 08o21’W

Y WD4 Grey brown podzol

Limestone glacial till BU

Dooary (3.39) Dooar Sitka spruce 52o56’N 07o15’W

Y WD4 Gleys Carboniferous shale CF

Moneyteige (3.40,3.41)

Money Sitka spruce 52o48’N 06o18’W

M WD4 Peaty podzol

Granite and sandstone BU

Quitrent (3.42,3.43)

Quitr Sitka spruce 52o16’N 08o27’W

Y WD4 Gleys Sandstone glacial till BU

Stanahely (3.44,3.45)

Swmid Sitka spruce 53o00’N 06o31’W

Y WD4 Peaty podzol

Granite and sandstone BU

52

3.3.2 Data collection methods

In order to link the diversity and functional diversity of macrofungi in these

forests, to the physical characteristics of the forests, a number of established

procedures were carried out in all of the forests surveyed.

Measurement of vascular plant species richness and canopy cover

The number of different tree, and plant species present in the plots was recorded

using the vertical cover index described in Ferris-Kaan et al. (1998) and tested in

Humphrey et al. (2003).

Table 3.3 Division of strata used in vertical cover index.

Code Description Height (meters)

S1 Field layer 0.1-2

S2 Shrub layer 2-5

S3 Lower canopy 5-15

S4 Upper canopy 15-20

The forest is divided into four strata (Table 3.3) and the percentage cover of each

plant/tree species is recorded in each of the strata. Open space is also recorded.

The method used was:

1. An arc using 3 pieces of string 10m in length was set up. This gave the

left, right and straight extent of the arc (Fig 3.2).

2. Starting with the field layer and working up, the identity and percentage

cover of each of the plant species that bisected the arc was recorded.

3. A number of estimations were taken for each hectare of forest. It was

recommended that 2 facing north of the plot and 2 facing south of the plot

were taken at 10m distance from each other.

The Cover index (CI) was calculated as:

CI= 1.9S1 + 3S2 +10 S3 + 5 S4

S1= % cover in the field layer S2= % cover in the shrub layer S3= % cover in the understory layer S4= % cover in the upper canopy

53

Fig.3.2 Schematic of the visual arc from observer, which was used in defining the structural characteristics of the sites according to Ferris-Kaan et al. (1998).

Measurement of canopy closure

The measurement of canopy closure using the canopy scope (Brown et al. 2000)

involved taking 8 random readings for canopy openness in the site (See colour

plate 1). The readings were taken by holding the canopy scope towards the largest

canopy gap and counting the number of points which fell inside the canopy gap.

The readings were averaged for the value for the entire site.

Coarse woody debris quantification

Coarse woody debris CWD was measured using the line intersect method (Kirby

et al. 1998). A 50m transect was laid down and all pieces of CWD which bisected

this transect had their diameter and decay stage recorded. The volume of CWD in

each decay stage was calculated using the formula:

V= nd2 π2 104/8t

V= volume of deadwood in m3 per Hectare n= number of times diameter class crossed line d= Diameter class of deadwood at point of intersection in meters t= Length of transect line in meters.

The diameter at breast height was measured for a random sample of trees

within each site following the method in Newton (2007). Stocking density as trees

per 100m2 was also recorded.

Soil variables measured

The litter layer was removed from the soil samples before soil analyses were

carried out. The pH of the sites was determined using a pH meter in a 5:1 H2O:

54

soil dilution (Anon. 2006b). Moisture content was calculated as the loss in mass

after 2 day at 105ºC (Anon. 2006b). Organic matter was calculated as the loss on

ignition after 1 day at 450ºC (Anon. 2006b). Bedrock description was estimated

from the use of geological maps of Ireland (Gardener and Radford 1980).

PRSTM-probes system (Plant Root Simulator probes; Western Ag

Innovations Inc., Saskatoon, Sask.) were used to measure the in-situ supply of

anions and cations through the soil in the sites (Colour plate 2). Four pairs of

probes (four cations and four anion) were placed in the soil in each site during the

sampling period (September to November 2008; and September to October 2009)

and a trench sliced around the probe to reduce competition from plant roots with

the probes. The probes were left for ca. 1 month and were then removed, washed

with de-ionised water and stored at 4ºC until they were returned to Canada for

analysis. The past land use category of the site was determined from the forester

and also from Ordnance Survey Ireland historic maps. Data for the meteorological

records for the seven weather stations closest to the sites was acquired from Met

Eireann (www.met.ie). The seven stations that were closest to the sites were

Birr/Oak park (Co. Offaly), Casement (Co. Dublin), Claremorris (Co. Mayo),

Cork airport (Co. Cork), Kilkenny (Co. Kilkenny), Shannon (Co. Clace) and

Valentia (Co. Kerry).

3.3.3 Statistical analysis

All environmental variables were tested for significant differences between the

forest types using a Mann-Whitney U test. Box plots were created to show the

distribution of the plots within a measured variable. Spearman’s rank correlation

(rs) was used to check for relationships between ordination axes and other

variables measured. The rs value is a measure of how strongly correlated two

variables are. Stepwise linear regression was used to analyse the relationships

between environmental variables and plant and tree species richness.

Simpson’s and Shannon’s diversity indices were calculated using the

program EstimateS (Colwell 2004) using similar settings as those used for the

macrofungal data in Section 4.3.5, Chapter 4.

The effects of site nominal variables (age, rotation stage etc.) on the

numbers of plant species in each forest type separately were tested using a

55

generalized linear model with Poisson error distribution and log-link function (for

analysis of non-normal data). For ash, the effect of site management (semi-

natural/managed) and forest age (young/mature) was tested for relationships with

species per plot visit. For oak, the effect of grazing large mammals (grazed/non-

grazed) on the numbers of species per plot visit. For Scot’s pine, the effect of

rotation stage (first/second rotation) and forest age (young/mature) was tested for

relationships with species per plot visit. For Sitka spruce forests, the effect of

rotation stage (first/second rotation) and forest age (young/mature) was tested for

relationships with species per plot visit.

Multivariate analysis

To analyse the plant species similarity between the different sites, the abundance-

based Jaccard similarity index (Chao et al. 2005) as calculated by Estimate-S

(Colwell 2004) was used. This index is a modified form of the Jaccard

presence/absence index which uses quantitative data and also estimates the

number of unseen taxa in the sample, thus combining a species richness estimator

and a similarity index.

Nonmetric multidimensional scaling (NMS) was used to compare

composition of plant vegetation communities the forest types. NMS is effective

with ecological data because it does not assume linearity of species’ responses to

gradients (McCune and Grace 2002). NMS uses rank order information in a

dissimilarity matrix that eliminates the ‘‘zero truncation’’ problem in most

ordination methods, and can use any distance measure. NMS was run on two data

sets in this chapter.

These were:

(a) Using the environmental variables as the main matrix (producing a 28 x 18

plot by variable matrix; Table 3.6) to examine if the plots grouped into their

respective forest types.

(b) Using the vascular plant presence/absence data as the main matrix (producing

a 28 x 68 plot by species matrix; Table 3.7), and the environmental variables

as the second matrix to examine if the vascular plant communities of the

56

forest types were useful in separating the plots into their respective forest

types.

The data was analysed using PC-Ord version 4.36 (MjM Software Design,

Gleneden Beach, OR). A Sørensen distance measure was used for all the data sets.

The Sørensen distance measure is an intuitively meaningful measure as it takes

into account the abundance of the shared species between two groups divided by

the abundances of species found in each sample (McCune and Grace 2002). A

random seed started each analysis and included 40 runs of real data and 50 runs of

randomized data for use in a Monte Carlo permutation procedure (McCune and

Grace 2002). The Monte Carlo permutation procedure is a randomization test that

determines if NMS is generating stronger axes than expected by chance. Results

of the Monte Carlo test and examination of the stress in a scree plot were used to

determine dimensionality. Overlays and correlations with axes were run in PC-

Ord to interpret the data.

To test if the sites were grouped more closely based on their plant

communities according to the forest type they were defined as, a multi-response

permutation procedure (MRPP), a non-parametric hypothesis test for multivariate

differences between groups was carried out in PC-ORD. MRPP constructs a

distance matrix, calculates average within-group distances, and compares these to

a Pearson’s type III continuous distribution of all possible partitions of the data

(Peck 2003). Groups were analyzed by forest type using PC-Ord. The Sørensen

distance measure was used for all analyses. With MRPP, the chance-corrected

within-group agreement (A) statistic describes effect size. Values for A range

from -1 to +1. When A = -1, there is less agreement between groups than expected

by chance. When A =0, groups are no more or less different than expected by

chance, and when A = 1, groups are identical. The more positive A is, the more

homogeneous groups are and the greater confidence in the p-value, especially

when the sample size is small (Peck 2003). The grouping variable forest type (ash,

oak, Scot’s pine and Sitka spruce) was used.

Indicator species analysis was used to identify plant species which are

indicative of a certain forest type. The method used was designed by Dufrene and

Legendre (1997). Indicator species analysis uses data on the concentration of

species abundances in a particular user set group (e.g. forest type), and also the

57

faithfulness of occurrence of a species to a particular group (McCune and Grace

2002). To identify plant species which are indicative of the habitat examined in

this project, Indicator values (IV) for each species were calculated in PC-ORD

using the method by Dufrene and Legendre (1997) with statistical significance

calculated by a Monte Carlo test with 1000 runs. Indicator values (IV) range from

0 (no indication) to 100 (perfect indication). Perfect indication means that

presence of a certain species points to a particular group without error (McCune

and Grace 2002).

58

3.4 Results

3.4.1 Soils physical and chemical attributes.

Soil physical attributes

The soils types of the 28 survey plots (Table 3.4) were classified according to

Gardener and Radford (1980). The principal soil type is the soil type which

makes up the majority of the soils in the area whilst the secondary soil type only

consists of minor proportions of the soil constituents.

Table 3.4 Survey plots and their principal and associated soil types along with the parent material of the soil. Also listed are the pH, organic matter (OM) and moisture content of the soils in the sites. Values in parenthesis are the percentages of the associations which make up the soil of the site. SP= Scot’s pine, SS= Sitka spruce.

Forest

type

Plot pH OM

(%)

Moisture

content

(%)

Principal

soils

Associated

soil

Parent

material

ASH Ballykillcavan 6.12 6.2 38 Grey brown podzol (80)

Gleys (20) Limestone gravely till

ASH Ross island 6.24 3.95 3.95 Brown podzolic (60)

Acid brown earths (20), Gley (20)

Sandstone

ASH Donadea park 6.91 6.4 2.7 Gleys (90)

Grey brown podzol (10)

Limestone glacial till

ASH St Johns wood 5.78 7.3 16.7 Brown earth and rendzinas (60)

Grey brown podzol (25),gleys (10), peat (5)

Limestone till

ASH Killough 6.54 7.9 23.9 Gleys (60)

Brown earths (20), peaty gleys (20)

Alluvium

OAK Raheen 4.83 24.5 33.3 Grey brown podzol (60)

Gleys (20), interdrumlin peat and peaty gley (20)

Limestone glacial till

OAK TomiesA 5.56 12.80 27.55 Brown podzolic (60)

Acid brown earths (20), Gley (20)

Sandstone

OAK TomiesB 5.86 35.80 43.67 Brown podzolic (60)

Acid brown earths (20), Gley (20)

Sandstone

59

OAK Abbeyleix 5.67 29.3 33.4 Grey brown podzol (80)

Gleys (10), Brown earths (10)

Stony Limestone glacial till

OAK Union oak 7.11 25.4 32.6 Grey brown podzol (60)

Gleys (20), interdrumlin peat and peaty gley (20)

Limestone glacial till

OAK Kilmacrea 5.01 31.3 19.4 Acid brown earths (75)

Gleys (15) Brown podzol )

Ordovician-Cambrian shale

SP Derryhogan 4.58 22.2 58.49 Basin peat

-- --

SP Gortnagowna 5.57 21.3 46.16 Brown podzolic (60)

Acid brown earth (20), gley (20)

Sandstone

SP Annagh 4.86 18.2 45.48 Basin peat

-- --

SP Ballygawley 5.14 20.2 17.32 Lithosol (80)

Rock outcrop and peat (20)

Granite and sandstone glacial till

SP Ballylug 5.11 24.8 24.9 Brown podzol (80)

Gleys (15), podzo (5)

Ordovician-Cambrian shale

SP Brittas 4.45 6.78 17.07 Peaty gley (70)

Blanket peat (20), peaty podzol (10)

Mica, quartzite, sandstone

SP Torc 4.55 25.4 68.38 Peaty podzol (75)

Lithosol (15), blanket peat (10)

Granite and sandstone

SP Bansha 4.41 46.40 31.92 Podzols (70)

Gleys (20), peat (10)

Sandstone, granite, mica

SS Stanahely west

5.29 8.32 25.4 Peaty podzol (75)

Lithosol (15), blanket peat (10)

Granite and sandstone

SS Cloonagh 5.76 11.6 66.19 Grey brown podzol (60)

Gleys (20), interdrumlin peat and peaty gley (20)

Limestone glacial till

SS Dooary 7.16 6.77 27.5 Gleys (75)

Acid brown earth (15), peat (10)

Carboniferous shale

SS Moneyteige nr nr nr Peaty podzol (75)

Lithosol (15), blanket peat (10)

Granite and sandstone

SS Bohatch 6.72 7.01 57.3 Blanket peat (high level)

-- --

SS Quitrent 4.46 7.00 4.16 Gleys (75)

Peaty gleys (25)

Sandstone glacial till

SS Chevy chase young

6.59 6.32 4.9 Gleys (75)

Peaty gleys (25)

Sandstone

60

SS BallyGawley 5.11 10.3 58.51 Lithosol (80)

Rock outcrop and peat (20)

Granite and sandstone glacial till

SS Chevy chase mature

6.63 7.42 13.76 Gleys (75)

Peaty gleys (25)

Sandstone

The medians for the pH, organic matter and moisture content of the forest types

were calculated (Table 3.5). Using a Mann-Whitney U test significant differences

were identified between the different forest types and highlighted in high-low

plots (Figs 3.3a,b,c).

Table 3.5 Summary statistics for the variables pH, organic matter (OM) and moisture content (moisture) from the different forest types. Numbers are median values. SP= Scot’s pine, SS= Sitka spruce.

Forest type pH OM (%) Moisture (%)

ASH 6.24 6.4 16.7

OAK 5.56 25.4 33.3

SP 4.58 23.5 31.92

SS 6.59 7.01 25.4

The pH of the forest type varies between the individual sites which make

up the forest type, especially in the case of the Sitka spruce forest type (Fig 3.3a).

The greatest variation is seen in the Sitka spruce sites, which has a range of 2.7

pH units, followed by oak, Scot’s pine and ash with ranges of 2.28, 1.6 and 1.13

respectively. The medians of pH for all forest types were significantly different

except for ash and Sitka spruce. Scot’s pine forests have a consistently low pH,

while ash forests normally have a high pH.

The median organic matter content was highest in oak plots, followed by

Scot’s pine, Sitka spruce and ash plots (Fig. 3.3b). It was found that all of the

forest types differed in the amount of soil organic matter except ash and Sitka

spruce, which had extremely low organic matter levels.

The soil moisture content of the different forest types can be seen to vary a

lot depending on the individual sites in that forest type. In Sitka spruce sites, the

soil moisture content values had a range over 62 percent points, with Scot’s pine

sites having a soil moisture content range of 51, followed by ash and oak with

moisture content ranges of 35 and 24 percent points (Fig. 3.3c). Ash forest soils

61

were found to have significantly lower moisture content than the other forest

types.

Soil chemical results

The anion and cation supply rate to the PRS probes in the soil over 30 days was

calculated (Table 3.6). Significant differences between the median values for soil

available nitrate nitrogen, ammonium nitrogen, calcium magnesium, potassium

and phosphorus in each of the forest types were found using a Mann-Whitney U

test (Figs 3.4a, b; 3.5a, b; 3.6a, b).

62

(a)

(b)

(c)

Fig. 3.3 a, b, c Median (circles) with 95% error confidence intervals (high-low whiskers) of the data for (a) pH: All medians are significantly different at P<0.05 except ash and Sitka spruce (b) organic matter %: All medians are significantly different at P<0.05 except ash and Sitka spruce and (c) moisture content of the soil from the forest types: the median value for ash was found to be significantly different (P<0.05) than all of the other forest types.

63

64

Table 3.6 Results from the PRS analysis of the plots. All values are the nutrient supply rate measured in µg/10cm2/month. SP = Scot’s pine, SS= Sitka spruce. * denotes samples where the actual amount of nutrient was below the minimum analysis capabilities of the PRS probes and so is not a reliable reading.

Tree

type

Plot Total

N

NO3-

N

NH4-

N

Ca Mg K P Fe Mn Cu Zn B S Pb Al Cd

ASH Ballykilcavan 430.4 430.4 0* 3758.0 79.6 8.6 1.8 132.6 3.8 7.2 2.8 1.6 85.2 7.2 30.8 0.2*

ASH Donadea 107.6 107.6 0* 3850.0 72.6 11.0 0.8 30.2 3.0 2.8 4.4 1.6 72.8 3.8 45.6 0.4

ASH St John 113.4 113.2 0.2* 3338.0 175.6 11.8 0.4 125.8 3.6 3.8 1.4 2.6 209.8 8.2 32.8 0*

ASH Ross Island 112.4 105.2 7.2 2105.3 387.3 15.1 0.2* 96.8 28.5 64.9 4.7 2.3 137.4 5.8 38.3 0*

ASH Killough 71.8 71.8 0* 3588.0 75.0 18.0 0.6 25.6 5.2 0.8 3.8 2.2 142.6 2.8 32.0 0.2*

OAK Abbeyleix 386.4 361.8 24.6 770.0 126.8 607.6 3.4 34.4 236.0 0.4 1.4 0.6 52.0 10.4 48.6 0*

OAK Kilmacrea 340.8 280.2 60.6 389.0 198.0 290.8 0.6 39.8 147.8 0.2* 3.6 1.0 24.6 4.2 86.2 0*

OAK TomiesA 54.6 0.4* 54.2 440.6 623.0 685.8 0.8 3.0 159.2 0.2* 3.4 0.2* 45.4 0.6 11.4 0*

OAK TomiesB 36.0 16.2 19.8 566.2 585.4 96.0 0.8 18.8 489.8 0.8 5.2 1.4 397.6 2.0 54.0 0*

OAK Union 350.2 340.2 10.0 1668.2 466.8 94.4 0.6 28.0 87.8 0.6 2.8 1.0 104.8 1.8 32.6 0*

OAK Raheen 85.2 0* 85.2 451.0 432.0 1273.2 7.2 2.4 108.6 0.2* 3.2 1.2 130.4 0.8 21.8 0*

SP Ballylug 33.2 0.8* 32.4 133.8 131.4 391.4 0.4 17.6 82.2 0.2* 2.0 0.6 34.4 1.8 63.0 0.2*

SP Torc 7.8 0.2* 7.6 479.4 649.8 57.2 0.2* 30.8 68.8 0.4 1.2 1.0 313.6 1.0 53.4 0*

SP Bansha 10.6 0.8* 9.8 276.2 276.8 214.2 0.4 3.0 76.8 0.2* 2.2 0.8 116.4 5.4 19.6 0*

SP BallygawleySP 5.4 0.8* 4.6 117.8 294.8 104.2 0.2* 18.8 1.4 0.4 0.8 0.8 56.4 1.2 40.8 0*

SP Derryhogan 17.2 15.0 2.2 698.6 747.8 40.6 0.2* 3.8 0.4 0.2* 0.6 1.4 52.0 0.8 36.8 0*

SP Annagh 231.8 109.6 122.2 575.4 220.4 138.8 0.4 11.0 23.2 0.2* 0.6 2.4 75.4 2.2 38.2 0*

SP Brittas 10.7 5.1 5.6 162.2 113.6 90.3 0.9 3.3 74.0 0.2* 0.7 2.8 30.1 0.5 47.1 0*

SP Gortnagowna 26.2 21.8 4.4 1379.8 697.0 24.8 0.4 7.2 12.4 0.4 1.2 1.4 33.8 0.4 47.2 0*

SS Dooary 92.2 81.6 10.6 2642.0 185.4 7.8 0.8 268.6 87.0 1.8 0.8 1.0 196.2 4.6 61.4 0*

SS Quitrent 15.6 0.6* 15.0 199.2 180.8 132.2 1.8 127.0 1.4 0.2* 2.4 2.2 265.8 0.4 103.2 0*

SS Moneyteige 240.8 220.2 20.6 117.4 135.8 127.4 0.6 82.8 54.4 11.8 1.8 0.4 81.4 3.4 113.2 0*

SS Stanahely 153.6 89.0 64.6 133.2 96.4 228.6 0.4 32.2 9.6 0.2* 2.0 0.6 188.8 1.8 110.6 0*

SS Cloonagh 238.6 231.2 7.4 2144.0 238.0 16.2 10.4 223.8 44.8 0.8 4.6 1.2 223.6 1.4 29.6 0*

65

SS Ballygawley 13.6 0* 13.6 640.6 576.6 96.0 0.6 7.2 5.6 0.2* 1.0 0.8 269.8 1.6 96.2 0.2*

SS Bohatch 114.4 16.4 98.0 248.0 250.0 71.8 1.8 11.6 24.0 0* 1.0 0.6 159.0 1.0 28.6 0*

SS Chevy chase M

30.0 16.6 13.4 830.8 274.8 71.8 2.8 19.0 30.6 0.6 2.2 2.8 24.0 3.6 47.8 0*

SS Chevy chase Y 22.8 13.8 9.0 876.8 199.6 180.6 0.6 20.8 13.4 0.2* 1.0 1.6 31.4 6.0 41.2 0*

66

(a)

(b)

(c)

(d)

(e)

(f)

Fig 3.4 a, b, c, d, e, f. Median (circles) with high–low values (whiskers) of the data for soil available (a) nitrate nitrogen, (b) ammonium nitrogen, (c) calcium, (d) magnesium, (e) potassium and (f) phosphorus in the different forest types. Medians which do not share letters are significantly different (P<0.05) according to a Mann-Whitney U test.

Soil nitrogen levels can be seen to vary a lot between the different forest

types (Table 3.6). One visible trend is the larger amounts of nitrate nitrogen

67

available in the broadleaf sites in comparison to the nitrate nitrogen available in

the coniferous forests (Fig 3.4a). Calcium levels are much higher in the ash sites

than they are in the other forest types (Fig. 3.4c). Magnesium levels show a very

wide range of values within the oak and Scot’s pine forest types, while there is

less of a spread in the values for magnesium in the ash and Sitka spruce forest

types (Fig. 3.4d). Potassium levels are rather low in all of the forest types except

the oak sites which show a wide spread of values for potassium depending on the

site examined (Fig. 3.4e).

3.4.2 Relationship between plots based on soil variables measured

A NMS ordination was carried out using the environmental variables (pH, %

organic matter, nitrate nitrogen, ammonium nitrogen, calcium, magnesium,

potassium, phosphorus) to ordinate the plots. Using the Sørensen distance

measure and 40 runs with real data and 50 runs with random data, a 2-dimensional

ordination (Fig. 3.5) was found to be the best solution by visually inspecting the

NMS scree plot. The final stress reported was 10.0 which indicated a fair solution

according to Kruskal’s rule of thumb (McCune and Grace 2002). A Monte Carlo

test showed that the probability of a similar final stress being obtained by chance

was P= 0.00001. Axis 1 represented 12.6% of the variation between the distances

in the original and ordination space while axis 2 represented 77% of the variation.

Axis 1 was found to be strongly correlated (according to Pearson’s correlation

coefficient rP) with annual precipitation (r= -0.647) and amounts of soil available

nitrogen (r= 0.704) and magnesium (r= -0.764). Axis 2 was correlated with pH

(r= 0.615) and soil available calcium (r= 0.935).

The resulting ordination (Fig. 3.5) separates the sites based on their forest

type. The ash sites come out as a separate group in the ordination, mostly due to

the high amounts of soil available calcium in the ash sites and high pH values.

68

69

Fig. 3.5 Non metric multidimensional scaling of axis 1 and 2 (r2= 0.896) showing the sites identified by the tree type of the site and ordinated using the environmental variables measured (Table 3.4; 3.6). r2 of axes: 1= 0.126, 2= 0.77. Blue squares= ash sites, green triangles= oak sites, red triangles= Scot’s pine sites and pink diamonds= Sitka spruce sites.

3.4.3 The species list

The vascular plant species list was created for each plot (Table 3.7).

70

71

Table 3.7 Vegetation species list from the sites. Numbers are presence/absence data. The column “Type” refers to the vegetation group of the species. H= herb, F= fern, T= tree, and M= moss.

Ash Oak Scot’s pine Sitka spruce

Species Typ

e

BK

IL

DO

NA

D

KIL

LO

RO

SS

ST

JON

AB

BE

Y

KIL

MA

C

RA

HE

E

TO

MIE

A

TO

MIE

B

UN

ION

AN

NA

BG

AW

SP

BL

UG

BN

SH

A

BR

ITT

DE

RR

Y

GO

RT

TO

RC

BG

AW

SS

BO

HA

T

CH

EV

M

CH

EV

Y

CL

OO

N

DO

OA

R

MO

NE

Y

QU

ITR

SW

MID

Arum

maculatum H 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Athyrium

felix femina F 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 1 0

Betula

pubescens T 0 0 0 0 0 1 0 0 0 1 0 1 0 0 0 0 1 1 1 0 1 0 1 0 0 0 0 0

Blechnum

spicant F 1 0 0 0 0 0 0 1 0 1 0 0 0 0 0 1 1 0 0 1 0 1 1 0 0 0 1 0

Calluna

vulgaris H 0 0 0 0 0 0 0 0 0 0 0 1 0 0 1 0 1 0 1 0 0 0 0 0 0 0 1 0

Carex

sylvatica H 0 0 0 0 0 1 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Carex

sylvatica H 0 1 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Chrysospleni

um

oppositifoliu

m

H 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Circaea

lutetiana H 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Climacium

dendroides M 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Climacium

sp. M 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0

Corylus

avellana T 0 0 1 0 1 0 1 1 0 1 1 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0

Crataegus

monogyna T 0 1 0 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Deschampsia

flexuosa H 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0

Dicranium

magus M 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0

Digitalis

purpurea H 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 1 0 0 0

Dryopteris

carthusiana F 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0

Dryopteris

dilatata F 0 1 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0

72

Dryopteris

filix-mas F 0 1 1 1 1 0 1 1 1 0 0 0 0 0 0 1 0 1 0 0 0 0 1 0 0 0 0 0

Dryopteris

psued- mas F 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Epilobium

sp. H 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0

Helleborine

sp. H 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Euonymus

europaeus H 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Fagus

sylvatica T 0 1 0 1 0 1 0 1 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Festuca H 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0

Fragaria

vesca H 0 0 1 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Fraxinus

excelsior T 1 1 1 1 1 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Geum

urbanum H 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Glechoma

hederacea H 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Hedera helix H 1 1 1 1 1 1 0 1 1 1 1 0 0 0 0 0 1 0 1 1 0 1 1 0 0 0 0 0

Heracleum

sphondylium H 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Hyacinthoide

s non-scripta H 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Ilex

aquifolium H 0 0 0 0 1 0 1 1 1 1 0 1 0 1 1 1 0 0 1 1 0 0 0 0 0 0 0 0

Larix

decidua T 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Lonicera

periclymenu

m

H 0 0 1 0 0 0 1 1 0 1 0 0 0 0 1 0 0 1 0 0 0 0 0 0 0 0 0 0

Luzula

sylvatica H 0 0 0 0 0 0 1 0 1 1 1 0 0 0 0 1 0 0 0 1 0 0 0 0 0 0 0 0

Oxalis

acetosella H 0 0 0 0 0 1 1 1 1 1 0 0 1 1 0 1 0 0 0 1 1 1 0 0 0 0 0 1

Phyllitis

scolopendriu

m

F 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Picea

sitchensis T 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 1 1 1 1 1 1 1 1 1

Pinus

contorta T 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0

Pinus

sylvestris T 0 0 0 0 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 0 0 0 0 0 0 0 0 0

Polystichum

setiferum M 0 0 1 0 1 0 0 1 1 1 0 1 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0

73

Polytrichum

sp. M 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0

Polytrichum

commune M 0 0 0 0 0 0 0 0 0 0 1 1 1 0 0 0 0 0 0 1 0 0 0 0 0 0 1 1

Potentilla

sterilis H 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Pseudotsuga

menziesii T 0 1 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0

Pteridium

aquilinum F 0 0 0 0 0 1 0 0 0 0 1 1 1 1 1 1 0 1 0 1 0 1 0 0 0 0 0 0

Quercus

petraea T 0 0 0 1 0 0 1 1 1 1 0 0 0 0 1 1 0 0 0 1 0 0 0 0 0 0 0 0

Quercus

robur T 0 0 0 0 0 1 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Rhododendro

n ponticum H 0 0 0 0 0 0 0 0 0 0 0 0 1 0 1 0 0 0 1 1 0 0 0 0 0 0 0 0

Rhytidiadelp

his sp. M 0 0 0 0 1 0 0 0 1 0 0 0 0 0 1 1 0 0 0 0 0 0 0 0 0 0 1 0

Rubus

fructicosus H 0 1 1 1 1 1 1 1 0 1 0 1 1 1 1 1 1 1 0 0 1 0 0 0 1 0 0 0

Rumex H 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Salix H 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 0 0 0 0 0 0 0 0 0 0

Sanicula

europaea H 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0

Sorbus

aucuparia T 0 0 0 0 0 0 1 0 0 1 0 0 0 0 1 1 1 1 0 0 1 0 1 0 0 0 0 0

Taraxacum H 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Taxus

baccata T 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0

Thuidium

tamariscinum M 0 0 0 0 1 0 1 0 1 1 1 1 1 1 1 1 1 1 1 1 1 0 1 1 1 0 0 1

Ulex

europaeus H 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0

Urtica dioica H 1 1 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0

Vaccinium

myrtillus H 0 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 0 0 1 0 0 0 0 0 1 0

Veronica

montana H 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Vibernum

opilus H 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Vicia sepium H 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0

Viola

riviniana H 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

74

A total of 68 vascular species were recorded in the 28 forest plots. The five

most common species in this project were Rubus fructicosus, Thuidium

tamariscinum, Hedera helix, Dryopteris filix-mas and Ilex aquifolium. All of these

species were found in each forest type in this study. There were a total of 26

species which were only found in one plot. Of these species there are no

nationally rare species present.

For comparison between the four forest types, data on total species

richness and diversity indices for each forest type was calculated (Table 3.8; Fig.

3.6). The large number of tree species in oak and Scot’s pine forests shows that

these are forest types that often share the canopy with other tree species in natural

ecosystems. Sitka spruce on the other hand, is a light demanding species, and

forms a tight closed canopy which does not allow many other tree species to

compete for light. In total, Scot’s pine and ash forests had the largest number of

vascular species present. This was followed by oak and Sitka spruce (Fig. 3.6).

Table 3.8 Number of tree and plant species present in the forest types. Also listed is the complement Simpson’s diversity index (1-D) and the Shannon’s diversity index (H’) with standard deviations (in parenthesis). SS= Sitka spruce, SP= Scot’s pine. Tree type Tree species other plants Simpson’s Shannon’s

Ash 6 32 0.898±0.047 2.376±0.482

Oak 8 23 0.912±0.015 2.44±0.18

SP 8 30 0.904±0.025 2.38±0.296

SS 5 19 0.654±0.375 1.482±0.905

75

ASH OAK SP SS0

10

20

30

40Tree speciesOther plants

Forest type

No

. o

f sp

ecie

s

Fig. 3.6 Numbers of tree and other plant species found in the different forest types. SP= Scot’s pine, SS= Sitka spruce. Dark bar= other plants, light bar = tree species.

3.4.4 The vegetation and classification of the individual sites

Ash sites

Certain trends in the ash plots are evident relating plant species richness to plot

structural variables (Table 3.9). The two most species-poor sites at Ballykilcavan

and Ross Island are both young managed ash forests. Donadea, Killough and St

John’s Wood were more species-rich then the previously named sites. Through

forest management for timber production, Donadea now has an upper canopy

composed totally of ash. The lack of other tree species in the canopy has allowed

fast spreading plant species like bramble and ivy to cover a large proportion of the

forest floor. Killough was the most species-rich ash site and it also has the highest

Simpson’s diversity index; it could be described as an old-unmanaged ash site.

The presence of an orchid species (Helleborine sp.) is an indicator that this site

has experienced little disturbance in the past. St John’s Wood is classed as one of

the most ancient woodlands in Ireland (Rackham 1995).

76

Table 3.9 List of ash sites with the tree and plant species richness along with the Simpson’s diversity index (1-D) and the Shannon’s diversity index (H’).

Site Tree species Other

plants

Simpson’s

diversity

Shannon’s

diversity

Ballykilcavan 1 5 0.83 1.79

Donadea 4 11 0.93 2.71

Killough 2 18 0.95 3.0

Ross Island 4 4 0.88 2.08

St John’s Wood 3 7 0.9 2.30

All ash sites 6 32 0.9 2.38

The ash sites at Ballykilcavan, Ross Island and Donadea are planted

woodland and are classified as mixed broadleaved woodland (WD1) in the Irish

Habitats Classification System (Fossitt 2000). The ash sites at St John’s Wood

and Killough are semi-natural woodlands and are of the type oak-ash-hazel

woodland WN2 in the Irish habitats classification system (Fossitt 2000).

Phytosociologically, St John’s, Ballykilcavan, Donadea and Ross Island all fall

into the group Carici remotae- Fraxinetum Koch. 1926. sub association typicum

(Kelly and Kirby 1982). The only ash site which differs from the previous

grouping is the woods at Killough which closely match the Carici remotae-

Fraxinetum Koch. 1926 sub-association veronicetosum (Kelly and Kirby 1982)

grouping.

Oak sites

The oak sites were quite similar in species richness and diversity index values

with a few exceptions (Table 3.10). Abbeyleix woodland is the only pedunculate

oak site (Quercus robur). All of the rest of the sites are dominated by sessile oak

(Quercus petraea). Of all the sites, Tomies A is the most species poor in all of the

different structural layers (Table 3.10). The effect of grazing exclusion at the

Tomies B site is evident in the increased Simpsons and Shannon’s diversity index

value over the unfenced site at Tomies A.

77

Table 3.10 List of oak sites with the tree and plant species richness along with the Simpson’s diversity index (1-D) and the Shannon’s diversity index (H’).

Site Tree species Other plants Simpson’s

diversity

Shannon’s

diversity

Abbeyleix 4 9 0.92 2.49

Kilmacrea 4 8 0.92 2.49

Raheen 3 9 0.92 2.49

Union 4 6 0.9 2.30

Tomies A 1 8 0.89 2.2

Tomies B 4 10 0.93 2.71

All oak sites 8 23 0.91 2.44

In phytosociological terms, the Tomies sites belong to the Blechno-

Quercetum Association (Braun-Blanquet and Tuxen 1952) subassociation

scapanietosum (Kelly and Moore 1975). The woods at Kilmacrea belong to the

Blechno-Quercetum Association (Braun-Blanquet and Tuxen 1952)

subassociation typicum (Kelly and Moore 1975). Both Raheen and Union oak fit

into the Blechno-Quercetum Association subassociation coryletosum (Kelly and

Moore 1975) while the site at Abbeyleix belongs to the Querco- Fagetea.

Scot’s pine sites

There are large differences between the individual sites within the grouping of

Scot’s pine forests (Table 3.11). Bansha and Brittas are the two most species rich-

sites. Both forests are structurally diverse with numerous tree species in each of

the canopy layers. All of the other SP sites are less species rich than these two

sites. The woodland at Torc is one of the most ancient SP woodlands in Ireland;

however its vegetation composition is very species-poor. This is probably due to

the invasive species Rhododendron ponticum which is aggressively spreading and

shading the native ground flora. Annagh, Derryhogan and Gortnagowna are all

relatively young forests planted on bog soil. Derryhogan differs in that it includes

a sizable proportion of Sitka spruce (20%) of the forest canopy. This inclusion of

Sitka spruce in the canopy has effects on the other plant species present.

78

Table 3.11 List of Scot’s pine sites with the tree and plant species richness along with the Simpson’s diversity index (1-D) and the Shannon’s diversity index (H’).

Site Tree species Other

plants

Simpson’s

diversity

Shannon’s

diversity

Annagh 2 7 0.89 2.2

Ballylug 1 7 0.89 2.2

Ballygawley 1 9 0.89 2.2

Bansha 4 14 0.94 2.89

Brittas 4 12 0.94 2.77

Derryhogan 4 7 0.91 2.4

Gortnagowna 3 7 0.9 2.30

Torc 3 6 0.88 2.08

All SP sites 8 30 0.90 2.38

The Scot’s pine sites at Bansha and Brittas are composed of a majority of

Scot’s pine in the upper canopies with scattered oak and birch trees in the

understory. The remaining Scot’s pine sites at Annagh, Derryhogan,

Gortnagowna, Torc, Ballylug and Ballygawley are predominantly single-species

stands. In phytosociological terms, these sites belong to the Vaccinio–Piceetea

(Br.–Bl. Et Vlieger 1939).

Sitka spruce sites

All of the Sitka spruce sites are vascular plant species-poor (Table 3.12). The

young Sitka spruce sites (Cloonagh, Dooary and Stanahely) are typically poor in

tree species and contain only the main tree species. The most species-rich site is

Ballygawley. This site is a mature Sitka spruce site with some scattered young oak

trees in the understory, although not within the plot.

The Sitka spruce sites examined in this project are all classified as conifer

plantation WD4 (Fossitt 2000). As is the case for the Scot’s pine sites, in

phytosociological terms all the spruce sites would belong to the Vaccinio –

Piceetea (Br. –Bl. Et Vlieger 1939).

79

Table 3.12 List of Sitka spruce sites with the tree and plant species richness along with the Simpson’s diversity index (1-D) and the Shannon’s diversity index (H’).

Site Tree species Other

plants

Simpson’s

diversity

Shannon’s

diversity

Ballygawley 2 9 0.91 2.4

Bohatch 3 4 0.86 1.95

Chevy chase M 1 7 0.86 2.08

Chevy chase Y 3 5 0.86 1.95

Cloonagh 1 1 0 0

Dooary 1 3 0.75 1.39

Moneyteige 1 1 0 0

Quitrent 2 7 0.89 2.2

Stanahely 1 3 0.75 1.39

All SS sites 5 19 0.8 1.75

3.4.5 Structural descriptions of the sites

Diversity in the different canopy layers

Examination of the values for each layer in the different sites (Table 3.13) gives

an idea of the structural complexity of the sites. On average, oak sites were highly

structurally complex, having many plant species in all of the layers examined.

Conversely, Sitka spruce normally had a species poor understory and shrub layer,

as Sitka spruce forms a closed canopy and solar radiation to the lower layers is

limited.

Vegetation structural descriptions

The structural characteristics of the sites were measured (Table 3.14; Figs 3.7a, b,

c, d) and significant relationships between age, diameter at breast height (DBH),

stocking density, and % cover in the four canopy layers tested for significance

using a Spearman’s correlation test. In general, all of the variables listed in the

below table are related to the age of the forest. As age increases, canopy openness

decreases, DBH increases, stocking density decreases, and the cover index for the

S3 and S4 canopy layers increase. This relationship was shown by correlation

analysis by the Spearman’s rank correlation values and significance values

between age and the variables DBH (rs= 0.741, P<0.001), stocking density (rs= -

0.512, P<0.001) and S4 %cover (rs= 0.7, P<0.001).

80

Table 3.13 Number of tree and other vascular plant species recorded in each of the canopy layers. S1 = field layer (10cm-1.9m), S2= shrub layer (2m-5m), S3= lower canopy (5.1m-15m) and S4=canopy layer (15.m+). FT= Forest type, TS= Tree species richness, OS= other plant species richness, SS= Sitka spruce, SP= Scot’s pine.

FT Site S1

TS

S1

OS

S2

TS

S2

OS

S3

TS

S3

OS

S4

TS

S4

OS

ASH Ballykilcavan 1 5 1 0 1 0 0 0

ASH Donadea 2 11 3 0 1 0 3 0

ASH Killough 0 17 1 1 2 0 1 0

ASH Ross island 0 4 2 0 2 0 3 0

ASH St Johns 1 6 2 1 3 1 1 0

OAK Abbeyleix 0 8 3 2 3 1 3 0

OAK Kilmacrea 1 9 3 1 2 1 1 0

OAK Raheen 0 9 2 2 1 2 1 0

OAK Union 0 5 1 1 2 0 2 0

OAK TomiesA 0 7 0 0 1 1 1 0

OAK TomiesB 0 10 3 1 3 1 1 0

SP Annagh 1 7 1 0 2 0 0 0

SP Ballylug 0 7 0 1 1 0 1 0

SP Ballygawley 0 9 1 0 1 0 1 0

SP Bansha 0 13 3 0 3 0 1 0

SP Brittas 0 13 2 1 1 1 2 0

SP Derryhogan 1 6 4 1 3 0 2 0

SP Gortnagowna 0 5 2 2 2 0 0 0

SP Torc 0 5 2 2 0 0 1 0

SS Ballygawley 0 7 1 3 2 0 1 0

SS Bohatch 0 4 3 0 0 0 1 0

SS Chevy chaseM

0 7 0 0 0 0 1 0

SS Chevy chaseY

0 4 1 1 2 0 0 0

SS Cloonagh 0 1 0 0 1 0 0 0

SS Dooary 0 3 0 0 1 0 1 0

SS Moneyteige 0 1 0 0 0 0 1 0

SS Quitrent 0 7 0 0 0 0 2 0

SS Stanahely 0 3 1 0 1 0 0 0

81

Table 3.14 List of the structural characteristics of the different sites. Canopy openness is measured in counted points on a canopy scope (Brown et al. 2000). S1-S4 refers to the different canopy strata which are measured in the vertical canopy structure assessment method by Ferris-Kaan et al. (1998). SP= Scot’s pine, SS= Sitka spruce, DBH= diameter at breast height, CI= Cover index. Site codes can be found in Table 3.2. Forest

type

Site Canopy

openness

DBH Stocking

density

(trees per

100m2)

S1

CI

S2

CI

S3

CI

S4

CI

Total

CI

ASH BkilCav 2.2±0.84 43±23 8 38 30 900 0 968

ASH Donad 3.4±1.67 104±91 9 190 75 100 500 865

ASH Killo 2±0.71 102±23 8 190 120 700 350 1360

ASH Rossi 1.6±0.89 83±47 10 120 30 600 475 1225

ASH StJon 1.2±0.45 112±33 3 160 195 850 100 1305

OAK Abbey 2.8±2.39 390±90 2 76 90 450 250 866

OAK Kilmac 1.4±0.55 104±19 5 160 165 500 400 1225

OAK Rahee 1.0±0 258±107 3 160 195 200 400 955

OAK Union 4±3.16 240±10 2 114 15 200 400 729

OAK TomieA 1.0±0 252±68 2 130 165 500 400 1195

OAK TomieB 1.0±0 297±43 2 190 180 500 200 1070

SP Anna 2±1.23 62±28 18 19 15 800 0 834

SP Blug 10±4.36 124±16 7 171 15 100 250 536

SP BgawSP 27.4±8.14

80±60 9 57 45 50 50 202

SP Bnsha 8±3.54 125±26 7 152 210 300 250 912

SP Britt 11±5.5 146±16 6 170 200 500 240 1110

SP Derry 1.8±1.1 80±10 14 160 60 800 50 1070

SP Gortna 1.4±0.89 114±66 8 150 60 900 0 1110

SP Torc 6.6±4.78 175±15 3 104 60 0 400 564

SS BgawSS 3.6±0.89 163±77 6 48 75 350 350 823

SS Bohat 1.0±0 89±56 16 29 45 0 400 474

SS ChevyM 2.4±1.67 139±49 8 50 0 0 475 525

SS ChevyY 1.0±0 55±35 18 30 15 950 0 995

SS Cloon 1.0±0 60±20 20 0 0 990 0 990

SS Dooar 2.6±0.55 74±34 6 133 0 600 150 883

SS Money 1.2±0.45 102±20 15 50 0 0 400 450

SS Quit 6.6±0.55 81±21 8 60 0 0 375 435

SS Swmid 4.4±1.34 70±20 13 50 150 980 50 1230

The majority of the variation in canopy openness was found in the Scot’s

pine forest type (Fig. 3.7a). Scot’s pine trees, which are known for their

characteristic “stags horn” tree structure, have a canopy which has many gaps

spread throughout it. This gives Scot’s pine forests a very patchy and open

canopy. The DBH data (Fig. 3.7b) highlights the massive size of the oak trees in

comparison to the other trees examined. The oak forests were the oldest sites in

the study. Higher stocking densities were found in the coniferous sites over the

deciduous sites (Fig. 3.7c). The cover index (Fig. 3.7d) was found to be similar

82

(P>0.05) between the different forest types according to a Mann-Whitney U test.

However, there was a wide spread of values within the Scot’s pine forest type and

the Sitka spruce forest type, which would indicate that in these forest types there

were sites which had a relatively low cover index.

(a)

(b)

(c)

(d)

Figs 3.7. High-low plots for the structural variables (a) canopy openness, (b) diameter at breast height (cm), (c) stocking density (trees per 100m2) and (d) Cover index. Circles are mean values with highest and lowest values indicated by whiskers.

Coarse woody debris quantification

Coarse woody debris quantity in each plot (Table 3.15) shows a positive

relationship with stand age according to Spearman’s correlation test.

83

Table 3.15 List of coarse woody debris results from the sites. All values are measured in m3/ha deadwood. *= sites which data missing and so amounts were replaced with mean values from across similar forests types. SP= Scot’s pine, SS= Sitka spruce.

Forest

type

Site CWD1 CWD2 CWD3 CWD4 Total CWD

ASH Ballykilcavan 18.12 7.02 1.48 0.00 26.62

ASH Donadea* 10.50 19.70 1.50 3.70 35.40

ASH Killough* 10.50 19.70 1.50 3.70 35.40

ASH Ross island* 10.50 19.70 1.50 3.70 35.40

ASH St Johns 2.96 32.41 1.48 7.39 44.24

OAK Abbeyleix 2.96 17.01 9.98 0.00 29.95

OAK Kilmacrea* 1.10 25.70 23.10 56.60 106.50

OAK Raheen* 1.10 25.70 23.10 56.60 106.50

OAK Union 0.00 59.40 1.48 68.65 129.53

OAK TomiesA 1.48 5.55 44.61 78.88 130.52

OAK TomiesB 0.00 20.95 36.48 78.88 136.31

SP Annagh 1.48 9.98 9.98 12.57 34.02

SP Ballylug 0.00 0.00 0.00 0.00 0.00

SP Ballygawley 7.02 136.68 56.32 5.55 205.57

SP Bansha 0.00 15.41 98.47 60.39 174.27

SP Brittas 0.00 15.41 98.47 60.39 174.27

SP Derryhogan 1.48 0.00 49.30 0.00 50.78

SP Gortnagowna 1.50 4.90 29.00 6.00 42.30

SP Torc 1.48 0.00 49.30 0.00 50.78

SS Ballygawley 0.00 0.00 1.48 115.23 116.71

SS Bohatch 9.70 64.00 24.00 48.00 145.70

SS Chevy chase M 0.00 52.01 0.00 0.00 52.01

SS Chevy chase Y 0.00 0.00 0.00 8.50 8.50

SS Cloonagh 0.00 0.00 0.00 0.00 0.00

SS Dooary 67.17 12.57 0.00 0.00 79.74

SS Moneyteige 0.00 126.57 81.34 16.64 224.55

SS Quitrent 39.07 79.86 14.05 60.39 193.37

SS Stanahely 0.00 5.55 30.81 80.36 116.71

3.4.6 Meteorological data

The met data used is based on the mean yearly values calculated from 1961-1990

(Table 3.16; 3.17). The climate of Ireland is generally wetter towards the west

coast (Fig. 3.8) and therefore sites closer to the west coast experience high

precipitation levels.

84

Table 3.16 Mean and standard deviations for daily temperature (n=12) and monthly total rainfall (n=12) for weather stations close to the sites used.

Weather

station

Corresponding sites Temperature Rainfall

Birr/Gurteen Annagh, Brittas, Gortnagowna, Derryhogan, Bansha

9.6 ± 3.23 95.21±27.11

Casement Donadea, Kilmacrea, Ballylug, Moneyteige, Stanahely

9.27±3.90 59.26±9.17

Claremorris St Johns, Union, Ballygawley SP, Cloonagh, BallygawleySS

8.87±3.74 94.7±23.20

Cork airport Quitrent 9.39±3.65 100.58±27.23

Kilkenny/Oak park

Ballykillcavan, Killough, Abbeyleix, Dooary

9.27±3.95 68.58±13.50

Shannon Airport Raheen, Bohatch, Chevy chase M, Chevy chase Y

10.15±3.88 77.23±16.47

Valentia Ross island, Tomies, Torc 10.42±3.11 119.19±33.87

Table 3.17 Mean and standard deviation (n=365) of daily temperature and rainfall for the weather stations over the years 2007, 2008 and 2009. Weather

station

Corresponding sites Temperature Rainfall

Birr/Gurteen Annagh, Brittas, Gortnagowna, Derryhogan, Bansha

2007 = 10.43±3.86 2008 = 9.7±4.36 2009 = 9.65±4.78

2007 = 2.37±4.16 2008 = 2.78±4.63 2009 = 2.92±4.67

Casement Donadea, Kilmacrea, Ballylug, Moneyteige, Stanahely

2007 = 10.3±3.9 2008 = 9.54±4.43 2009 = 9.68±4.67

2007 = 2.22±4.42 2008 = 2.69±5.62 2009 = 2.53±5.42

Claremorris St Johns, Union, Ballygawley SP, Cloonagh, BallygawleySS

2007 = 10.18±3.8 2008 = 9.51±4.17 2009 = 9.42±4.6

2007 = 3.09±4.75 2008 = 3.73±5.12 2009 = 3.61±5.18

Cork airport Quitrent 2007 = 10.50±3.47 2008 = 9.64±3.76 2009 = 9.43±4.11

2007 = 2.89±5.29 2008 = 3.68±6.32 2009 = 4.28±7.18

Kilkenny/Oak park

Ballykillcavan, Killough, Abbeyleix, Dooary

2007 = 10.43±3.97 2008 = 9.7±4.42 2009 = 9.6±4.76

2007 = 2.28±4.22 2008 = 2.6±5.11 2009 = 3.1±5.38

Shannon Airport

Raheen, Bohatch, Chevy chase M, Chevy chase Y

2007 = 11.15±3.74 2008 = 10.5±4.09 2009 = 10.5±4.64

2007 = 2.53±4.36 2008 = 3.5±5.72 2009 = 3.2±4.61

Valentia Ross island, Tomies, Torc

2007 = 11.8±2.97 2008 = 11.2±3.19 2009 = 11.0±3.67

2007 = 3.70±6.24 2008 = 4.6±7.33 2009 = 6.0±8.73

Daily met data was also acquired from the nearest weather stations for the

years of the study 2007, 2008 and 2009. The mean and standard deviations of the

data for each weather station were calculated (Table 3.17).

85

Fig. 3.8 Map of Ireland showing the average precipitation levels over the 30 year period 1960-90. Source MET Eireann www.met.ie.

3.4.7 Community structure of the vegetation in the plots

Similarity analysis of the forest sites

Using the modified Jaccard similarity index, the vegetation communities from the

different plots were compared (Table 3.18). The median, minimum and maximum

of the Jaccard index values between the different forest types were also calculated

(Table 3.19).

There is a high degree of similarity between the different forest types.

What is evident is that, on average, a plot is more similar to another plot if they

share the same forest type than it is to a plot of a different forest type (Table 3.19).

There is an extremely high degree of similarity between the Scot’s pine plots

(median JI= 1.0), followed by the oak plots (median JI= 0.9) and ash plots

(median JI= 0.66± 0.33), with the Sitka spruce plots showing high degree of inter-

plot variation in the Jaccard index (median JI= 0.43). The other forest type pair to

show a high degree of similarity were oak and Scot’s pine (median JI= 0.56)

(Table 3.19).

86

87

Table 3.18 The number of shared species (clear cells) and the abundance based Jaccard similarity index (grey cells) which ranges from 0 (not similar) to 1 (very similar) of the vegetation data between the corresponding sites. Blue cells correspond to ash sites, green cells to oak sites, orange cells to Scot’s pine sites and brown cells to Sitka spruce sites.

BKIL DONAD KILLO ROSS STJON ABBEY KILMAC RAHEE TOMIE1 TOMIE2 UNION ANNA BGAWSP BLUG

BKIL X 0.71 0.17 0.36 0.30 0.26 0.00 0.26 0.11 0.22 0.30 0.00 0.00 0.00

DONAD 4.00 X 0.49 1.00 1.00 0.73 0.17 1.00 0.19 0.15 0.40 0.06 0.07 0.19

KILLO 2.00 4.00 X 0.55 1.00 0.29 0.56 1.00 0.32 0.83 0.31 0.15 0.06 0.05

ROSS 2.00 6.00 4.00 X 1.00 0.54 0.54 1.00 0.69 0.44 0.63 0.09 0.10 0.09

STJON 2.00 5.00 6.00 5.00 X 0.21 1.00 1.00 1.00 1.00 1.00 1.00 0.27 0.59

ABBEY 2.00 4.00 3.00 3.00 2.00 X 0.19 0.90 0.22 0.73 0.93 0.51 0.54 0.93

KILMAC 0.00 2.00 4.00 3.00 5.00 2.00 X 1.00 1.00 1.00 0.48 0.51 0.92 0.93

RAHEE 2.00 5.00 6.00 5.00 6.00 4.00 7.00 X 1.00 1.00 0.48 0.51 0.23 0.51

TOMIE1 1.00 2.00 3.00 3.00 6.00 2.00 6.00 6.00 X 1.00 0.59 0.64 0.28 0.64

TOMIE2 2.00 2.00 5.00 3.00 6.00 4.00 9.00 9.00 7.00 X 0.75 1.00 0.44 0.74

UNION 2.00 3.00 3.00 3.00 4.00 4.00 3.00 3.00 3.00 4.00 X 0.59 0.63 0.25

ANNA 0.00 1.00 2.00 1.00 4.00 3.00 3.00 3.00 3.00 5.00 3.00 X 1.00 1.00

BGAWSP 0.00 1.00 1.00 1.00 2.00 3.00 4.00 2.00 2.00 3.00 3.00 5.00 X 1.00

BLUG 0.00 2.00 1.00 1.00 3.00 4.00 4.00 3.00 3.00 4.00 2.00 5.00 5.00 X

BNSHA 0.00 2.00 2.00 2.00 4.00 2.00 6.00 4.00 4.00 6.00 2.00 6.00 5.00 6.00

BRITTAS 1.00 2.00 5.00 3.00 7.00 3.00 9.00 8.00 8.00 11.00 4.00 6.00 5.00 6.00

DERRY 2.00 2.00 2.00 2.00 3.00 4.00 3.00 3.00 2.00 6.00 2.00 5.00 3.00 3.00

GORT 0.00 2.00 3.00 2.00 3.00 3.00 5.00 3.00 2.00 5.00 2.00 5.00 4.00 4.00

TORC 1.00 1.00 1.00 1.00 3.00 2.00 2.00 2.00 3.00 4.00 2.00 5.00 3.00 3.00

BOHAT 0.00 1.00 1.00 1.00 2.00 3.00 5.00 2.00 2.00 5.00 1.00 3.00 4.00 3.00

MONEY 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00

QUITR 1.00 1.00 0.00 0.00 1.00 0.00 1.00 1.00 1.00 1.00 1.00 2.00 2.00 0.00

BGAWSS 2.00 1.00 1.00 2.00 3.00 3.00 5.00 5.00 6.00 7.00 5.00 4.00 5.00 4.00

CHEVM 3.00 3.00 1.00 1.00 1.00 4.00 1.00 3.00 2.00 3.00 2.00 1.00 2.00 2.00

CLOON 0.00 0.00 0.00 0.00 1.00 0.00 1.00 0.00 1.00 1.00 1.00 1.00 1.00 1.00

DOOAR 0.00 1.00 1.00 1.00 2.00 1.00 2.00 1.00 1.00 2.00 1.00 2.00 2.00 2.00

SWMID 0.00 0.00 0.00 0.00 1.00 1.00 2.00 1.00 2.00 2.00 2.00 2.00 3.00 2.00

CHEVY 2.00 2.00 2.00 2.00 3.00 2.00 3.00 3.00 3.00 5.00 2.00 2.00 1.00 1.00

88

89

BNSHA BRITT DERRY GORT TORC BOHAT MONEY QUITR BGAWSS CHEVM CLOON DOOAR SWMID CHEVY

BKIL 0.00 0.07 0.28 0.00 0.11 0.00 0.00 0.11 0.28 0.84 0.00 0.00 0.00 0.40

DONAD 0.13 0.14 0.17 0.18 0.07 0.07 0.00 0.06 0.06 0.44 0.00 0.08 0.00 0.21

KILLO 0.11 0.84 0.14 0.31 0.06 0.06 0.00 0.00 0.05 0.06 0.00 0.06 0.00 0.16

ROSS 0.17 0.41 0.25 0.27 0.10 0.11 0.00 0.00 0.25 0.10 0.00 0.13 0.00 0.33

STJON 0.62 1.00 0.51 0.54 0.63 0.28 0.00 0.08 0.51 0.09 0.12 0.34 0.11 0.69

ABBEY 0.15 0.34 0.94 0.48 0.23 0.57 0.00 0.00 0.45 0.92 0.00 0.10 0.10 0.25

KILMAC 1.00 1.00 0.45 1.00 0.23 1.00 0.00 0.07 1.00 0.08 0.10 0.29 0.29 0.57

RAHEE 0.62 1.00 0.45 0.48 0.23 0.25 0.00 0.07 1.00 0.54 0.00 0.10 0.10 0.57

TOMIE1 0.62 1.00 0.23 0.25 0.69 0.31 0.00 0.09 1.00 0.28 0.13 0.12 0.38 0.76

TOMIE2 1.00 1.00 1.00 1.00 0.73 1.00 0.00 0.06 1.00 0.44 0.08 0.23 0.23 1.00

UNION 0.16 0.70 0.22 0.24 0.27 0.09 0.00 0.08 1.00 0.27 0.12 0.11 0.34 0.28

ANNA 1.00 1.00 1.00 1.00 1.00 0.76 0.00 0.27 1.00 0.09 0.13 0.38 0.38 0.31

BGAWSP 0.89 1.00 0.58 1.00 0.77 1.00 0.00 0.28 1.00 0.31 0.15 0.43 0.80 0.11

BLUG 1.00 1.00 0.54 1.00 0.69 0.76 0.00 0.00 1.00 0.28 0.13 0.38 0.38 0.10

BNSHA X 1.00 0.91 1.00 0.89 0.38 0.00 0.36 0.91 0.06 0.07 0.19 0.07 0.18

BRITTAS 8.00 X 1.00 1.00 0.41 0.68 0.00 0.18 1.00 0.41 0.08 0.22 0.22 0.68

DERRY 5.00 5.00 X 1.00 1.00 1.00 0.09 0.54 1.00 0.58 0.27 0.58 0.31 1.00

GORT 6.00 6.00 6.00 X 0.63 1.00 0.00 0.00 0.22 0.09 0.12 0.34 0.11 1.00

TORC 5.00 3.00 5.00 3.00 X 0.33 0.00 0.09 1.00 0.10 0.15 0.13 0.13 0.85

BOHAT 3.00 4.00 5.00 4.00 2.00 X 0.14 0.31 0.62 0.33 0.43 0.91 0.91 1.00

MONEY 0.00 0.00 1.00 0.00 0.00 1.00 X 0.11 0.09 0.13 0.50 0.25 0.25 0.14

QUITR 3.00 2.00 3.00 0.00 1.00 2.00 1.00 X 0.54 0.28 0.13 0.12 0.38 0.31

BGAWSS 5.00 7.00 4.00 2.00 4.00 3.00 1.00 3.00 X 1.00 0.27 0.31 0.91 0.99

CHEVM 1.00 3.00 3.00 1.00 1.00 2.00 1.00 2.00 5.00 X 0.15 0.43 0.43 0.85

CLOON 1.00 1.00 2.00 1.00 1.00 2.00 1.00 1.00 2.00 1.00 X 0.75 0.75 0.43

DOOAR 2.00 2.00 3.00 2.00 1.00 3.00 1.00 1.00 2.00 2.00 2.00 X 0.78 0.48

SWMID 1.00 2.00 2.00 1.00 1.00 3.00 1.00 2.00 4.00 2.00 2.00 2.00 X 0.48

CHEVY 2.00 4.00 6.00 4.00 3.00 4.00 1.00 2.00 4.00 3.00 2.00 2.00 2.00 X

90

91

Table 3.19 Medians and minimum/maximum values (in parenthesis) for the Jaccard similarity index between the vegetation communities of the forest types based on inter-plot values (shaded cells). The median number of shared species per pair of plots between the different forest types are also given (clear cells).

Forest type Ash Oak Scot’s pine Sitka spruce

Ash 0.63 (0.17/1) 0.34 (0/1) 0.27 (0/1) 0.19 (0/0.84)

Oak 3 0.9 (0.19/1) 0.56 (0.15/1) 0.23 (0/1)

Scot’s pine 2 3 1 (0.41/1) 0.28 (0/1)

Sitka spruce 1 2 2 0.43 (0.09/1)

Ordination based on vascular plant communities.

A NMS ordination was carried out on the vegetation data (presence/absence) data

from each of the plots. Using Sørensen distance measure, 40 runs with real data,

and 50 runs with random data. A 3 dimensional ordination (Figs 3.11, 3.12, 3.13;

3.14) was found to be the best solution by visually inspecting a NMS scree plot.

The final stress reported was 15.7 (Fig. 3.9) which indicated an intermediate

solution according to Kruskal’s rule of thumb (McCune and Grace 2002). A

Monte Carlo test showed that the probability of a similar final stress being

obtained by chance was P= 0.0196. Axis 1 explained 30% of the variance with a

further 35.6% of the variation explained by axis 2. Axis 3 explained the remaining

10.2% of the variation to give the ordination a cumulative r2 value of 75.8%.

There were a total of 19 attributes entered into the NMS ordination. Of

these, six were found to be correlated with the ordination axes from the NMS.

Axis 1 was found to be correlated with organic matter, axis 2 with calcium and

axis 3 with pH and precipitation. Using Spearman’s rank correlation (rs) the

relationships between the environmental variables and the ordination axes were

examined (Table 3.20).

92

0

20

40

60

1 3 5

FunctionalBio

Dimensions

Stre

ss

Randomized Data

Maximum

Mean

Minimum

Real Data

Fig. 3.9 NMS scree plot to show the stress levels associated with the increment in dimensions used for NMS. The “elbow” (McCune and Grace 2002) appears at x= 3 y= 15.7. This indicates that a three dimensional ordination would have a final stress of 15.7. Table 3.20 Significant Spearman’s rank correlation (rs) values for the relationships between the ordination axes and the environmental variables. OM= organic matter, NO3= nitrate nitrogen, Ca= calcium and CWD4= the volume of Coarse woody debris in the site of diameter greater than 40cm. *= P<0.05, **=P<0.01.

pH OM Precipitation NO3 Ca CWD4

Axis 1 0.337* -0.517**

Axis 2 -0.582**

Axis 3 0.492** -0.474* -0.375* -0.607**

Using linear regression the variables organic matter, total nitrogen and

available phosphorus were found to be significantly related to plant species

richness (Figs 3.10a, b, c). The regression statistics were organic matter and plant

species richness are positively related (F1,26= 5.205, r2= 0.17, P<0.05). Total

nitrogen and plant species richness were positively related (F1,26= 11.73, r2= 0.31,

P<0.01). Phosphorus and plant species richness were positively related (F1,26=

5.799, r2= 0.18, P<0.05).

93

0 10 20 30 40 500.0

2.5

5.0

7.5

10.0

12.5

15.0

17.5

20.0

22.5

OM%

Pla

nt

sp

ecie

s

0.00 0.25 0.50 0.75 1.00 1.25 1.500.0

2.5

5.0

7.5

10.0

12.5

15.0

17.5

20.0

22.5

N total

Pla

nt

sp

ecie

s

0 1 2 3 4 5 6 7 8 90.0

2.5

5.0

7.5

10.0

12.5

15.0

17.5

20.0

22.5

P

Pla

nt

sp

ecie

s

(a)

(b)

(c)

Figs. 3.10 a, b and c. Regression curves for plant species richness and (a) organic matter (F1,26= 5.205, r2= 0.17, P<0.05), (b) total nitrogen (F1,26= 11.73, r2= 0.31, P<0.01) and (c) phosphorus (F1,26= 5.799, r2= 0.18, P<0.05).

94

95

BKIL

DONAD

KILLO ROSS

STJON

ABBEY

KILM AC

RAHEE

TOOM 1

TOOM 2

UNION

ANNA

BGAWSP

BLUG

BNSHA

BRITTAS

DERRY

GORT

TORC

BOHAT

M ONEY

QUITRBGAWSS

CHEVM

CLOON

DOOAR

SWM ID

CHEVYOM

Ca

NMS ordination

Axis 1

Axi

s 2

Tree

AshOakSPSS

Fig. 3.11 Non metric multidimensional scaling of axis 1 (r2= 0.301) and 2 (r2= 0.355; cumulative r2= 0.656) showing the plots identified by the forest type of the site and a biplot of environmental and physical variables. Axis 1 shows correlations with organic matter (OM: r2= 0.306). Axis 2 shows correlations with calcium (Ca: r2= 0.379). SP = Scot’s pine, SS= Sitka spruce.

96

97

BKIL

DONAD

KILLO

ROSS

STJON

ABBEY

KILM AC

RAHEE

TOOM 1

TOOM 2

UNION

ANNA

BGAWSP

BLUG

BNSHA BRITTAS

DERRY

GORT

TORC

BOHAT

M ONEY

QUITR

BGAWSS

CHEVM

CLOON

DOOAR

SWM ID

CHEVY

OM

Precip

CWD4

NMS ordination

Axis 1

Axi

s 3

Tree

AshOakSPSS

Fig. 3.12 Non metric multidimensional scaling of axis 1 (r2= 0.301) and 3 (r2= 0.102; cumulative r2= 0.403) showing the plots identified by the forest type of the site and a biplot of environmental and physical variables. Axis 1 shows correlations with Organic matter (OM: r2= 0.306). Axis 3 shows correlations with precipitation (Precip: r2= 0.255) and Volume of Coarse woody debris size of size greater than 40cm diameter (CWD4: r

2= 0.33). SP = Scot’s pine, SS= Sitka spruce.

98

99

BKIL

DONAD

KILLO

ROSS

STJON

ABBEY

KILM ACRAHEE

TOOM 1

TOOM 2

UNION

ANNA

BGAWSP

BLUG

BNSHA

BRITTAS

DERRY

GORT

TORC

BOHAT

M ONEY

QUITR

BGAWSS

CHEVM

CLOON

DOOAR

SWM ID

CHEVYPH

Precip

Ca

CWD4

Axis 2

Axi

s 3

Tree

AshOakSPSS

Fig. 3.13 Non metric multidimensional scaling of axis 3 (r2= 0.102) and 2 (r2= 0.355; cumulative r2= 0. 468) showing the plots identified by the forest type of the site and a biplot of environmental and physical variables. Axis 2 shows correlations with calcium (Ca: r2= 0.379) and pH (pH: r2= 0.116). Axis 3 shows correlations with ph (pH: r2= 0.185), precipitation (Precip: r2= 0.255) and Volume of Coarse woody debris size of size greater than 40cm diameter (CWD4: r

2= 0.33). SP= Scot’s pine, SS= Sitka spruce.

100

101

The resulting ordination from axis 1 and 2 (Fig. 3.11) shows good separation

between the sites based on the forest type of the site. The correlation of axis 1 with

organic matter is useful in separating the sites as the Scot’s pine and oak sites had

higher OM content than the ash or the Sitka spruce. Axis 2 is useful in separating the

sites as it shows correlations with calcium. The ash sites were found to have more

calcium present than any other tree type. Two points to note are the grouping of the

Scot’s pine site Brittas with the oak sites, and the of the ash site at St John’s wood

with the oak sites. Both of these sites are relatively undisturbed sites which have rich

understory species richness. The site at Brittas has some oak trees in the understory

and so this has no doubt contributed to grouping this site in the oak grouping.

The resulting ordination using axes 1 and 3 (Fig. 3.12) does not show good

separation of the sites based on forest type. The correlation of axis 1 with organic

matter is useful in separating the Scot’s pine sites which normally had high OM%

from the ash and Sitka spruce sites.

The ordination showing axis 2 and 3 (Fig. 3.13) produces good separation of

the sites based on the forest type of the sites. The values for pH and calcium are very

useful in separating the sites, as ash sites had a very large calcium and hence pH

value. pH was useful in separating the usually low (~4) pH coniferous sites from the

medium to high pH (5-6) deciduous sites.

Indicator species analysis

To identify plant species which are indicative of the habitat examined in this project,

Indicator values for each species were calculated in PC-ORD using the method by

Dufrene and Legendre (1997) with statistical significance calculated by a Monte

Carlo test with 1000 runs. Eight species were found to be indicators of a specific

forest type (Table 3.21).

The results of the indicator species analysis show that Calluna vulgaris and

Pteridium aquilinum could be useful indicator species of Scot’s pine forests.

Crataegus monogyna could be a useful indicator species for ash habitats in Ireland.

Oak habitats would be indicated by the presence of Luzula sylvatica and Oxalis

acetosella. Sitka spruce habitats were not strongly indicated by the presence of any

plant species (Table 3.21).

102

Table 3.21 The species which indicated a particular forest type with statistical significance (P<0.05) are listed along with the strength of their relationship with the habitat (Indicator value). Significance values were calculated by using 1000 Monte Carlo simulations of the untransformed site x species data.

Indicator

value

Indicator

value

Indicator

value

Indicator

value

Species Ash Oak SP SS Significance

Calluna

vulgaris

0 0 41 2 <0.05

Corylus

avellana

13 37 1 0 <0.05

Crataegus

monogyna

60 0 0 0 <0.01

Hedera helix 41 29 3 5 <0.05

Luzula

sylvatica

0 49 2 1 <0.01

Oxalis

acetosella

0 42 9 12 <0.05

Pteridium

aquilinum

0 9 43 4 <0.05

Thuidium

tamariscinum

2 18 39 18 <0.01

The ordination showing the sites and the indicator species present (Fig. 3.14)

is useful in giving a picture of the diversity of the communities in each of the forest

types. The Sitka spruce sites cluster to one side of the ordination. No vascular species

were shown to be strong indicators of this forest type. The ash sites form a grouping

along axis 2 (Fig. 3.14) which is correlated with increasing pH. The indicator species

analysis found Crataegus monogyna to be a useful species which marks the ash forest

type. The Scot’s pine sites cluster very tightly, except for the site at Brittas which is

closer to the oak sites than the other Scot’s pine sites (Fig. 3.14). The oak sites form a

close grouping based on forest type, except for the site at Union which was one of the

most species poor oak sites (Table 3.10).

103

Fig. 3.14 Non metric multidimensional scaling of axis 1 (r2= 0.301) and 2 (r2= 0.355; cumulative r2= 0.656) showing the plots identified by the forest type of the site. Also plotted are the plant species shown to be indicator species of the forest types (Table 3.21). Abbreviations: SP=Scot’s pine, SS= Sitka spruce.

104

105

Multi response permutation procedure

Plant communities did differ significantly between forest types, and this was

confirmed by MRPP analysis (P<0.001, A=0.178). What this means is that sites

with the same forest type were grouped closer together in the ordination than was

expected by chance based on their plant communities. This analysis backs up the

ordination and community similarity analysis by confirming that the different

forest types have significantly different vascular plant communities.

3.4.8 Forest management related to plant species richness

Although the multivariate analysis showed that the forest types had significantly

different plant species communities, there was variation in the communities

between plots of the same forest type. Possible effects of forest structural and

management variables on plant species richness were examined using generalized

linear modelling.

The generalized linear model for oak found that the model explained a

significant amount of the variation between the oak forest plots (likelihood ratio

chi-square= 18.878, df=1, P<0.001). This means that the model explained

significantly more variation between the oak sites than the null model. The model

found that ungrazed oak plots had significantly higher plant species richness than

grazed plots (mean difference [ungrazed –grazed]= 0.187, df= 1, P<0.001).

In the ash forests the model was significant at describing the numbers of

species per plot visit (likelihood ratio chi-square= 43.837, df=3, P<0.001). It

found that vascular plant species richness per plot was much higher in un-

managed plots than in managed plots (mean difference [unmanaged–managed]=

0.654, df= 1, P=0.001).

In Scot’s pine forests, the model was significant at explaining the numbers

of plant species per plot visit (likelihood ratio chi-square= 8.723, df=2, P<0.05).

The model found that rotation stage was significant at explaining some of the

variation within the Scot’s pine plots with 2nd rotation plots having higher

vascular species richness per plot than 1st rotation plots (mean difference [2nd

rotation-1st rotation]= 0.35, df= 1, P<0.05). Age was not found to have a

significant effect on plant species richness.

106

The model describing the plant species richness in Sitka spruce forests was better

at explaining the variation in the dependent variable (litter decay species per plot

visit) than the null model (likelihood ratio chi-square= 43.728, df=2, P<0.001).

The model found that in Sitka spruce plots, age explained a significant proportion

of the variation between the vascular species richness of the Sitka spruce plots.

Mature spruce plots were found to have significantly higher plant species richness

than young plots (mean difference [Mature –young]= 0.6, df= 1, P<0.001).

107

3.5 Discussion

3.5.1 Significantly different plant communities across the four forest types

The analysis in this chapter has shown that there are distinctive plant communities

in each of the four forests types, and that these communities are significantly

different from each other. Previous studies in Irish native (Perrin et al. 2008) and

plantation forests (French et al. 2008) have shown by ordination, that the vascular

plant species present in a plot can be used to separate the plots based on a higher

classification, such as the dominant tree species or woodland classification type of

the plot. It is also known that the macrofungal communities of forests can be

related to the vascular plant communities, and this relationship is examined in

more detail later (Section 3.5.3, this Chapter; and Section 4.5.5, Chapter 4).

However, there was some variation in the similarity of the communities

within plots of similar forest type. Based on (1) the loose ordination of the group

sites by NMS, (2) the failure of indicator species analysis to identify species

specific to the group and (3) the variability of the site structural and edaphic

variables, Sitka spruce forests are the most variable with regard to their biological

and abiotic variables. The forest type that showed the least variability was the oak

forest type. The reason for the variation in vascular plant communities within the

Sitka spruce and to a lesser extent Scot’s pine forest types, was found to be due to

the inter-plot variation in age and rotation stages of the coniferous plots. In

contrast to this, all of the oak plots were of similar age and rotation stage; and

many of the ash plots were of similar age and rotation stage. The climax plant

communities that develop in old growth forests have much more vascular species

rich communities than forests that are managed for short rotation wood production

(Ferris et al. 2000b), and their communities also vary considerably.

Previous studies in Irish forests have identified other forest variables that

affect vascular plant community composition and species richness (Smith et al.

2007; Iremonger et al. 2006). Among the variables identified are canopy gaps,

open spaces, and retention of patches of forest beyond financial maturity. The

2006 BIOFOREST report (Iremonger et al. 2006) concluded that the current

forest regulation of providing 5-10% of the forest as open space (Anon. 1996),

108

may not be sufficient for the conservation and enhancement of species diversity of

vascular plants in Irish plantation forests.

In order for plantation forests to provide worthwhile returns in a short time

period and also conserve biodiversity and therefore ecosystem function, previous

studies (Peterken et al. 1992; Ferris et al. 2000b; Humphrey 2005) have noted that

biodiversity in forests might benefit more from the retention of patches of forest

beyond financial maturity than the increase which might arise from extended

growth rotations of larger areas of forest. By retaining forest patches, a mosaic of

different habitats would be created within each forest stand. Mosaic patterns and

structural diversity in forests has been identified as a positive indicator of

biodiversity (see Lindenmayer et al. 2000 for review), and so management

techniques which increase the patchwork pattern of forest structure in forests

should be promoted.

3.5.2 The plant communities of the four forest types

The overall plant species richness was 68 with 38, 31, 38 and 24 species found in

ash, oak, Scot’s pine and Sitka spruce forests respectively. The distribution data

for all of the species ranges from 500 to 1300+ records in Ireland (NBDC 2010)

and none of the species are on the checklist of protected species in Ireland (Curtis

and McGough 1988). Each forest type was found to have a vascular plant

community which was significantly different from the other forest types.

Irish ash forests

Overall the ash sites were the 3rd most diverse with respect to their plant

communities. Indicator species analysis identified Corylus avellana, Crataegus

monogyna and Hedera helix as possible vegetation indicators of ash forests in

Ireland table 3.19. Of these three, Crataegus monogyna is the only plant species

which has a high IV and so shows high fidelity to the ash forest communities.

French et al. (2008) also identified Hedera helix as an indicator species for ash

forests in Ireland. Soil available calcium and pH were found to be higher in the

ash forest type than in the other forest types. These variables would have a strong

effect on the plant species community composition of the ash plots. Irish ash

forests have been noted as having very high soil available calcium levels (Smith et

109

al. 2005) and this is due to the high requirements and basic soils necessary for

good growth of ash in Ireland (Horgan et al. 2004).

The effect of inter-forest type variation in site characteristics on the

vegetation community of the ash plots is evident within the ash forest type.

Generalized linear modelling identified forest management as having negative

effects on plant species richness with the unmanaged sites having higher plant

species richness than the managed sites. Through management the understory

plant composition of the managed plots has been modified from the perhaps

expected oak-ash-hazel WN2 (Fossitt 2000) to a type which might fit the highly

modified broadleaf woodland WD1 (Fossitt 2000).

The ash plot at Killough is in an unmanaged condition with little

disruption and forest management (apart from small amounts of coppice removal)

as can be seen by the presence of an orchid species (Helleborine sp.), which are

known to be negatively affected by habitat change and eutrophication (Dixon et

al. 2003). The plot at St John’s wood also had high species diversity and would be

classified as oak-ash-hazel WN2 (Fossitt 2000). This site is noted as being in one

of the oldest and most pristine forests in Ireland (Rackham 1995).

French et al. (2008) classified ash plantation forests in Ireland into 4

groups: grassland and heath (mostly remnant vegetation from a grassland, acidic

grassland and heath (grassland vegetation with more acid loving species),

basophilic forest (vegetation closely corresponding to that of Corylo-Fraxinetum

(Braun-Blanquet and Tuxen 1952) and bramble dominated plantations. The plots

at St John’s wood and Killough would be classified as the basophilic group

according to French et al. (2008) descriptions. The remaining ash sites all fit into

the broad classification of bramble dominated plantation forests. Similar to the

study by French et al. (2008) fast spreading species like bramble and ivy are very

common in the ash sites, as even when a closed canopy is formed there is still

sufficient light for growth reaching the forest floor.

Irish oak forests

The oak sites had the third highest species richness of vascular plants of the 4

forest types with a total of 31 species present. According to the diversity indices,

oak was the most species diverse of the forest types. The species most commonly

found in the oak sites in this project were all previously found to be common in

110

oak sites throughout Ireland (Kelly and Iremonger 1997). Indicator species

analysis identified Corylus avellana, Hedera helix, Luzula sylvatica and Oxalis

acetosella as possible indicators of oak forests. Of these indicator species Luzula

sylvatica and Oxalis acetosella show high indicator values which would indicate a

high degree of faithfulness of the species to the oak forest grouping. Kelly and

Iremonger (1997) utilised the distribution of Oxalis acetosella in TWINSPAN

analysis to separate their oak dominated groupings from the other forest types

they examined.

Ordination of the oak sites by vegetation species (presence/absence data)

produced a tight cluster of data points with the plots at Abbeyleix and Union

shown as outliers. Abbeyleix is the only pure pedunculate oak site in this project

and is known to be quite old being the remnants of a demesne woodland

(Rackham 1995). Its ground flora is very typical of Quercus robur sites, with a

good representation of bluebells and wood sanicle. It has been investigated by

Kelly and Iremonger (1997) and Kelly (2005) and described as an acidophilous

oak woodland dominated by of Quercus robur. It is best described as oak-ash-

hazel woodland WN2 (Fossitt 2000).

Union wood is classified as old (>200 years) sessile oak wood by a

previous study (Anon. 2005a). It is based on calcareous soils with very high

calcium availability, a neutral pH and is subject to grazing by Sika deer (Cervus

nippon Temminck) which has led to greater woodrush becoming a very dominant

species in the forest floor layer (Anon. 2005a). The preferential grazing by the

deer on oak saplings reduces the regeneration of the oak floristic component

(Kelly 2002) and causes a decline in the tightly closed oak dominated canopy. The

increase in light availability to the ground floor gives bushy quick-growing herbs,

like woodrush, a competitive advantage over slow growing shade tolerant herb

species. This change in the ground flora is a large reason that the Union site did

not group closely with the other oak sites in the project.

Raheen woodland has many species common to sessile oak woodlands

WN1 (Fossitt 2000) in Ireland including hard fern and honeysuckle. Kilmacrea

woodland clusters between the oak and Scot’s pine sites. It is an acidophilous

sessile oak wood (Perrin et al. 2006a) and has a history of oak trees in the site

since 1840. It is one of the only oak sites to have bilberry present, although

bilberry is listed as a common composite of the oak-birch-holly WN1 (Fossitt

111

2000). It is the only oak site to be situated on acid brown earth and along with the

low calcium levels, may have more acidic loving plant species present than the

other oak sites.

The effect of browsing by Sika deer on oak regeneration was highlighted

in this chapter by generalized linear modelling and also by the lack of oak

seedlings in the lower canopy layers in plots which suffer from deer browsing.

The plots at Union, Abbeyleix and TomiesA show almost no oak trees in the

lower canopy layers. Deer have been found to preferentially eat

oak>ash>beech>birch (Peterken and Jones 1989) and previous studies have

highlighted grazing by deer as a problem in the woodlands at Tomies wood (Kelly

2002; Perrin et al. 2006b) and Union woods (Anon. 2005a). Possible steps may

include exclusion of the deer by fencing or removal or control of the deer

population by increased culling (Heinze et al. 2010).

An interesting method of controlling the rampant red deer Cervus elaphus

is documented in Beschta and Ripple (2010). In order to help regenerate the

natural cotton wood plant communities of Yellowstone national park, scientists at

Oregon State University re-introduced the extinct predator grey wolf Canis lupis

into the park in 1995. Due primarily to the predation of the wolves on the deer, the

park has seen large increases in tree species regeneration since the re-introduction

of the “top predator”. The grey wolf was also a native species to Ireland up to ca.

1780 when it was made extinct following years of hunting and destruction of its

habitat (Hickey 2000). Although the re-introduction of wolves is most certainly

not an option for Irish or other heavily populated countries for many reasons not

least public opinion (Agarwala et al. 2010), the recent re-introduction of eagles

into Ireland in Donegal (O’Toole et al. 2002) and Killarney (White tailed eagle

project 2010) may have a small effect on the deer populations as eagles have been

known to kill young sheep and other small mammals.

Irish Scot’s pine forests

The Scot’s pine sites had the tied highest species richness of plants, along with

ash sites with 38 plant species. A past study by Roche et al. (2009) also identified

many of the species found in Scot’s pine habitats in this study as being commonly

found in Scot’s pine sites in Ireland. Indicator species analysis identified

Thuidium tamariscinum, Calluna vulgaris and Pteridium aquilinum as indicator

112

species of Scot’s pine forests. Both Calluna vulgaris and Pteridium aquilinum

were identified as indicator species of Scot’s pine habitats in Ireland, with

Pteridium aquilinum being a constant species in any pine habitats and Calluna

vulgaris being indicative of the Calluna vulgaris–Eriophorum vaginatum sub-

group (Roche et al. 2009). Humphrey and Coombs (1997) investigated a Scot’s

pine plantation forest in Scotland and found 37 vascular plant species of which 10

species were common to the Scot’s pine forests in this study.

The Scot’s pine plots formed a close grouping according to ordination

with the two Scot’s pine plots at Brittas and Bansha falling slightly outside this

grouping. The plot at Brittas has a number of oak planted in it. Although not as

dominant in the canopy as the Scot’s pine, oak trees are represented well in the S2

and S3 canopy layers, which has effects on the ground flora species and makes the

site similar to the oak-birch-holly WN1 habitat (Fossitt 2000). Bansha also has an

amount of oak (Quercus petraea) in the lower canopy layers within 10m of the

plot, which along with the presence of the oak-birch-holly indicator species,

honeysuckle and bilberry would classify it as a WN1 habitat (Fossitt 2000). The

rest of the Scot’s pine plots cluster very well together due to their similar species

composition which is best described as the Semi natural bog woodland WN7

(Fossitt 2000).

The Scot’s pine plots are some of the most open canopy plots of the

project. The typical crown form of the Scot’s pine trees leaves large open spaces

in the canopy which allows a lot of solar radiation to the ground layer, often

resulting in very high vegetation cover of fast growing light dependent species

(bracken, bramble, willow and birch) in the S1 and S2 layers.

The study by Roche et al. (2009) examined the vegetation of Scot’s pine

forests in Ireland. They found that Scot’s pine supported a high number of native

plant species (116). Using ordination, they divided the vegetation types of Scot’s

pine forests into 4 groups. Group 1 the V. myrtillus-I. aquifolium corresponds to

the plots at Brittas, Bansha and Torc. This vegetation type is the most similar to

that of semi-natural forest communities especially that of oak-birch-holly WN1

(Fossitt 2000). Group 2 (sensu Roche et al. 2009) the C. vulgaris-E. vaginatum

grouping matches the vegetation at Annagh, Ballygawley, Derryhogan and

Gortnagowna and is similar to the Semi natural bog woodland WN7 grouping

(Fossitt 2000). All of the Scot’s pine sites (except Gortnagowna) have a

113

component of peat in their soil make-up. Scot’s pine is often planted on peat sites

due to its tolerance to acid and waterlogged conditions, and also due to its tap

rooting system which ensures stability on soft peaty soils (Horgan et al. 2004).

Group 4: the G. saxatile-A. capillaries grouping matches the vegetation at

Ballylug and has been said to resemble that of acid grasslands GS3 (Fossitt 2000)

more so than any woodland vegetation community. The largely grassland type

plant communities of the Ballylug Scot’s pine plot could be ascribed to the high

light availability in Scot’s pine woodlands allowing colonisation of the forest

floor by the surrounding vegetation types, in this case acid grassland type

(Rodwell and Cooper 1995).

Irish Sitka spruce forests

Sitka spruce forests had the lowest total species richness and diversity index value

of all the forest types examined. The maximum of 11 plant species is very low

compared to the other forest types. The low species richness of Irish Sitka spruce

forests, especially in comparison to native oak forests is a well established fact

(Fahy and Gormally 1998; Boyle and Farrell 2004). All of the plant flora found in

the spruce forests in this study were identified as being found in Sitka spruce

plantations in a past study (French et al. 2008).

Indicator species analysis was not very successful at identifying indicator

species in Sitka spruce forests, but it identified Oxalis acetosella and Thuidium

tamariscinum as indicators of Sitka spruce forests albeit only weak indicators

(based in indicator value). Ferris et al. (2000b) and Wallace et al. (1992) also

noted a lack of ground flora species in young plantation Sitka spruce stands in the

U.K., especially when there was not a history of woodland establishment on the

site. Age was identified as significant in explaining the variability in plant species

between the Sitka spruce plots. Young plots were found to have less vascular

species richness than mature plots in this project. Ferris et al. (2000b) found that

as age increased, so did total species and vascular plant species cover. This result

can be explained by the dense canopy which is formed in mid-rotation Sitka

spruce and other coniferous forests that cause a decline in ground floor species

richness due to shading (Sakura et al. 1985; Hill 1986).

Ordination of the Sitka spruce plots shows a clear separation of the spruce

plots from the other forest types. As Sitka spruce is not a native tree species, there

114

has been a delay in designating it as a habitat in Ireland. Under the Fossitt (2000)

habitat guide, four of the Sitka spruce plots are designated as conifer plantation

WD4, while the remaining five could be classified as oak-birch-holly WN1

habitats. A difficulty in applying this system to non native vegetation, is that some

of the key species which designate a habitat grow in many different habitats, even

previously unrecognised ones. These “new biotic assemblages” (sensu Hobbs et

al. 2006) consist of groupings of flora and fauna which have not been documented

before or are not recognised as an established community. Hard fern Blechnum

spicant is an indicator species of the oak-birch-holly WN1 grouping (Fossitt

2000) and grows in 5 of the 9 Sitka spruce plots, although only 2 of these plots

have either oak, birch or holly present in the canopy. The study by French et al.

(2008) was one of the first to examine Sitka spruce vegetation in Ireland. They

classified Sitka spruce plots into 5 groups, grassland and heath groups (none in

this study), two age classes of closed canopy forest (closest fit for all of the Sitka

spruce plots examined in this project) and bramble dominated forest (no plots in

this study).

The spruce sites also had very low quantities of coarse woody debris,

especially lacking large diameter CWD in the plots. This anomaly has been

identified as a problem in Irish plantation forests (Sweeney et al. 2010a) and may

have implications for biodiversity in these forests. When examining the number of

tree species present in the canopy of Sitka spruce forests, it is noticeable that

often, Sitka spruce dominates the canopy especially in mature sites where it can

be the only tree species present in the plot. The highly competitive growth form of

Sitka spruce has been noted in the past and along with its heavy shading of the

forest floor may lead to low plant species diversity in the sites (Ferris et al.

2000b).

3.5.3 Relationship of these findings to possible fungal species richness

As a precursor to the Chapter 4, the possible fungal species richness of the sites

was estimated from the plant species richness recorded in the plots. It has been

accepted in the past in most ecosystems that the total fungal to plant species

richness ratio stands at 7:1 (Hawksworth 1991). In this project only macrofungi

115

were of interest, and therefore total fungal species richness was not being

measured. Using data from a well studied area (the Hebrides) the recalculated

ratio for macrofungi to vascular plants is 1259:860 or 1.46:1 (Hawksworth 1991;

Dennis 1986). The previous figure is quite similar to one proposed by Villeneuve

et al. (1989) for data from Canadian deciduous and coniferous forests, which was

2:1 macrofungi to vascular plants. In a study of both deciduous and coniferous

forests in the drier region of South Dakota, Gable and Gable (2007) found a

macrofungi to plant species relationship of 1:1.5. Using a meta-analysis of the

total fungal and vascular plant records from Europe, Schmit and Mueller (2007)

found a lower ratio of 1:0.55 macrofungal to plant species. The estimate by

Schmit and Mueller is undoubtedly a lower limit estimate as it is accepted that

there are still many undiscovered macrofungal species in European habitats

(Rodriguez 2000 as cited in Hawksworth and Mueller 2005), which would affect

the ratio if included in the analysis.

If the relationship between vascular plants and macrofungi in Irish forests

is similar to that from a well studied region in Canada (Villeneuve et al.1989),

then the richness of vascular plant species in the examined forest habitats would

be directly related to the species richness of macrofungi in these forests. Table

3.22 and Fig. 3.15 identify the possible macrofungal species richness estimates for

the examined plots calculated from the vascular species richness of the plots.

116

Table 3.22 Vascular species richness and estimated macrofungal species richness of the sites. Also given in column 1 are the plot codes which are used in Fig. 3.15 of the data. Numbers in parenthesis in column 4 are the fungal species excluding the ECM species which would not occur in pure ash plots. Column 5 lists other factors which have been found to (+) increase or (-) decrease fungal richness in forest ecosystems that are present in the sites.

Plot name (code) Tree type Vascular

richness

Estimated

macrofungal richness

Other factors which increase fungal

diversity

Ballykilcavan (Ash1) Ash 6 12 (7.2) - low tree species diversity - first rotation forest

Donadea (Ash2) Ash 15 30 (18) - low tree species diversity

Ross island (Ash3) Ash 8 16 (9.6) - low tree species diversity

Killough (ash4) Ash 20 40 (24) +hazel in understory

St John’s wood (Ash5) Ash 10 20 (12) +hazel in understory + undisturbed forest

Abbeyleix (Oak1) Oak 13 26 + undisturbed forest + high amounts of CWD

Kilmacrea (Oak2) Oak 12 24 + undisturbed forest + high amounts of CWD

Raheen (Oak3) Oak 12 24 + undisturbed forest + high amounts of CWD

Union (Oak4) Oak 10 20 + high amounts of CWD

Tomies woodA (Oak5) Oak 9 18 + high amounts of CWD

Tomies woodB (Oak6) Oak 14 28 + high amounts of CWD

Annagh (SP1) Scot’s pine 9 18

Ballylug (SP2) Scot’s pine 8 16 - low tree species diversity - first rotation forest

Ballygawley (SP3) Scot’s pine 10 20 - low tree species diversity - first rotation forest

+ high amounts of CWD

Derryhogan (SP4) Scot’s pine 11 22

Gortnagowna (SP5) Scot’s pine 10 20

117

Bansha (SP6) Scot’s pine 18 36 +multiple tree species + undisturbed forest

+ high amounts of CWD

Brittas (SP7) Scot’s pine 16 32 +multiple tree species + undisturbed forest

+ high amounts of CWD

Torc (SP8) Scot’s pine 9 18

Ballygawley (SS1) Sitka spruce 11 22

Bohatch (SS2) Sitka spruce 7 14 +multiple tree species

Chevy mature (SS3) Sitka spruce 8 16 - low tree species diversity - first rotation forest

Chevy young (SS4) Sitka spruce 8 16 +multiple tree species

Cloonagh (SS5) Sitka spruce 2 4 - low tree species diversity - first rotation forest

Dooary (SS6) Sitka spruce 4 8 - low tree species diversity - first rotation forest

Moneyteige (SS7) Sitka spruce 2 4 - low tree species diversity - first rotation forest

+ high amounts of CWD

Quitrent (SS8) Sitka spruce 9 18 + high amounts of CWD

Stanahely (SS9) Sitka spruce 4 8 - low tree species diversity - first rotation forest

118

Ash

1A

sh2

Ash

3A

sh4

Ash

5O

ak1

Oak

2O

ak3

Oak

4O

ak5

Oak

6SP1SP2SP3SP4SP5SP6SP7SP8SS1SS2SS3SS4SS5SS6SS7SS8SS9

0

5

10

15

20

25

30

35

40

45

50

55

Macrofungal species

Removed ectomycorrhizal species

Plot

No

. o

f sp

ecie

s

Fig. 3.15 The estimated macrofungal species richness of the plots based on the 2:1 fungi to vascular plant ratio (Villeneuve et al. 1989). The plot codes used are explained in Table 3.22. In the case of ash1-5 the extra bar segment (mycorrhizal) refers to the 40% of the macrofungal species which should be removed from the estimate as ectomycorrhizal fungi would not be found in a forest of pure ash composition. The arrows above the bar refer to other plot characteristics which may increase (up arrow) or decrease (down arrow) relative to the arrow length, the fungal richness and correspond to the +/- information given in Table 3.22.

119

According to this simple calculation, the oak forest would be estimated to

be the most fungal species rich with a mean species richness of 23 (±3.7sd)

species, followed by Scot’s pine, ash and Sitka spruce forests with 23 (±7.2), 14

(±6.8) and 12 (±6.4) fungal species. The relationship between estimated species

richness and actual species richness of macrofungi in these plots is examined in

the next chapter (Section 4.5.5, Chapter 4).

120

3.6 Conclusions

• The forest types were well differentiated based on their vegetation

communities, with plots of the same forest type having a more similar

vascular plant community.

• The ash and Scot’s pine plots were found to be the most species rich

followed by the oak and Sitka spruce forests. Sitka spruce forests were

found to have low levels of species richness and age of the plot was found

to be positively related to the species richness. Future management of

plantation forests should incorporate management techniques which

increase the diversity of plants in these forests, similar to

recommendations made in Iremonger et al. (2006).

• The native oak forests of Ireland are at risk due to large herbivore

browsing. Management techniques along the lines of those proposed by

Perrin et al. (2006) need to be examined in order to promote natural

regeneration in oak forests.

• Possible macrofungal species richness estimates according to the ratio of

vascular plants to macrofungi indicate that oak forests would be the most

macrofungal species rich forest type followed by Scot’s pine, ash and

Sitka spruce forests.

121

Chapter 4: Macrofungal species richness and

diversity in the different forest types

122

123

4.1 Introduction

4.1.1 Fungal diversity in temperate forest ecosystems

Of the estimated 1.5 million fungal species (Hawksworth and Rossman 1997),

only 7% (100,000) have been described so far (Kirk et al. 2008). The total number

of species recorded for Ireland and the U.K. alone, currently stands at over 16,500

(FRDBI 2010). Rodriguez (2000), as cited in Hawksworth and Mueller (2005),

estimate that there are over 25,000 species of fungi (including epigeous and

hypogeous macrofungi, yeasts and lichen-forming fungi) in Europe with another

40,000 yet to be described. Fungi are most diverse in forest ecosystems, and it has

been estimated that the fungal biomass in forest soils exceeds the biomass of all

other soil organisms combined, except roots (Paul 2007). The link between fungi

and forests is so strong that fungal species richness has been shown to increase

with increasing forest area (Peay et al. 2007; Strong and Levin 1975), and the

number of tree species present (Ferris et al. 2000a; Schmit et al. 2005). Forests

also support more rare and threatened fungi, with over 70% of threatened fungi

being found primarily in forests (Arnolds and de Vries 1993).

The fungal biota of forests is often found to be dominated by

ectomycorrhizal species (Humphrey et al. 2000), which normally account for ca.

40% of the macrofungal species richness in forest ecosystems (Watling 1995).

The most diverse ectomycorrhizal genera found in temperate forest ecosystems

are Cortinarius, Russula and Lactarius which have 193, 133 and 67 species in

England alone (Legon and Henrici 2005). The genus Cortinarius contains almost

2000 described species worldwide (Garnica et al. 2005).

An evolving area of conservation research is the area of biodiversity

indicators (Lindenmayer et al. 2000). It has been suggested that fungi could be

used as surrogate indicators for overall biodiversity in forest ecosystems (Ferris

and Humphrey 1999). In the past, fungi have been used as indicators of lichen and

bryophyte species-rich habitats (Norden et al. 2007), and also as indicators of the

presence of other red-listed fungal species (Halme et al. 2009). Forest fungi could

possibly be used as an indicator of collembolan diversity in forests, as past studies

have found that the community structure of Collembola and other

124

microarthropods is clustered on or near fungal sporocarps (O’Connell and Bolger

1997, 2007).

4.1.2 Fungal diversity in Irish forests and forests similar in tree species

composition and climate

In Ireland, there have been very few investigations into the diversity of fungi in

forests and no systematic study of macrofungal diversity in Irish forests. The

Atlantic oak woods in Co Kerry are probably the best studied woodland habitat in

Ireland (Gill et al. 1948; Ramsbottom 1936; Muskett and Malone 1978, 1980).

Grasslands (Carter 1998; McHugh et al. 2001; Mitchel 2006, 2007, 2008, 2009)

along with other habitat types such as sand dune ecosystems (Hassell 1953; Landy

2001) and the relict communities of forest macrofungi of the Burren (Harrington

1996, 2003; Harrington and Mitchell 2002) have also been documented.

Previous studies have commented on the wealth of macrofungi found in

Irish oak forest habitats (Gill et al. 1948; Ramsbottom 1936). The British

Mycological Society’s visit to the sessile oak woods of Killarney, Ireland, in

1935, identified 39 macrofungal species from one woodland area from a one-day

foray. The macrofungi of forests in the Counties Wicklow and Dublin have also

received attention with a number of published lists of macrofungi for these vice-

counties (Pim 1989; Muskett and Malone 1978; 1980). Counties Wicklow (with

the largest percentage of woodland cover) and Kerry (with the largest percentage

cover of semi-natural woodlands), have the first and second highest numbers of

macrofungal records in Ireland. The macrofungi of other woodland areas in

Ireland have received little or no attention. O’Hanlon and Harrington (in press)

estimated that 25 of the 26 vice counties in Ireland have less than 50% of their

macrofungal species recorded, and that Irish woodlands may reveal many

unrecorded macrofungal species. The work by Harrington in the Burren area,

although not investigating macrofungi in forest ecosystems, does give an idea of

the diversity of forest macrofungi which at one time existed in Irish native

woodlands: 88 macrofungi were recorded from three years sampling (Harrington

2003). Thirty-five percent of these species were putative ectomycorrhizal species

that were ectomycorrhizal on mountain avens (Dryas octopetala), in the absence

of host tree species in the Burren. There was also a number of woodland

125

saprotrophs e.g. Lepista nuda, Mycena spp. present. It has been proposed that the

ectomycorrhizal and woodland saprotrophic macrofungi present in the area are

relict species from the extensive pine forests which grew in the area over 2000

years ago (Harrington 1996).

The macrofungal diversity of British native and plantation woodlands have

received much more study that of Irish forests, and three recent systematic studies

have revealed that plantation woodlands in the U.K. support levels of macrofungal

richness comparable to those of semi-natural woodlands (Ferris et al. 2000a;

Humphrey et al. 2000; Newton et al. 2002). The first study investigated the fungal

biota of Scot’s pine and Norway spruce habitats in the U.K. and found that there

was a high richness of fungi (343 species) in these plantation forests. The second

study compared the macrofungal richness and community structure of plantation

forests of exotic (Sitka spruce) and native tree species (Scot’s pine and oak). They

also found that plantation forests can provide a supplementary habitat for native

fungal species. The final study, by Newton et al. (2002), examined the distribution

of the threatened group of macrofungi known as stipitate hydnoid fungi in

Scottish native and plantation woodlands. They found that plantation woodlands

could support a number of these endangered macrofungi, and agreed with the

studies by Ferris et al. (2000a) and Humphrey et al. (2000), that plantation

woodlands in the U.K. can be managed to provide a sufficient habitat for many

native forest macrofungi. In its home range of North America, Sitka spruce is

known to show low levels of ECM specificity, and its ability to form

ectomycorrhizas with many different ECM fungi is noted as a possible reason for

the tree species survival and success in its native habitat, which is normally very

tree species rich (Alexander and Watling 1987). The ability of Sitka spruce to

form ectomycorrhizal linkages with many different fungi may increase its fitness,

by allowing access to many different common mycorrhizal networks and a share

in the carbon and limiting nutrients which they carry (Smith and Read 2008).

4.1.3 Factors affecting macrofungal diversity in forests

Forests are now seen as one of the key reservoirs for biodiversity in Europe (EEA

2010). Conserving biodiversity is important as: (a) declining species richness

leads to declines in overall ecosystem functioning, (b) at least one species per

126

functional group is essential for ecosystem functioning and (c) the nature of an

ecosystem’s response to declining biodiversity depends on the species lost

(Naeem 1999). Functional diversity (defined as a group of species sharing similar

ecological features that can influence their performance in nature; Keil et al.

2008) has also been shown to be an important factor affecting ecosystem

functioning (see Gaston and Spicer 2007; Bolger 2001). Humphrey et al. (2000)

and Ferris et al. (2000a) examined the functional diversity of fungi in forests in

the U.K., and found that the numbers of species within different functional groups

varied between forest types, and that certain structural and abiotic variables were

responsible for these functional group variations.

There have been numerous factors identified which are known to affect the total

numbers of macrofungal species, and species within defined functional groups

found in temperate forest ecosystems. These are:

• Deciduous forests have been found to harbour a larger number of fungal

species than coniferous forests (Bills et al. 1986; Humphrey et al. 2000).

• It has been found that the number of tree species present in a forest is

positively related to the number of macrofungal species in the forest (Ferris et

al. 2000a; Schmit et al. 2005). This is possibly due to the increasing number of

symbiotic partners available in multi-tree species forests for ectomycorrhizal

fungi (Ishida et al. 2007) and due to the substrate specificity demonstrated by

many litter-decay fungi (Ludley et al. 2008). Certain fungal parasites are tree

species specific (e.g. Heterobasidion annosum only attacks coniferous trees)

and therefore increases in the number of tree species can also cause an increase

in parasitic fungi present.

• Stand age has been shown to be related to the number of macrofungal species

recorded in coniferous forests (O’Dell et al. 1999; Smith et al. 2002). It has

been noted that as some coniferous forests age, they acquire a large number of

macrofungal species that fruit infrequently, thus giving the appearance of

reduced macrofungal species richness if sporocarp sampling is used to record

the fungal biota.

127

• The quantity of CWD present in a forest has been shown to be positively

related to the species richness of wood-decay macrofungi (Ferris et al. 2000a).

The decay stage and overall quality of the CWD has also been linked to the

species richness of wood-decay macrofungi present in forest ecosystems

(Sippola and Renvall 1999).

• The proximity of fungal inoculum sources (as spores or mycelium) has been

identified as important for increasing the species richness of macrofungi in

temperate forests. Distance to retained mature and ectomycorrhizally diverse

trees has been shown to allow for the inoculation of young trees with existing

ectomycorrhizal fungi (Jumpponen et al. 2002; Ashkannejhad and Horton

2006; Cline et al. 2005).

4.1.4 Difficulties in estimating fungal diversity in forests

Creating an inventory of the entire fungal biota of a forest area is a difficult task.

The ephemeral nature of fruitbodies and the effects of weather on fruitbody

production make fungal diversity studies difficult (Ohenoja 1993; Gulden et al.

1992; Carrier 2003; Krebs et al. 2008). Sporadic fruiting means that long-term

studies are required. Many long-term studies (>20 years) were still finding new

species (Straatsma and Krisai-Greilhuber 2003; Tofts and Orton 1998; Orton

1987). Often the overwhelming number of species which can be found in even a

relatively small area also present difficulties (for review see Watling 1995). These

obstacles have been an impediment to macrofungal diversity studies, but recent

molecular techniques such as pyrosequencing are revealing extraordinarily high

fungal diversity from small sample areas (for review see Hibbett et al. 2011).

In cases where the total species richness cannot be recorded, richness

estimators can be used to extrapolate from existing data, and provide an estimate

of population diversity to which confidence intervals can be assigned (Colwell

and Coddington 1994). Species richness estimators have been used in the past for

many species groups, including fungi (Schmit et al. 1999; Unterseher et al. 2008).

Many of the techniques for estimating species richness are based on models

devised by Anne Chao and her colleagues (Chao 1984, 1987; Chao et al. 2000).

Some of the estimation methods most commonly used are Chao1 (Chao 1984),

Chao2 (Chao et al. 2000), ACE and ICE (Chao and Lee 1992), jackknife

128

estimates (Burnham and Overton 1979), bootstrap richness estimator (Smith and

Belle 1984) and the Michaelis–Menten richness estimator (Raaijmakers 1987).

The Chao1, Chao2, jacknife, bootstrap, ICE, ACE and Michaelis- Menten

estimators have been used in fungal diversity studies (Schmit et al. 1999;

Unterseher et al. 2008). Many of the older estimators such as Chao1 and Chao2

use the ratio of singletons (species only found in one sample) to doubletons

(species found in two samples) to calculate the species estimate. The most recent

estimators, the coverage estimators ICE and ACE, use the ratio of rare to frequent

species to estimate the richness. In this study, rarefaction was employed to

estimate the number of species which would have been discovered had a lesser

sampling intensity been used. The purpose of this is to equalise the sampling

intensity between the least and the most intensive sample and thus make them

statistically comparable. Species richness estimation was carried out using the

Chao2, ACE and ICE species richness estimators.

129

4.2 Aims of this chapter

The aims of this chapter are:

• To enumerate the macrofungal species richness of forest plots belonging to

four principal forest types that are represented in Ireland. It was expected that

there would be considerable differences in terms of species richness,

functional group richness and fungal community composition (addressed in

Chapter 5). Between the forest types it was expected that native forest types

would have greater species richness than non-native plantation conifers, as

would forests composed of ectomycorrhizal host trees compared to non-

ectomycorrhizal hosts. On this basis, a working hypothesis was that oak

forest would have the highest species richness and ash the least.

• To examine if variables within each forest type, such management history and

stand age (as described in Ch. 3), have a significant influence on macrofungal

species richness and functional group richness. As forest age has been shown

to be negatively related to the numbers of macrofungal species in some

temperate coniferous forests (O’Dell et al. 1999; Smith et al. 2003), it was

expected that the macrofungal species numbers would be higher in younger

coniferous forest sites than in the mature coniferous forest sites.

• To examine if plots with a history of forest use (for example, second rotation

stands on sites that had been previously forested) would have higher species

richness of macrofungi than forests which did not have a history of use as

forests (first rotation forests). In previous work, the availability of fungal

inoculum (either as spores of mycelium) has been shown to be positively

related to fungal species richness in forests (Jones et al. 2003; Dickie and

Reich 2005; Outerbridge and Trofymow 2004).

130

4.3 Materials and Methods

4.3.1 Sites

Details of the sample plots within the forest sites sampled are given in Section

3.3.1, Chapter 3 (Table 3.2). Plots were visited from 2007-2009 inclusive during

the autumn (August – November). Most plots were sampled on at least three dates

in each year. Each of the forest types was represented by at least two replicate

plots that were selected on the basis of similarity of vegetation cover, stand age

and management history.

4.3.2 Macrofungal assessment

In each forest site a 100m2 permanent plot (2x50m) was set up. The plot was

divided into 5 sub-plots of 10m x 2m. All macrofungi inside the plots were

identified in situ where possible, and sporocarps of unidentified species were

retained for later identification. Species were identified using standard general

texts (Breitenbach and Kranzlin 1984-2005; Phillips 2006; Moser 1983) and more

specialized texts for Lactarius (Heilmann-Clausen et al. 1998), Entoloma

(Noordeloos 1992; 2005), Cortinarius (Brandrud et al. 1990-1997), Boletus

(Muñoz 2005) and Tricholoma (Riva 2003). Specimens of rare or critical species

were dried and retained as herbarium samples at the Department of Life Sciences,

University of Limerick, Limerick, Ireland. Sporocarps of ectomycorrhizal species

were frozen at -80ºC for molecular analysis and sequencing (see Chapter 6).

Digital photographs were taken of the macro- and micro-morphological features

of rare species (Folder 2, Appendix 1). Species were assigned to functional groups

based on their primary mode of nutrition as used in Newton and Haigh (1998),

Ferris et al. (2000a) and Humphrey et al. (2000); i.e.

1. Litter-decay species found on forest floor.

2. Ectomycorrhizal fungi.

3. Parasitic on living plant material, or very fresh dead material.

4. Wood-decay species found on decaying wood.

The abundance of fungal sporocarps in each plot was recorded by counting

the numbers of sporocarps of each species in each sub-plot. The number of

fruiting bodies was recorded in each sub-plot up to a maximum of 12 fruitbodies.

131

Species outside of the plot, but within the same stand i.e. generally within 50 m of

the perimeter of the plot were also recorded, but only on a presence/absence basis.

4.3.3 Statistical analysis

For each plot and forest type, a number of parameters of macrofungal species

abundance were determined for all macrofungal species collectively, and for each

functional group of macrofungi:

• Total species richness = the cumulative number of macrofungal species

found in a plot or forest type over the entire sampling period of three

years. The total species richness per forest type = Σxi where x is the total

species richness of each plot in forest type i

• Mean species richness per plot visit = cumulative number of macrofungal

species found in a plot over the entire sampling period, divided by the

number of visits to that plot. By extension, the mean species richness per

forest type = cumulative number of macrofungal species found in all plots

of a specific forest type over the entire sampling period, divided by the

number of visits to all plots within that forest type.

• Log number of fruitbodies per plot visit = the natural logarithm of the

number of fruitbodies found on a single plot visit to a specified plot.

The forest plot visit matrix was created to use the qualitative presence/absence

data for species in each plot visit as a quantitative measurement. It was created by

pooling all of the macrofungal presence/absences from each plot visit into a site

by species matrix. Therefore these matrices were 219 species by 25, 36, 31 and 55

plot visits for ash, oak, Scot’s pine and Sitka spruce forests respectively.

132

4.3.4 Species richness comparisons and the estimation of fungal diversity in

the forest types and sites

Rarefaction analysis

To compare the species richness of the different plots and forest types at a similar

sampling intensity, sample-based rarefaction curves with 95% confidence

intervals were calculated, using the computer program EstimateS (Colwell 2004).

In these analyses, plots were sampled randomly with re-placement, over 500

permutations of the data, because otherwise confidence intervals are meaningless

in the upper end of the rarefaction curve (Colwell et al. 2004). The forest plot

visit matrix was used in these analyses. Sample-based rarefaction was used

instead of individual-based rarefaction, as fungal species have been shown to be

non-randomly distributed (Taylor 2002, Tedersoo et al. 2003). Rarefaction curves

read from right to left and are used for estimating species richness at smaller

sample sizes. Most cannot be used to estimate total species diversity of a

community, but work by Colwell et al. (2004) has allowed the estimation of

species richness from rarefaction curves. In this project, rarefaction was only used

to compare species richness in plots and forest types at lower sampling intensity.

Estimation of total species richness was conducted through species richness

estimators.

Species richness estimation

In order to statistically estimate the undiscovered species richness in the forest

types, and therefore rate the efficiency of a three year study to discover

macrofungal richness in temperate forests, species richness estimators were used

in EstimateS. The most recent estimators, the coverage estimators ICE and ACE

use the ratio of rare to frequent species to estimate the richness. The ICE estimator

is useful as it only needs presence/absence data, is fairly robust to patchy

distributions (Chazdon et al. 1998), and does not need large sample sizes to reach

an asymptote (Magurran 2004). The ACE richness estimator uses the ratio of

abundances of the common species to the rare species to estimate the total species

richness of the ecosystem. It has the drawback of requiring abundance data, and

has been shown to be strongly affected by patchy distributions. The use of

abundance-based richness estimators on fungal data may not be tenable, as these

estimators assume a larger degree of homogeneity between samples than the

133

incidence-based estimators. Fungal populations have been shown to be non-

randomly distributed (Taylor 2002) in nature and so abundance based estimators

may fail to return likely estimates.

The forest plot visit data matrices from the different forest types were used

for the species richness estimators. As rareness is an arbitrary figure, depending

on the organism group being investigated, the ecosystems examined and the

overall purpose of the investigation, cut-off points for rareness of fungal species

were created for each forest type. Using Gaston’s (1994) definition of rareness,

the cut-off point was set at the lower quartile of species abundance values from

samples. In all of the forest types and sites, this was set at two plot visits. This

meant that any species that was found on less than two occasions was defined as

rare. The Chao2 diversity estimator was also calculated for the sites and forest

types. This richness estimator is a presence/absence based estimator that uses the

ratio of species found in only one sample to species found in two or more

samples. It is useful as a minimum richness estimate is cases where species

richness estimates from other estimators have not reached an asymptote (Longino

et al. 2002). The % discovered species richness of the plots and the forest types is

calculated as:

=Actual species richness of forest typeX / Estimated species richness of forest typeX *100

4.3.5 Species diversity and evenness analysis of the sites

Diversity indices were applied to the forest types and plots to compare their

species diversity. The indices were calculated in EstimateS for the different forest

types using the sporocarp abundance data from the plot visits to the forest plots.

The Shannon diversity index, although having many critics (see Magurran 2004),

is one of the most popular diversity indices available and its values were

computed to facilitate comparison between this project and past projects.

Shannon-Wiener index of diversity:

H' = -ΣPi lnPi

134

where Pi is the probability of sampling the ith species among all species (Shannon

and Weaver 1949). This index has been used in similar studies (Bills et al. 1986;

Kranabetter 2004). It is popular because it takes into account the total number of

species and also the evenness with which the individuals are distributed between

each species (Zak and Willig 2004). The index can be modified to give a measure

of the evenness of the sample by dividing H’ by the natural log of the total species

found in the sample or H’ ÷ ln s.

Simpson’s diversity index (Simpson 1949) is another commonly used

index of diversity. It measures the probability that two individuals randomly

selected from a sample will belong to the same species. Its use is advocated by

Magurran (2004) as it is intuitively meaningful and does not make assumptions

about the degree of sampling.

Simpson’s diversity index:

D= Σ ni(ni-1) N(N-1)

Where ni is the number of individuals of the ith species and N is the total number

of individuals.

4.3.6 Species richness of functional groups in the different forest types

The mean and standard deviation of total species richness, mycorrhizal, litter-

decay, wood-decay and parasitic species richness per plot visit was calculated. As

the numbers of fungal species in the different functional groups were found to be

normally distributed, the parametric statistical test one-way analysis of variance

(ANOVA) with a Hochbergs GT2 post-hoc test using PASW version 18 (SPSS

Inc., Chicago, Illinois) was used on untransformed species per plot visit data. The

data was examined to test for significant differences between (a) mean species

richness of the forest types, and (b) mean species richness of the functional groups

in the forest types. The Hochbergs GT2 test is useful when sample sizes vary

between variables (Day and Quinn 1989).

135

4.3.7 Functional group frequencies across the forest types

To test if the relative numbers of species in each of the functional groups was

similar across oak, Scot’s pine and Sitka spruce forests, a one-way classification

chi-square test (following Hawkins 2009) was employed. The species richness in

the different functional groups was compared for each forest type pair separately.

Any relationships between the species richness in each of the functional

groups was tested using Pearson’s correlation coefficient on the species richness

of the functional groups per plot visit.

4.3.8 Influence of site variables on species richness

The effects of site nominal variables (age, rotation stage and other variables; see

Table 4.11) on the mean species richness and the mean richness of species in each

functional group per plot visit in each forest type were tested using a generalized

linear model with Poisson error distribution and log-link function (for analysis of

species count data). For ash, the effect of site management (semi-

natural/managed) and forest age (young/mature) was tested; for oak, the effect of

grazing by large mammals (grazed/non-grazed) was examined; for Scot’s pine the

effect of rotation stage (first/second rotation) and forest age (young/mature) was

tested and for Sitka spruce forests the effect of rotation stage (first/second

rotation) and forest age (young/mature) was tested for relationships with the mean

species richness and the mean richness of species in each functional group per

plot visit.

The generalized linear model procedure was carried out in PASW version

18. The dependent variables species richness, litter decay, ectomycorrhizal or

wood decay species richness per plot visit was entered and tested separately. The

Poisson log-linear model was selected as the model type, following Humphrey et

al. (2000). The factors entered in the model differed depending on the forest type

being tested e.g. presence of grazing in the oak forest type. A custom model was

created using the type 3 sum of squares option and only the main effects of the

model were examined. The effect of interaction between the nominal variables

was not examined as there was no a priori reason to do so (Hawkins 2009). The

Pearson’s chi-square option was chosen as the scale parameter method as

136

McCullagh and Nelder (1989) found that the Pearson’s chi-square estimate can

obtain more conservative variance estimates and significance levels. All of the

levels within each factor (i.e. grazed vs. non-grazed) were examined pairwise

using the sequential Sidak adjustment for their effect on the dependent variable.

Once the model was created, the likelihood ratio chi-square test identified

if the model could explain more variation in the dependent variable than variance

explained by the null model. Next, the influence of the factor on the dependent

variable was identified by examining the difference in the means of the dependent

variable (i.e. species richness) across all levels of the factor (i.e. grazed vs. non-

grazed). If statistically significant according to sequential Sidak adjustment for

pairwise comparisons, the factor level with the higher mean difference for the

dependent variable can be taken as having a positive effect on the dependent

variable.

4.3.9 Seasonal effects on fungal phenology

The numbers of fruitbodies were not normally distributed and therefore they were

natural-log transformed. Log transformations of biological variables often results

in better correlations with other variables (Burton 1998). To analyse the

relationship between meteorological variables (rainfall and temperature)

Spearman’s correlation coefficient was carried out between the species richness,

species richness of each functional group, log total number of fruitbodies and log

total number of fruitbodies in each functional group per plot visit and the

meteorological variables mean annual rainfall, mean annual temperature, total

annual rainfall, mean daily temperature in June; July and August, and mean daily

temperature and rainfall 1, 2, 3 and 4 weeks prior to sampling the plots. If

variables were found to be significantly correlated, then linear regression was

carried out in PASW version 18 using stepwise removal of independent variables

to find the linear model which best explained the variation in the dependent

variable. Both regression and correlation were carried out in PASW.

137

4.4 Results

4.4.1 Macrofungal species in the forest plots

A total of 409 species were identified from the sites over the three-year sampling

period (Table 4.2), with 215 different species photographed (Folder 2, Appendix

1). Of these, 219 species were found within the permanent plots. The remaining

190 were found within the same stand and generally with 50 m radius of the plot.

Of the species found within the plots, 56 were found in ash forests, 113 in oak

forests, 89 in Scot’s pine forests and 144 in Sitka spruce forests. Of the species

found in the plots, 17 species were found in all forest types, 34 species were found

in three forests types, 54 were found in two forest types and the remainder found

in only one forest type. Ash and oak had 41 species in common; ash and Scot’s

pine shared 21 species while ash and Sitka spruce shared 30 species. Oak and

Scot’s pine shared 44 species and oak and Sitka spruce shared 65 species. Sitka

spruce and Scot’s pine shared 58 species.

138

Table 4.2 Species list of macrofungi found during the FUNCTIONALBIO project. Listed are the numbers of plot visits to a forest type on which the species was found inside (and outside) the plot over the three years. * denotes presence on the UK Red data list. FG= functional group, L= litter-decay, M= Ectomycorrhizal, P= Parasitic and W= wood-decay fungi. Ash Oak SP SS

Plots visits: 25 36 31 55

Species name FG

Agaricus langei (F.H. Møller & Jul. Schäff.) Maire 1952

W 0 0 0 0 (1)

Amanita citrina Pers. M 0 1 (1) 0 0

Amanita citrina v.alba (Gillet) E.-J. Gilbert M 0 0 (1) 0 0 (1)

Amanita excelsa Fr. M 0 0 0 0 (1)

Amanita fulva Schaeff. M 0 1 2 2

Amanita phalloides Fr. M 0 1 0 0

Amanita rubescens (Pers.) Gray M 0 1 (1) 1 3 (1)

Amanita vaginata (Bull.) Fr.. W 0 0 0 0

Amanita virosa Gonn. & Rabenh. M 0 0 0 0

Ampulloclitocybe clavipes (Pers.) Redhead, Lutzoni, Moncalvo & Vilgalys 2002

W 0 0 0 0

Arcyria denudata (L.) Wettst., 1886 W 0 1 0 0

Armillaria cepistipes Velen. 1920 P 0 0 0 0

Armillaria mellea (Vahl) P. Kumm P 1 4 1 1

Armillaria ostoyae (Romagn.) Herink P 0 0 0 0

Asterophora parasitica (Bull.) Fr. P 0 0 (1) 0 0

Auriscalpium vulgare Gray (1821), W 0 0 1 (1) 0

Badhamia sp.1 W 0 1 0 0

Baeospora myosura (Fr. ex Fr.) Sprig W 0 0 2 0

Bertia moriformis (Tode) De Not W 0 0 0 0

Bisporella citrina (Batsch : Fries) Korf & Carpenter W 1 1 0 3

Boletus appendiculatus Schaeff. 1763 M 0 0 0 0

Boletus badius Pers. M 0 2 0 2

Boletus calopus Pers. M 0 0 0 0

Boletus chrysenteron Bull. M 0 1 0 0

Boletus edulis Bull. M 0 0 0 1 (1)

Boletus luridiformis Rostk. M 0 0 0 0

Boletus porosporus (Imler ex Bon & G. Moreno) Watling

M 0 0 0 0

Boletus pruinatus Fr. & Hök M 0 0 0 0

Boletus subtomentosus Pers. M 0 0 0 1

Calocera sp. W 0 0 0 0

Calocera viscosa (Persoon : Fries) W 0 0 0 4

Camarophyllopsis atropuncta (Pers.) Arnolds * L 0 (4) 0 0 0

Cantharellus aurora (Batsch) Kuyper * M 0 0 0 0

Cantharellus cibarius Fr. M 0 1 (1) 1 1 (1)

Cantharellus tubaeformis (Bull.) Fr. M 0 0 1 1

Ceratostomella ampullasca (Cooke) Sacc. 1882 L 0 1 0 0

Chalciporus piperatus (Bull.) Bataille M 0 0 0 0

Chamaemyces fracidus(Fr.) Donk L 0 0 0 0

Chlorociboria aeruginascens (Nyl.) Karst. ex Ram W 1 2 0 0

Chlorophyllum rhacodes (Vittad.) Vellinga L 0 0 0 0

Chroogomphus rutilus (Schaeff.) O.K. Mill. M 0 0 0 (1) 0

139

Cineromyces lindbladii (Berk.) Gilb. & Ryv W 0 0 0 0 (1)

Clavaria acuta Fr. L 0 0 0 0

Clavaria fragilis Holmsk. L 0 0 0 1

Clavulina cinerea (Bull.) J. Schröt. L 0 0 0 1

Clavulina coralloides (L.) J. Schröt. L 0 1 1 1

Clavulina rugosa (Bull.) J. Schröt. L 1 1 1 4 (1)

Clavulinopsis helvola (Pers.) Corner M 0 0 0 0

Clavulinopsis umbrinella (Sacc.) Corner * L 0 0 0 1

Clitocybe brumalis (Fr.) Gillet L 0 0 0 0

Clitocybe fragrans Sowerby L 0 0 0 1 (1)

Clitocybe geotropa (Bull.) Fr. L 0 0 0 0

Clitocybe gibba (Pers.) P. Kumm. L 0 1 0 1

Clitocybe nebularis (Batsch) Quél. L 0 0 0 0

Clitocybe vibecina (Fr.) Quél. L 0 1 3 3 (1)

Clitopilus prunulus (Scop.) P. Kumm. 1871 L 0 0 0 1

Collybia confluens (Pers.) P. Kumm. L 0 0 0 0

Collybia distorta (Fr.) Quél. 1872 L 0 0 0 (1) 0

Collybia dryophila (Bull.) P. Kumm. L 0 1 1 2

Collybia erythropus (Pers.) P. Kumm. L 0 0 0 0

Collybia fusipes (Bull.) QuTl. L 0 0 0 0

Collybia peronata (Bolton) P. Kumm. 1871 L 0 0 0 1

Conocybe blattaria (Fr.) Kühner L 0 0 0 0

Coprinopsis laanii (Kits van Wav.) Redhead W 0 0 0 1

Coprinus lagopus (Fries, 1821) W 1 0 0 0

Coprinus plicatilis (Curt.: Fries) W 0 0 0 0

Ophiocordyceps forquignonii (Quél.) G.H. Sung, J.M. Sung, Hywel-Jones & Spatafora

P 0 0 0 0

Elaphocordyceps longisegmentis (Ginns) G.H. Sung

P 0 1 1 1

Cortinarius acutus (Pers.) Fr. 1838 M 0 3 2 (1) 3 (2)

Cortinarius anomalus Fr. M 0 0 1 3

Cortinarius bolaris (Pers.) Fr. M 0 1 0 0

Cortinarius brunneus (Pers.) Fr. M 0 0 0 0

Cortinarius camphoratus (Fr.) Fr. * M 0 0 0 0

Cortinarius caninus Fr. M 0 0 0 (1) 0

Cortinarius cinnamomeus (L.) Fr. M 0 0 0 (1) 4 (1)

Cortinarius decipiens (Pers.) Fr. M 0 0 0 0

Cortinarius delibutus Fr. M 0 0 0 0

Cortinarius evernius Fr. M 0 0 1 1 (2)

Cortinarius flexipes Fr. M 0 4 (1) 2 3

Cortinarius flexipes var. flabellus (Fr.) Lindst. & Melot

M 0 0 0 1

Cortinarius flexipes var. inolens H. Lindstr. 1998 M 0 2 0 0

Cortinarius hemitrichus (Pers.) Fr. M 0 0 1 0

Cortinarius hinnuleus Rob. Henry M 0 0 0 0

Cortinarius imbutus Fr. 1838 M 0 0 0 0 (1)

Cortinarius limonius (Fr.) Fr. 1838 * M 0 0 0 0 (1)

Cortinarius malachius (Fr.) Fr. M 0 0 0 0

Cortinarius obtusus Fr. M 0 0 0 4 (1)

Cortinarius ochroleucus (Schaeff.) Fr. M 0 0 0 0

Cortinarius privignus (Fr.) Fr. M 0 0 0 0

Cortinarius rubellus Cooke M 0 0 0 2 (2)

Cortinarius sanguineus (Wulfen) Fr. M 0 1 0 (1) 1

Cortinarius saturninus (Fr.) Fr. M 0 0 0 0

140

Cortinarius scandens Fr. M 0 0 0 1

Cortinarius semisanguineus (Fr.) Gillet M 0 0 1 (1) 0

Cortinarius stemmatus Fr. 1838 M 0 0 0 (1) 0

Cortinarius stillatitius Fr. M 0 2 1 1

Cortinarius torvus (Fr.) Fr. M 0 0 0 0

Cortinarius umbrinolens P.D. Orton M 1 1 1 (1) 1 (1)

Cortinarius venetus (Fr.) Fr. 1838 M 0 0 0 1

Cortinarius vernus H. Lindstr. & Melot M 0 0 0 0

Cortinarius vibratilis Fr. M 0 0 0 0

Craterellus cornucopioides (L.) Pers. M 0 0 0 0

Craterellus lutescens (Fr.) Fr. 1838 M 0 0 0 0

Crepidotus applanatus (Persoon, 1796) W 0 0 0 0

Crepidotus luteolus (Lambotte) Sacc., 1887 W 0 0 0 1

Crepidotus mollis (Fries) Staude W 2 (1) 2 0 0

Crepidotus variabilis (Pers. ex Fr.) Kummer W 3 4 0 1

Cudoniella aciculare (Bull.) J. Schröt. L 1 1 0 0

Cylindrobasidium laeve (Pers.) Chamuris P 0 0 0 1

Cystoderma amianthinum (Scop.) Fr. L 0 0 0 (1) 4 (1)

Dacrymyces chrysocomus (Bull.) Tul. 1860 W 0 1 0 0

Dacrymyces stillatus Nees ex Fr. Syn W 1 1 0 2

Daedalea quercina L. ex Fr. Syn W 0 0 0 0

Daldinia concentrica (Bolt. ex Fr.) W 0 (1) 0 0 0

Lachnum niveum (R. Hedw.) P. Karst. W 0 0 0 0

Elaphomyces granulatus Fr. M 0 0 1 1

Enteridium lycoperdon (Bull.) M.L. Farr W 1 1 0 1

Entoloma cetratum (Fr.) M.M. Moser M 0 0 4 (1) 3 (1)

Entoloma clypeatum (L.) P. Kumm. M 0 0 0 0

Entoloma conferendum (Britzelm.) Noordel. M 0 2 3 (1) 2 (2)

Entoloma conferendum var. pusillum (Velen.) Noordel.

M 0 0 0 0

Entoloma cuspidiferum Noordel. 1980 M 0 0 0 1

Entoloma lividocyanulum Noordel M 0 0 0 0

Entoloma nitens (Velen.) Noordel. 1979 M 0 0 0 0

Entoloma nitidum Quél. M 0 0 0 0

Entoloma queletii (Boud.) Noordel M 0 0 0 0

Entoloma rhodopolium (Fr.) anon M 0 0 0 1

Entoloma sericellum Fr. M 0 0 0 0

Entoloma turbidum (Fr.) M 0 0 0 0

Entoloma vinaceum var. fumosipes (Scop.) Arn. & Noord.

M 0 0 0 0

Fistulina hepatica (Schaeff.) With. P 0 1 0 0

Flammulaster granulosus (J.E. Lange) Watling W 0 0 0 0

Flammulina velutipes (Curtis) Singer W 0 0 0 0

Fomes fomentarius (L.) J.J. Kickx P 1 0 0 0

Galerina atkinsoniana A.H. Sm L 0 0 0 0

Galerina calyptrata P.D. Orton 1960 L 0 0 0 0 (1)

Galerina marginata (Batsch) Kühner L 1 1 (1) 0 0

Ganoderma applanatum (Pers.) Pat W 0 0 0 0

Geastrum triplex Jungh. L 0 0 0 0

Geoglossum cookeanum Nannf. L 0 0 0 0

Geoglossum fallax E.J. Durand L 1 0 0 (1) 0

Hyphodontia crustosa (Pers.) J. Erikss W 0 0 0 0

Gymnopilus bellulus (Peck) Murrill 1917 W 0 1 1 1

141

Gymnopilus junonius (Fr.) P.D. Orton W 0 0 0 0

Gymnopilus penetrans (Fr.) Murrill W 0 0 1 (1) 1

Gymnopilus picreus (Pers.) P. Karst. 1879 W 0 0 0 (1) 0

Handkea excipuliformis (Scop.) Kreisel W 0 0 0 0

Handkea utriformis (Bull.) Pers W 0 0 0 0

Hebeloma crustuliniforme (Bull.) Quél. M 0 0 0 0

Hebeloma laterinum (Batsch) Versterh M 0 0 0 0

Hebeloma mesophaeum (Pers.) Quél. M 0 0 0 0

Hebeloma pusillum J.E. Lange M 0 0 0 0

Hebeloma sinapizans (Paulet) Gillet 1874 M 0 0 0 0

Helvella crispa (Scop.) Fr. L 0 0 0 0

Helvella lacunosa Afzel. L 0 (1) 0 0 1

Helvella macropus (Pers.) P. Karst. L 1 0 0 0

Hemimycena gracilis (Quél.) Singer L 0 0 0 3

Hemimycena lactea (Pers.) Singer L 0 0 0 0

Hericium sp. 0 (1) 0 0 0

Heterobasidion annosum (Fr.) Bref. P 0 0 2 2 (1)

Hydnum repandum L. 1753 M 0 0 (2) 2 1

Hydnum rufescens Pers. M 0 0 1 0

Hygrocybe ceracea (Wulfen) P. Kumm. L 0 0 0 0

Hygrocybe conica (Scop.) P. Kumm. L 0 0 0 0

Hygrocybe laeta var. laeta (Pers.) P. Kumm. L 0 0 0 0

Hygrocybe quieta (Kühner) Singer L 0 0 0 0

Hygrocybe virginea (Wulfen) P.D. Orton & Watling L 1 0 0 (1) 0

Hygrocybe vitellina (Fr.) P. Karst., L 0 0 1 0

Hygrophoropsis aurantiaca (Wulfen) Maire M 0 0 1 0 (1)

Hymenochaete cruenta (Pers.) Donk 1959 W 0 0 0 1

Hymenoscyphus calyculus (Sowerby) W. Phillips W 0 0 0 0

Hyphoderma praetermissum (P. Karst.) W 0 0 0 0

Hypholoma capnoides (Fr.) P. Kumm W 0 0 0 3

Hypholoma fasciculare (Huds.) P. W 2 (1) 3 0 (1) 7 (1)

Hypholoma marginatum (Pers.) J. Schröt. W 0 0 0 2

Hypomyces chrysospermis Tul. & C. Tul. P 0 1 (1) 0 0

Hypoxylon fragiforme (Scop.) J. Kickx f. W 0 1 0 0

Hypoxylon fuscum (Pers.) Fr W 1 1 0 0

Hypoxylon multiforme (Fr.) Fr. W 1 0 0 0

Hypoxylon rutilum Tul. & C. Tul. 1863 W 0 0 0 0

Inocybe adaequata (Britzelm.) Sacc. M 0 0 0 0

Inocybe asterospora (Quél.) M 0 0 0 0

Inocybe bongardii (Weinm.) Quél. M 0 0 0 0

Inocybe cincinnata (Fr.) Quél. M 0 0 0 0

Inocybe geophylla (Fr.) P. Kumm. 1871 M 0 0 2 (1) 1

Inocybe geophylla var. lilacina (Peck) Gillet M 1 0 0 1

Inocybe godeyi Gillet M 0 0 0 0

Inocybe hystrix (Fr.) P. Karst M 0 0 0 0

Inocybe lanuginosa Cooke M 0 1 2 1

Inocybe lanuginosa var. ovatocystis (Boursier & Kühner) Stangl

M 0 0 0 0

Inocybe maculata Boud. M 0 0 0 0

Inocybe napipes J.E. Lange M 0 2 2 0

Inocybe nitidiuscula (Britzelm.) Sacc. 1895 M 0 0 0 0

Inocybe phaeodisca Kühner M 0 0 0 0

Inocybe posterula (Britzelm.) Sacc M 0 0 0 0

142

Inocybe proximella P. Karst., M 0 0 0 0 (1)

Inocybe rimosa (Bull.) P. Kumm. 1871 M 0 0 0 1 (1)

Inocybe sindonia (Fr.) P. Karst. M 0 0 0 0

Inonotus radiatus (Sowerby) P. Karst. M 0 0 0 0

Kuhneromyces mutablis (Schaeff.) W 0 (1) 1 0 0

Laccaria amethystina Cooke M 1 6 2 5 (2)

Laccaria bicolor (Maire) P.D. Orton M 0 0 0 1 (1)

Laccaria laccata (Scop.) Cooke 1884 M 1 5 5 7

Lacrymaria lacrymabunda (Bull.) Pat L 0 (1) 0 0 0

Lactarius azonites (Bull.) Fr. M 0 0 0 0

Lactarius blennius (Fr.) Fr. M 0 0 (2) 0 1

Lactarius camphoratus (Bull.) Fr. M 0 2 2 1

Lactarius deliciosus (L.) Gray 1821 M 0 0 0 1

Lactarius deterrimus Gröger M 0 0 1 (1) 1 (2)

Lactarius flexuosus (Pers.) Gray M 0 0 0 0

Lactarius fluens Boud. M 0 0 0 0

Lactarius fuliginosus Fr. M 0 0 0 0

Lactarius glyciosmus(Fr.) Fr. M 0 0 0 1

Lactarius hepaticus Plowr. M 0 1 4 1

Lactarius mammosus Fr. M 0 0 0 0

Lactarius piperatus (Scop.) Fr. M 0 0 0 0

Lactarius pyrogalus (Bull.) Fr. M 1 1 0 1

Lactarius quietus (Fr.) Fr. M 0 4 2 2

Lactarius rufus (Scop.) Fr. M 0 1 3 0

Lactarius serifluus (DC.) Fr. M 0 0 0 0

Lactarius subdulcis (Bull.) Fr. M 0 1 0 0

Lactarius tabidus Fr. 1838. M 1 3 (1) 3 2

Lactarius torminosus (Schaeff.) Fr. M 0 0 0 0

Lactarius turpis (Weinm.) Fr. M 0 0 0 0

Lactarius vellereus (Fr.) Fr. M 0 0 0 0

Lactarius vietus (Fr.) Fr. M 0 1 0 0

Leccinum scabrum (Bull.:Fr.) Gray M 0 0 0 (1) 0

Leocarpus fragilis (Dicks.) Rostaf W 0 0 0 1 (1)

Leotia lubrica (Scop.) Pers. L 1 2 (1) 1 (1) 2

Lepiota cristata (Bolton) P. Kumm. 1871 M 0 0 0 0

Lepista nuda (Bull.) Cooke M 0 0 0 1

Leucocoprinus brebissonii (Godey) Locq. L 0 0 1 0

Leucopaxillus giganteus (Sowerby) Singer M 0 0 0 0

Lichenomphalia umbellifera (L.) Redhead L 0 0 0 0

Lycoperdon molle Pers. L 0 1 0 1

Lycoperdon nigrescens Pers. L 0 3 0 4 (1)

Lycoperdon perlatum Pers. L 1 3 0 0 (1)

Lycoperdon pyriforme (Schaeff.) Pers. W 2 0 0 2 (1)

Macrocystidia cucumis (Pers.) Fr. L 0 0 0 0

Marasmiellus candidus (Bolton) Singer L 0 0 0 0

Marasmiellus ramealis (Bull.) Singer L 2 3 1 2 (1)

Marasmius androsaceus (L.) Fr. L 2 (2) 3 2 3

Marasmius epiphylloides (Rea) Sacc. & Trotter L 0 2 1 0

Marasmius hudsonii (Pers.) Fr. L 2 (1) 5 1 2

Marasmius rotula (Scop.) Fr. L 0 0 0 0

Marasmius wynnei Berk. & Broome L 0 0 0 0

Megacollybia platyphylla (Pers.) Kotl. & Pouzar L 0 0 0 0

143

Melanoleuca cognata (Fr.) Konrad & Maubl. W 0 0 0 0 (1)

Melanoleuca exscissa (Fr.) Singer 1935 M 0 0 0 0 (1)

Melanoleuca turrita (Fr.) Singer M 0 0 0 0

Microglossum viride (Pers.) Gillet L 0 0 0 0

Micromphale brassicolens (Romagn.) P.D. Orton L 0 0 0 0

Mollisia cinerea (Batsch) P. Karst. W 0 2 0 0

Mutinus caninus (Huds.) Fr. 1849 W 0 0 0 0

Mycena acicula (Schaeff.) Fr. L 0 0 0 1

Mycena aetites (Fr.) Quél. L 1 2 2 0

Mycena alphitophora (Berk.) Sacc L 1 1 0 0

Mycena amicta (Fr.) Quél. L 0 1 2 1

Mycena arcangeliana Bres L 0 0 (1) 3 3

Mycena capillaripes Peck L 0 0 2 2

Mycena cinerella (P. Karst.) L 0 0 0 2

Mycena epipterygia (Scop.) Gray L 1 1 (1) 2 4

Mycena erubescens Höhn. L 0 1 0 0

Mycena filopes (Bull.) P. Kumm. W 2 1 1 4

Mycena flavescens Velen. L 0 0 0 0

Mycena flavoalba (Fr.) Quél. L 0 0 0 0

Mycena galericulata (Scop.) Gray L 1 (1) 0 0 0

Mycena galopus (Pers.) P. Kumm L 2 4 (1) 4 2

Mycena galopus var.candida J.E. Lange L 0 1 0 1

Mycena galopus var. nigra Rea L 0 0 1 0

Mycena inclinata (Fr.) Quél. W 0 1 0 0

Mycena leptocephala (Pers.) Gillet W 1 4 (1) 8 7

Mycena metata (Fr.) Quél. L 0 2 1 7 (1)

Mycena mirata Peck L 0 0 0 0

Mycena picta (Fr.) Harmaja * L 0 0 0 0

Mycena polygramma (Bull.) Quél. L 2 3 1 2 (1)

Mycena pura (Pers.) Sacc. L 0 1 0 2

Mycena rorida (Scop.) Quél. L 2 0 4 5

Mycena rosella (Fr.) P. Kumm L 0 (1) 0 0 2

Mycena sanguinolenta (Alb. & Schwein.) Quél. L 1 0 0 3 (1)

Mycena speirea (Fr.) Gillet L 0 0 0 1

Mycena stipata Maas Geest. & Schwöbel L 0 0 0 1

Mycena stylobates (Pers.) Quél. L 1 2 1 4

Mycena vitilis (Fr.) Quél. L 2 2 (1) 2 7 (1)

Mycoacia uda (Fr.) Donk L 0 0 0 0

Nectria cinnabarina (Tode) Fr W 0 0 0 0

Neobulgaria pura (Pers.) Petr. W 0 1 0 1

Omphalina pyxidata (Bull.) Quél. W 0 0 0 0

Orbilia xanthostigma (Fr.) Fr. L 0 0 0 0

Otidea bufonia (Pers.) Boud W 0 0 1 0

Oudemansiella mucida (Schrad.) Höhn. W 0 0 0 0

Panaeolina foenisecii (Pers.) Maire W 1 0 0 0

Panaeolus acuminatus (Schaeff.) Gillet L 0 0 0 0

Panaeolus fimicola (Pers.) Gillet L 0 0 1 0

Panaeolus papilionaceus v.parvisporus Ew. Gerhardt

L 0 0 0 0

Panellus mitis (Pers.) Singer L 0 0 0 0

Panellus stipticus (Bull.) P. Karst. W 0 (1) 2 0 1

Paxillus involutus (Batsch) Fr. W 0 0 (1) 3 0

Peniophora cinerea (Pers.) Cooke M 1 0 0 0

144

Peniophora lycii (Pers.) Höhn. & Litsch. W 0 (2) 0 0 0

Peziza badia Pers. W 0 (1) 0 (1) 0 0

Phaeolus schweinitzii (Fr.) Pat. W 0 1 0 1

Phallus impudicus L. L 0 1 (1) 0 1 (1)

Phlebia radiata Fr. W 0 0 0 0

Phleogena faginea (Fr.) Link 1833 L 0 0 0 0

Pholiota flammans (Batsch) P. Kumm W 0 0 0 0

Pholiota squarrosa (Weigel) P. Kumm. W 0 (1) 0 0 0

Pholiota tuberculosa (Schaeff.) P. Kumm L 0 0 0 0

Phylloporia ribis (Schumach.) Ryvarden L 0 0 0 0

Physarum polycephalum Schwein. 1822 W 0 2 0 2 (1)

Piloderma bicolor (Peck) Jülich 1969 W 0 1 0 0

Piptoporus betulinus (Bull.) P. Karst. W 0 0 2 0

Pleurotus pulmonarius (Fr.:Fr.) Quelet P 0 0 0 0

Pluteus cervinus (Schulzer) Massee W 0 1 (1) 0 1

Pluteus podospileus Sacc. & Cub W 1 0 0 0

Pluteus romellii (Britzelm.) Lapl. W 0 0 0 0

Pluteus salicinus (Pers.) P. Kumm W 0 0 0 0

Polyporus badius (Pers.) Schwein. W 0 0 0 0

Porotheleum fimbriatum (Pers.) Fr. 1818 W 0 0 0 0

Postia caesia(Schrad.) P. Karst. W 0 0 1 (1) 5 (3)

Postia fragilis (Fr.) Jülich 1982 W 0 0 0 0

Postia stiptica (Pers.) Jülich W 0 0 0 1

Postia subcaesia (A. David) Jülich W 3 0 (1) 0 0

Psathyrella candoliana (Fr.) G. Bertrand W 1 1 0 0

Psathyrella microrhiza (Lasch) Konrad & Maubl. L 0 0 0 0

Psathyrella obtusa (Pers.) A.H. Sm., L 0 0 0 0

Psathyrella pennata (Fr.) Konrad & Maubl. L 0 1 0 0

Psathyrella prona Kits van Wav. L 0 0 1 0

Pseudohydnum gelatinosum (Scop.) P. Karst. L 0 0 (1) 0 1

Psuedoclitocybe cyanthiformis (Bull.) Singer W 1 1 1 0

Terana caerulea (Lam.) Kuntze, W 0 0 0 0

Purple stalk species W 2 2 0 0

Pyrenula nitida (Weigel) Ach. 1814 W 0 (1) 0 1 0

Ramariopsis subtilis (Pers.) R.H. Petersen 1978 W 0 0 0 0

Rhytisma acerinum (Pers.) Fr. L 0 0 0 0

Ricknella fibula (Bull.) Raithelh. W 0 0 3 1 (1)

Collybia butyracea var. butyracea (Bull.) Lennox 1979

L 0 1 1 3

Russula aeruginea Fr. L 0 2 0 0

Russula aquosa Leclair, M 0 0 0 0

Russula atropurpurea (Krombh.) Britzelm. M 0 1 0 1

Russula betularum Hora M 0 1 0 (1) 0

Russula brunneoviolacea (Crawshay) Bon M 0 0 0 0

Russula caerulea (Pers.) Fr. M 0 0 1 (1) 1

Russula cyanoxantha (Schaeff.) Fr. M 0 0 0 1

Russula delica Fr. M 0 0 0 0

Russula densifolia Secr. ex Gillet M 0 0 0 0

Russula emetica (Schaeff.) Pers. 1796 M 0 0 0 1 (1)

Russula fellea Fr. M 0 0 1 0

Russula illota Romagn. 1954 M 0 2 0 1

Russula fragilis (Pers.) Fr. M 0 0 2 1

Russula gracillima Jul. Schäff. M 0 0 0 0

145

Russula graveolens Romell M 0 0 0 0

Russula heterophylla Fr. M 0 0 1 1

Russula nigricans (Bull.) Fr. M 0 0 0 1

Russula nobilis Velen. 1920 M 0 2 0 4

Russula ochroleuca (Pers.) Fr. M 0 4 (1) 6 (1) 5 (1)

Russula olivacea (Schaeff.) Fr. M 0 0 0 0

Russula parazurea Jul. Schäff. M 0 1 0 0

Russula puellaris Fr. M 0 0 0 0

Russula queletii Schulz M 0 0 0 1 (1)

Russula risigallina (Batsch) Sacc M 0 0 0 0

Russula sardonia Fr. M 0 0 1 0

Russula versicolor Jul. Schäff M 0 0 1 0

Russula vesca Fr. M 0 0 0 0

Russula violeipes Quél. M 0 0 0 0

Russula xerampelina (Schaeff.) Fr. M 0 0 0 0

Rutstroema firma (Pers.) Dumont M 0 2 0 0

Rutstroemia sydowiana (Rehm) W.L. White 1941 W 0 1 0 0

Schizopora paradoxa (Schrad.) Donk W 0 0 0 0

Scleroderma areolatum Ehrenb. W 0 1 (3) 0 0

Scleroderma citrinum Pers. L 0 2 (1) 0 0

Scleroderma verrucosum (Bull.) Pers. L 0 0 0 0

Scutellinia scutellata (L.) Lambotte 1887 L 0 0 0 0 (1)

Sebacina incrustans (Pers.) Tul. & C. Tul. W 1 1 0 0

Stemonitis splendens (Rex) Lister L 0 2 0 2

Stereum gausapatum (Fr.) Fr. W 0 0 0 0

Stereum hirsutum (Willd.) Gray W 2 6 1 0 (1)

Stereum rugosum (Pers.) Fr. W 0 (1) 0 0 0

Strobilurus esculentus (Wulfen) Singer 1962 W 0 0 0 0

Strobilurus tenacellus (Pers.) Singer L 0 0 1 0

Stropharia semiglobata (Batsch) Quél. L 0 0 0 0

Subulicystidium longisporum (Pat.) Parmasto L 0 0 0 0

Suillus bovinus (Pers.) Kuntze W 0 0 0 (1) 0

Suillus luteus (L.) Gray M 0 0 0 0

Tapesia fusca (Pers.) Fuckel M 1 2 0 1

Tapinella atrotomentosa (Batsch) Šutara W 0 0 0 0

Trametes versicolor (L.) Pilát W 0 1 1 2

Tremella mesenterica Retz. W 0 0 0 0

Trichaptum abietinum (Dicks.) Ryvarden W 0 1 5 1

Trichoderma sp. W 2 3 0 3

Trichoglossum hirsutum (Pers.) Boud. W 0 0 0 0

Tricholoma album (Schaeff.) Fr. L 0 (1) 1 0 1

Tricholoma columbetta (Fr.) P. Kumm. 1871 M 0 1 0 0

Tricholoma fulvum (DC.) Sacc. M 0 0 0 0

Tricholoma inamoenum (Fr.) Gillet M 0 0 0 0

Tricholoma sulphureum (Bull.) Fr. M 0 0 0 0

Tricholoma ustale (Fr.) Quél. M 0 0 0 0 (1)

Tricholoma ustaloides Romagn. M 0 0 0 0

Tricholoma virgatum (Fr.) P. Kumm M 0 0 0 1

Tricholomopsis decora (Fr.) Singer M 0 1 0 0

Tricholomopsis rutilans (Schaeff.) Singer W 0 0 3 1

Tubaria furfuracea (Pers.) Gillet, W 0 0 1 0

Tubulicrinis borealis J. Erikss. 1958 W 0 0 0 0

146

Tylopilus felleus (Bull.) P. Karst. W 0 0 0 0

Vuilleminia comedens (Nees) Maire W 0 0 0 0

Xylaria carpophila (Pers.) Fr. W 2 1 0 0

Xylaria hypoxylon (L.) Grev. W 4 5 0 2

Xylaria longipes Nitschke W 0 0 0 0

Xylaria polymorpha (Pers.) Grev. W 0 (1) 0 0 0

New species recorded in the Republic of Ireland

Of the total 409 species recorded, 119 are recorded as “common” in Ireland, 37

are “occasional” (Legon and Henrici 2005), and 205 although previously found in

Ireland are of uncertain distribution. The remaining 48 species are possible new

records for Ireland and are awaiting confirmation by referees (Table 4.3).

Camarophyllopsis atropuncta (Folder 2, Appendix 1) was not found until 2010,

when it was found in two of the semi-natural ash sites (Killough and St John’s

wood). It was found multiple times during 2010 in these two sites from late

August to late October. The habitat in which it was found closely matches the

known habitats for the species in the U.K. (Legon and Henrici 2005). Cortinarius

camphoratus (Folder 2, Appendix 1) was found in 2010 outside of the

Derryhogan plot under a >50 year-old spruce/pine mixture of trees planted on a

bog. The habitat matches that of the known habitats from U.K. It was only found

once during the study. Cortinarius limonius (Folder 2, Appendix 1) was found

outside the Bohatch plot (Site: Bohat) under Sitka spruce in 2008. It was only

found once during the study. Its habitat is described by Legon and Henrici (2005)

as “pine on boggy soils”. Mycena picta was found in a beech site at Carrowgarrif

in County Clare in 2007, which was not subsequently visited in the study. The

habitat where it was found matches the habitat for the species in the U.K. as listed

by Legon and Henrici (2005).

Five species in the total list (Cantharellus aurora, Camarophyllopsis

atropuncta, Cortinarius camphoratus, Cortinarius limonius and Mycena picta) are

listed as of conservation importance in the British Red Data List (Ing 1992; Evans

et al. 2003). Cantharellus aurora was found under Sitka spruce and Scot’s pine

outside the plot at Bansha Wood (Site Bansha). While it was found only once

during the study, anecdotal evidence suggests that it is more widely distributed in

Ireland than in the U.K.

147

Table 4.3 Species found during the study that have not been previously reported from the Republic of Ireland (FRDBI 2010). The species presence on the UK list (Legon and Henrici 2005) is given in column 2. C=Common, O=Occasional, P= Present in the UK, DU= Distribution unknown, LC=Lacking confirmation, NL= Not on list/Not recorded in UK. Column 3 gives information about the habitat in which the species was found in this project. †= species awaiting confirmation, ¥= species molecular sequence sent to genbank, *= image of species present in appendix 4.

Species UK

List

Habitat

Camarophyllopsis atropuncta (Pers.) Arnolds†¥*

DU Found in unmanaged ash forests at Killough and St John’s wood. Growing on soil.

Coprinus laani Kits van Wav. (1968) †*

DU Found in microfungi experiement from Dooary Sitka spruce site growing on wood.

Ophiocordyceps forquignonii (Quél.) G.H. Sung, J.M. Sung, Hywel-Jones & Spatafora ¥*

P Found in the oak site at Raheen and outside of the ash Site at St John’s wood. Ascomycete fungus growing on dead dipteran.

Cortinarius camphoratus (Fr.) Fr. †*

DU Found outside of the Derryhogan plot under spruce trees. Growing on soil.

Cortinarius evernius Fr. *

LC Found growing under Sitka spruce and Norway

spruce in the Bohatch plot. Growing on soil.

Cortinarius limonius (Fr.) Fr. 1838 *

P Found growing under Sitka spruce and Norway spruce in the Bohatch plot. Growing on soil.

Cortinarius privignus (Fr.) Fr. †*

P Found growing under Sitka spruce in the Chevy chase Y plot. Growing on soil.

Cortinarius scandens Fr. V.* DU Found growing under Sitka spruce in the Ballygawley Spruce plot. Growing on soil.

Cortinarius stemmatus Fr. 1838 †*

P Found growing under mature Scot’s pine trees at Torc plot. Growing on soil.

Cortinarius vernus H. Lindstr. & Melot.

DU Found growing outside the Chevy chase M plot with Beech. Growing on soil.

Elaphocordyceps longisegmentis (Ginns) G.H. Sung, J.M. Sung & Spatafora *

P Found under Sitka spruce at the Chevy chase Y plot and with mature Scot’s pine at the Torc plot. Growing on Elaphomyces granulatus fruitbody.

Entoloma conferendum var.

pusillum (Velen.) Noordel. DU Found growing under oak at the Raheen plot.

Growing on soil.

Entoloma lepidissimum (Svrček) Noordel. †*

NL Found under oak outside the St Johns wood plot. Growing on soil.

Entoloma nitens (Velen.) Noordel. †*

DU Found under oak at Union and Under Sitka spruce at Bally gawley Sitka spruce plot. Growing on soil.

Exidia recisa (Ditmar.) Fr.

DU Found in microfungi experiment. Growing on wood.

Flammulaster granulosa (J.E. Lange) Watling *

DU Found outside Abbeyleix oak plot. Growing on deciduous wood.

Galerina atkinsoniana A.H. Sm. DU Found outside Ballylug Scot’s pine plot on grassy margin. Growing on soil.

Geoglossum fallax E.J. Durand. *

P Found outside Chevy chase M plot. Growing on soil.

Gymnopilus bellulus (Peck) Murrill. 1917 †

DU Found outside of Bansha Scot’s pine plot. Growing on wood.

Gymnopilus picreus (Pers.) P. Karst. 1879 †

DU Found outside of Bohatch Sitka spruce plot. Growing on wood.

Hymenochaete cruenta (Pers.) Donk

P Found outside of Dooary and Ballykilcavan sites. Growing on wood.

148

Hypoxylon rutilum Tul. & C. Tul. *

P Found during microfungi experiment. Growing on oak wood.

Lachnum niveum (R. Hedw.) P. Karst. 1871

C Found during microfungi experiment on oak wood.

Lactarius azonites (Bull.) Fr. DU Found outside of the plot at St Johns wood. Growing on soil.

Lactarius mammosus Fr. † DU Found outside of Meentinny (Sitka spruce) plot (Co. Cork, Ireland). Growing on soil.

Leucocoprinus brebissonii

(Godey) Locq.* O Found outside of the Scot’s pine plot at

Gortnagowna. Growing on soil.

Marasmiellus candidus (Bolton) Singer.

DU Found on wood.

Marasmius setosus (Sowerby) Noordel.

O Found on wood.

Melanoleuca turrita (Fr.) Singer *

DU Found growing outside of the Gortnagowna plot under spruce. Growing on soil.

Mycena alphitophora (Berk.) Sacc. *

DU Found in microfungi experiment from Ross island and Tomies wood. Growing on wood.

Mycena erubescens Höhn.

DU Found outside of the Union oak plot. Growing on wood.

Mycena picta (Fr.) Harmaja. DU Foud outside of the Carrowgarrif Beech plot (Co. Galway). Growing on wood.

Peniophora pini (Schleich. & DC.) Boidin.

DU Found growing on wood.

Pezicula livida Berk. & Broome. O Found outside of the Ross island plot. Growing on wood.

Phleogena faginea (Fr.) Link 1833

O Found growing on wood.

Pholiota tuberculosa (Schaeff.) P. Kumm.

O Found outside of the Glengort plot (Co. Cork). Growing on wood.

Pleurotus pulmonarius (Fr.:Fr.) Quelet. *

O Foud outside of the Raheen oak plot. Growing on ivy stem.

Pluteus podospileus Sacc. & Cub.

O Found outside of Killough ash plot. Growing on wood.

Porotheleum fimbriatum (Pers.) Fr.

DU Found during microfungi experiment. Growing on wood.

Psathyrella pennata (Fr.) Konrad & Maubl.

O Found outside of the Kilmacrea oak plot. Growing on soil.

Russula badia Quél. 1881 DU Found outside of Derryhogan plot under Spruce. Growing on soil.

Russula illota Romagn. †¥* DU Found outside of Raheen oak plot. Growing on soil.

Russula olivacea (Schaeff.) Fr. O Found outside of the plot at Union. Growing on soil.

Russula pectinatoides Peck * DU Found outside of the plot at Chevy chaseM. Growing on soil.

Russula subfoetens Wm.G. Sm. DU Found outside of the plot at Chevy chaseM. Growing on soil.

Skeletocutis carneogrisea A. David 1982

O Found in microfungi experiment from Kilmacrea oak wood. Growing on oak wood.

Terana caerulea (Lam.) Kuntze *

O Found outside of the ash plot in St Johns wood. Growing on ash wood.

Tomentellopsis echinospora

(Ellis) Hjortstam. C Found during the microfungi experiments. Found

growing on wood.

Xenasma pruinosum (Pat.) Donk 1957

DU Found growing on wood.

149

Of the 48 new records to Ireland in this project, 16 were found within the

plots. Of these 16, one species was found in ash plots, seven in oak, six in Scot’s

pine and eight in Sitka spruce forests. The new species that were found outside of

the plots were most commonly found in the roadside verges of the forest sites,

which often comprise of many more tree species than areas within a forest stand.

An expectation of this project was that the investigation would reveal

some new records to Ireland and many more new vice-county records for the

counties which were examined (Table 4.4). O’Hanlon and Harrington (in press)

estimated that 25 of the 26 vice counties in Ireland have had less than half of their

macrofungal species richness recorded and that much of the undiscovered

macrofungal species are to be found in the many unexamined forests of Ireland. It

is a reasonable assumption that the majority of the fungal species recorded in this

study are new vice-county records, and in the extreme case, this project has

doubled the macrofungal species records for the vice county of Tipperary (Table

4.4).

Table 4.4 Vice counties examined in this project and the numbers of fungal records from that vice county (FRDBI 2010). Column 3 lists the numbers of fungal species found from the vice county in this project.

Vice County/ Region Records in Ireland Records in this project

Tipperary 40 90

Sligo 128 71

Roscommon 183 35

Laois 238 83

Kildare 350 9

Cork 469 47

Clare & Aran islands 545 129

Offaly 690 50

Kerry 978 66

Wicklow 1064 110

Species sequenced

DNA of selected species was extracted (for method, see Section 6.2.3, Chapter 6)

and the internally-transcribed spacer region of the rDNA gene (ITS-rDNA) was

amplified using the primer pair ITS1-F and ITS 4. Amplicons were sequenced and

sequences were submitted to Genbank, through the National Centre for

Biotechnology Information NCBI website (Altschul et al. 1997) (Table 4.5).

150

Table 4.5 Species which were sequenced and their associated accession number on the Genbank website http://www.ncbi.nlm.nih.gov/genbank/.

Species name GenBank accession number

Camarophyllopsis atropuncta HQ662165

Ophiocordyceps forquignonii HQ662164

Cortinarius rubellus HQ662166

Russula fellea HQ703018

Russula illota HQ677769

Common species and genera

All of the most common species found during this project (Table 4.6) are noted as

being very common or common in Ireland and Britain except Mycena metata,

Cortinarius flexipes, Mycena rorida, Lactarius hepaticus and Marasmius

hudsonii, which are reported as uncommon in Britain (Legon and Henrici 2005).

In this study, the first five species were often found in plantation woodlands.

Marasmius hudsonii is common in Irish sessile oak woods with a holly

understory, where it is found on holly leaves. Other species on the list are rather

specialised as regards substrate and habitat. The wood-decay species Xylaria

hypoxylon and Mycena polygramma for example are much more common in

deciduous forests than they are in coniferous forests (Table 4.6).

Cortinarius was the most diverse genus with 33 species recorded from the

forest plots over the three year period (Table 4.7), followed by Mycena, Russula

and Lactarius. The rest of the genera comprise less than 20 species each. The

majority of genera comprising six species or more were putatively ECM except

Mycena, Clitocybe and Hygrocybe. One major difference between the most

common genera in the forest types, is the low numbers of ectomycorrhizal genera

recorded in the ash forests. Perhaps a more meaningful difference is the large

number of Mycena spp. found in the Sitka spruce forests.

151

Table 4.6 List of the 20 most common macrofungi in each forest type sorted from most common to less common according to their plot visit presence (%) based on all plot visits in all forest types. Also listed is the rank (1-20) of the species in the most common species list from that forest type based on plot visit frequency. If present, but not in top 20 fungi = +, nf =not found in that forest type. Species ASH OAK SP SS Plot visit presence (%)

Mycena leptocephala 1 2 2 2 20

Laccaria laccata + 3 3 3 18

Russula ochroleuca nf 1 1 1 15

Laccaria amethystina + 4 4 4 14

Mycena vitilis 3 6 5 6 13

Hypholoma fasciculare 2 5 + 5 12

Mycena galopus 4 7 6 7 12

Xylaria hypoxylon 5 8 nf + 11

Mycena rorida 8 nf + 11 11

Stereum hirsutum 7 9 + nf 10

Mycena metata nf + + 10 10

Marasmius hudsonii 10 11 + + 10

Marasmius androsaceus 11 13 + + 10

Cortinarius flexipes nf 10 + + 9

Lactarius tabidus + 15 12 + 9

Mycena polygramma 12 14 + + 8

Mycena epipterygia 6 + 7 8 7

Armillaria mellea 9 12 + 12 7

Lactarius hepaticus nf + 8 + 6

Collybia butyracea nf + + 9 5

Table 4.7 Species-diverse genera found in different forest types from both inside and outside the plots. Only genera with more than 6 species found are included. SP= Scot’s pine, SS= Sitka spruce.

Genera OAK SP SS Total species

Cortinarius 7 8 13 33

Mycena 15 14 19 30

Russula 7 7 11 29

Lactarius 8 6 9 22

Inocybe 3 3 4 19

Entoloma 1 2 4 13

Boletus 2 0 3 9

Amanita 2 2 2 8

Tricholoma 2 0 2 8

Clitocybe 2 1 3 6

Hygrocybe 0 1 0 6

152

Most species occurred only sporadically and on only one plot visit. In the

ash, oak, Scot’s pine and Sitka spruce forest types, 68, 58, 54 and 57%

respectively of the species were found on only one plot visit (Fig. 4.1).

Ash

1 2 3 4 5+0

10

20

30

40

50

60

70

No. of plot visits

% o

f sp

ecie

s

Oak

1 2 3 4 5+0

10

20

30

40

50

60

No. of plot visits

% o

f sp

ecie

s

SP

1 2 3 4 5+0

5

10

15

20

25

30

35

40

45

50

55

No. of plot visits

% o

f sp

ecie

s

SS

1 2 3 4 5+0

10

20

30

40

50

60

No. of plot visits

% o

f sp

ecie

s

Fig. 4.1 Frequency of occurrence of macrofungal species in plot visits to the different forest types. SP= Scot’s pine, SS= Sitka spruce.

The mean species richness per plot visit was highest in oak, followed by Sitka

spruce, Scot’s pine and ash (Table 4.8). The untransformed means from the

different forest types were compared by ANOVA. The mean species richness per

plot visit in the ash forest type was significantly lower than that of oak, SP or SS

(P<0.05), but there were no significant differences between the other forest types

(Fig. 4.13a). The maximum species per plot visit was 19 species, which occurred

in Gortnagowna Scot’s pine forest in 2009.

153

Table 4.8 Mean species richness per plot visit (± standard deviation) and maximum species recorded in a plot visit in each of the forest types. Significant differences were identified by ANOVA with Hochberg’s GT2 post-hoc test. Means that do not share a similar superscript letter are significantly different. SP= Scot’s pine, SS= Sitka spruce.

Forest type Mean species richness Maximum recorded

Ash 3.74 (±3.1)a 12

Oak 7.65 (±4.3)b 18

SP 6.87 (±4.5) b 19

SS 7.38 (±3.9) b 18

4.4.2 Species richness estimation of fungal diversity in the forest types

New records to the macrofungal species list were added continuously over the

three years of the project (Fig. 4.2). With regular sampling of the plots, it would

be expected that the numbers of new species found would decline with each year.

This is evident in all the forest types except Scot’s pine, which received a more

intensive search in 2009 to bring the numbers of plot visits to parity with the other

forest types. In the cases of ash, oak and Sitka spruce forest types there were

fewer new species discovered in 2009 than in 2007 (Fig. 4.2).

ASH OAK SP SS0

25

50

75

100

125

150

175

200

225200720082009

Forest type

No

. o

f sp

ecie

s

Fig.4.2 Numbers of macrofungal species found in the forest types over the sampling years 2007-2009. SP= Scot’s pine, SS= Sitka spruce.

Of the total 219 species found within the plots, 40 were found in every

year of the study, 62 were found in two of the three years of the study, and the

154

remaining 117 were only found in one year of the study. Almost all of the species

which fruited in every year of the study would be considered as very common, for

example, Armillaria mellea, Calocera viscosa, Collybia butyracea, Laccaria

laccata and L. amethystina. Of the species which only fruited in one year, many

were small delicate species such as Mycena rorida and other Mycena species.

However, there were also larger species that fruited in only one year, but when

they fruited they were locally common. The best example of this is

Camarophyllopsis atropuntca. This species was not found until 2010, when it

fruited in great numbers and over the entire fruiting season in two sites. The sites

were the two semi-natural ash sites at Killough and St John’s Wood. Another

good example of a species which fruited rarely but in large numbers was Inocybe

napipes. This species was not found until 2010, when it was discovered in six

sites. These were the oak sites Tomies and Abbeyleix, the Scot’s pine sites

Ballygawley and Gortnagowna, and a spruce site Kinnity. Inocybe rimosa also

followed this trend of sporadic but locally common fruiting; it fruited in Chevy

Chase Mature and St John’s Wood in 2007 and it was also found in the

Ballygawley Scot’s pine in 2009.

On the basis of the three year’s sampling and species recorded inside and

outside plots, Sitka spruce sites produced the greatest number of macrofungal

species (144), followed by oak (113), Scot’s pine (89) and ash (56). These totals

are based on quite different numbers of sampling units in the case of each forest

type (25, 36, 31 and 55 plot visits respectively). To compensate for the difference

in sampling effort and to compare species richness of the different sites and forest

types at a similar sampling intensity, sample-based rarefaction curves with 95%

confidence intervals were calculated for the four forest types (Figs. 4.3; 4.4). The

curve of randomized data is created by sampling and then randomly resampling

the plot data while the actual accumulation of new species is shown by the curve

titled “actual”. Levelling off of the randomized curves indicates that the number

of species found is getting close to the actual total number of species in the area.

Levelling off cannot be seen in any of the curves (Fig. 4.3) and so it may be

concluded that more species would be found with continued sampling.

According to the rarefaction estimates of macrofungal species richness in

the forest types, at a similar sampling intensity (23 plot visits) the species richness

of the forest types are: ash was estimated to be 34 species, oak 71, Scot’s pine 74

155

and Sitka spruce 76 species. The estimate for ash forest was significantly less

(P<0.05) than that of all of the other forest types at this sampling intensity, based

on non-overlapping confidence intervals (Fig. 4.4). According to the rarefaction

estimates, Sitka spruce would be the most species-rich forest type, followed by

Scot’s pine, oak and ash, notably so at higher sampling intensities.

156

2007

2008

ASH

0 5 10 20 25

2007

2008

0

10

20

30

40

50

Ploty visits

Cum

ula

tive s

pecie

s

2007

2008

OAK

0 10 20 25 30 35

2007

2008

0

10

20

30

40

50

60

70

80

90

100

110

120

Plot visits

Cum

ula

tive s

pecie

s

2008

SP

0 5 10 15 20 25 30 35 40

2008

0

10

20

30

40

50

60

70

80

90

100

110

Plot visits

Cum

ula

tive s

pecie

s

SS

0 10 20 30 40 50 60

0

10

20

30

40

50

60

70

80

90

100

110

120

130

140

Plot visits

Cum

ula

tive s

pecie

s

Fig. 4.3 Macrofungal species accumulation curves for ash, oak, SP (Scot’s pine) and SS (Sitka spruce) forests, and sample-based rarefaction estimates of species richness. The black jagged line is the actual cumulative species numbers for the forest type over the sampling period, dotted lines at x-axis indicate the end of a sampling year. Smoothed grey lines are the sample-based rarefaction curve for the cumulative species from that forest type with the dotted line indicating the 95% upper and lower confidence bounds.

157

0 10 20 30 40 50 600

25

50

75

100

125

Ash

SPSS

Oak

No. of plot visits

No

. o

f sp

ecie

s

Fig. 4.4 Sample-based rarefaction curves of mean species (symbols) with 95% upper and lower confidence intervals (whisker bars) for ash (square symbols), oak (triangle symbols), SP= Scot’s pine (cross symbols) and SS= Sitka spruce (Circular symbols). Plots were sampled randomly with replacement using 500 permutations for each sample size.

158

Species richness estimates

Rarefaction was used in this study to estimate the number of species that would

have been found, at a lower but equal number of plot visits to the different forest

types. Species richness estimation uses the available data to statistically estimate

the hypothetical species richness, which would have been found if the entire

species richness was recorded. A number of estimators were employed to

statistically estimate the likely species richness of the forest types, based on the

accumulated number of species from the three year’s sampling. The data for the

different estimators versus sampling intensity was calculated (Figs. 4.5; 4.6).

From the overall trend of the curves (rising with sampling effort), it is likely that

the total species richness would increase with increasing sampling effort.

Depending on the estimator used, taking the final estimate of species richness in

each forest type gives values for species richness in ash forests ranging from 77 to

161, in oak from 166 to 208, in Scot’s pine from 135 to 203 and in Sitka spruce

from 186 to 218 species (Fig. 4.6). The actual number of species found in these

forests during this study was 56, 113, 89 and 144 respectively.

ASH

-ICE

ASH

-CHAO

2

ASH

-ACE

OAK-IC

E

OAK-C

HAO2

OAK-A

CE

SP-IC

E

SP-C

HAO2

SP-A

CE

SS-IC

E

SS-C

HAO

2

SS-A

CE

0102030405060708090

100110120130140150160170180190200210220

Nu

mb

er

of

sp

ecie

s

Fig. 4.5 The results for the species richness estimation of the ICE= Incidence based coverage estimator, ACE= Abundance based coverage estimator, CHAO2= Chao 2 richness estimator for each forest type are plotted. SP= Scot’s pine, SS= Sitka spruce. The grey lines indicate the actual species richness recorded in the ash, oak, Scot’s pine and Sitka spruce forest types.

159

ASH

0 5 10 15 20 250

10

20

30

40

50

60

70

80

90

100

110

120

130

140

150

160

Plot visits

Nu

mb

er

of

sp

ecie

s

OAK

0 5 10 15 20 25 30 350

102030405060708090

100110120130140150160170180190200210

Plot visits

Nu

mb

er

of

sp

ec

ies

SP

0 5 10 15 20 25 30 35 400

102030405060708090

100110120130140150160170180190200

Plot visits

Nu

mb

er

of

sp

ecie

s

SS

0 10 20 30 40 50 600

25

50

75

100

125

150

175

200

ICEACECHAO2Actual

Plot visitsN

um

ber

of

sp

ecie

s

(a) (b)

(c) (d)

Fig. 4.6 Estimates of macrofungal species richness versus sampling intensity for different species richness estimators in ash, oak, Scot’s pine (SP) and Sitka spruce (SS) forest types. Cross= Incidence based coverage estimator (ICE), triangles= abundance-based coverage estimator (ACE), diamond= Chao2 richness estimator (CHAO2) and square= smoothed sample based rarefaction curve of actual species richness with standard deviation (dotted line). Plots were sampled randomly without replacement using 500 permutations for each sample size.

160

Of the three species richness estimators tested, none showed signs of reaching an

asymptote in all of the forest types. The oak and Sitka spruce forest types appear

to be reaching an asymptote, although more sampling would be required to reach

a stable estimate. Reaching an asymptote before 100% sampling effort is one of

the main characteristics looked for in a richness estimator. Static values for

species richness estimation indicate that further sampling will not change the

estimate and therefore a realistic estimate has been reached.

As none of the estimators showed signs of reaching an asymptote, the

Chao2 estimator will be used as an estimate of the lower boundary for species

richness in the forest types. Therefore, taking the Chao2 as a realistic estimate of

the possible macrofungal diversity in the forest types, 45% of the fungal species

richness in the ash sites, 65% of the fungal species richness in the oak sites, 52%

of the fungal species richness in the Scot’s pine sites and 77% of the fungal

species richness in the Sitka spruce sites has been realized (Fig. 4.7).

161

Fig. 4.7 Percentage of species discovered (dark grey segment) and undiscovered (light grey segment) according to Chao2 species richness estimator from the ash, oak, Scot’s pine (SP) and Sitka spruce (SS) plots.

4.4.3 Species richness estimation: variation within forest types

Overall estimates of species richness in the forest types hide the fact that there

was considerable variation in species richness between individual plots within the

same forest type (Table 4.9). As all of the plots were not sampled to an equal

degree, the effect of sampling intensity was assessed for the plots using sample-

based rarefaction, to estimate the species richness of the sites at the lowest site

sampling intensity for each forest type (Fig. 4.8). This was at four plot visits for

the ash, oak and Sitka spruce sites and 3 plot visits for Scot’s pine sites. ICE was

chosen as the species-richness estimate as it was shown to reach an asymptote at

the lowest sampling intensity for the majority of the forest types. ICE species

162

richness estimates, total species recorded, rarefaction data and the percentage of

the estimated species richness which was realised in the three years sampling were

calculated (Table 4.9).

Ash

BKilcav Donad Killo RossI StJon0

5

10

15

20

25

30

35

No

. o

f sp

ecie

s

Oak

Abbey Kilmac Rahee Toom Union0

10

20

30

40

SP

Anna

Blu

g

BGaw

SPBrit

t

Der

ryGort

Torc

Bnsh

a

0

10

20

30

Plot name

No

. o

f sp

ecie

s

SS

Bgaw

SS

Bohat

Chev

M

Chev

Y

Clo

on

Dooar

Money

Qui

t

Swm

id

SSYoung

SSMat

ure

0

10

20

30

40

50

60

70

80

90

Plot name

Fig. 4.8 Estimates of species richness of individual sites within each forest type based on rarefaction. Also given is the rarefaction estimate for young and mature Sitka spruce plots calculated from the pooled data for the spruce plots. Sample-based rarefaction was generated using 500 randomized samples with replacement from the plot based sample results. Also shown are the upper and lower 95% confidence limits. SP= Scot’s pine, SS= Sitka spruce.

163

Table 4.9 Plot list showing the total number of species recorded from the plot, the mean species richness estimate according to the Chao2 estimator (± standard deviation) and the percentage of the estimated species richness which was discovered during the three years sampling. SP= Scot’s pine, SS= Sitka spruce. a= rarefied to 4 plot visits, b= rarefied to 3 plot visits.

Forest

type

Plot name Total species

recorded

Species

richness

estimate

(Chao2)

Rarefied

species

richness

% estimated

species

realised

ASH Ballykilcavan 14 39 13a 36

ASH Donadea 8 12 8 a 67

ASH Killough 18 60 13 a 30

ASH Ross Island 6 14 3 a 43

ASH St John’s Wood 32 49 21 a 65

OAK Abbeyleix 38 91 26a 42

OAK Kilmacrea 43 78 23 a 55

OAK Raheen 54 66 23 a 82

OAK TomiesA 30 59 17 a 51

OAK Union 18 90 17 a 20

SP Annagh 16 21 15b 76

SP Ballylug 10 12 6 b 83

SP BallygawleySP 20 30 16 b 67

SP Brittas 25 41 17 b 61

SP Derryhogan 21 31 19 b 68

SP Gortnagowna 24 33 18 b 73

SP Torc 17 37 10 b 46

SP Bansha 19 69 13 b 28

SS BallygawleySS 26 30 22 a 87

SS Bohatch 43 47 26 a 91

SS Chevy chaseM 25 62 16 a 40

SS Chevy chaseY 33 48 23 a 69

SS Cloonagh 9 17 5 a 53

SS Dooary 34 55 29 a 62

SS Moneyteige 19 37 16 a 51

SS Quitrent 29 73 19 a 40

SS Stanahely 43 66 27 a 65

Species richness of the ash plots

The average rarefaction estimate of macrofungal species richness of the ash plots

was 12 species. Ross Island and Donadea were found to have a species richness

value lower than the average. The species richness estimators for these sites

indicate that with more sampling the increases in species richness would be

significant. The semi-natural ash plot at Killough and St John’s (Fig. 4.9) are the

second and first most species-rich plots according to the rarefied and Chao2 data.

The plot at Killough would also see significantly large increases in total species

164

richness with continued sampling effort according to the % of the estimated

species realised from three years sampling.

Species richness of oak plots

The average rarefaction estimate of macrofungal species richness for the oak plots

is 21 species. The plots at Union and Tomies A were below this value. Four of the

five oak plots have ≤50% of their estimated species richness discovered and so

would likely see large increases in species richness with continuing sampling.

Within the oak forest type, there are no significant differences (P=0.05) in rarefied

species richness between any of the plots (Table 4.9).

Species richness of the Scot’s pine plots

The average rarefaction species richness estimate of the Scot’s pine plots was 14

species per plot. The plot at Ballylug had very low rarefied species richness with

only six species in total after 3 plot visits. The most species-rich Scot’s pine plot

according to rarefaction was Derryhogan, but with sufficient sampling to identify

all fungal species in the Scot’s pine plots (according to Chao2 estimate), the plot

at Bansha would be the most species rich Scot’s pine plot. Only 2 of the Scot’s

pine plots have less than 60% of their species richness left to be recorded

according to species richness estimation. In the Scot’s pine forest type, Ballylug

(Blug) had significantly lower (P=0.05) species richness than either Derryhogan

(Derry) or Gortnagowna (Gort) according to rarefaction (Table 4.9).

Species richness of the Sitka spruce plots

Overall, the Sitka spruce plots had an average rarefaction estimate of macrofungal

species richness of 20 species per plot. However, there was a very large variation

in the rarefaction estimate of species richness of the different Sitka spruce plots,

ranging from 5 to 29 species per plot. The plot with the lowest estimate,

Cloonagh, was one of the youngest Sitka spruce plots examined and was the first

rotation of forest which was planted on bog. The Sitka spruce plot at Stanahely

(Fig. 4.9) had the joint second highest actual species richness of all plots. Sitka

spruce plots showed the most variability in rarefaction estimates of macrofungal

species richness. The plot at Cloonagh (Cloon) had significantly lower (P=0.05)

species richness than the plots at Ballygawley (BgawSS), Bohatch (Bohat), Chevy

165

chase Y (ChevY), Dooary (Dooar), Quitrent (Quit) and Stanahely (Swmid) (Table

4.9).

The rarefaction analysis also found that the rarefied species richness of the

young versus the old Sitka spruce plots were not significantly different (Fig 4.8).

StJon (Ash) Rahee (Oak) Britt (SP) Swmid (SS)0

10

20

30

40

50

60

70

80 Discovered species richnessUndiscovered species richness

Plot name

No

. o

f sp

ecie

s

Fig. 4.9 Discovered and undiscovered species richness in each of the most species rich (actual species richness) plots from the four forest types. Plots were sampled randomly without replacement using 500 permutations for each plot. StJon= St John’s Wood plot, Rahee= Raheen wood, Britt= Brittas plot, Swmid= Stanahely plot; SP= Scot’s pine site, SS= Sitka spruce.

4.4.4 Species diversity and evenness

Sporocarp abundance data from the plot visits for each of the plots over the three

years was used to calculate diversity and evenness estimates for each of the plots

and for the different forest types (Table 4.10; Fig. 4.10). According to Simpson’s

diversity index, the most diverse plots were Dooary (Sitka), Raheen (oak), Chevy

chase Y (Sitka), Stanahely (Sitka) and Kilmacrea (Table 4.10). In contrast,

Donadea (ash), Ross Island (ash), Ballylug (Scot’s pine) and Cloonagh (Sitka) had

very low diversity index values, indicating that the sporocarp count tended to be

dominated by a limited number of species. Pooled plot visit data from all sites

within a forest type was used to estimate diversity within forest types using

166

Shannon’s and Simpson’s diversity indices. Scot’s pine sites were found to be the

most diverse, followed by Sitka spruce, oak and ash sites (Table 4.10).

Table 4.10 Three diversity indices for the site and forest type species data. H’ = Shannon’s diversity index, 1/D = Reciprocal form of the Simpsons diversity index, E1/D = Simpsons evenness measure. SP= Scot’s pine, SS= Sitka spruce.

Forest type Site name Total species H’ 1/D E1/D

ASH Ballykilcavan 14 1.95 6.81 0.49 ASH Donadea 8 1.41 3.71 0.46 ASH Killough 18 1.93 5.42 0.30 ASH Ross island 6 1.06 3.02 0.50 ASH St John’s wood 32 2.44 7.41 0.23 OAK Abbeyleix 38 2.49 8.1 0.21 OAK Kilmacrea 43 2.8 11.63 0.27 OAK Raheen 54 2.97 14.29 0.26 OAK TomiesA 30 2.31 6.95 0.23 OAK Union 18 2.09 6.74 0.37 SP Annagh 16 2.01 6 0.38 SP Ballylug 10 1.57 4.04 0.40 SP BallygawleySP 20 2.39 9.18 0.46 SP Brittas 25 2.28 7.36 0.29 SP Derryhogan 21 2.27 7.84 0.37 SP Gortnagowna 24 2.27 6.7 0.28 SP Torc 17 2.03 6.04 0.36 SP Bansha 19 2.17 7.31 0.38 SS BallygawleySS 26 2.22 6.82 0.26 SS Bohatch 43 2.44 6.61 0.15 SS Chevy chaseM 25 2.4 8.84 0.35 SS Chevy chaseY 33 2.74 11.95 0.36 SS Cloonagh 9 1.28 2.86 0.32 SS Dooary 34 2.9 14.43 0.42 SS Moneyteige 19 1.93 5.25 0.28 SS Quitrent 29 2.46 8.5 0.29 SS Stanahely 43 2.81 11.71 0.27

All ash 56 2.9 12.5 0.22 All oak 113 3.52 20.91 0.19 All SP 89 3.61 24.59 0.28 All SS 144 3.72 24.21 0.17

167

BK

ilcav

Do

nad

Killo

Ro

ssI

StJ

on

Ab

bey

Kilm

ac

Rah

ee

To

om

Un

ion

An

na

Blu

g

BG

aw

SP

Bri

tt

Derr

y

Go

rt

To

rc

Bn

sh

a

Bg

aw

SS

Bo

hat

Ch

evM

Ch

evY

Clo

on

Do

oar

Mo

ney

Qu

it

Sw

mid

0.0

2.5

5.0

7.5

10.0

12.5

15.0

17.5

ASH OAK SP SSPlot name

Sim

pso

n's

div

ers

ity i

nd

ex

Fig. 4.10 Simpson’s diversity index values for each of the sites. The bars represent the mean Simpson’s value calculated from 500 permutations with random selection of samples with replacement using the plot visit data from each of the plots over three years. Error bars = standard deviations of the index values.

4.4.5 Functional groups of macrofungi

The functional groups used in this project were based on the normal mode of

nutrition used by the fungal species. Species were grouped into one of four

functional groups, litter-decay, ectomycorrhizal, parasitic or wood-decay species.

The percentage of macrofungal species in the different functional groups across

all forest types was 29% litter-decay, 38% mycorrhizal, 3% parasitic and 30%

wood-decay species. The species richness in the four different functional groups

recorded per plot visit differed between the forest types (Fig. 4.11) and the

relative proportions of species in the different functional groups varied as well

(Fig. 4.12).

Ectomycorrhizal species were largely absent from the ash sites with the

exception of Laccaria laccata found in the St John’s Wood site. This species was

168

most certainly present on the roots of a nearby hazel Corylus avellana. There was

no significant difference in the species richness of ectomycorrhizal macrofungi

recorded per plot visit between oak, Sitka spruce or Scot’s pine forest types (Fig.

4.11c). Similarly, analysis of the below-ground ectomycorrhizal morphotype

richness did not reveal significant differences between these three forest types

(Chapter 6). The species richness of litter-decay macrofungi recorded per plot

visit were found to be significantly higher in Sitka spruce plots than in ash plots

(P<0.001), but there were no significant differences found between any of the

other forest types (Fig. 4.11b). Parasitic species were poorly represented in all

forest types. The species richness of wood-decay macrofungi per plot visit was

significantly higher in oak than in Scot’s pine forests (P<0.01), but there were no

significant differences between oak and ash, or between Scot’s pine and ash or

Sitka spruce (Fig. 4.11d).

The relative proportions of the different functional groups also differed

between the forest types (Fig. 4.12 and Table 4.12). Ectomycorrhizal species

predominated slightly in Sitka spruce and Scot’s pine forests, comprising 42 and

44% of the species respectively, while in oak forests there were proportionally

more wood-decay species than the other forest types with 34% in oak vs. 19 and

23% of the species in Scot’s pine and Sitka spruce forests respectively being

wood-decay species.

169

(a)

(b)

(c)

(d)

Fig. 4.11 a, b, c, d. Mean numbers of species recorded per plot visit for: (a) all macrofungi, (b) litter-decay (c) ectomycorrhizal species, (d) wood-decay species, from the different forest types. Mean, lower and upper quartiles are displayed along with highest records (circles). Box plots with the same letter have means which are not significantly different at P<0.05. * indicate possible outliers.

170

ASH OAK SP SS0

10

20

30

40

50

60

70

80

90

100LD ECM P WD

Forest type

% o

f sp

ecie

s

Fig. 4.12 Percentage of species from the different functional groups in the four forest types. SS= Sitka spruce, SP = Scot’s pine, LD= Litter-decay, ECM= ectomycorrhizal, P= Parasitic and WD= Wood-decay fungi.

Table 4.12 Functional group species richness in the different forest types. Numbers in parenthesis refer to the percentage of the total species richness in a forest type. LD= litter-decay, ECM= ectomycorrhizal, P= parasitic and WD= wood-decay species.

LD species ECM

species

P species WD

species

Total

species

Ash 22 (39) 8 (14) 2 (4) 24 (43) 56

Oak 33 (29) 39 (34) 4 (4) 37 (33) 113

Scot’s pine 28 (32) 39 (44) 3 (3) 19 (21) 89

Sitka

Spruce

47 (32) 60 (42) 4 (3) 33 (23) 144

Chi-square analysis was used to test whether the proportions of species in

the three functional groups were similar in the oak, Scot’s pine and Sitka spruce

forest types (ash forest was excluded from this analysis because of the absence of

ectomycorrhizal species in this forest type). The parasitic functional group was

pooled with the wood-decay functional group in these analyses as the parasitic

functional group has values of less than five for species richness and so were not

compatible with a chi-square test. This means that each test has 3 degrees of

freedom (denoted by the subscript 3 in χ23). The results of the chi-square analysis

show that the proportions of species in the different functional groups across the

different forest types were significantly different from each other; oak and Scot’s

171

pine (χ23 = 27.08; P= < 0.001), oak and Sitka spruce (χ2

3 = 12.67; P= < 0.01) and

Scot’s pine and Sitka spruce (χ23 = 22.00; P= < 0.001). Notably, ectomycorrhizal

species were better represented in Scot’s pine and Sitka spruce sites than in oak;

and wood-decay species were better represented in the oak than in the former

types.

Correlation analysis (Pearson’s) was used to see if there was a relation

between the numbers of species in the different functional groups across the forest

types (Table 4.13). There was a significant positive relationship (P< 0.05)

between litter-decay species richness and ectomycorrhizal and wood-decay

species richness. There was also a slight but significant (P<0.05) negative

relationship between ectomycorrhizal species and wood-decay species.

Table 4.13 Correlation coefficients (Pearson's r) between numbers of species in different functional groups recorded in the oak, Scot’s pine and Sitka spruce plots. Sign indicates the relationship between the functional groups, *= significant at P<0.05, **= Significant at P<0.01. LD= litter-decay, ECM= ectomycorrhizal, P= parasitic and WD= wood-decay species.

LD ECM P

ECM 0.187**

P -0.069* 0.029

WD 0.205** -0.08* -0.015

4.4.6 Influence of site variables on species richness

The effects of site nominal variables (Table 4.11) on the numbers of species and

species in each functional group were tested using a generalized linear model with

Poisson error distribution and log link function.

172

Table 4.11 Data entered into generalized linear models for each forest type. a= model generated was significant at explaining variation.

Tree type Functional groups analysed Variables entered

Sitka

spruce

Total species, litter-decay species a, ectomycorrhizal species, wood-decay species

Rotation stage (first vs. second rotation), forest age (young vs. mature)

Scot’s

pine

Total speciesa, litter-decay species, ectomycorrhizal speciesa, wood-decay speciesa

Rotation stage (first vs. second rotation), forest age (young vs. mature)

Oak Total species, litter-decay species, ectomycorrhizal speciesa, wood-decay species

Grazing (grazed vs. non-grazed)

Ash Total speciesa, litter-decay speciesa, wood-decay speciesa

Forest management (managed vs. not managed), forest age (young vs. mature)

Eight of the generalized linear models tested showed significant results,

with the model being more explanatory than the null model (table 4.11). It was

found that the age of the Sitka spruce forest was significant at explaining the

variation of litter decay species per plot visit. The model describing the total

species richness in Sitka spruce plots was better at explaining the variation in the

dependent variable (litter-decay species per plot visit) than the null model

(likelihood ratio chi-square= 8.403, df=2, P<0.05). The likelihood ratio chi-square

is a value based on the chi-square distribution which shows that the model

explained the litter-decay fungal species richness in Sitka spruce plots better than

was accounted for by chance. The pairwise comparisons between the two age

groups of Sitka spruce forests highlighted that young forests had higher litter

decay species per plot visit than mature forests (mean difference [young –mature

forests]= 0.449, df= 1, P<0.05). Therefore from these results it could be stated that

young Sitka spruce plots have a greater species richness of litter-decay

macrofungi than older plots.

In Scot’s pine forests it was found that forest age had a significant effect

on the numbers of ectomycorrhizal and wood-decay fungi in Scot’s pine plots.

The GLM also revealed that total species richness and ectomycorrhizal species

richness per plot visit was related to the rotation stage of the Scot’s pine forest.

The model was significant at explaining the numbers of fungal species per plot

173

visit (likelihood ratio chi-square= 13.812, df=2, P<0.001). It was found that

macrofungal species increased in 2nd rotation forests and in younger stands. The

model statistics for the variables were rotation stage (mean difference [2nd

rotation-1st rotation]= 0.966, df= 1, P<0.05) and age (mean difference [young-

mature]= 0.1358, df= 1, P<0.01). Ectomycorrhizal species were also aptly

modelled by the generalized procedure with the model predicting more variation

than the null model (likelihood ratio chi-square= 21.507, df=2, P<0.001).

Ectomycorrhizal species increased with rotation stage (mean difference [2nd

rotation-1st rotation]= 0.1372, df= 1, P<0.01) and decreased with age (mean

difference [young-mature]= 0.1816, df= 1, P<0.001). The model also highlighted

that the diversity of wood-decay species per plot visit Scot’s pine sites was

sufficiently related to forest age (likelihood ratio chi-square= 7.619, df=2,

P<0.05). Younger Scot’s pine sites had a higher diversity of wood-decay species

per plot visit than mature sites (mean difference [young-mature]= 0.1376, df= 1,

P<0.05). Overall, in Scot’s pine forests it can be stated that total species,

ectomycorrhizal species and wood-decay species all decreased with increasing age

and that total species and ectomycorrhizal species richness both increased in

second rotation Scot’s pine forests.

It was found that in oak forests, the presence of large grazing mammals

had a significant effect on the species richness of litter-decay fungi per plot visit.

In the oak forests, the generalized linear model was not significant (P<0.07) at

describing the richness of litter-decay species per plot visit. The model generated

by the generalized linear model was better at explaining the variation in the

dependent variable (litter-decay species per plot visit) than the null model

(likelihood ratio chi-square= 3.397, df=1, P<0.65). The pairwise comparisons

between grazed and un-grazed oak forests revealed that un-grazed forests had

marginally significantly higher litter-decay species richness than grazed forests

(mean difference [un-grazed –grazed forests]= 0.155, df= 1, P<0.07). The number

of cases of ungrazed and grazed examined was 20 and 15 respectively. The

generalized linear model explained a significant amount of the variation in the

ectomycorrhizal species per plot visit (likelihood ratio chi-square= 32.592, df=1,

P<0.01). The model found that grazed oak forests had significantly higher

ectomycorrhizal species richness per plot than un-grazed plots (mean difference

[grazed –ungrazed]= 0.34, df= 1, P<0.001).

174

The GLM procedure identified relationships between forest management

and the total species and litter-decay species richness per plot visit in ash forests.

The modelling procedure also identified a relationship between forest age and the

numbers of wood decay species per plot visit in ash forests. In the ash forests the

model was significant at describing the numbers of species per plot visit

(likelihood ratio chi-square= 10.772, df=3, P<0.05). It was found that forest

management type was a important variable in determining the fungal species

richness of the ash sites with unmanaged forests having significantly higher

fungal species richness per plot visit than managed forests (mean difference

[unmanaged–managed]= 0.1457, df= 1, P=0.012). The model also predicted the

numbers of litter-decay (likelihood ratio chi-square= 10.815, df=1, P<0.05) and

wood-decay (likelihood ratio chi-square= 8.074, df=3, P<0.05) species per plot

visit. Litter-decay species increased in unmanaged ash forests (mean difference

[managed–unmanaged]= 0.1472, df= 1, P<0.05) while wood-decay species

decreased as forests aged (mean difference [young-mature]= 0.1609, df= 1,

P<0.08) In ash forests it can be stated that the total species and litter-decay species

per plot visit are higher in unmanaged ash forests than in managed ash forests. It

can also be stated that there are more wood-decay species in older ash forests than

there are in young ash forests.

Although only eight examples of relationships between site chronological

or historical variables and fungal species richness could be identified through the

generalized linear model analysis, certain general descriptions of the sites and

their fungal diversity can be stated. These are divided into the different forest

types below.

Ash plots

The ash plots can be divided into two groups regarding their land use history. The

semi-natural ash woodlands at Killough and St John’s wood are unmanaged and

as pristine ash forests as can be found in Ireland. The three remaining ash forests

at Donadea, Ballykilcavan and Ross Island are plantation ash forests. The two

plots which are classed as semi-natural ash woodlands were significantly more

species rich than the managed plots (P<0.05). Generalized linear modelling

revealed that management type had an effect on total species and litter-decay

175

species richness while age had an effect on wood-decay species richness in ash

plots.

Oak plots

There is no difference in the recent land use history of any of the oak plots, as

circa 1913 all of the plots were oak dominated forests. Based on structural

characteristics of the oak plots (Section 3.4.5, Chapter 3) there are not many

differences between the oak sites. One difference highlighted in Chapter 3 is the

presence or absence of large mammal grazing in the oak sites. The sites at

Raheen, Tomiesb and Kilmacrea are free from signs of grazing by deer whilst the

sites at Union, Tomies A and Abbeyleix are affected by year round browsing

grazing by deer. The effect of grazing was not significant (P<0.07) at lowering

litter-decay fungal species richness per plot and increasing ectomycorrhizal

species per plot visit (P<0.001).

Scot’s pine plots

Regarding past land use, it seems that Scot’s pine first rotation plantations do not

support as high a richness of fungal species (P<0.05) or ectomycorrhizal species

(P<0.01) as 2nd rotation sites. Age was found to affect the mycota of Scot’s pine

plots whereby younger plots were found to have a higher richness of total fungal

species (P<0.01), ectomycorrhizal (P<0.001) and wood-decay species (P<0.05)

per plot visit.

Sitka spruce plots

The Sitka spruce plots were the most numerous of all of the forest types. A central

aim of this project was to examine the diversity and functional group diversity of

fungi between different types of Sitka spruce forests. According to the generalized

linear modelling, only litter-decay species showed a significant relationship with

the variables examined. Younger sites were found to have a higher litter decay

species richness than older sites (P<0.05). The results of the generalized linear

modelling from Sitka spruce sites points to the fact that the examined variables

may not be sufficient at explaining the high inter-site variability with regard to the

fungal and functional group species richness in Sitka spruce forests in Ireland.

176

4.4.9 Effect of weather variables on fungal fruiting

To analyse the relationships between fungal species richness and abundance and

weather variables, stepwise linear regression was carried out. Non-parametric

Spearman’s correlation (rs) was also carried out to identify relationships between

fungal phenology and weather variables. The two fungal variables (species

richness and natural log of fruitbody abundance) investigated in this analysis were

found to be highly positively related (Pearson’s correlation r683= 0.525, P<0.001;

Fig. 4.14).

Fig. 4.14 Mean values per quadrat of species richness and fruitbody abundance from the study. Light grey area = natural log of fruitbody numbers, dark grey area= species richness.

Ash plots

In Ash plots the number of litter-decay species found was negatively correlated

with the mean daily temperature in July (r2=23.5%). The abundance of litter-

decay fruitbodies show a negative relationship with the mean daily temperature in

July (r2= 36.6%; Fig. 4.15a).

The abundance of wood-decay fruitbodies show a positive relationship

with the mean temperature 3 weeks previous to fruiting (r2= 33.4%; Fig. 4.15b).

177

Total fungal fruitbody counts were negatively correlated with the temperature 2

weeks previous to sampling (r2= 27.2%).

The main factors which are necessary to have high species richness in an

ash site seem to be a period of low rainfall 2-3 weeks prior to sampling. A mild

summer (especially July) with low to medium temperatures and low rainfall

throughout the autumn season will also help with the increase in species richness.

The main factors which influence the amount of fruitbodies of fungi in ash sites

are low rainfall and low temperatures 3-4 weeks previous to sampling. Low to

moderate temperatures in June, July and to a lesser extent August will lead to an

increase in fruitbodies as will low temperatures and low precipitation during the

autumn season.

Oak plots

Litter-decay fruitbodies were found to be negatively related to rainfall 2 and 4

weeks previous and positively related to rainfall 1 week previous (r2= 37.9%).

Litter-decay fruitbodies were shown to have a positive relationship with the

maximum average temperature in July (r2=41%; Fig. 4.15c). Ectomycorrhizal

fruitbodies were positively correlated with annual total precipitation (r2= 19.5%;

Fig. 4.15d). Ectomycorrhizal fruitbodies were positively correlated with

temperature 4 weeks and precipitation 3 weeks prior to sampling (r2= 35.1%).

The inclusion of the ectomycorrhizal species in oak makes it more

complex than the ash sites, but there are some similarities. Litter-decay and wood-

decay species richness show the same pattern with rainfall 3 and 4 weeks prior to

sampling. A dry spell is followed 3 weeks later by an increase in species richness

of litter-decay and wood-decay fungi. A warm spell is followed by an increase in

ectomycorrhizal species richness 3 weeks later.

It is less obvious than in ash, but warm mild summer favours species

richness of litter-decay fungi. The amounts of litter- and wood-decay fruitbodies

in oak plots show similar correlations as in ash plots. A dry spell is followed 3

weeks later by an increase in litter- and wood-decay fruitbody numbers.

Conversely, ectomycorrhizal fruitbodies seem to experience increased production

with increasing rainfall up to 2 weeks prior to sampling. Temperature is highly

correlated with ectomycorrhizal fruitbodies in oak sites. It appears that constant

warm weather from week 4 right up to sampling causes an increase in the

178

ectomycorrhizal fruitbody numbers in oak sites. Mild to high temperatures and

low rainfall in July and August can cause an increase in litter-decay fruitbodies.

High rainfall in July and August can also increase the number of ectomycorrhizal

fruitbodies found. Rainfall during the June, July, August, September and October

season increases the amount of ectomycorrhizal fruitbodies found whilst

decreasing the amount of litter-decay fruitbodies found.

Scot’s pine plots

Litter-decay fruitbodies were negatively related to precipitation 3 and 4 weeks

previous to sampling and positively related to temperature 4 weeks previous to

sampling (r2= 31.2%). Litter-decay fruitbodies are also negatively related to the

temperature in August and positively related to July temperatures (r2=36.8%)

Sitka spruce plots

Wood-decay fruitbodies were positively related to rainfall levels in July (r2=

14.3%). It was found that few variables were correlated with species richness in

Sitka spruce plots. The temperature 4 weeks prior to sampling was found to have

a weak effect of the number of ectomycorrhizal species found. The amount of

rainfall 1 week prior to sampling is positively related to litter-decay fungal

fruitbody numbers.

All forest types

Litter-decay and ectomycorrhizal fungi were found to be the most controlled by

the weather variables examined. With regard to litter decay species richness and

fruitbody production, a spike in rainfall 3 and 4 weeks prior to sampling led to

increased litter-decay species and fruitbody production (Figs. 4.16; 4.17). Rainfall

spikes such as those at x coordinates 09/09/08 and 21/09/09 correspond to low

values for both litter-decay species richness and fruitbody production (Figs. 4.16;

4.17). The relationship also proved to be statistically significant (Fig. 4.18).

Ectomycorrhizal species richness and fruitbody production were related to

temperature in oak and Sitka spruce plots but not Scot’s pine plots. The large

clusters of Scot’s pine plots geographically, mean that there was little variation in

weather variables from the Scot’s pine plots which may have caused the

relationship between fungi in Scot’s pine plots and weather variables to be lost.

The temperature from 1-4 weeks prior to sampling has a large effect on the

179

ectomycorrhizal species richness and ectomycorrhizal fruitbody numbers in oak

and Sitka spruce forests. Wood-decay fungi are related and controlled by fewer

environmental variables than either ectomycorrhizal or litter-decay. Perhaps this is

due to the more stable micro environment which the wood-decay fungi are found

in.

(a)

(b)

(c)

(d)

Figs 4.15 Regression curves between: (a) LnLDFrb (Natural log litter-decay fungal fruitbodies) and TempAveJuly (mean temperatures for July) in ash plots, (b) LnWDFrb (Natural log wood-decay fruitbodies) and Temp3Weeks (Mean daily temperature 3 weeks prior to sampling) in ash plots, (c) LnLDFrb (Natural log of litter-decay fungal fruitbodies) and TempMaxAveJuly (Mean maximum daily temperature during July) in oak plots, (d) LnMYCFrb (Natural log of mycorrhizal fruitbodies) and TotRainfall (the total yearly rainfall) in oak sites.

180

Fig. 4.16 The abundance of fruitbodies (natural log of values) of litter-decay species with precipitation levels 4 weeks prior to sampling. Black line = Precipitation 4 weeks previous to sampling, bar= mean of litter-decay fruitbodies per plot.

Fig. 4.17 The mean species richness of litter-decay fungi to the mean precipitation levels 4 weeks prior to sampling. Black line = Precipitation 4 weeks previous to sampling, bar= mean of litter-decay species per quadrat.

181

Fig. 4.18 Regression curve of relationship between the natural log of litter-decay fungal fruitbodies (LnLDFrb) and the mean daily precipitation 4 weeks prior to sampling (Prec4Week) in all plots.

182

4.5 Discussion

4.5.1 Fungal species richness and the effect of forest type

The total number of fungal species identified in this project was 416, which is on

par with similar studies in the U.K., mainland Europe and North America (Table

4.16). The most closely comparable studies are those of Ferris et al. (2000a) and

Humphrey et al. (2000) in the U.K. Ferris et al. (2000a) recorded 343

macrofungal species in Scot’s pine and Norway spruce sites in England, while

Humphrey et al. (2000) reported 419 species in forests of Sitka spruce, Scot’s

pine and oak in the U.K. (England and Scotland). In a study of oak forests in

Sweden, Ruhling and Tyler (1990) identified 470 macrofungal species over a 3

year study. A long-term survey (>7 years) by Straatsma and Krisai-Greilhuber

(2003) identified 886 taxa of macrofungi to species level in woodlands near

Vienna, that were composed of oak, beech, pine and ash. In North America, Smith

et al. (2003) found 263 taxa of fungi from old growth Douglas fir forests in

Oregon, U.S.A. Outerbridge (2002) found a total of 277 taxa of macrofungi

working in coniferous forests (Sitka spruce, Douglas fir, Western hemlock and

Western red cedar) on Vancouver Island, British Columbia. Another longer study

on Vancouver Island identified 551 taxa of fungi (Roberts et al. 2004) from six

different habitats.

183

Table 4.16 Functional group species richness from macrofungal studies in forests in 9 different regions. Values in columns 3-7 are numbers of species in particular functional group, values in parenthesis are percentage of total species which functional group represents where given.

Country/

region

Tree species S LD ECM P WD Reference

Ireland Fe, Qr, Qp, Ps, Pi 221 65 (29) 83 (38) 7 (3) 66 (30) This project

England Pa, Ps 343 100 (29) 99 (29) 7 (2) 133 (39) Ferris et al. 2000a

U.K. Ps, Pi, Qr 419 157 (37) 174 (42) 12 (3) 76 (18) Humphrey et al. 2003

U.K. Ps, Bp, Pi or Pa 662 338 (51) 324 (49) Orton 1987

Scotland Pi, Bp 109 37 (34) 71 (65) 1 (1) Alexander and Watling 1987

Switzerland Fs, Qp, Pa, Pm, Ps, Pt, Ld

408 125 (31)a 265 (65) 18(4) Straatsma et al. 2001

Finland Ps 152 62 (41) 90 (59) Vare et al. 1996

Finland Ps 170 72 (42) 97 (57) 1 (1) Tarvainen et al. 2003

Finland Ps, Pa, Bp 316 190 (60) 126 (40) Salo 1993

Portugal Qs 171 96 (56) 74 (43) 1 (1) Azul et al. 2009

Portugal Qs 73 13 (18) 60 (82) Baptista et al. 2010

Spain Ps 164 20 (12) 144 (88) Bonet et al. 2004

Spain Fs, Qr 125 25 (20) 69 (55) 3 (2) 28 (22) Sarrionandia et al. 2009

Spain Qi, Qs 838 457 (55) 381 (45) Ortega and Lorite 2007

Canada Pi, Pc, Th, Tp, 551 263 (48) 272 (49) 15 (3) Roberts et al. 2004

Canada Pi, Pm, Th, Tp 277 114 (41) 97 (35) 2 (1) 64 (23) Outerbridge 2002

Canada Pi, Pm, Th, Tp 62 28 (45) 9 (15) 1 (2) 24 (38) Berch et al. 2001

Austrailia Em 149 60 (40) 89 (60) Anderson et al. 2010 S= total species, LD= litter-decay species, ECM= ectomycorrhizal species, P= parasitic species, WD= wood-decay species. Tree species codes: Bp= Betula

pubescens, Ca= Coryllus avellana, Cm= Crataegus monogyna, Cs= Castanea sativa, Em= Eucalyptus marginata, Fe= Fraxinus excelsior, Fs= Fagus sylvatica, Ld= Larix decidua, Ps= Pinus sylvestris, Pi= Picea sitchensis, Pc= Pinus contorta, Pa= Picea abies, Pm= Pseudotsuga menziesii, Pt= Pinus strobes, Qr= Quercus robur, Qp= Quercus petraea, Qs= Quercus suber, Qi= Quercus ilex, Th= Tsuga heterophylla, Tp=Thuja plicata.

184

Sitka spruce forest yielded the greatest species richness of macrofungi. A

total of 144 fungi were found in Sitka spruce forests, with 50 of these species

found only in these forests. This high species richness may reflect the ability of

Sitka spruce to accommodate a wide range of macrofungi (Alexander and Watling

1987). That Sitka spruce was also the most species rich according to rarefaction

analysis, shows that the high species richness of the Sitka spruce forest type was

not an artefact of the study design, which included more Sitka spruce plots.

Humphrey et al. (2000) found 269 macrofungal species in forests primarily

composed of Sitka spruce, with 73 of these species found only in this forest type.

Outerbridge (2002) investigated Sitka spruce in its home range on Vancouver

Island and found that in supported 37 fungal species which were not found in any

of the other forest types examined. To compare these previously mentioned

studies on similar scales, species per m2 data were calculated (Table 4.17). The

results from this study fall between those of the U.K. (Humphrey et al. 2003) and

Canadian studies (Outerbridge 2002). The study listed by Roberts et al. (2004)

used transect sampling, and so meaningful comparisons are difficult to draw. At

0.16 fungal species per m2, Sitka spruce forests are the second most species rich

forests (after oak) per unit area investigated in this study.

The high numbers of macrofungal species found in Sitka spruce

plantations in Ireland and the U.K. is a welcome fact for macrofungal

conservation, especially as the levels of Sitka spruce plantations are set to increase

throughout Europe (UNECE 2003). Several studies have shown that plantation

forests in Ireland can provide a habitat for a range of native flora and fauna such

as: vascular plants (French et al. 2007; Smith et al. 2007), spiders (Oxbrough et

al. 2006), beetles (Oxbrough et al. 2010), birds (Sweeney et al. 2010b) and mites

(Arroyo et al. 2010). The results of this study supports the views of Lindenmayer

et al. (2008) and Brockerhoff et al. (2008) that the afforestation with exotic tree

species can have positive effects on the diversity of other taxa, as long as native

woods are not removed to make way for the plantations (Quine and Humphreys

2010). Rarefaction analysis did not find significant differences between the total

species richness in young versus mature Sitka spruce plantations. Previous

research has shown a significant effect of age on species richness (Dighton et al.

1986), but the age groupings used in this study are not identical to those used in

the study of British Sitka spruce plantations by Dighton et al. (1986). It would be

185

expected that the community composition of the two age groupings of spruce

plots would be different and this is investigated later (Chapter 5).

The relatively lower numbers of macrofungal species found in ash forests

is not surprising as ash does not support ectomycorrhizal fungi. Those

ectomycorrhizal species that were found in ash plots were probably due to the

presence in the plots of roots from nearby trees of hazel or oak. In Ireland at least,

forests of pure ash are difficult to find. It is one of the most nutrient-demanding

tree species (Horgan et al. 2004) and it most often found planted in mixed

broadleaf plantations. Ash forest was the least species-rich forest type examined

per unit area with only 0.11 species per m2 (Table 4.17). There are very few

published studies on the macrofungi of ash forests. One such is the seven-year

study in a variety of Swiss forest habitats by Straatsma and Krisai-Greilhuber

(2003); they investigated the fungal communities in a very large mixed ash and

elm (Ulmus minor) plot. Over the seven years, they found a much larger number

of macrofungal species (228 species) than was found in this study (56 species).

However, when the two studies are equated in plot size by converting the results

to fungal species per m2, this study has a higher number of species at 0.11 species

per m2, while the Swiss study has only 0.02 species per m2 sampled (Table 4.17).

The study by Hering (1966) in an English ash woodland on limestone revealed

similar total species richness as this project with 59 species found over three

years. However, Hering undertook much more intense sampling of his plots,

visiting each one up to nine times a year for three years. The other two ash forest

studies (Table 4.17) used small-scale sampling strategies, and comparison

between these and this project are not meaningful.

Oak, one of two native tree types examined in this study, showed a very

high species richness of macrofungi (113 species) and the highest richness per m2

(0.23 species per m2) (Table 4.17). The oak forests of Ireland have been noted as

mycologically rich habitats (Ramsbottom 1936), especially the annexed habitat:

old sessile oak woods with Ilex and Blechnum in the British Isles (EUROPA

2009). The study by Humphrey et al. (2003) of British woodlands, found a total of

284 macrofungal species in oak forests (all pedunculate oak Quercus robur),

which equates to 0.32 species per m2 sampled, the most species-rich forest type on

an area basis. The study by Sarrionandia et al. (2009) found 99 species in oak (Q.

186

robur) forests in Northern Spain, or 0.06 species per m2. Differences in oak

species, the climate and in forest composition are possible reasons for the lower

species richness in the Spanish study. The study by Straatsma and Krisai-

Greilhuber (2003) examined a 16,000m2 plot with sessile oak as its dominant tree

species, and found 132 macrofungal species over seven years of sampling which

equates to 0.008 species per m2. Although a positive relationship has been

identified between forest size and macrofungal species richness (Peay et al. 2007)

it is unlikely that this relationship is linear with constant increases in species

richness as forest size increases. In the case of the oak plot in Straatsma and

Krisai-Greilhuber (2003), it may be possible that similar levels of species richness

could have been identified using a smaller plot size and therefore the values for

species/m2 are unrealistically low.

Of the two coniferous forest types Scot’s pine and Sitka spruce, the former

type was expected to have a higher overall diversity of macrofungi, given that 3 of

the Scot’s pine sites were in old stands and on previously forested areas, and due

to Scot’s pine being present in Ireland longer than Sitka spruce. However, this

was not the case, as Sitka spruce forest type yielded 144 macrofungal species

compared to Scot’s pine’s 89 species, and spruce forests were also more species

rich according to rarefaction.

A comparable study of plantation Scot’s pine stands in the U.K. (Ferris et

al. 2000a) identified 253 macrofungal species over a period of three years.

Humphrey et al. (2003) found 357 species of fungi in Scot’s pine forests in

England and Scotland. These forests were also plantation forests, but with

differing land use history, including six plots which were once native pine forests

and nine plots which were not native pine forests. It might be expected that the

plots which were once native stands would harbour greater macrofungal species

richness, but this was not the case as the non-pine past land use plots contained

285 fungal species while the plots used as native stands in the past contained 168

macrofungal species (C. Quine, unpub. data). The number of plots sampled was

not equal between the two land use histories, and so these comparisons of species

richness may not be realistic. In this study, the opposite result was found as

second rotation Scot’s pine forests were shown (by GLM) to have a higher species

richness than first rotation plots. Although Ferris et al. (2000a) did not investigate

the effect of past land use on fungal species richness, they did hypothesise that it

187

may be a key variable in explaining some of the variation in species richness

between their pine plots.

Studies on Scot’s pine macrofungi in Spain carried out by Oria de Rueda

et al. (2010) and Bonet et al. (2004) identified 84 and 164 species respectively,

while Tarvainen et al. (2003) found 168 species in Scot’s pine forests in Finland.

On a species richness per area basis, this study averaged 0.1 fungal species per

m2, which was lower than all of the studies listed in Table 4.17 with the exception

of the Oria de Rueda (2010) study. This low number of species per m2 is most

likely not a true representation of macrofungal diversity in Irish Scot’s pine

forests. Mature forests (as 4 of the 8 Scot’s pine plots could be regarded in this

study) are known to have a high number of rarer species that infrequently produce

sporocarps (O’Dell et al. 1999; Smith et al. 2003), which would therefore require

a longer period of sampling to discover. Morevover, since there are no surviving

native Scot’s pine forests in Ireland today (Mitchell 1988; Cross 1998), inoculum

sources for Scot’ pine forest fungi are not available as would be the case in

Scotland.

188

Table 4.17 Comparison of studies on macrofungal diversity in a range of forest types. S= total species found, S/m2 is the average number of species found per metre square sampled. SP= Scot’s pine, SS= Sitka spruce. a= Number of branches sampled, b Transect sampling was used in this study.

Tree

type

Area

surveyed

(m2)

S S/m2 Reference

Ash 500 56 0.11 This study

Ash 12000 228 0.02 Straatsma and Krisai-Greilhuber 2003

Ash 300 59 0.2 Hering 1966

Ash 18a 12 N/a Boddy et al. 1987

Ash 25a 12 N/a Unterseher et al. 2005

Oak 500 113 0.23 This study

Oak 880 284 0.32 Humphrey et al. 2003

Oak 1600 99 0.06 Sarrionandia et al. 2009

Oak 16000 132 0.008 Straatsma and Krisai-Greilhuber 2003

SP 900 89 0.1 This study

SP 1200 357 0.3 Humphrey et al. 2003

SP 560 253 0.45 Ferris et al. 2000a

SP 900 84 0.09 Oria de Rueda et al. 2010

SP 360 164 0.46 Bonet et al. 2004

SP 1000 168 0.17 Tarvainen et al. 2003

SS 800 144 0.16 This study

SS 1280 269 0.21 Humphrey et al. 2003

SS 864 127 0.15 Outerbridge 2002

SS >20000b 211 0.01 Roberts et al. 2004

4.5.2 Rare and common fungi in Irish forests

This three-year study recorded 48 species which are new to Irish forest habitats,

suggesting that additional and longer studies may reveal more new species in

Irelands largely uninvestigated forest macrofungal communities (O’Hanlon and

Harrington in press). There is currently no red-data list for fungi in Ireland, while

a list does exist for the U.K. (Ing 1992; Evans et al. 2003), although it is not

without its critics (Orton 1994). In this study, the U.K. red-data list was applied to

the species found in this project as at this time it is the best option for identifying

areas which may be in need of conservation status. Six red-listed macrofungal

species were found during this study and 198 species were found whose

distributions are not fully known. Confirmation of the presence of these species in

Ireland is a significant addition to the knowledge base of macrofungi in Ireland.

189

The list of the most species rich genera found from Irish forests (Fig. 4.19)

shares many similarities with the most commonly recorded fungi from Ireland

(O’Hanlon and Harrington in press) and the U.K. (FRDBI 2010). A more

focussed surveying of the grassland fungi in Ireland in the past, has led to high

levels of species richness in common grassland genera such as Entoloma and

Hygrocybe, which are not reflected in the genera richness levels found during this

study (Fig. 4.19).

Of the top 20 most commonly recorded fungi from this project, six are

present in the most common list for Ireland (O’Hanlon and Harrington in press),

five species are present in the U.K. top 20, five more are in the U.K. top 100 and

another five are within the top 150 species in the U.K. Seven of the top 20 most

commonly-found species in this project were also in the top 20 macrofungal

species in the study of plantation forests in England and Scotland (Humphrey et

al. 2003). Species that were common in this study but are not common in either

the list of common Irish species or U.K. woodland species (Humphrey et al. 2003)

include Cortinarius flexipes, Marasmius hudsonii, and Mycena rorida. Both

Mycena rorida and Cortinarius flexipes are listed as occasional in the U.K. by

Phillips (2006). According to the checklist of Basidiomycetes in Britain and

Ireland, Cortinarius flexipes is present in the U.K. and Ireland but its distribution

is not well known. Throughout this project Cortinarius flexipes was found in 13

oak plot visits, 4 Scot’s pine plot visits and 8 Sitka spruce plot visits. Taking these

values as an indication of its distribution in Ireland, it could be said to be

occasional to common in Irish oak forests and conifer plantations. Mycena rorida

is listed as occasional in the U.K. according to the Checklist of Basidiomycetes in

Britain and Ireland (Legon and Henrici 2005), whilst its distribution is unknown

in Ireland. In this study it was collected in 7 ash plot visits, 5 Scot’s pine plot

visits and 10 Sitka spruce plot visits. In the ash sites it was most commonly found

on Rubus fruticosus stems, while in the pine and spruce plots its substrate was

usually small conifer twigs. The expansion of Ireland’s plantation forests (Anon.

1996) may have the effect of making species like Cortinarius flexipes and Mycena

rorida more common in Ireland. Marasmius hudsonii is a saprobe of holly leaves.

It seems to be more common in Atlantic oak woods. It was not found in a survey

of Scottish western oak woods (Watling 2005), but has been found on a number of

the British Mycological Society’s forays (FRDBI 2010). Its distribution data show

190

that it is mainly found in coastal forests in Ireland and England and so may be

restricted mainly to oak woodlands that experience a mild wet climate; as many of

the Irish oak woods do.

The genera that were most species rich in this study (Fig. 4.19) are ones

which are also very diverse in similar studies of forests elsewhere. The 5 most

species-rich genera were Cortinarius, Mycena, Russula, Lactarius and Inocybe. In

Scot’s pine, oak and Sitka spruce plantation forests in the U.K., (Humphrey et al.

2003; C. Quine, unpub. data) recorded 47 species of Cortinarius, 22 species of

Inocybe, 24 species of Lactarius, 47 species of Mycena and 33 species of Russula

with Cortinarius, Galerina, Inocybe, Lactarius, Mycena and Russula being the 1st

to 6th most diverse genera. Smith et al. (2002) found Cortinarius, Inocybe and

Russula to be the most diverse genera in Douglas fir forests in Oregon. Studies in

the Sitka spruce, Douglas fir, Western red cedar and Western hemlock forests of

Vancouver Island, found Mycena to be the most diverse genus, followed by

Cortinarius, Inocybe, Lactarius and Russula (Outerbridge 2002; Roberts et al..

2004). Molina et al. (1992) noted that the genus Cortinarius has very few species

(<20%) which are restricted to specific hosts with which they can form

ectomycorrhizas. Inocybe and Russula are similar in their host specificity, but

Lactarius has a significantly higher proportion (40%) of host-specific species

(Molina et al. 1992).

191

Cort

inar

ius

Myc

ena

Russ

ula

Lacta

rius

Inocy

be

Entolo

ma

Bole

tus

Am

anita

Tricholo

ma

Clit

ocybe

Hyg

rocy

be

05

101520253035404550556065

Decreasing species richness in this project

No

. o

f sp

ecie

s i

n I

rela

nd

Fig. 4.19 Comparison of species rich-genera from this project and from the Republic of Ireland. The genera are ordered from left to right along the x-axis according to decreasing species richness found in this project. The values on the y-axis are the total species richness for that genus in the Republic of Ireland (FRDBI 2010).

4.5.3 Functional groups of fungi in Irish forests

The relative proportions of species in the different functional groups across all

forest types were: 29% litter-decay, 38% ectomycorrhizal, 3% parasitic and 30%

wood-decay species. These proportions follow the commonly-found trend from

other macrofungal diversity studies (Table 4.16). Ectomycorrhizal fungi typically

constitute 30-50% of the fungal species in forest assessments (Watling 1995). The

exception in this study was for ash forest because ash does not support

ectomycorrhizal fungi.

Sitka spruce plots had significantly higher proportions of litter-decay

species, primarily Mycena species (20 species), than the other forest types. Sitka

spruce plots also had higher numbers of Collybia and corraloid fungi. Spruce

forests are known to support a high diversity and sporocarp abundance of Mycena

species (Dighton et al. 1986; Senn-Irlet and Bieri 1999; Outerbridge 2002). Sitka

192

spruce forests harbour large numbers of Mycena species in its home range in

Canada; Outerbridge (2002) found 45 species of Mycena on the thick layer of

needles in her Sitka spruce sites. A total of 28 species of Mycena were found in

Sitka spruce plots in the U.K., the second highest number of Mycena species after

Scot’s pine plots found in the study of Biodiversity in Britain’s planted forests (C.

Quine, unpub. data). In this study, it was found that the age of the spruce forest

had a significant effect on the litter decay species richness of the plots. Litter

decay species richness was seen to be negatively related to forest age. This finding

is similar to that made by Dighton et al. (1986) in a study of Sitka spruce

plantations in the U.K., when the age groupings used in the U.K. study are

amalgamated into similar groupings as used here.

Sitka spruce had the second highest proportion of ectomycorrhizal species

after Scot’s pine. The total numbers of ectomycorrhizal species recorded with

Quercus, Picea and Pinus in the U.K. are 233, 151, and 201 respectively (Newton

and Haigh 1998), and therefore it was expected that the oak forest type would

have a higher proportion of ectomycorrhizal species than the coniferous forest

types. However, this was not the case, as oak had a lower ectomycorrhizal species

richness per plot visit than Sitka spruce (although not significantly different).

Another factor which must be taken into account in explaining the lower number

and proportion of ectomycorrhizal species in oak than in Sitka spruce plots, is the

number of other mycorrhizal trees which were part of the plots. It was attempted

to position the plots in such a way to exclude tree species other than the tree

species of interest. This was difficult in the coniferous sites, as there were often

randomly placed birch trees in the plots. Species such as Lactarius glyciosmus, L.

torminosus and L. quietus were all present in the coniferous plots, although almost

definitely in association with birch for the first two species and with oak trees in

the plots for the latter species.

Oak and Sitka spruce forest contained the highest number of wood-decay

species, but in relative terms this group were most abundant in ash sites, where

they comprised 45% of the total number of species. Average wood decay species

per plot visit were significantly higher in oak than in Scot’s pine sites. This

difference is most likely due to the presence of more coarse woody debris (CWD)

in oak sites than in Scot’s pine sites (Table 3.15, Chapter 3). The lack of quantity

and high quality of CWD in Irish plantation forests has been highlighted by

193

Sweeney et al. (2010a), and they proposed a diversity of log sizes and qualities

that should be left on site, which would increase the invertebrate and fungal

diversity of the sites. As CWD quantity has been shown to be positively related to

the number of wood-decay species (Ferris et al. 2000a), increasing the quantity

and quality of CWD on site would also improve the number of fungal species.

The species richness of the macrofungi was similar between the plots of a

similar forest type. Larger differences in species richness values between sites of

the same forest type were most noticeable in the Sitka spruce forest type. It was

noted in two of the Sitka spruce plots that large slash piles bisecting the plots were

often the habitat for a number of wood decay species e.g. Mycena species (R

O’Hanlon pers. observation). Also, plots which had other ectomycorrhizal tree

species present near the plots had higher ectomycorrhizal species richness. This

finding is not surprising, as it is known that fungal species richness shows a

positive relationship with the number of tree species present (Ishida et al. 2007).

The amount of CWD present in sites has been shown to increase the richness of

decay fungi (Ferris et al. 2000a), and increasing the number of tree species causes

an increase in the number of fungi present (Ferris et al. 2000a; Schmit et al.

2005). Therefore, the guidelines put in place to govern the planting and

management of plantation forests in Ireland (Anon. 2000) which state that

minimum levels of CWD must be present, and that main tree species cannot

exceed 80% of the plantation, are both beneficial to overall macrofungal species

richness and diversity of functional groups.

4.5.4 Species and fruitbody abundance correlated to weather parameters

It is well known that macrofungi are highly affected by weather conditions

(Ohenoja 1993; Gulden et al. 1992; Carrier 2003). In this study, the main effects

of climatic conditions differed between the different forest types. One of the few

conclusions which hold true for all of the forest types investigated is that litter-

decay fruitbodies are negatively affected by large amounts of rainfall 3-4 weeks

previous to sampling. It has been shown by experimentation that excess water can

stimulate the growth of the mycelium to the detriment of fruiting body formation

(Manachere 1980; Pinna et al. 2010). In the present study, ectomycorrhizal fungal

194

fruitbodies were found to be more abundant in sites with a higher annual rainfall.

Similar results were found by Straatsma et al. (2001) and O’Dell et al. (1999),

where mean annual rainfall was found to be positively correlated to

ectomycorrhizal species richness and fruitbody abundance. This led O’Dell et al.

(1999) to state that ectomycorrhizal fungi may show a unimodal response to

precipitation levels.

4.5.5 Estimating fungal diversity in Irish forest habitats

The rarefied species accumulation curve had not levelled off for any of the forest

types by the end of the three year stud. It is highly probable that the number of

species from the plots would continue to increase with more sampling effort. An

important point to note from Fig. 4.3 is that the rarefaction estimate of the species

richness of the forest types is lower than the actual species richness at the

maximum sampling intensity. This is an artefact of the random sampling with

replacement sampling method used in the analysis. As samples are chosen

randomly and then replaced back into the data set, some samples can be chosen

twice, and some are not chosen at all. This has the effect of making the rarefaction

estimate fall below the actual estimate (Colwell 2009). This artefact however is an

acceptable pitfall of sample-based rarefaction with replacement, as without

replacement of samples confidence intervals would be meaningless in the upper

end of the rarefaction.

From examination of the species richness estimates the most complete list

of fungi for the forest types was from the Sitka spruce followed by the oak, Scot’s

pine and ash forest types with an estimated 77, 65, 52 and 45% of their fungal

species richness realized respectively. Schmit et al. (1999) tested the application

of a number of species richness estimators to macrofungal survey data from a

mixed oak and maple forest that included 177 identified species. They produced

richness estimates ranging from 193 to 348 species or somewhere between 51 and

91% of the species richness realised from two year’s sampling. They concluded

from their data that at that point in time there was no diversity estimator which

was suitable for use on fungal diversity data, but they did not include the ICE or

ACE estimators in their study. Unterseher et al. (2008) also tested fungal data

195

using species richness estimators. Out of the total 146 species they identified

(including 109 macrofungi) they produced diversity estimates ranging from 178 to

210 species or between 70 and 85% of the species discovered from their selected

wood samples. They concluded that the Chao2 estimator was the most applicable

as it gave a realistic figure for species richness and reached an asymptote before

100% sampling effort was reached.

From the data and analysis in this study, it can be concluded that, with

three years of sampling of the macrofungal communities in these temperate forest

types, none of the estimators fulfil the selection criteria set out by Magurran

(2004). The Chao2 estimator is used here as it has been identified as a useful

lower limit richness indicator when other estimators fail to meet the selection

criteria (Longino et al. 2002). The Chao2 which uses qualitative data was chosen

over similar estimators which take species abundance into account, because

fungal species, particularly ectomycorrhizal species, are known to be non-

randomly distributed (Taylor 2002; Tedersoo et al. 2003), and so the assumption

of homogeneous distribution of fungal species is not met. Patchy distributions

have been shown to severely diminish the ability of abundance-based

(quantitative) estimators such as ACE and Chao1, but the Chao2 estimator has

been shown to be rather robust when used on moderately patchy distributions

(Chazdon et al. 1998). It has been noted that richness estimates in hyperdiverse

communities such as ectomycorrhizal communities (Tedersoo et al. 2006), are

problematic when species richness estimators are used, due to the large number of

rare species. In these cases, the species richness estimate should be taken as a

lower estimate of diversity (Mao et al. 2005).

Comparing rarefied species richness values between the sites reveals that

Sitka spruce plots are amongst the five most species rich plots. The top five most

species-rich sites are Abbeyleix (oak), Dooary (SS), Raheen (oak), Bohatch (SS)

and Stanahely (SS). The fact that Sitka spruce sites rank in the top five most

species-rich sites shows the potential of Sitka spruce plantations as habitats for

forest fungi in Ireland, as has already been shown for spruce plantations in the

U.K. (Humphrey et al. 2000).

Also worthy of note is that the St John’s wood (ash) site ranks as the

seventh most species-rich site based on rarefied species richness. This site, which

is one of the few remaining semi-natural forest habitats in Ireland shows that even

196

though the site has very few ectomycorrhizal fungi, mature semi-natural ash

forests can have a very high diversity of macrofungi present. Straatsma and

Krisai-Greilhuber (2003) showed the high diversity of macrofungi which can be

present in old growth ash forests in Switzerland.

Another unexpected result to come from the analysis of rarefied species

richness values for the sites was that the old growth Scot’s pine forest at Torc,

which is believed to be one of the oldest in Ireland (Mitchell 1988) had such a low

species richness. The invasion of rhododendron in this site may well have

negative effects on the fungal species richness of this site. The presence of this

invasive plant in this plot makes comparisons between similar sites difficult

because based on vegetation present, the Torc site is most similar to the pine sites

at Annagh and Ballygawley. However, these sites are planted on different soil

types and are also much younger than the Torc site so comparisons of fungal

diversity between the sites are not reasonable. The Biodiversity in Britain’s

planted forests project (Humphrey et al. 2003; C. Quine, unpub. data) investigated

a total of 15 Scot’s pine plots, and found that their Scot’s pine plot at Strathspey

had the lowest species diversity of all of their pine plots with 48 species recorded.

Their site at Strathspey is the most similar to the Torc site in this project, an old

pine plot with a vegetation community consisting of dominant moss cover with

Calluna vulgaris also present corresponding to Vaccinium myrtillus-I. aquifolium

(Br. –Bl. Et Vlieger 1939).

The most species-rich Scot’s pine site in this project was the site at Brittas

which had 25 species. This site had oak (Quercus petraea) trees in the understory.

It is worth mentioning that in the Humphrey et al. (2003) study, the most species

rich Scot’s pine site was found to be a ~70 year old Scot’s pine site at New Forest

which also had a number of oak trees in its understory. Examination of the data

from the U.K. study (C. Quine, unpub. data) study show just how species rich

Scot’s pine plots can be if thorough systematic sampling is carried out. They

found an average of 68 species per Scot’s pine plot, with a range of species

richness from 48 to 124 species. However, as there are no pristine Scot’s pine

woodlands present in Ireland, the species richness of the Scot’s pine plots could

not be expected to be as high as those from the U.K. Humphrey et al. (2000)

found that the closer a forest plantation was to a pristine forest of similar tree type,

the higher the number of rare species found in that forest so it could be concluded

197

that the lower species richness in Irish Scot’s pine forests may be due to the lack

of pristine pine forests in Ireland.

Species richness (the total number of species in a given area) has become

the common currency in biodiversity research (Gaston and Spicer 2007). Species

richness is a straightforward measure, but does not convey any information about

the relative frequency of occurrence of species in an area or the probability of

finding a particular species in the area. For this, an abundance-based diversity

index is required. A conceptual difficulty with using such indices in assessment

of macrofungal diversity using sporocarps is that there is no correlation between

the occurrence of a sporocarp and an “individual” fungal mycelium. Nevertheless

such indices are often used in assessments of fungal diversity (Azul et al. 2009;

Berch et al. 2006; Humphrey et al. 2000; Gebhardt et al. 2007; Straatsma et al.

2001). One of the main conclusions that can be made from the diversity index

analysis of different forest types is that, overall Sitka spruce plantations can be

very diverse, almost equalling the fungal diversity of oak forests. However, they

are also the most uneven in terms of species distributions across the sites in this

study. Taking the values for the Simpson’s diversity index as the best and most

meaningful estimate of diversity (Magurran 2004), the Scot’s pine forest type is

the most diverse followed by the oak, Sitka spruce and ash forests.

Another method often used to estimate total macrofungal species richness

is to relate macrofungal and plant species richness (Villeneuve et al. 1989;

Hawksworth 1991; Watling 1995; Schmit and Mueller 2007). Estimates of

possible macrofungal species richness were made using an established 2:1 fungal

to plant species ratio in Section 3.3.5, Chapter 3.

198

Ash

1

Ash

2

Ash

3

Ash

4

Ash

5

Oak

1

Oak

2

Oak

3

Oak

4

Oak

5SP

1SP

2SP

3SP

4SP

5SP

6SP

7SP

8SS

1SS

2SS

3SS

4SS

5SS

6SS

7SS

8SS

9

0

5

10

15

20

25

30

35

40

45

50

55 Estimated macrofungal species richness

Actual macrofungal species richness

Plot name

No

. o

f sp

ecie

s

Fig. 4.20 Estimated and actual macrofungal species richness of the forest plots. Macrofungal species richness was estimated using the 2:1 (Villeneuve et al. 1989) ratio of macrofungal to plant species from temperate forests in Canada. Abbreviations for the plot names can be found in Table 3.22, Chapter 3.

In general, the estimates based on macrofungal to plant species richness

are quite accurate for the ash and Scot’s pine forest types (Fig 4.20). The

estimates from the oak and Sitka spruce plots are not accurate, with 4 out of 5 of

the oak plots being underestimated for macrofungal species richness, and all of

the Sitka spruce plots being underestimated according to the plant to macrofungal

species richness ratio (Fig. 4.20). From the findings of this project, the ratio for

macrofungi to plant species in the oak and Sitka spruce forests is over 2:1. For the

oak forest type the average across all plots is 3.2 (±1sd):1 macrofungal to plant

species richness. The Sitka spruce forest type has an even higher ratio at 5.8

(±3sd):1 macrofungal to plant species ratio. In these forest types the established

ratios do not hold true, a finding that would indicate that estimates across large

regions using this method should be taken as a lower bound for macrofungal

species richness (Schmit and Mueller 2007).

199

4.6 Conclusions

• This study identified a large number of macrofungal species over a relatively

short period. A total of 416 species were identified, including 48 possible new

records to the Irish fungal flora and 6 species on the British red data list.

• In this study, Sitka spruce plantations were found to have the highest total

species richness and rarefied species richness of macrofungi of all the forest

types. Sitka spruce also contained a high proportion (35%) of species not

found in any other forest type in this project.

• Spruce plantations were found to have high total species richness estimates,

but also high variability between the different plots examined. Neither

rotation stage nor chronological stage was useful in explaining these intra-

forest type differences.

• Plantation Scot’s pine forests were also found to be relatively species-rich

habitats for macrofungi and this was made up of a high proportion of

ectomycorrhizal species.

• It was found that the species richness of the different functional groups were

different across all forest types. Oak forests had higher wood decay species

richness than the coniferous forest types and this was most likely due to the

higher amounts of coarse woody debris present in the oak forests.

• This study found that that species richness estimators did not fulfil the

necessary criteria for the successful application of an estimator, and so their

future use on macrofungal data from studies of this duration and in these

forests types is questionable.

• The use of macrofungal to plant species richness ratios were not accurate for

two of the four forest types examined in this project. The established ratio of

2:1 macrofungal to plant species underestimated macrofungal species

richness in the oak (actual ratio 3.2:1) and Sitka spruce (actual ratio 5.8: 1)

forest types.

200

201

Chapter 5: Macrofungal communities of

the forest types

202

203

5.1 Introduction

5.1.1 Mycocoenological study methods

The primary aim of this chapter is to determine if there are discernable differences

in the distribution of macrofungi between the different forest types, such that

distinct assemblages or communities of macrofungi can be recognised. The study

of macrofungal communities (mycocoenology) of forest ecosystems is based on

the methods used to describe the plant communities (phytocoenology) of those

ecosystems (Arnolds 1981; Barkman 1976), and treats a fungal sporocarp as a

recording unit just as an individual plant may be considered a unit. The

application of community concepts and sampling methodologies to fungal

assemblages faces a number of difficulties, which are unique to fungi and

macrofungi in particular:

1) Although next generation DNA identification methods such as

pyrosequencing have significantly increased the ability to identify the large

diversity of macrofungal species present in ecosystems (Hibbett et al. 2011),

the identification of fungi by taxonomic examination of their sporocarps is

still the most common method. A pitfall of using sporocarps is that species

that do not produce macroscopic sporocarps or fruit irregularly are not

considered or are likely to be overlooked.

2) The lack of consistent patterns between the abundance of sporocarps above-

ground and the distribution of mycelium below-ground (Horton and Bruns

2001; Peter et al. 2001) means that it is not possible to quantify the actual

abundance of a species in real ecosystems.

3) The sporadic and variable annual fruiting patterns of macrofungi necessitate

extended studies spanning a number of years. Three years has been

recommended as a minimum (Hering 1966), but much longer periods may be

required for a full inventory of the macrofungi (Arnolds 1988; Watling 1995).

The effectiveness of a study is a function of duration by sampling frequency.

The former may be reduced if the latter is increased or vice versa for a

particular level of effectiveness. Sampling frequency should take account of

differences in the lifespan of different sporocarps. Intervals of 3-4 days have

204

been suggested (Watling 1995) if the entire macrofungal community is to be

described including the very transient sporocarps such as those of the genus

Mycena (Richardson 1970).

4) The concept of community when applied to fungi, poses a number of

conceptual and practical difficulties. The term fungal community is analogous

to the term plant community, i.e. it is a neutral term for any concrete

assemblage of fungi that grow together in a uniform habitat, without

considering any differences in resource exploitation among the assemblages’

members (Arnolds 1992). For example, the fungal community is not

functionally homogenous because it contains symbiotic, parasitic and

saprotrophic elements. The habitats occupied by these different groups could

hardly be considered uniform at a population level and most members of the

assemblage of the “community” probably do not interact in any significant

fashion. Dividing the fungal species into functional groups based on their

primary mode of nutrition (Section 4.3.2 Chapter 4) is one method used to

address this conceptual difficulty.

5) Difficulties also arise in the synthesis of mycocoenological data for

comparison between studies and even within studies. Because plots are

sampled several times per year it is necessary to calculate a synthetic measure

of species abundance. Unfortunately there is no general agreement on what

this measure should be and this has hampered comparison of results from

different mycocoenological studies. The range of synthetic measures has been

reviewed by Winterhoff (1984), Barkman (1987), Arnolds (1981) and more

recently by Zak and Willig (2004) and include measures of frequency of

occurrence and measures of average sporocarp production.

5.1.2 Macrofungal communities of forest ecosystems

A prime requirement for the formation of distinct assemblages of organisms or

communities is that populations of different species share a similar habitat, which

is defined by distinct abiotic and biotic conditions. The formation of communities

is also strongly influenced by the ecological amplitudes of individual species. In

205

the case of fungi, some have quite specific ecological requirements, while others

are more general in their requirements. For example, the ectomycorrhizal genus

Naucoria is largely confined to wet alder woods, because of its association with

Alnus (Molina et al. 1992), while the species Suillus grevillei associates

exclusively with larch. Marasmius hudsonii, which is very restricted worldwide,

grows only on fallen holly (Ilex aquifolium) leaves and is therefore restricted to

woodlands containing holly, such as Atlantic oak forests (Gill et al. 1948). In

contrast, ectomycorrhizal species such as Laccaria laccata and Russula

ochroleuca associate with a wide range of tree hosts and are consequently found

in a range of forest habitats. Such species would tend to be less useful in

discriminating macrofungal communities. The general consensus is that restriction

of a fungal species to a single habitat is rare, and it seems that in general

endemism in macrofungi is not common, at least in temperate regions (Mueller et

al. 2007).

For reasons outlined above, the macrofungal communities of forests

cannot be defined using restricted species alone. Wilkins et al. (1937) devised

guidelines to help define the “characteristic mycological flora” of any ecosystem.

The two grouping which can be applied to the macrofungi found are:

1) Relatively constant species

This group can be further sub-divided into (a) common species but are found

constantly in every community (e.g. Laccaria laccata with oak, Scot’s pine, Sitka

spruce and beech), (b) characteristic species which are found constantly in one

community and less frequently in other communities (e.g. Lactarius quietus with

oak) and (c) restricted species which are found in only one community (e.g.

Daldinia concentrica with ash).

2) Relatively inconstant species

These are the species which can appear sporadically in different ecosystem types

and thus their specific ecological requirements are difficult to determine. These

species are usually not specifically commented on, as their abundances in studies

are too low to draw any major conclusions concerning their ecology.

206

Using multivariate statistics and species abundance patterns from a

number of different forest types, recent studies (Humphrey et al. 2000; Ferris et

al. 2000a; Ishida et al. 2007; Buee et al. 2011) have shown that the fungal

communities of temperate forests can be clearly defined and delineated.

5.1.3 Past studies of macrofungal communities in temperate oak, ash, Scot’s

pine and Sitka spruce forest habitats

There have been no previous systematic studies of macrofungal communities in

Irish forest habitats. Habitats which have received intensive fungal community

analysis are the Burren area of western Ireland (Harrington and Mitchell 2005)

and Irish grasslands (Mc Hugh et al. 2001), and their fungal communities have

been well described. The closest studies examining macrofungal communities

have been carried out in the U.K.

The macrofungal communities of oak forests in Britain have been

described by several workers (Hering 1966; Watling 1974; Wilkins et al. 1937;

Humphrey et al. 2000). The list of ectomycorrhizal fungi from oak woods is often

very large. The macrofungal species Lactarius quietus, Russula cyanoxantha and

R. nigricans were found to be abundant in sessile oak forests while absent from

nearby ash (Fraxinus) forests by Hering (1966), and therefore are important

macrofungal components of these oak woods in Britain. Watling (1974) identified

several species from the ectomycorrhizal genera Boletus and Leccinum as being

somewhat restricted to oak forests in the U.K. Wilkins et al. (1937) found that

over 30% of the species found in their pedunculate oak woods were not found in

either their coniferous or beech plots, and these included species from the genus

Lactarius including L. glyciosmus, L. mitissimus, L. pyrogalus and L. vellereus.

The most recent study of macrofungal communities in British oak forests

(Humphrey et al. 2000) produced a clear separation of the of oak forests from the

other forests based on the ordination of the macrofungal species found. The

authors identified a distinctive group of Russula species (R. betularum, R.

nigricans, R. fragilis and R. cyanoxantha; C. Quine, unpub. data) as being

abundant in their oak plots but less common in their Scot’s pine, Sitka spruce and

Norway spruce plots.

207

Ash forests due to their lack of the ectomycorrhizal fungal element have

received much less study than other forest types in temperate regions. Studies

which examined the ash macrofungal communities have been conducted in Britain

(Hering 1966; Boddy et al. 1987) and Switzerland (Straatsma and Krisai-

Greilhuber 2003). The study by Boddy et al. (1987), examined the fungi which

formed on ash branches at different stages of decay. They identified species such

as Daldinia concentrica and Hypoxylon rubiginosum as being very common

macrofungal species on decaying ash wood. In a study of British deciduous

forests, Hering (1966) found that his ash plot had almost as high species richness

as his sessile oak plot and included the ash restricted species Coprinus micaceus

and Marasmius epiphyllus. The long term (seven years) study in the Swiss plot

(Straatsma and Krisai-Greilhuber 2003), identified 228 macrofungal species

present in an ash and elm forest plot. Although distinctive species were not

commented on, the species seem to follow the trend of other macrofungal

communities, in that they fruit infrequently and vary considerable from year to

year, with 41% of the species being found in only one year of the study and so

could be classified as relatively inconstant species according to the Wilkins et al.

(1937) framework above.

The macrofungal ecology of Scot’s pine forests has not been investigated

directly in Ireland, but the study by Harrington (2003) investigated the

macrofungal communities in Dryas habitats in the Burren, western Ireland. The

fungal community found in this habitat is said to have persisted in association

with Dryas roots since the ancient forests of Scot’s pine became extinct from the

area (Harrington 1996). The community is composed of profusely fruiting species

such as Craterellus lutescens and a number of Cortinarius species (including the

nationally rare C. odorifor, C. calochrous and C. infractus). Other common

woodland species are the normally forest based Mycena leptocephala and other

Mycena species, which function as wood- and litter-decay fungi in forest

ecosystems. The closest analysis of plantation Scot’s pine forests was conducted

in England (Ferris et al. 2000a). They found that the macrofungal communities of

lowland Scot’s pine forests was distinct, and separated the Scot’s pine plots from

Norway spruce forest plots based on the ordination of the macrofungal

community. Related data from the Biodiversity in Britain’s planted forests study

208

(C. Quine unpub. data), found that the macrofungal species Auriscalpium vulgare,

Hygrophorus hypothejus, Russula caerulea, Chroogomphus rutilus and

Gomphidius roseus were found frequently in the Scot’s pine forests and never in

any of the other investigated (oak, Norway spruce, Sitka spruce, Corsican pine)

forest types, possibly indicating that these macrofungal species are semi-restricted

to this forest type.

The fungal communities of Sitka spruce have been investigated in the

U.K. by Humphrey et al. (2000). They discovered that spruce plantations formed

a distinct group in ordination, according to the fungal species present and the

abundance of these species. They found that there were 130 species which were

only found in Sitka spruce habitats during the study, with three of these species

(Cortinarius bataillei, Micromphale perforans and Gloeophyllum sepiarium)

being found frequently in their Sitka spruce forests. In its home range of British

Columbia, Sitka spruce forests can be characterised as having high richness and

abundance of Mycena species including Mycena tenax, M. rosella and M. galopus

(Outerbridge 2002). Roberts et al. (2004) also noted a large abundance of Mycena

species (e.g. M. clavicularis, M. tenax and M. vulgaris) in Sitka spruce habitats in

British Columbia, and characterised the habitat as also having a high abundance

of Cortinarius (e.g. Cortinarius calochrous and C. duracinus) and Russula (e.g.

Russula sororia and R. brunneola) species.

The fungal community analysis of Irish natural and plantation forests is

necessary, as anthropogenic and other factors which disturb the natural cycle of

forests, may damage the habitats available for native fungi. Given that Ireland

does not currently have a red-data list, a comprehensive inventory of fungal

species or ecological information on the scarcity or commonness of the different

fungal species in its forests (O’Hanlon and Harrington in press), information on

fungal communities in Irish woodlands will be very informative and add to an

aging information base.

209

5.2 Aims of this chapter

The aims of this chapter are:

• To examine if the fungal community of the ash, oak, Scot’s pine and Sitka

spruce forests vary significantly. It is expected that plots of the same forest

type would be more similar than plots of a different forest type based on their

macrofungal communities (Humphrey et al. 2000). It may also be expected

that certain macrofungal species would be much more common and possibly

semi-restricted to a single forest type in this study, and thus may have

potential for use as indicator species of a certain forest type.

• To examine if relationships between certain macrofungal species and

environmental variables exist, as found in Ruhling and Tyler (1990). It is

expected that the fungal community of the sites will be affected by the

dominant tree type of the forest and also by abiotic soil variables. It may be

expected that pH will show significant correlations with the fungal

communities described (Ferris et al. 2000a).

• To examine if the macrofungal communities of the exotic conifer Sitka spruce

are composed of many species which are present in other forests types,

namely oak and Scot’s pine; as Sitka spruce has been noted as an

ectomycorrhizal generalist (Alexander and Watling 1987). Sitka spruce in

Ireland was imported as seed and therefore it has not had the opportunity of

bringing its usual macrofungal communities with it.

210

5.3 Materials and methods

5.3.1. Site list

The site list and map for the sites used can be found in Chapter 3 (Table 3.2). The

sites were visited from 2007-2009 inclusive during the autumn (August–

November), with a maximum of three visits per season attempted.

5.3.2 Macrofungal assessment

For macrofungal sampling methodology see Section 4.3.2 Chapter 4.

5.3.3 Abiotic variables

The abiotic variables recorded were pH, organic matter %, moisture content,

coarse woody debris amount and soil nutrient availability. The methods used to

record and measure these variables are given in Chapter 3, Section 3.2.2.

Climatic variables such as mean and total precipitation along with mean

temperature per month was acquired from the Irish meteorological service. The

weekly data for rainfall, temperature mean, temperature high and temperature low

for the period June 2007-december 2009 was also recorded.

5.3.4 Stand structural attributes

The number of tree and other plant species was recorded in each plot. Ground

floor vegetation was recorded including vascular plants, bryophytes and mosses.

The structure of the forest in each of its canopy layers was recorded according to

the method in Chapter 3 Section 3.2.2. The rotation stage, age and historical use

of the land were acquired from the local forester and from historical maps of

Ireland (www.osi.ie).

5.3.5 Statistical and multivariate analysis

A quantitative measure of species abundance (as distinct from sporocarp

abundance) was created by recording the presence/absence of the species in the

five sub-plots within each plot. This gives a maximum abundance of 5 per plot

visit. The sub-plot frequency of occurrence (range 0-1) of SpeciesX in Siten was

calculated as:

=Number of sub-plots containing SpeciesX in Siten/ Total number of sub-plots examined in Siten.

211

The data set used for this analysis was 213 species x 28 sites and will be referred

to as the species abundance matrix (Spreadsheet 1 in Folder 3, Appendix 1).

To assess relative macrofungal distributions within the macrofungal

communities, rank abundance plots (Whittaker 1965) were constructed following

Magurran (2004). The relative frequency of SpeciesX in forest typen was

calculated as:

=Number of appearances of SpeciesX in all sub-plots in forest typen / Number of appearances of all Species in all sub-plots in forest typen.

The fitting of the log-normal model was attempted in Microsoft Excel

version 12 (Microsoft, Redmond, WA) following Magurran (2004) after visual

inspection of the curves. A Kolmogorov-Smirnov goodness of fit test (Sokal and

Rohlf 1995) was used to test if the observed rank abundance distributions differed

significantly from the expected log-normal distributions. To analyse for

significant differences in the distributions of the rank abundance values between

the forest types, a Kolmogorov-Smirnov two sample test was used as described by

Magurran (2004). Relative cumulative frequency tables were calculated for each

forest type. Following Sokal and Rohlf (1995), the D statistic and D0.05 statistic

were calculated and compared. If D≥ D0.05 then the difference is significant at

P<0.05. The Kolmogorov-Smirnov goodness of fit test and the Kolmogorov-

Smirnov two sample test were both calculated following the methods in Magurran

(2004) using Microsoft Excel.

5.3.6 Multivariate analysis

Community similarity analysis

Similarity of the macrofungal community composition between plots was

determined by calculating the abundance-based Jaccard similarity index (Chao et

al. 2005) in Estimate-S (Colwell 2004) using the species abundance matrix. This

index is a modified form of the Jaccard index (JI), uses quantitative data and also

estimates the number of unseen taxa in the sample thus combining a species

richness estimator and a similarity index (Chao et al. 2005). Median values for the

Jaccard index (JI) were calculated by pooling the index values for individual sites

based on forest type. Median JI values and the minimum and maximum values

212

were calculated within forest types and between forest types. For example the

median value for similarity between oak and ash was calculated as the median

value of the JI index between all combinations of oak and ash plots. This was used

to highlight if certain forest types were similar based on their macrofungal

communities. The numbers of shared species were also calculated between all plot

combinations.

Ordination of macrofungal species data

Nonmetric multidimensional scaling (NMS) was used to compare composition of

the fungal communities in the forest types. NMS was run on the species

abundance matrix using PC-Ord version 4.36 (MjM Software Design, Gleneden

Beach, OR). A Sørensen distance measure was used for the data set. A random

seed started each analysis and included 40 runs of real data and 50 runs of

randomized data for use in a Monte Carlo permutation procedure (McCune and

Grace 2002). Results of a Monte Carlo test and examination of stress in a scree

plot were used to determine dimensionality. The amount of variance represented

by the ordination axes was calculated as the coefficient of determination (r2)

between the distances in the ordination space and the distances in the original

space (McCune and Grace 2002). The species abundance matrix was entered

along with the site variables into PC-ORD. Histograms and normal probability

plots showed that abiotic soil variables and the structural variables did not fit a

normal distribution and were incompatible with parametric statistics (Kent and

Coker 1992). The non-parametric Spearman’s rank correlation coefficient was

calculated for environmental variables, species richness and ordination scores in

PASW version 18 (SPSS inc., Chicago, Illinois). In ordinating the sites, NMS

ignores the environmental variables. Therefore the axes used in the ordination of

the sites were created using only the macrofungal species abundance data with

environmental data analysed then checked against these site ordinations to

identify significant correlations.

Multi-response permutation procedure

The homogeneity and distinctiveness of the macrofungal community of each

forest type was assessed using a multi-response permutation procedure (MRPP;

McCune and Grace 2002) in PC-Ord. MRPP is a non-parametric approach for

testing the hypothesis of no differences between two or more groups of entities;

213

i.e. in this case differences in fungal community composition between the

different forest types. The working hypothesis was that macrofungal species were

not randomly distributed between the different forest plots, but rather formed

distinct assemblages according to the dominant tree type of the forest plot. MRPP

constructs a distance matrix, calculates average within-group distances, and

compares these to a Pearson’s type III continuous distribution of all possible

partitions of the data (Peck 2003). With MRPP, the chance- corrected within-

group agreement (A) statistic describes effect size. A ranges from -1 to +1. When

A = -1, there is less agreement between groups than expected by chance. When A

=0, groups are no more or less different than expected by chance, and when A = 1,

groups are identical. The more positive A is, the more homogeneous groups are

and the greater confidence in the p-value, especially when the sample size is small

(Peck 2003). Groups were analyzed by forest type using PC-Ord. The Sørenson

distance measure was used for all analyses.

Indicator species analysis

Indicator species analysis was used to identify macrofungal species which are

indicative of a certain forest type. The method used was designed by Dufrene and

Legendre (1997). Indicator species analysis uses data on the concentration of

species abundances in a particular user set group (e.g. forest type) and also the

faithfulness of occurrence of a species to a particular group (McCune and Grace

2002). To identify macrofungal species that are indicative of the habitats

examined in this project, Indicator values (IV) for each species were calculated in

PC-ORD using the method by Dufrene and Legendre (1997) with statistical

significance calculated by a Monte Carlo test with 1000 runs. Indicator values

range from 0 (no indication) to 100 (perfect indication). Perfect indication means

that presence of a certain species points to a particular group without error

(McCune and Grace 2002).

5.3.7 Year to year variation

To test if the fungal communities were similar over the different years, the sub-

plot frequency of occurrence data from the sites which were sampled in the three

years of the project (Killough, St John’s Wood, Kilmacrea, Raheen, Bohatch,

Chevy chase M, Chevy chase Y, Quitrent and Stanahely) was analysed using a

214

Mantel test. The species x site matrix consisted of 155 species from 9 sites.

Similarity matrixes were created using the Sørenson distance index and were

compared using PC-Ord and evaluated for significant using a Monte Carlo test

with 9999 randomizations. If significantly different, the standardized Mantel

statistic (r; Sokal and Rohlf 1995) was used as a measure of effect size (McCune

and Grace 2002). The standardized Mantel statistic ranges from -1 to +1 and is a

measure of the correlation between the matrixes. A value of +1 indicates that the

two matrixes (communities) are very similar while a negative r statistic indicates

that the matrixes (communities) are very dissimilar.

215

5.4 Results

5.4.1 Species abundance patterns

In total 219 fungal species from 98 genera were found within the permanent plots.

The sub-plot frequency of occurrence of these species was calculated using the

species abundance matrix. Throughout the project a total of 2622 fruitbodies were

recorded, comprising 591, 891 and 1140 sporocarps in 2007, 2008 and 2009

respectively. Mycena leptocephala and Russula ochroleuca were the most prolific

producers of fruitbodies (Table 5.1). The majority of the most common

fruitbodies are of small delicate (typically <4cm height) species e.g. Mycena spp.

Many of the most commonly encountered species were also the most abundant

fruitbody producers. Fourteen of the most prolific sporocarp producers (Table 5.1)

were also amongst the most commonly found fungi (Table 4.6, Chapter 4). The

three species that produced large numbers of fruitbodies but were not common

across all forest types show varying levels of restriction to certain forest types i.e.

Hemimycena gracilis with Sitka spruce, Scleroderma citrinum with oak, and

Crepidotus mollis with oak and ash.

Table 5.1 The most prolific frutibody producers in each forest type. Column 6 (Total) is the total number of fruitbodies from that species recorded over all forest types. SP= Scot’s pine, SS= Sitka spruce. Species Ash Oak SP SS Total

Mycena leptocephala 2 6 200 263 471

Russula ochroleuca 0 33 115 251 399

Laccaria laccata 3 123 58 101 285

Hemimycena gracilis 0 0 0 279 279

Hypholoma fasciculare 15 30 34 186 265

Laccaria amethystina 1 134 48 53 236

Mycena vitilis 8 9 63 77 157

Xylaria hypoxylon 94 34 0 28 156

Mycena galopus 37 30 13 59 139

Collybia butyracea 0 1 8 127 136

Cortinarius flexipes 0 58 41 35 134

Marasmius hudsonii 6 58 36 4 104

Scleroderma citrinum 0 93 0 0 93

Lactarius hepaticus 0 6 78 1 85

Mycena rorida 19 0 6 38 63

Mycena polygramma 21 5 1 34 61

Crepidotus mollis 19 14 0 0 33

216

Rank abundance plots of macrofungi in different forest types

To analyze species abundance patterns, the relative frequency of macrofungal

species over a number of plot visits was used to construct rank abundance plots.

In order to test if the distributions followed the log-normal distribution,

Kolmogorov-Smirnov goodness-of-fit tests tests were carried out to discover if the

expected and observed frequencies differed significantly (Fig. 5.1). All plots were

found to follow the log-normal distribution. The rank abundance plots were

constructed for the macrofungi in each forest type based on their relative

frequency of occurrence in the sub-plots (Fig. 5.2). Sporocarp numbers were not

used because it was felt that a fungus which produces small sporocarps can

produce more sporocarps than a fungus with large sporocarps using the same

amount of resources; therefore, comparisons of species abundance patterns based

on number of sporocarps would be distorted in favour species with smaller

sporocarps.

All of the rank abundance curves follow the typical form for fungal

communities with a small number of common species (based on frequency of

occurrence in sub-plots) and a longer list of less common to rare species (Figs.

5.2). The most common species in the ash forest type were very different to the

most common species in the other forest types. Ash shared only one most

common species (Mycena galopus) with oak, and none with the Scot’s pine or

Sitka spruce forest types. Russula ochroleuca and Laccaria laccata were among

the five most common species of the oak, Scot’s pine and Sitka spruce forest

types. Scot’s pine and Sitka spruce forests also shared Mycena leptocephala in

their most common species lists.

To analyse for significant differences in the rank abundance curves

between the forest types, a Kolmogorov-Smirnov two-sample test was carried out

on the relative cumulative frequency for each forest type; with significant

differences (P<0.05) identified if D≥ D0.05. The results indicate that the rank

abundance distribution were similar between oak and Scot’s pine, oak and Sitka

spruce, and ash and Scot’s pine, but were different between ash and oak, ash and

Sitka and Scot’s pine and Sitka spruce (Table 5.2). In practical terms, what this

analysis signifies is that forest types with similar curves have similar relative

abundances of their most common species.

217

Table 5.2 Comparison of rank abundance distribution patterns of macrofungi in different forest types using the two-sample Kolmogorov-Smirnov test. Columns indicate which forest types were compared. n1= number of sub-plot samples in forest type 1, n2= number of sub-plot samples in forest type 2. D= largest unsigned difference between the two relative cumulative frequency distributions from the two forest types. D0.05= Critical significance value for two-sample Kolmogorov- Smirnov test calculated for that number of samples. SP= Scot’s pine, SS= Sitka spruce.

ASH-OAK ASH-SP ASH-SS OAK-SP OAK-SS SP-SS

n1 136 136 136 405 405 402

n2 405 402 726 402 726 726

D 0.161 0.0978 0.189 1.1x10-15 0.054 0.097

D0.05 0.135 0.135 0.127 0.096 0.084 0.085

Significantly different Yes No Yes No No Yes

Ash

1 2 3 4 5 6 7 8 9

10-1

515

+

0

10

20

30

Rela

tive f

req

uen

cy

Oak

1 2 3 4 5 6 7 8 9

10-1

515

+

0

5

10

15

20

25

30

35

40

45

50

55

SP

1 2 3 4 5 6 7 8 9

10-1

515

+

0

5

10

15

20

25

30

35

Rank group

Rela

tive f

req

uen

cy

SS

1 2 3 4 5 6 7 8 9

10-1

515

+

0

510

1520

2530

3540

45

50

55ObservedExpected

Rank group

Fig 5.1 Observed and expected distributions of fungal species in each of the forest types. The expected distribution is based on the fungal data fitted to the log-normal distribution as detailed in Magurran (2004). All of the forest types were found to be similar to the expected log-normal distribution curve. SP= Scot’s pine, SS= Sitka spruce, grey bar = observed distributions, black bar= expected distributions.

218

ASH

0 10 20 30 40 50 600.001

0.01

0.1Xylaria hypoxylon

Mycena roridaMycena polygramma

Mycena galopus

Marasmius hudsonii

Rank

Rela

tive f

req

uen

cy

OAK

0 25 50 75 100 1250.001

0.01

0.1

Scleroderma citrinum

Laccaria laccataStereum hirsutum

Mycena galopus

Laccaria amethystina

Rank

Rela

tive f

req

uen

cy

SP

0 10 20 30 40 50 60 70 80 90 1000.001

0.01

0.1Russula ochroleuca

Laccaria laccataLactarius hepaticus

Mycena leptocephala

Trichaptum abietinum

Rank

Rela

tive f

req

uen

cy

SS

0 25 50 75 100 125 150 1750.001

0.01

0.1Russula ochroleuca

Mycena leptocephalaLaccaria laccata

Hypholoma fasciculare

Collybia butyracea

Rank

Rela

tive f

req

uen

cy

Fig 5.2 Rank abundance plot of macrofungal species in ash, oak, Scot’s pine (SP) and Sitka spruce (SS) forests. Also shown are the names of the top five most frequently encountered species.

219

The species were assigned to functional groups and were plotted on rank

abundance graphs for each forest type (Fig 5.3 a, b). A two sample Kolmogorov-

Smirnov test was used to identify significant differences in the distributions of the

different fungal functional groups between the forest types (Table 5.3, 5.4). The

most common ectomycorrhizal species in the oak, Scot’s pine and Sitka spruce

forest included Laccaria laccata and L. amethystina in the top five most common

species in all tree forest types (Fig. 5.3 a). Oak differed by having a much greater

proportion of species that are semi-restricted to oak forests (Scleroderma citrinum

and Lactarius quietus) than either of the coniferous forest types. Scot’s pine had

two Lactarius species (L. hepaticus and L. tabidus) in its five most common

ectomycorrhizal fungi. In Sitka spruce forests, Russula ochroleuca and

Cortinarius obtusus were in the top five species, with C. flexipes listing as the 6th

most common ectomycorrhizal species in spruce forests.

The rankings of the most common litter- and wood-decay species showed

much less similarity between forest types (Fig. 5.3 b). Ash forests had no species

in common with the lists from the coniferous forests types and only one in

common with the oak forest types (Mycena galopus). Oak forests had no species

in common with the lists from the coniferous forests. Mycena leptocephala is the

only species common to both of the lists from the coniferous forests. For litter-

and wood-decay fungi, significant differences (P<0.05) were found between the

distributions of the in Scot’s pine versus Sitka spruce forests (Table 5.3). This was

due to the relatively lower records for wood-decay species in Scot’s pine forests

than in Sitka spruce forests. Mycena leptocephala was the most common decay

fungi in both coniferous forest types, but the Scot’s pine forests had fewer overall

litter and wood decay species with 38 versus 69 in spruce forests. Genera such as

Mycena and Lycoperdon were much more species rich in Sitka spruce forests than

in Scot’s pine forests with 20 vs. 14 species and 4 vs. 0 species in the respective

forest types.

220

221

0 10 20 30 40 50 60 700.003

0.03

0.3

SPSS

OAK

Laccaria amethystina

Laccaria laccata

Scleroderma citrinum

Cortinarius flexipes

Lactarius quietus

Russula ochroleuca

Lactarius hepaticus

Laccaria laccata

Laccaria amethystina

Lactarius tabidus

Russula ochroleuca

Laccaria laccata

Laccaria amethystina

Cortinarius obtusus

Russula emetica

Rank

Rela

tive f

req

uen

cy

0 10 20 30 40 50 60 70 80 900.001

0.01

0.1

ASHOAKSPSS

Xylaria hypoxylon

Mycena galopus

Mycena polygramma

Mycena rorida

Marasmius hudsoni i

Stereum hirsutum

Mycena galopus

Crepidotus variabil is

Armillaria mel lea

Lycoperdon perlatum

Mycena leptocephala

Trichaptum abietinum

Mycena epipterygia

Rickenel la fibula

Tricholomopsis ruti lans

Mycena leptocephala

Hypholoma fasciculare

Col lybia butyracea

Mycena vi ti l is

Mycena metata

Rank

Rela

tive f

req

uen

cy

a.

b.

Fig. 5.3 Rank abundance plots of (a) ectomycorrhizal species in oak, Scot’s pine (SP), and Sitka spruce (SS) plots and (b) litter and wood decay species in ash, oak, Scot’s pine and Sitka spruce plots. Also shown are the five most common species for each forest type. Blue text and symbols refer to ash forests, green refers to oak forests, orange refers to Scot’s pine forests and brown refers to Sitka spruce forest type.

222

223

Table 5.3 Comparison of rank abundance distribution patterns of litter- and wood-decay macrofungi in different forest types using the two-sample Kolmogorov-Smirnov test. Columns indicate which forest types were compared. n1= number of sub-plot samples in forest type 1, n2= number of sub-plot samples in forest type 2. D= largest unsigned difference between the two relative cumulative frequency distributions from the two forest types. D0.05= Critical significance value for two-sample Kolmogorov- Smirnov test calculated for that number of samples. Sp= Scot’s pine, SS= Sitka spruce.

Litter- and wood-decay species

ASH-OAK ASH-

SP

ASH-

SS

OAK-SP OAK-SS SP-SS

n1 131 131 131 234 234 187

n2 234 187 446 187 446 446

D 0.111 0.008 0.096 0.004 0.016 0.141

D0.05 0.148 0.155 0.135 0.133 0.11 0.118

Significantly different No No No No No Yes

For ectomycorrhizal macrofungi (Table 5.4) significant differences

(P<0.05) were found in the distributions of oak and Scot’s pine; and oak and Sitka

spruce. In Scot’s pine and Sitka spruce the ectomycorrhizal fungus Russula

ochroleuca was the most abundant species and made up a large number of the

species records in these forests. In oak forests, Laccaria amethystina and L.

laccata were the two most commonly recorded fungi and were almost equal in

relative frequency of occurrence in sub-plots.

Table 5.4 Comparison of rank abundance distribution patterns of ectomycorrhizal macrofungi in different forest types using the two-sample Kolmogorov-Smirnov test. Columns indicate which forest types were compared. n1= number of sub-plot samples in forest type 1, n2= number of sub-plot samples in forest type 2. D= largest unsigned difference between the two relative cumulative frequency distributions from the two forest types. D0.05= Critical significance value for two-sample Kolmogorov- Smirnov test calculated for that number of samples. Sp= Scot’s pine, SS= Sitka spruce.

Ectomycorrhizal species

OAK-SP OAK-SS SP-SS

n1 161 161 202

n2 202 264 264

D 0.147 0.144 0.088

D0.05 0.144 0.136 0.127

Significantly different Yes Yes No

224

What the results of two-sample Kolmogorov-Smirnov tests indicate, is that

forest types that have similar rank abundance plots have similar relative

frequencies of their most common species. For example, it can be seen that the

most common decay macrofungi in oak forests have a higher relative abundance

than the most common decay species in either Scot’s pine or Sitka spruce (Fig

5.3b).

5.4.2 Analysis of macrofungal community similarity between the sites

The number of shared species and abundance-based Jaccard similarity index (JI)

values were calculated in PC-ORD for all site combinations (Table 5.6) using the

species abundance matrix. The sites were coded for ease of reading (Table 5.5).

Median (and minimum/maximum) Jacard similarity values were calculated for

each forest combination (Table 5.7).

Table 5.5 Coding used for plots listed in the similarity analysis Table 5.6. Ash forest type= plots 1-5 and are colour coded blue, oak forest type= plots 6-11 and are colour coded green, Scot’s pine forest type= plots 12-19 and are colour coded orange and Sitka spruce forest type= plots 20-28 and are colour coded brown.

1 Ballykilcavan 15 Bansha

2 Donadea 16 Brittas

3 Killough 17 Derryhogan

4 Ross island 18 Gortnagowna

5 St John’s wood 19 Torc

6 Abbeyleix 20 Bohatch

7 Kilmacrea 21 Moneyteige

8 Raheen 22 Quitrent

9 TomiesA 23 Ballygawley

10 TomiesB 24 Chevy ChaseM

11 Union 25 Cloonagh

12 Annagh 26 Dooary

13 Ballygawley 27 Stanahely

14 Ballylug 28 Chevy ChaseY

225

Table 5.6 Matrix of number of shared species (clear cells), and abundance-based Jaccard similarity values (grey cells) between the plots. Coloured cells in first column and first row correspond to forest type. Blue cells= ash, green cells= oak, orange= Scot’s pine, brown= Sitka spruce. Site codes are found in Table 5.5. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

1 X

.08

.15

.14

.21

.13

.11

.07

.09

.13

.15 0

.02 0 0

.03 0 0 0 0

.08 0

.01

.07 0 0

.04

.01

2 2 X .1 0

.22

.05

.11

.15

.03

.16

.12

.03

.13

.17

.15

.05

.09

.16

.02

.03

.05

.24

.17

.06

.09

.21

.18

.13

3 2 2 X

.21

.19

.04

.09

.15 .1

.19

.08 0

.06

.07

.09

.04 0

.03 0

.07

.02

.05

.06

.12

.06

.09 .1 .1

4 2 0 2 X

.12 .1

.08

.08

.11

.13

.11 0

.03 0

.05 0 0 0 0 0 0 0 0

.05 0 0

.02

.01

5 6 4 5 2 X

.25

.23

.21

.31

.23

.28

.02

.07

.01

.18

.17

.05

.09 0 .1

.09 .1

.13

.26

.01

.17

.19

.16

6 6 3 3 3 11 X

.35

.18 .3

.21 .3

.06

.18

.06

.26

.19

.12

.14

.01

.17

.07

.18

.21

.19

.05

.16

.16

.18

7 5 4 6 3 12 13 X

.28

.32

.25 .3

.11

.28

.05

.29

.21

.15

.14

.02

.24

.09

.21

.26

.14

.01

.12 .2

.16

8 4 5 7 3 15 12 19 X

.29

.18 .3

.04

.11

.01

.15

.18

.07

.07

.06

.12

.11

.16

.18

.13

.09

.16

.21

.25

9 3 1 5 4 12 9 12 13 X .6

.29

.16

.16

.07

.31

.22

.11

.19

.02

.25

.07

.18

.32

.15

.03

.14

.08

.36

10 3 3 4 3 6 8 10 8 13 X

.14

.14

.13

.08

.19

.15 .1

.18

.02

.15

.07

.18

.28

.12

.04

.15

.14

.23

11 5 2 4 3 7 7 10 11 5 4 X 0

.08 0 .2

.16

.07

.07 0 .1

.06 .1

.15

.28

.03 .1

.14

.11

12 0 1 0 0 2 4 6 3 6 5 0 X

.27

.06

.06

.07

.25 .5

.05

.16

.01

.15 .2

.01

.03

.04

.08

.12

13 1 3 2 1 3 5 11 7 5 4 3 6 X

.27

.31

.07

.24

.45

.05

.28

.02

.35

.27

.03

.07

.16 .2

.11

14 0 2 2 0 1 4 2 1 2 2 0 2 4 X

.18

.08

.09

.19

.02

.22 0

.27

.21

.05 .1

.11

.15

.12

15 0 3 3 1 6 7 12 7 6 6 3 3 8 5 X

.18

.12

.16

.03 .3

.08

.42

.35 .1

.12

.23

.31

.19

16 1 2 2 0 8 7 9 14 6 5 4 3 3 2 5 X

.07

.15

.09

.14

.09

.17

.15

.25

.05

.15

.21

.13

17 0 2 0 0 2 6 6 5 4 3 2 6 7 2 4 4 X

.31

.06

.16 0

.25

.23

.09 .1

.12

.14

.11

18 0 3 2 0 4 5 6 7 7 5 2 10 11 4 5 6 8 X

.08

.24 0

.26

.33

.03

.09

.13

.16

.16

19 0 1 0 0 0 1 2 6 2 1 0 2 2 1 2 3 2 3 X

.08 0

.06 .1

.01

.12

.03

.08 .1

20 0 2 4 0 7 7 14 12 11 6 2 7 10 6 9 7 6 8 5 X

.09

.31

.27 .1

.03

.21

.37

.21

226

227

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 21

3 1 1 0 5 5 5 8 3 2 2 1 1 0 3 2 0 0 0 5 X .2 .1 .11

.07

.14

.25 .1

22 0 4 2 0 6 6 10 11 5 4 5 4 8 4 9 7 6 6 2 11 9 X

.34

.15

.19

.28 .4 .2

23 1 3 3 0 7 6 10 11 9 7 4 5 6 2 8 6 6 8 3 10 6 6 X

.12

.12

.13

.27 .3

24 2 2 4 1 11 7 7 10 5 4 7 1 2 3 5 8 4 2 1 6 5 8 7 X

.07 .1

.17 .1

25 0 1 1 0 1 3 1 4 1 1 1 1 1 1 2 2 3 1 2 2 2 4 3 4 X

.16

.14

.08

26 0 4 4 0 9 7 9 12 5 5 4 2 7 3 9 8 6 6 2 13 6 11 7 6 3 X

.25

.26

27 3 5 6 2 12 11 15 16 6 6 7 3 8 5 12 9 6 7 3 15 8 14 10 10 3 12 X

.17

28 1 3 6 1 9 8 8 15 11 8 5 5 4 3 7 6 4 7 5 9 5 8 10 6 3 13 11 X

228

229

Of the plots that had a high degree of species similarity (JI ≥0.3), there

were a total of 22 instances. One ash-oak combination, 5 oak-oak, 1 SP-oak, 2

oak-SS, 4 SP-SP, 5 SP-SS and 4 SS-SS pairs of plots had JI values above 0.3.

There was a high variability within some forest type pairs with regard to the JI,

for example, some of ash plots show very low (0) JI values indicating a high

variability in species composition between the plots. The most similar plots based

on JI were the two oak plots at Tomies (TomiesA and TomiesB; JI=0.6). These

plots were also the two plots which were least geographically separated forest

plots examined in the project. One would expect that since the plots were

replicated, similarity values would be higher between plots of the same forest type

than different forest types. This was the case for each of the forest types except

Scot’s pine; the Scot’s pine plots showed a higher median similarity level

(JI=0.12) to the oak plots as they did to each other (JI= 0.1; Table 5.7). This was

likely due to the large inter-plot variation between the pine plots. Omitting the two

plots (Bansha and Brittas), shown to be outliers from the Scot’s pine grouping

produced by ordination (Fig. 5.6), raised the median JI to 0.186, much higher than

the value for Scot’s pine – oak (Table 5.7).

Table 5.7 Medians and minimum/maximum values (in parenthesis) for the Jaccard similarity index values between the forest types based on inter-plot values (shaded cells). Number of shared species between the different forest types are also given (clear cells).

Tree type Ash Oak Scot’s pine Sitka spruce

Ash 0.15 (0/0.22) 0.12 (0.03/0.31)

0.03 (0/0.18) 0.06 (0/0.26)

Oak 41 0.29 (0.14/0.6) 0.12 (0/0.31) 0.12 (0.01/0.36)

Scot’s pine 21 44 0.1 (0.02/0.5) 0.12 (0/0.42)

Sitka spruce 30 65 58 0.17 (0.03/0.4)

The Sitka spruce plots were divided into two age groups; young (4 sites)

and mature (5 plots). The inter-plot JI values were also compared for these plots

(Table 5.8). Young plots were more similar to the mature plots than they were to

other young plots (median JI: 0.17 vs. 0.18; Table.5.8). Likewise, the mature plots

were more similar to the young plots than they were to other mature plots (median

230

JI: 0.18 vs. 0.14; Table.5.8). Although there were many species shared between

the two Sitka spruce age groups in this study (45 species), there were also species

that were only found in one of the age groups, 40 species in the young aged group

and 48 species in the mature age group. In the young age group, Collybia

butyracea, Mycena pura and Lactarius deterrimus were the most common. In the

mature age group, Hypholoma capnoides, Heterobasidium annosum, Cortinarius

rubellus and Hemimycena gracilis were the most commonly found macrofungi.

Table 5.8 Medians and minimum/maximum values (in parenthesis) for the Jaccard similarity index for the two age groups of Sitka spruce forests. Column 3 gives the median number of shared species between the two age groups of Sitka spruce based on inter-plot values.

Sitka spruce age group Median JI Median shared species

Young vs. young 0.17 (0.08/0.26) 7

Mature vs. mature 0.14 (0.09/0.34) 6.5

Young vs. mature 0.18 (0.03/0.4) 7.5

5.4.3 Ordination of the forest plots.

NMS ordination was carried out using the species abundance matrix. NMS found

a three-dimensional solution (Figures 5.4, 5.5, 5.6) with a final stress of 14.38,

which indicated an intermediate solution according to Kruskal’s rule of thumb

(McCune and Grace 2002). A Monte Carlo test showed that the probability of a

similar final stress being obtained by chance was 0.001. The amount of variance

represented by the 3 axes was calculated as the coefficient of determination (r2)

between the distances in the ordination space and the distances in the original

space. Axis 1 represented 26.4% of the variance with a further 22.4% of the

variation represented by axis 2. Axis 3 represented the remaining 28.2% of the

variation to give the ordination a cumulative r2 value of 77%. Axis 1 was not

correlated with any of the environmental variables (Table 5.9). Axis 2 showed

significant correlations with species richness, soil available calcium and coarse

woody debris volume. Axis 3 was correlated with pH, % soil organic matter,

mean annual precipitation, soil total available nitrogen, soil available nitrate

nitrogen, soil available magnesium and phosphorus (Table 5.9).

231

Table 5.9 Spearman’s rank correlations (rs) for ordination axes, environmental variables and species richness. %OM= Organic matter, precipitation= mean annual precipitation, N total = total soil available nitrogen, NO3= total soil available Nitrate Nitrogen, CWD total= total volume of Coarse woody debris in m2/ha, ns= non-significant correlation. Significance levels *= P<0.05, **= P<0.01.

Variable Axis 1 Axis 2 Axis 3

Species richness ns -0.461* ns

pH ns ns -0.432*

%OM ns ns 0.431*

Precipitation ns ns 0.424*

N total ns ns -0.396*

NO3 ns ns -0.394*

Calcium ns 0.594** ns

Magnesium ns ns 0.54*

Phosphorus ns ns -0.459*

CWD total ns -0.621** ns

232

233

Fig 5.4 Biplot of NMS ordination of macrofungal communities in different forest types showing axes 2 and 3 (cumulative r2= 0.506) of the ordination showing the sites identified by the tree type of the site and a biplot of the environmental and physical variables that showed >0.2 correlation coefficients with the axes. SP = Scot’s pine, SSa= Sitka spruce young age group, SSb= Sitka spruce mature age group. Ca= available soil calcium, Mg= available soil magnesium and CWD-tot= total volume of coarse woody debris in the site (m2/ha).

234

235

Fig 5.5 Biplot of NMS ordination of axis 1 and 2 (cumulative r2= 0.488) showing the sites identified by forest type on a biplot of environmental and physical variables that were correlated (r>0.2) with the ordination axes. SP = Scot’s pine, SSa= Sitka spruce young age group, SSb= Sitka spruce mature age group. CWD-total= total volume of coarse woody debris in the site (m2/ha), Ca= soil available calcium.

236

237

The ordination of axis 2 vs. axis 3 (Fig 5.4) is the most useful in separating

the ash plots from the other forest types as the axes used are correlated to soil

available calcium (axis 2= positive correlation) and pH (axis 3= negative

correlation). The Scot’s pine plots are also well delineated from the other forest

types based on their fungal communities. The two Scot’s pine plots (Britt= Brittas

and Bnsha= Bansha) are grouped closer to the oak plots than to the other Scot’s

pine plots. These two plots have oak trees in their understory and were also shown

to group closely with the oak plots based on their vegetation communities (Fig.

3.13; 3.16, Chapter 3). It is also noticeable that the Sitka spruce plots are

intimately mixed with the oak and Scot’s pine plots indicating a high level of

macrofungal community similarity between these forest types.

The NMS biplot using axes 1 and 2 (Fig. 5.5) is useful in producing close

groupings of the oak and the ash plots. The correlation of axis 2 with calcium and

coarse woody debris separates the ash plots from the other forest types as the ash

plots are characterised by high soil calcium levels and low coarse woody debris

quantity. It can be seen that the ash plot at St John’s Wood (STJON) is grouped

closely with the oak sites.

The ordination using axes 1 and 3 (fig. 5.6) produced noticeable groupings

of all of the plots into their respective forest types based on their macrofungal

communities. The ash plots (except Donadea= DONAD) all group in the bottom

left of the diagram along an increasing pH and soil available calcium gradient.

The Scot’s pine plots form a group at the top of the ordination along a gradient of

decreasing pH and soil calcium and increasing soil magnesium. As in Fig. 5.4, the

Scot’s pine plots at Bansha (BNSHA) and Brittas (BRITT) are grouped close to

the oak plots indicating a high level of community similarity between these plots

and the oak plots. The three Sitka spruce plots at Moneyteige (MONEY),

Cloonagh (CLOON) and Chevy Chase mature (CHEVM) are shown to be outliers

from the main Sitka spruce grouping.

5.4.4 Indicator species analysis

To identify fungal species which may be associated with particular forest types,

Indicator Values (IV) sensu Dufrene and Legendre (1997) were calculated in Pc-

238

Ord for each species were calculated with statistical significance identified by a

Monte Carlo test. The species abundance matrix was used in this analysis.

Seventeen species were found to have IV values that were significant

(Table 5.10). In the case of ash, the species are Postia subcaesia and Xylaria

hypoxylon, both of which were found primarily in ash forests in this study. The

macrofungal community of oak forest was marked by the presence of Lycoperdon

perlatum, Crepidotus variabilis, Lactarius quietus, Stereum hirsutum and

Laccaria amethystina. The macrofungal community of Scot’s pine sites were

distinguished by the presence of Trichaptum abietinum and Lactarius hepaticus.

Young Sitka spruce habitats were marked by the presence of Cystoderma

amianthinum, Hemimycena gracilis, Postia caesia and Russula emetica. The

mature Sitka spruce sites differed in that they had the species Collybia butyracea,

Mycena metata and Mycena pura as indicator species.

Table 5.10 The fungal species which were found to be indicator species of a particular forest type. SP= Scot’s pine, SSa= Sitka spruce young age grouping, SSb= Sitka spruce mature age grouping.

Indicator values

Species Ash Oak SP SSa SSb Significance

Collybia butyracea 0 0 0 0 71 0.005

Crepidotus variabilis 14 44 0 0 3 0.05

Cystoderma amianthinum 0 0 0 42 8 0.05

Hemimycena gracilis 0 0 0 60 0 0.01

Hypholoma capnoides 0 0 0 60 0 0.01

Laccaria amethystina 0 60 4 6 5 0.01

Lactarius hepaticus 0 1 47 1 0 0.05

Lactarius quietus 0 44 4 7 0 0.05

Lycoperdon perlatum 3 44 0 0 0 0.05

Mycena metata 0 6 0 16 54 0.05

Mycena pura 0 1 0 0 47 0.05

Postia caesia 0 0 1 52 6 0.05

Postia subcaesia 60 0 0 0 0 0.01

Russula emetica 0 4 0 40 5 0.05

Stereum hirsutum 4 82 1 0 0 0.001

Trichaptum abietinum 0 0 56 2 0 0.01

Xylaria hypoxylon 51 18 0 2 1 0.05

239

The indicator species analysis also takes into account the abundance of a

species in a habitat. For example, Mycena metata shows indicator values of 6, 16

and 54 for oak, young Sitka spruce and mature Sitka spruce habitats respectively.

Although this species was found in all three habitats, it is very linked to the

mature Sitka spruce habitat due to it having a greater abundance in the sub-plots

of mature Sitka spruce habitats. The species found to be indicator species were

plotted on the ordination of axis 1 and 3 (Fig. 5.6).

Fig. 5.6 Biplot of NMS ordination of axis 1 and 3 (cumulative r2= 0.546) showing the sites identified by forest type, and a biplot of environmental and physical variables that were correlated (r>0.2) with the ordination axes. The indicator species of the different forest types are also marked on the ordination (species name and black dot marker). Groupings of sites based on forest type are indicated by the coloured lines enclosing sites. SP = Scot’s pine, SS= Sitka spruce, Ca= available soil calcium, Mg= available soil magnesium.

240

241

The NMS ordination based on macrofungal communities differentiated the

plots into groups that corresponded approximately to the original classification of

the plots according to dominant tree type (Fig. 5.6). The NMS ordinations

separated the plots into the following groups based macrofungal community

assemblages:

1. The ash plots formed a fairly distinct group in the ordination (Fig. 5.6). The

macrofungal community in most of the ash plots is species-poor and lacking in

ectomycorrhizal species. The ash plots were ordinated along axis 2 which was

correlated with decreasing species richness (Table 5.9). The ash plots had the

highest levels of soil calcium, soil pH and low levels of coarse woody debris.

Macrofungal species such Postia subcaesia and the ascomycete Xylaria

hypoxylon were indicator species of this forest type. Other wood decay species

such as Crepidotus variabilis, Marasmius androsaceus, M. hudsonii and Stereum

hirsutum were also abundant species in some of the ash forests. The almost total

lack of ectomycorrhizal species in the ash forests also separated the ash sites from

the other forest types. The four ectomycorrhizal species (two Laccaria species and

two Lactarius species) found in the St John’s wood plot are almost certainly due

the presence in the plot of roots of nearby hazel trees. Other species which were

restricted to the ash sites but were found outside of the plots were the wood decay

ascomycete Daldinia concentrica and the litter decay species Camarophyllopsis

atropuncta. These two species were only found in the un-managed or semi-natural

ash plots at St John’s Wood and Killough.

2. The oak plots formed a distinct group in the centre of the ordination (Fig. 5.6),

indicating that the oak plots possessed a macrofungal community that was distinct

from that of the other forest types. Ectomycorrhizal species such as Laccaria

amethystina and Lactarius quietus were indicator species of the oak forest type.

The wood decay species Crepidotus variabilis and Stereum hirsutum and the litter

decay species Lycoperdon perlatum were also strong indicators of the oak forest

type. Other species found commonly in the oak sites were from the

ectomycorrhizal genera Amanita, Cortinarius, Inocybe, Lactarius and Russula

with 3, 7, 3, 8 and 7 species respectively in the oak plots. The earthball

242

Scleroderma citrinum was also abundant in all of the oak sites except Union,

which in general was the most species-poor oak plot and this is reflected in its

ordination at a slight distance from the rest of the oak plots. Marasmius hudsonii

was a common litter decay fungus found in the oak sites. This species only forms

on holly leaves, and holly was a common understory element in four of the oak

plots (Table 3.10, Chapter 3). Russula ochroleuca was very common in the oak

plots, along with Laccaria laccata and these species often fruited in great

abundance. The only hypogeous macrofungus found in this project, Elaphomyces

granulatus was found in four of the six oak plots. Hypogeous fungi were not

especially investigated in this project and were only encountered due to the

digging involved in ectomycorrhizal root sampling. That these truffle-type

macrofungi were encountered in 4 of the 6 plots suggests that they (and possibly

other hypogeous fungi) may be more common in Irish oak woods than previously

thought.

3. The Scot’s pine plots formed a discrete grouping in the ordination (Fig. 5.6).

Two of the Scot’s pine plots (Bansha: BNSHA and Brittas: BRITT) were grouped

closer to the oak sites probably because of the presence of the ectomycorrhizal

species Laccaria amethystina and Lactarius quietus in these plots. The finding of

these macrofungi in these plots is most likely due to the presence of sapling oak

trees in the understory of both of these plots. Most of the Scot’s pine plots

positioned at the extreme of Axis 3; which was negatively correlated with soil pH

and available calcium. The macrofungal species Lactarius hepaticus and

Trichaptum abietinum were identified as indicator species of the Scot’s pine forest

type (Table 5.10). In general the Scot’s pine plots were characterised by having

high numbers of the ectomycorrhizal species Cortinarius, Lactarius and Russula

with 9, 6 and 8 species respectively. Distinctive species such as Cortinarius

sanguineus, Lactarius rufus and the copper spike Chroogomphus rutulis were

only found in Scot’s pine plots. The first and last of these species were found in

pine plots at Brittas and Bansha, which had the most diverse vegetation of all the

Scot’s pine plots (Table 3.11, Chapter 3). Lactarius rufus was found in abundance

in the Scot’s pine plantations on boggy soil at Annagh, Derryhogan and

Gortnagowna.

243

4. The ordination of the macrofungal community in the Sitka spruce plots formed

a loose cluster in the ordination which indicated that this forest type has a fungal

community that differs in some ways from the other forest types examined (Fig.

5.6). It also indicates that macrofungal communities of Sitka spruce plots share

some similarities with those of the oak and Scot’s pine plots. The species-poor

Sitka spruce plots at Cloonagh (CLOON), Moneyteige (MONEY) and Chevy

Chase Mature (CHEVM) did not align with the other Sitka spruce plots, and can

be seen as outliers from their respective young and mature Sitka spruce site

groupings. These sites had the lowest levels of species richness of all the Sitka

spruce sites with a total of 9, 19 and 25 macrofungal species respectively. Another

factor these three sites had in common was a lack of species that were common in

other Sitka spruce plots. The species Russula ochroleuca, Cortinarius obtusus and

Mycena metata were common in the other spruce plots yet lacking in these three

plots and had an effect on the ordination of these plots.

With regard to the environmental variables, the spruce sites were grouped

between the ash and the Scot’s pine sites which corresponded to intermediate

levels of pH and soil calcium. They were also separated from the ash and the oak

sites by having high levels of coarse woody debris. Indicator species found to

signify Sitka spruce forests were Cystoderma amianthinum, Hemimycena gracilis,

Hypholoma capnoides, Postia caesia, Russula emetica, Collybia butyracea,

Mycena metata and M. pura with the first five species indicating young Sitka

spruce forests and the latter three species indicating mature Sitka spruce forests.

Sitka spruce forests in Ireland were very species-rich in the following genera,

Mycena (20 species), Cortinarius (13 species), Lactarius (10 species) and Russula

(11 species). The macrofungal species Collybia butyracea was found to be a very

abundant and widespread macrofungus in Sitka spruce plots. This species along

with Hypholoma fasciculare and Russula ochroleuca were probably the

macrofungal species with the highest biomass recorded in the Sitka spruce plots.

Multi-response permutation procedure MRPP analysis

The MRPP analysis found that fungal communities were significantly different

between the different forest types examined (Table 5.11). The significant P-value

244

and positive A value prove the starting hypothesis: that macrofungal species were

not randomly distributed between the different forest plots, but rather formed

distinct assemblages according to the dominant tree type of the forest plot. What

this means is that plots of the same forest type were grouped closer together in the

ordination than was expected by chance. The results of this analysis back up the

previous identification of distinct macrofungal assemblages in each of the forest

types through ordination (Figs 5.4, 5.5, 5.6), indicator species analysis (Table

5.10) and fruiting body abundance data (Table 5.1).

Table 5.11 Results of MRPP analysis of the macrofungal communities in the different forest types. A= Chance-corrected within-group agreement, P= probability of relationship being found by chance.

Grouping variable A P

Forest type 0.0687 0.001

Relating ordinations of species to the ecology of species

The relationships between individual macrofungal species and the ordination axes

were examined to reveal information about the ecology the species. Species

whose distributions were significantly correlated (r2>0.2%) with one or more of

the NMS axis were identified (Table 5.12).

Table 5.12 Macrofungal species which showed strong (r2>0.2) correlations with ordination axes. N.s. = non-significant correlations.

Species Axis 1 Axis 2 Axis 3

Collybia butyracea 0.225 n.s. n.s.

Crepidotus mollis 0.262 n.s. n.s.

Lactarius hepaticus n.s. n.s. 0.286

Lactarius rufus n.s. n.s. 0.269

Mycena leptocephala 0.256 n.s. n.s.

Paxillus involutus n.s. n.s. 0.275

Ricknella fibula n.s. n.s. 0.373

Stereum hirsutum 0.242 n.s. n.s.

Trichaptum abietinum n.s. n.s. 0.483

Trichoderma sp. n.s. n.s. 0.371

Xylaria hypoxylon 0.278 n.s. 0.303

245

All of the species correlated with axis 1 are litter or wood decay species,

which indicates that plots situated in the upper end of this axis had sufficient

conditions for the growth of these functional groups of fungi. As axis 3 was

correlated with both macrofungal species abundances and environmental

variables, therefore it is useful for the examination of the responses of these

species to the measured variables. A total seven species (3 ectomycorrhizal, 1

litter decay and 3 wood decay species) were correlated with axis 3. Axis 3 was

also correlated with decreasing pH, increasing organic matter (%), decreasing soil

available nitrogen and decreasing soil available phosphorus. It could be

extrapolated from these correlations that the species correlated with axis 3 show

some significant relationships with one or more of these environmental variables.

Alternatively, as the environmental variables listed were useful in separating the

deciduous from the coniferous forest types then the correlations may merely

reflect that the species are normally coniferous or deciduous forest species or are

normally found in one forest type over another.

5.4.5 Change in fungal community over sample years

The relationships between the forest sites revealed by the NMS analysis in the

previous section were based on pooled data for the three years of sampling.

Considerable variation in the year-to-year abundance of certain species was

observed in individual plots. Mantel tests were used to test whether the Sørenson

similarity matrices for individual years were more similar than expected merely

by chance. This would give an indication of the continuity of the macrofungal

community from year to year. To test the effect of sampling year on fungal

communities, a Mantel test was used on a species abundance matrix created for

each year of sampling which listed the sub-plot frequency of occurrence of the

species in each of the plots. The hypothesis being tested was that there was no

relationship between the similarity matrices of the sites between the different

years.

246

2007 and 2008

The Monte Carlo test showed that the relationship between the two matrixes may

have been due to chance (P= 0.137). The effect size r (according to the

standardized Mantel statistic) was -0.169.

2007 and 2009

The Monte Carlo test showed that the relationship between the two matrixes was

likely due to chance (P= 0.45). The effect size r was 0.017.

2008 and 2009

The Monte Carlo test showed that the relationship between the two matrixes was

not due to chance (P<0.05). The effect size r was 0.321. The effect size is low to

medium at r= 0.321 and positive which shows that the communities are positively

related.

The Mantel test only identified a significant relationship between the years

2008 and 2009. This means that the similarities between the plots based on

macrofungal species abundances were statistically replicated between 2008 and

2009, suggesting a continuity of community composition between the years, but

not between 2007 and later years. In order to test if the use of a quantitative

measure (Sørenson) of community composition could have caused a type II error

(the failure to reject a null hypotheses when it was true) a qualitative measure

(Jaccard’s) of community similarity was also used in the Mantel tests. Similar to

the Mantel test using the Sørenson index, the only significant relationships found

were between 2008 and 2009 (r= 0.335, P<0.05). Therefore the fungal

communities of the plots did not show similar patterns over the years 2007 and

2008; and 2007 and 2009.

247

5.5 Discussion

5.5.1 Evidence for the existence of distinctive communities of macrofungi in

Irish forests

This study has shown that distinctive macrofungal communities exist in each of

the four forest types. By analyzing abundance data calculated from species

frequency over a number of sub-plots, it was found that plots of a similar forest

type were more similar to each other than to plots of a different forest type. It was

found that species with high fruiting abundance and species which were more

common in one forest type than in the others, were more important in defining the

communities than forest type restricted species. Wilkins et al. (1937) also found

that species which are truly restricted are so few in number that they are not useful

in defining macrofungal communities, and may often be missed unless the

sampling frequency is high.

Other studies of temperate forests have also found that the macrofungal

communities of forests can be used to identify plots, based on the dominant tree

species of the plots. In the U.K., Humphrey et al. (2000) found that the data on

macrofungal communities collected over three years could be used to produce

distinctive groupings of their Scot’s pine, Sitka spruce and oak plots, which

corresponded to the forest type of the plot. Buee et al. (2011) collected

macrofungal data over 7 years in coniferous and deciduous plots in France, and

found that the ordination of the recorded sporocarp abundances separated their

plots into their respective forest types. In forest plots in Switzerland, Straatsma

and Krisai-Greilhuber (2003) found that cluster analysis of the macrofungal

communities grouped plots with similar dominant tree species closer to each other

than plots that had a different dominant tree species. In North America, it has been

found that the macrofungal communities of hardwood and coniferous forests are

very different (Villeneuve et al. 1980; Bills et al. 1986), and these differences

may be due to host preferences of ectomycorrhizal fungi (Newton and Haigh

1998) or substrate specificity which is demonstrated by many decay macrofungi

(Boddy 2001; Ludley et al. 2008).

248

This study also indicated that there is a level of continuity in macrofungal

species between some of the sample years. The macrofungal communities of

forests are made up of two components:

a) Species that consistently produce sporocarps in every sample year. Forest

macrofungi are known to show large variation in species appearances and

abundances over different sampling years (Krebs et al. 2008; Courty et al.

2008), with the long term study by Straatsma et al. (2001) highlighting this

variation by finding that only 8 of the 408 species found during the 21-year

duration appeared in all years. Of the 8 species, 6 were ectomycorrhizal

(Lactarius blennius, Russula cyanoxantha, R. fellea, R. nobilis as “R.

fageticola”, R. ochroleuca, and Xerocomus badius) and 2 were litter-decay

species (Collybia butyracea var. asema and C. dryophila). It is known that

some fungi from the genera Lactarius and Russula (Redecker et al. 2001);

and also the species Laccaria amethystina (Gherbi et al. 1999) rely heavily

on spores for propagation, and therefore for their yearly continuity in an area

it may be vital that they produce sporocarps every season.

b) Species that infrequently produce sporocarps over some sample years. The

vast majority of forest macrofungal species do not produce sporocarps in

every year (Straatsma et al. 2001) and these are likely responsible for the

non-significant relationship between community composition in some of the

years. It has been found that the majority of forest macrofungal species

cannot produce bumper crops two years in a row (Krebs et al. 2008), and

therefore relationships in species abundances between subsequent years may

not exist.

If both species group a (annual fruiting species) and the majority of group b

(inconsistently fruiting species) above produced sporocarps, then relationships in

macrofungal community composition between years (as in 2008 and 2009) is not

surprising.

249

5.5.2 The macrofungal communities of the forest types

Macrofungal community of Ash forests

Ash shared only one of its most frequent species (Mycena galopus) with another

forest type (oak). This indicates that ash forests have a rather specific group of

common macrofungi, which are not as common in the other forest types, although

these fungi are not restricted to ash forests either. Examples of some macrofungal

species which are more commonly found in ash forests, but not restricted, are

Daldinia concentrica and Terana caerulea which are most often found on ash

woody debris, but are also known from other tree species (Phillips 2006). Ash

plots were more similar to other ash plots than they were to oak, SP or SS plots

based on fungal community composition. The ordination analysis produced a

clearly delineated grouping for the ash plots, which corresponded with the ash

forest type grouping. Postia subcaesia and Xylaria hypoxylon were indicator

species of ash forests, while Crepidotus variabilis, Marasmius androsaceus,

Stereum hirsutum, Camarophyllopsis atropuncta, and Daldinia concentrica were

found to be more common in ash forests than the other forest types. Macrofungal

studies in ash forests are rare, in one of the few studies examining this forest type,

Hering (1966) found that Mycena galopus and M. polygramma were two of the

most common fungi of the 59 recorded in total over three years. These species

were also the second and third most commonly recorded fungi from ash forests in

this study. The species Camarophyllopsis atropuncta, a new record to the

Republic of Ireland, was found only in the two unmanaged ash forests. It is listed

from Northern Ireland as being found commonly with hazel (NIFG 2010) and is

noted as being found in moist basic forest soils (Bas et al. 1990). The total

number of species recorded from ash forests in England (Hering 1966) is very

similar to the total for the same forest type in this study (56 species).

Macrofungal community of Oak forests

Oak sites had the highest inter-plot Jaccard similarity and shared a high number of

macrofungal species between its plots than the other forest types. High intra-forest

type similarity values indicate that the oak forests have a more distinct fungal

community than the other forest types. Ordination of the plots based on

macrofungal community assemblages produced a close grouping of the oak plots,

250

which was related to their forest type classification. The litter- and wood-decay

macrofungi Crepidotus variabilis, Lycoperdon perlatum and Stereum hirsutum,

and the ectomycorrhizal fungi Laccaria amethystina and Lactarius quietus were

shown to be indicators of the oak forest macrofungal community. Lactarius

quietus is also associated with oak forests in the U.K., as it was found in almost

twice the number of quadrats in oak than in any other forest type (C. Quine,

unpub. data). L. quietus is one of the few species pointed out as being a typical

oak forest fungus by Watling (2005) in Scotland and in England by Wilkins et al.

(1937) and Hering (1966). Heilmann-Clausen et al. (1998) list it as being a

Quercus associate in Europe.

In oak plots, Laccaria laccata and L. amethystina were very common and

were also the most prominent sporocarp producers. These two species had equally

high abundances in the oak forest type, a finding similar to oak forests in the U.K.

(Watling 1974; Humphrey et al. 2003), that highlights the similar ecological

requirements of both species. Stereum hirsutum and Mycena galopus were the 3rd

and 5th most common fungi in Irish oak forests in this study and of similar

abundance in oak forests in the U.K. (Humphrey et al. 2003; C. Quine, unpub.

data).

The earthball Scleroderma citrinum was not found in any of the oak plots

in the U.K. (Humphrey et al. 2003; C. Quine, unpub. data), while it was very

common in the Irish oak forests. S. citrinum was listed by Hering (1966) in

pedunculate oak woods in the U.K. under the previous name of Scleroderma

aurantium, although it was not one of the more common fungi. It is listed as being

a common associate of oak, beech and birch in the British Isles (Pegler et al.

1995). This species was common to both the pedunculate and sessile oak plots in

this project. Scleroderma species produce long distance exploration hyphae, and

therefore are more successful at colonising root space in areas of low root density

(Peay et al. 2011). The oak plots examined in the U.K. study were much younger

and probably more densely stocked than the plots examined in this project, so the

tree density of the plots may be the explanation for the lower frequency of

Scleroderma species in the oak forests in Humphrey et al. (2000) study than in

this study of mature oak plots.

Other species only found in oak forests in this project are Daedalea

quercina and Fistulina hepatica, both of which are noted as typical oak forest

251

fungi in Scotland (Watling 2005). Both of these species were only found in the

pedunculate oak forest (Abbeyleix), which might indicate that they are more

common in pedunculate than in sessile oak forests in the Republic of Ireland.

Fistulina hepatica was found in only one pedunculate oak plot in the study in the

U.K. (C. Quine, unpub. data), and the plot was also the oldest oak plot examined

at >177 years old. The oak plot where Fistulina hepatica was found in this study

has been noted as one of the oldest (>200 years old) oak stands remaining in

Ireland (Rackham 1995), and this may indicate a preference for old pedunculate

oak woods by the bracket fungus Fistulina hepatica.

Watling (1974) lists 29 species from Boletales that are common

ectomycorrhizal macrofungi found with oak in the U.K., but Humphrey et al.

(2003) only found 3 species of Boletus (Boletus appendiculatus, B. chrysenteron

and B. porosporus) in their pedunculate oak woods. This study agrees with the

previous study by Humphrey et al. in that very few boletes (4), were found in oak

plots and they had no preference for oak forests. The clade Boletales has been

noted as being under-recorded if not species poor in the Republic of Ireland

(O’Hanlon and Harrington in press), and therefore this study may indicate that the

clade is actually species poor in Ireland. The records from Northern Ireland also

indicate that the genus Boletus may be species poor in Ireland, with only 23

species of Boletus recorded (NIFG 2010) compared to 40 species in England

(O’Hanlon and Harrington in press).

A specific group of Russula species were identified as indicative of oak

forests in the U.K. (Humphrey et al. 2000; C. Quine, unpub. data), including the

species (R. betularum, R. nigricans, R. fragilis and R. cyanoxantha), but no such

grouping was found in this study for oak forests. R. nigricans was found outside

of the plots a number of times during this study while R. cyanoxantha was found

outside of the plots on even less occasions during the three years. Both of these

species show widespread distribution in Northern Ireland (FRDBI 2010) and

therefore their scarcity in this project may be due to infrequent fruiting pattern of

these species. Hering (1966) found these two species to be specific to their sessile

oak woods, although they were only found as 17 and 4 fruitbodies respectively

over three years of sampling.

In a macrofungal community study of British pedunculate oak woods

(Wilkins et al. 1937), the authors identified the genus Lactarius as very species

252

rich in pedunculate oak woods in Britain, especially the species Lactarius

glyciosmus, L. mitissimus, L. pyrogalus and L. vellereus, but no such grouping

was found in the oak forest type in this project indicating that these species may

be sessile oak specific macrofungi; or associated with one of the other tree species

present in the examined plots (e.g. L. pyrogalus with the hazel in the plots). The

parasitic macrofungus Armillaria mellea is a species that was frequent in the oak

plots of this project and in British pedunculate (Wilkins et al. 1937) and sessile

(Hering 1966) oak-woods. It is listed as a parasite of deciduous trees in Ireland

and Britain (Legon and Henrici 2005), being common in deciduous forests up to

Scotland, where it is less common.

Macrofungal community of Scot’s pine forests

Scot’s pine plots had the lowest inter-plot similarity values because of the large

inter-site variation in macrofungal species abundance. The macrofungal

communities of Scot’s pine plots on average had more in common with those of

oak plots than with other Scot’s pine plots, although this was shown to be an

artefact of two of the pine plots having oak saplings in their understory.

Nevertheless, NMS separated the pine plots into a clear grouping which

corresponded loosely to their forest type determined by dominant tree species.

In total Scot’s pine plots shared 21, 44 and 58 species with ash, oak and

Sitka spruce plots respectively. Russula ochroleuca and Laccaria laccata were

very common in Scot’s pine, oak and Sitka spruce plots. The litter-decay fungus

Myena leptocephala was very common in Scot’s pine and Sitka spruce plots, thus

highlighting the non-specific requirements of these forest macrofungi. These three

species (Russula ochroleuca, Laccaria laccata and Mycena leptocephala), along

with Lactarius hepaticus, were some of the most prominent sporocarp producers

in Scot’s pine plots. The first three of these species were some of the most

commonly found macrofungi in British forests (Humphrey et al. 2003), with

Lactarius hepaticus being noted as one of the few ectomycorrhizal species which

is increasing in abundance in Europe in recent years (Heilmann-Clausen et al.

1998).

Lactarius hepaticus and Paxillus involutus were two of the species which

showed significant correlations with the ordination axis 3. This axis was also

correlated with the available nitrogen in the plot. That these two species were

253

identified as ectomycorrhizal species which are broadly tolerant of high nitrogen

(Outerbridge 2002; Kranabetter et al. 2009) may indicate that they have a

preference for habitats high in soil available nitrogen.

The indicator species analysis for Scot’s pine forests identified L.

hepaticus and Trichaptum abietinum as species which show some fidelity to the

Scot’s pine forest type. Lactarius hepaticus is listed as being an ectomycorrhizal

associate of Pinus, but also known to occur with Picea in Europe (Heilmann-

Clausen et al. 1998), while Trichaptum abietinum is known to grow on Picea,

Pinus and Abies (Breitenbach and Kraenzlin 1985). Both of these species, along

with two species known to colonize pine cones, Baeospora myosura and

Auriscalpium vulgare, were also common to all of the plantation Scot’s pine sites

examined in England by Ferris et al. (2000a). The latter two species were only

found in the old growth sites at Torc, Bansha and Brittas in this project.

Auriscalpium vulgare is restricted to pine cones, while being found less often on

spruce cones in the U.K. (Pegler et al. 1997) and is known to be only rarely

collected across its entire range (Peterson and Cifuentes 1994). Baeospora

myosura is restricted to the cones of pine and spruce trees (Breitenbach and

Kraenzlin 1991) in Europe. Microhabitat-restricted fungi such as these pine cone

fungi, are restricted in distribution in Ireland probably because of the sporadic

occurrence of patches of old-growth pine forests, which act as the spore bank for

such fungi.

The species Russula emetica, Amanita cecilia (listed as Amanita inaurata)

and Hypholoma marginatum were found to be very common in plantation Scot’s

pine forests of England by Richardson (1970), while they were found to be totally

lacking from the pine plots in this project. It is likely that the first two species

were locally more common in the area investigated in the English study as

opposed to being absent from Scot’s pine forests in Ireland. The latter species

(Hypholoma marginatum) may be restricted in Irish plantations by the lack of

suitable coarse woody debris available in some Irish plantations (Sweeney et al.

2010a).

The work examining Dryas macrofungi in the Burren by Harrington

(2003) can be used as an indication of what macrofungi existed with Scot’s pine

in the Burren area of Ireland in the past. A large Cortinarius element (~10%

species) is present in the macrofungal communities of both this study and the

254

study of the Burren heaths (~18% species) by Harrington (2003), although the

individual species differ markedly with the Cortinarius species from Scot’s pine

forests in this study being found throughout Ireland while the species in the

Burren are very restricted in their distribution across Ireland.

Macrofungal community of Sitka spruce forests

The hypothesis that Sitka spruce forests would share many fungal species with the

other forest types was confirmed in this research, due to the high levels of

similarity between spruce and the other forest types and due to the ordination of

the spruce plots close to the oak and Scot’s pine plots. The Sitka spruce plots in

Humphrey et al. (2000) study were also ordinated close to the pine and oak plots,

indicating that spruce forests shared many similarities in their macrofungal

communities with oak and Scot’s pine forests.

The ordination and indicator species analysis indicate that there are

differences in the macrofungal communities of young and mature Sitka spruce

forests. As coniferous sites age, the build up of litter on the forest floor is likely to

affect many of the abiotic characteristics of the forest floor. This litter build-up

has been shown to have mixed effects on the growth of some ectomycorrhizal

fungi (Barr et al. 1994; Baar 1996). Dighton et al. (1986) examined a

chronosequence of Sitka spruce forests in England, and found that, although

species diversity (Shannon index) was not significantly related to forest age, there

were marked differences in the community composition in the different age

groups, and these differences were related to the successional stages of

ectomycorrhizal colonisation and with the build up of the litter layers.

This study is the second systematic published study on the macrofungi of

Irish Sitka spruce forests. The one other published description of macrofungal

assemblages in Irish Sitka spruce forests by Heslin et al. (1992) identified 17

macrofungi in their mixed species spruce plantations from one sampling occasion,

although quantifying of macrofungal diversity was not the main aim of their

investigation. The pure spruce plots in the Heslin et al. study only produced four

species of macrofungi (Cortinarius obtusus, C. croceus, C. tubarius as

“Dermocybe sphagneti” and an unidentified Cortinarius species). The first species

(C. obtusus) was found in four of the nine spruce plots in this project while the

255

other species were not found in the project. Of the species from the mixed Sitka

spruce/Japanese larch and Sitka spruce/lodgepole pine plots in the Heslin et al.

study, Amanita rubescens, Laccaria laccata, L. amethystina, Marasmius

androsaceus and Tricholomopsis rutilans were common to spruce plots in this

study.

Sitka spruce forests shared 30, 65 and 58 species with ash, oak and Scot’s

pine plots respectively. The most prominent sporocarp producer in Sitka spruce

forests (Hemimycena gracilis) was found in three of the nine spruce sites. The

second and third most abundant sporocarp producers (Mycena leptocephala and

Russula ochroleuca) were also abundant in oak and Scot’s pine forests.

Hypholoma fasciculare and Collybia butyracea are very common in Irish Sitka

spruce forests (4th and 5th most found in Sitka spruce sub-plots). Hypholoma

fasciculare was also common in the Sitka spruce forests of the U.K (Humphrey et

al. 2003). Collybia butyracea was found more often with Norway spruce than

with Sitka spruce in the U.K. study (C. Quine unpub. data) and was found in all of

the Norway spruce sites examined by Ferris et al. (2000a) in England.

Sitka spruce forests were found to be more similar to each other than to

sites of the other forest types. NMS clearly distinguished a Sitka spruce

macrofungal community and also detected differences in community composition

between the young and mature plots. The spruce plots at Cloonagh, Moneyteige

and Chevy Chase Mature were outliers from the spruce group in the ordination,

and the reason for this is that these plots lacked a number of species which were

abundant in the other spruce plots. Examples of species which were absent from

the communities of Cloonagh, Moneyteige and Chevy Chase mature while being

present in the majority of the remaining Sitka spruce plots are Mycena

archangelina, M. rorida, M. epipterygia, Russula ochroleuca, Lycoperdon

nigrescens, Marasmius androsaceus, Entoloma cetratum, Clitocybe vibecina and

Amanita rubescens. The first two of these plots were first rotation spruce plots

planted on previously unforested land, therefore the lack of some species from

these plots may be due to a lack of inoculum sources such as spores or mycelium

in the soil of the plots. The lack of sufficient spore and vegetative inoculum has

been shown to affect fungal communities in plantation forests (Kranabetter and

Wylie 1998; Jones et al. 2003). The apparent low species richness of the mature

site at Chevy Chase might be explained by the diverse but infrequently fruiting

256

ectomycorrhizal macrofungal found in many late stage coniferous forests (O’Dell

et al. 1999; Smith et al. 2003). Although the belowground examination of

ectomycorrhizal morphotypes also identified this plot as poor in ectomycorrhizal

species (Table 6.4, Chapter 6).

Indicator species analysis identified Collybia butyracea, Cystoderma

amianthinum, Hemimycena gracilis, Hypholoma capnoides, Mycena metata, M.

pura, Postia caesia and Russula emetica as indicators of the Sitka spruce

macrofungal community. None of these species except Mycena pura (Roberts et

al. 2004) were found with Sitka spruce in its home range in British Columbia

(Outerbridge 2002). However, Mycena rosella, Lycoperdon nigrescens, L.

pyriforme, L. perlatum and Clavulina rugosa, which were identified by

Outerbridge (2002) as indicator species from the Vancouver Sitka spruce forests,

were present in the Sitka spruce forest type in this project, but were also found in

at least one of the other forest types. The majority of the very frequent species in

Sitka spruce forests in British Columbia are not found in Sitka spruce forests in

Ireland. The species Cantharellus formosus, Inocybe sororia, Mycena tenax, M.

aurantiidisca, Guepiniopsis alpina, Xeromphalina campanella and X. fulvipes

were some of the most commonly found from quadrats of Sitka spruce in the

Outerbridge (2002) study. Only two of the aforementioned species (Guepiniopsis

alpina and Xeromphalina campanella) are present in the Checklist of

Basidiomycetes of Britain and Ireland (Legon and Henrici 2005).

There were strong similarities between Sitka spruce forests in different

locations in respect of certain biodiverse genera. Mycena, Cortinarius and Russula

were very species rich in this project (20, 13 and 11 species) and in similar studies

from the U.K. (28, 35 and 16 species; C. Quine, unpub. data) and Canada (22, 14

and 13 species; Roberts et al. 2004). Of the Cortinarius species found with spruce

during this project, Cortinarius flexipes, C. rubellus, C. evernius and C. stillatitius

were all found fruiting in large quantities, but showed very local distributions in

some cases. C. rubellus has been recorded with Sitka spruce in Ireland

(Harrington 1994), and is thought to be a species which has undergone a “host

shift” moving from its normal ectomycorrhizal host of Scot’s pine to Sitka spruce

in Scotland (Watling 1982) and Ireland (Harrington 1994). It is a known associate

of spruce and pine in Europe (Brandrud et al. 1990-1998) and has been recorded

in Northern Ireland with Sitka spruce and Birch (NIFG 2010), and from England

257

and Wales in association with pine and spruce in plantations (Legon and Henrici

2005).

Two Clavulina species (Clavulina rugosa and C. corraloides) were often

found fruiting in large abundances in the Sitka spruce forests in this project,

although they were not found to be indicator species. Both of these species were

also found in Sitka spruce plots in British Columbia (Outerbridge 2002), while C.

corraloides was very abundant in red spruce plots in Virginia (Bills et al. 1986),

thus highlighting this species, and possibly genera, as having a preference for

spruce/coniferous forests as indicated by Breitenbach and Kraenzlin (1985) from

Swiss data.

Of the edible fungi found to be capable of growing in Sitka spruce forests,

Cantharellus cibarius, C. tubaeformis and Boletus edulis are of note. Cantharellus

cibarius was found in a spruce plantation, near birch trees which probably

indicates that it was ectomycorrhizal with the birch. Both Cantharellus species are

known from Spruce forests in the U.K. (Pegler et al. 1997). The macrofungi C.

tubaeformis and Boletus edulis were found in spruce forests in areas with only

Sitka spruce present. B. edulis is known to be found in both deciduous and

coniferous forests around Europe including Norway spruce Picea abies (Muñoz

2005). In this project and from anecdotal records, a relationship has been observed

in the close proximity of sporocarps of B. edulis and of Amanita muscaria in the

field. This relationship has also been noticed before by many collectors in the

forests of America (Arora 1986) and New Zealand (Wang and Hall 2004). A

similar relationship between Chalciporus piperatus and Amanita muscaria was

recorded in this project and elsewhere (Tedersoo et al. 2010). A possible

explanation for the fruiting of these species in close proximity as proposed by

Wang and Hall (2004), is that they are in a biotrophic relationship. It has been

recorded that Boletus edulis and Amanita muscaria can form composite

ectomycorrhizal structures on roots, and the sharing of resources which may

follow such events could be vital for sporocarp production of either or both

species (Wang and Hall 2004). Cantharellus tubaeformis was often found

growing in great abundance in Sitka spruce forests in this project and a related

Cantharellus species, C. formosus was also noted as being a prolific sporocarp

producer in Sitka spruce forests in the Vancouver sites (Outerbridge 2002).

258

5.5.3 Host preference in macrofungi

It was found that species such as Mycena leptocephala, M. vitilis and M. galopus

were common to all forest types examined in this project. Other species with a

seemingly cosmopolitan distribution are Russula ochroleuca, Cortinarius obtusus,

C. flexipes, Hypholoma fasciculare, Laccaria laccata and L. amethystina, being

found in large abundances in a number of plots from different forest types. Five of

the species listed above were also in the most common list of fungi from the study

of British forests (Humphrey et al. 2003). The species Laccaria amethystina, L.

laccatta, Hypholoma fasciculare, Russula ochroleuca and Mycena galopus have

already been noted as species that show very low levels of host preference in

temperate forests (Buee et al. 2011). The vast majority of the species recorded in

this project would fall into the relatively inconstant species according to Wilkins

et al. (1937).

Relatively constant species found often in ash plots were the species

Marasmius androsaceus and Stereum hirsutum, while the species

Camarophyllopsis atropuncta, Daldinia concentrica, Postia subcaesia and

Terana caerulea were ash restricted species. Other studies have shown that all of

the previously listed species are also found with other hosts in Europe (Phillips

2006; Breitenbach and Kraenzlin 1985; Whalley and Watling 1982; Kuyper et al.

1990) and therefore are not entirely restricted to ash forests.

In oak plots the species Crepidotus variabilis, Lycoperdon perlatum,

Lactarius quietus, Stereum hirsutum and Armillaria mellea showed a preference

for oak plots, while the species Scleroderma citrinum was restricted to the oak

forest type. Scleroderma citrinum, Lycoperdon perlatum, Stereum hirsutum and

Armillaria mellea have been noted as showing a low level of forest type

specificity in temperate forests in France, while Lactarius quietus showed a very

high level of host specificity to oak forests (Buee et al. 2011).

In the pine plots, the species L. hepaticus and Trichaptum abietinum

showed a preference for the habitat, fruiting more in pine than in any other forest

type. The species Baeospora myosura, Auriscalpium vulgare, Lactarius rufus,

Hygrophoropsis aurantiaca and Chroogomphus rutilus were restricted to pine

plots in this project. The species Lactarius hepaticus, Trichaptum abietinum,

259

Baeospora myosura and Auriscalpium vulgare were found in all of the Scot’s pine

plots in the U.K. study (Ferris et al. 2000a), and therefore may show a high level

of specificity to Scot’s pine in Western Europe.

In the spruce plots the species Collybia butyracea, Hypholoma capnoides,

Mycena metata, M. pura, Postia caesia, Russula emetica, Clavulina corraloides

and C. cristata showed a preference for Sitka spruce forests over the other forests

types, while the species Cystoderma amianthinum, Cortinarius obtusus, C.

rubellus, and Hemimycena gracilis were restricted to the Sitka spruce plots. An

example of a possible host shift (Watling 1995) may be the association of

Cystoderma amianthinum with Sitka spruce in Ireland. In Sweden, Cystoderma

amianthinum has been identified as a strong indicator of oak (Quercus robur)

forests (Tyler 1992). In agreement with this project, the study of the U.K. forests

(C. Quine, unpub. data) found that C. amianthinum was much more common in

coniferous forests. Another example of a host shift is likely in the relationship

between Cortinarius rubellus and Sitka spruce in Ireland. Such a relationship has

already been proposed in Scotland (Watling 1982), where it is said that C.

rubellus has migrated from its usual host Scot’s pine to the exotic species Sitka

spruce.

260

5.6 Conclusions

• This study shows that specific macrofungal communities exist for Irish ash,

oak, Scot’s pine and Sitka spruce forests, and these communities are linked to

the dominant tree species of the forest.

• Species which were very abundant in one forest type compared to other forest

types were more important in defining the fungal communities than species

that were restricted to a single forest type.

• This study found that the Sitka spruce macrofungal community is similar in

many ways to that of oak and Scot’s pine forests, thus corroborating with

other studies which identified Sitka spruce as an ectomycorrhizal generalist in

western European forests.

• Mixed relationships were found between the macrofungal communities over

the sample years, and these were due to the inconsistent fruiting patterns of

the majority of macrofungal species.

• Information as to the lack of host preference is presented for nine

macrofungal species with high levels of host preference identified for several

macrofungal species.

261

Chapter 6: Ectomycorrhizal morphotype richness and community analysis of the

forest types

262

263

6.1 Introduction

6.1.1 Ectomycorrhizas in forest ecosystems

Up to 85% of the known angiosperm species alive today are mycorrhizal, with

some 72% of those species being obligatory mycorrhizal (Wang and Qui 2006).

Within the other groups of land plants, gymnosperms, pteridophytes and

bryophyte species, 90, 52 and 46% respectively of their species show mycorrhizal

associations. Indeed, the majority of land plants depend on mycorrhizal fungi with

80% of known plant species forming mycorrhizal associations (Wang and Qui

2006). Ectomycorrhizas are known to be highly diverse, even in relatively small

geographic areas (Ishida et al. 2007; Morris et al. 2008). Explanations for the

large diversity of ectomycorrhizal types on the limited space available in the root

zone are numerous, and have been proposed by numerous reviews on the subject

(Bruns 1995; Taylor 2002; Kennedy 2010), such as: ectomycorrhizas are both

horizontally (Dickie and Reich 2005; Pickles et al. 2010) and vertically (Dickie et

al. 2002) distributed in the soil layers, ectomycorrhizas show large temporal

variation (Courty et al. 2008) and certain species also show distinct preferences

toward different substrates such as ectomycorrhizas from the clade Thelephorales

with CWD and certain ascomycetes with mineral soil layers (Tedersoo et al.

2003). Ectomycorrhizal (ECM) assemblages may change over chronological

stages of the forest cycle (Tweig et al. 2009) in what has been referred to by some

workers as “EM succession” (Deacon and Fleming 1992). However, young trees

planted near established trees have been shown to form ectomycorrhizal linkages

with so called “late stage” ectomycorrhizal fungi (Kranabetter 1999). Soil abiotic

variables such as nitrogen (Avis et al. 2003) and pH (Kjoller and Clemonsson

2009) have large affects on the species richness and community structure of the

ectomycorrhizas present.

Ectomycorrhizas are particularly important for plant hosts that grow in

nutrient-poor soil conditions (Smith and Read 2008), for example boreal

coniferous forests, and also where tree species are introduced into new habitats

where they may not be suited to the particular soil and environmental conditions

264

(Nunez et al. 2009). Ectomycorrhizas have been shown to link multiple trees of

different chronological stages by common mycorrhizal networks (CMN) (Beiler

et al. 2010). These linkages may facilitate carbon transfer between connected trees

(Simard et al. 1997; Simard and Durall 2004) and allow for the absorption of

limiting nutrients (Smith and Read 2008) and water (Lehto and Zwiazek 2011) by

connected trees. These two functions of CMN promote structural heterogeneity in

forest ecosystems, by allowing young seedlings to survive in the presence of

much larger trees. Many ectomycorrhizas have also been shown to promote

disease resistance in trees (reviewed in Whipps 2004).

When ECM assemblages (communities) on roots are assessed by

examining relative numbers on roots tips, they are generally dominated by a few

species accompanied by a long tail of less common species (Taylor 2002; Jonnson

et al. 1999; Peter et al. 2001). Cenococcum geophilum is known to be extremely

common on tree roots in temperate forests in Europe (Richard et al. 2005; Buee et

al. 2007; Courty et al. 2008 Gebhardt et al. 2007) and North America (Valentine

et al. 2004; Walker et al. 2005; Morris et al. 2008). Other species are known to be

highly competitive for space and nutrients; inter-species ectomycorrhizal

competition has become an area of recent focus in ECM research (Kennedy 2010;

Pickles et al. 2010; Peay et al. 2011). An example of competitive interactions

between ectomycorrhizas has been identified between the hypogeous species pair

Rhizopogon occidentalis and R. salebrosus. In this interaction R. salebrosus was

found to be competitively inferior to R. occidentalis, only being present on roots

that were not already colonised by R. occidentalis; indicating the timing of

colonization as having a strong effect on ECM success (Kennedy and Bruns

2005).

6.1.2 Quantifying and recording ectomycorrhizal fungi in forests

Traditionally, studies of ectomycorrhizas in forests recorded presence either by

sporocarp studies or by morphological identification of EM root tips, based on

descriptions of ECM morphotypes by Agerer (1997-2002), Ingleby et al. (1990),

Goodman et al. (2000) and others. Assessments of ECM diversity based on

sporocarps alone usually cannot adequately describe ECM communities, because

many species may sporulate irregularly or not at all. There may in fact, be little

265

correlation between above- and below-ground ECM communities; species that are

poorly-represented underground on roots may sporulate prolifically and may be

over-represented above-ground as sporocarps, and vice versa (Gardes and Bruns

1996; Dahlberg et al. 1997; Jonsson et al. 1999; Peter et al. 2001; Durall et al.

2006). In addition, it can be difficult to distinguish species using ECM

morphotyping methods (Sakakibara et al. 2002), because many described

morphotypes (Agerer 1997-2002) have not been identified to species, and because

different species may have similar morphology, such as ECM mantle structure

which can lead to mis-identification of some morphotypes (Sakakibara et al.

2002).

PCR-based methodology has allowed investigators to overcome many of

these problems in recent years, and has permitted researchers to examine in

greater detail the diversity, distribution and ecology of ectomycorrhizal

communities (Horton and Bruns 2001). Combining molecular methods with

traditional morphotyping may be the most time and cost effective method of

garnering information about the below-ground communities in forest ecosystems

(Sakakibara et al. 2002). Combinations of morphological and molecular methods

have been used to reveal extremely high species richness in forest ectomycorrhizal

communities (Jonsson et al. 1999; Tedersoo et al. 2006; Morris et al. 2008).

6.1.3 Ectomycorrhizal research in Ireland and other temperate countries

There have been relatively few studies carried out examining the diversity of

ectomycorrhizas on roots of forest trees in Ireland. Some studies carried out in

Ireland which examined ECM fungi on roots by morphological classification are

the examination of pure culture synthesis of Sitka spruce ectomycorrhizal in the

80s by Schild et al. (1988) and the study of Heslin et al. (1992) which examined

the ectomycorrhizas present on the roots of Sitka spruce in a plantation forest. The

Heslin et al. study found that the ECM richness of Sitka spruce plantations was

rather poor when compared to its richness in its home range (Alexander and

Watling 1987) and that it shared many of its ectomycorrhizas with other

coniferous tree species, thus indicating that it may have a generalist

ectomycorrhizal mycota in Irish plantations. The distribution of forest ECM

species on the roots of Dryas octopetala and other shrubs in the Burren, Western

266

Ireland, has been investigated by Harrington and Mitchell (2002a; 2002b; 2005a;

2005b) and Liston and Harrington (unpublished).

In the U.K., there have been many more published studies examining the

below-ground ectomycorrhizal component of forests, with many of the older

studies listed in Harley and Harley (1987). The study by Palfner et al. (2005),

identified 13 morphotypes on the roots of Sitka spruce trees in England, and

reasons such as the non-native status of Sitka spruce and the destructive short

rotation cycles were proposed for the low ectomycorrhizal richness and

dominance of the community by a single morphotype (Tylospora fibrillosa).

Taylor and Alexander (1989) carried out one of the first examinations of the

ectomycorrhizas in plantation Sitka spruce stands in the U.K. and found that of

the 17 morphotypes present, the Tylospora fibrillosa morphotype was very

abundant, colonizing almost 70% of root tips while the rest of the morphotypes

were much less abundant. Thomas et al. (1983) investigated the changes in ECM

assemblages that occur after outplanting from nurseries to forest sites. They found

24 ECM morphotypes present in their mature Sitka spruce plots, including C.

geophyllum, Russula ochroleuca, Laccaria amethystina, Lactarius rufus and

Paxillus involutus; with P. involutus, Russula ochroleuca and Lactarius rufus

being present on both mature and young trees, indicating a lack of successional

stage specificity in these species. All of these species, except Cenococcum

geophyllum, are also common above-ground in Sitka spruce forests in the U.K.

(Humphrey et al. 2000).

The ECM assemblages present in sessile oak forests have been examined

in France by Buee et al. (2007) and Courty et al. (2008). Buee et al. (2007) found

thirty six ectomycorrhizal morphotypes, of which the majority (12 morphotypes)

were from the genus Tomentella, with lesser numbers from the genera Lactarius

(four morphotypes) and Russula (five morphotypes). Their study also found that

ECM assemblages vary according to niche effects in the forest. They found that

morphotypes from the genera Lactarius and Tomentella were much more common

and abundant in coarse woody debris, while morphotypes from the genus Russula

were more common in the upper soil layers. Courty et al. (2008) examined the

same oak forest, and found that the four morphotypes, Lactarius quietus,

Tomentella sublilacina, Cenococcum geophyllum and an unidentified Russula

morphotype were very common throughout the samples. Although the above-

267

ground sporocarps were not examined in either of the two studies by Buee et al.

(2007) or Courty et al. (2008); the genera Russula and Lactarius are known to be

very species rich and abundant in similar forests in England (Hering 1966),

indicating that these genera may show similar levels of abundance and diversity

above- and below-ground. The genus Tomentella is rarely found as sporocarps in

the U.K. (Humphrey et al. 2003), because it forms minute fruiting bodies on

coarse woody debris and may often be overlooked.

The ECM assemblages of Scot’s pine forests were investigated in Scotland

by Pickles et al. (2010). They identified 24 different ECM morphotypes, the

majority (8 morphotypes) of which belonged to the genus Cortinarius, with 2

morphotypes from the genera Russula and Suillus (S. bovinus and S. variegatus)

also present. They found that Cenococcum geophyllum was one of the most

abundant ectomycorrhizas in the samples, being found in >98% of the samples.

The above-ground analysis of a similar site to the Pickles et al. (2010) site, carried

out by (Humphrey et al. 2000), found that Cortinarius species were the most

abundant followed by Russula and also the two Suillus species (S. bovinus and S.

variegatus).

268

6.2 Aims of this chapter

The aims of this chapter are:

• To examine if the ECM assemblage of the oak, Scot’s pine and Sitka spruce

forests vary significantly. It is expected that plots of the same forest type

would be more similar than plots of a different forest type based on their

ECM assemblages. It may also be expected that certain ECM morphotypes

would be much more common and possibly semi-restricted to a single forest

type in this study and thus may have potential for use as indicator species of a

certain forest type.

• To examine if there is a relationship between above- and below-ground

species richness. A plot that is very species rich in ECM sporocarps might be

expected to have high levels of below-ground ECM morphotype richness.

• To investigate if the ECM assemblage present in Sitka spruce forests shares

many ECM morphotypes in common with oak and Scot’s pine forests. Sitka

spruce is known to be an ECM generalist in the UK with many known ECM

associates but few species restricted to spruce (Newton and Haigh 1998).

269

6.3 Methods

6.3.1 The plots

The three forest types were composed of (1) oak - four plots, Raheen, Tomies,

Kilmacrea (Q. petraea) and Abbeyleix (Quercus robur); (2) Scot’s pine Pinus

sylvestris – four plots Annagh, Brittas, Torc and Bansha; (3) Sitka spruce Picea

sitchensis – four plots Dooary, Bohatch, Chevy Chase (Young) and Chevy Chase

mature . Full details of the plots are given in Table 3.2, Section 3.3.1.

6.3.2 Sampling and enumeration

The sampling method was adapted from the intensive method of ECM sampling

and anlaysis employed by Menkis et al. (2005) in that it examined a large number

of roots from a small number of plants. A total of five soil cores, one from each

sub-plot, were taken from the upper soil layer in each plot using a 10cm diameter

soil corer which measured 15cm in length. The samples were kept at 4ºC until

counting and analysis could be undertaken. The cores were soaked in tap water for

15 min and washed through a 1mm sieve. Senescent roots, coarse woody debris

and large root segments were removed manually. The remaining roots were cut

into 2cm segments and divided randomly into 6 sub-samples, each in a separate

petri-dish. The length of the ECM roots in each petri dish was estimated using the

line intersect method (Tenant 1975). Each petri dish was examined at 30x

magnification using a Nikon stereo microscope and morphologically similar

morphotypes were distinguished. Morphotype richness (per plot) is the number of

distinct ectomycorrhizal morphotypes recorded from the 30 sub-samples (5 cores

x 6 sub-samples per core) taken from the plot. Morphotype abundance was

estimated in two ways:

(i) Morphotype abundance per forest type; the number of appearances of the

morphotype in the sub-sample from each forest type.

(ii) % frequency (per plot); this was calculated as:

Presence of the morphotype in 30 subsamples/30

* 100 = % frequency of morphotype in a plot

270

Each morphotype was microscopically examined at x 400 to x 1000 by

taking pieces of the mantle and describing the inner- and outer-mantle according

to Agerer (1987-2002), digital images of the outer surface and basal layers of the

mantle were used in identification with references to descriptions by Agerer

(1987-2002). Secondary morphological characteristics such as clamp connections

and presence of specialised cells such as cystidia were also noted. Terminology

for the descriptions of ectomycorrhizas follows Agerer (1987-2002). Samples of

the mycorrhizal root tips were stored in 2% glutaraldehyde for future

morphological examination, and at -80ºC for molecular analysis at a later date at

the Department of Life Sciences, University of Limerick. Types which could not

be identified morphologically or by molecular methods were given a unique code

e.g. 3 a1 and are referred to here after as their code name.

The percentage similarity between below-ground and above-ground (sporocarps) ECM species assemblages was calculated using a modified form of the Jaccard index:

Species in common

* 100 = percentage similarity

Species in common +Species only found aboveground + Species only found belowground

6.3.3 Molecular identification of ECM types

DNA was extracted from ECM sporocarps and ECM root tips using the Qiagen

Plant Mini Kit (Qiagen, California, U.S.A.). The manufacturer’s instructions were

followed without modifications.

The ITS region of rDNA was amplified using the method described by

Gardes and Bruns (1993) using combinations of the primers ITS1-F, ITS4 and

ITS4-B (White et al. 1990; Gardes and Bruns 1993). The primer pairs ITS1-F and

ITS4; and ITS 1-F and ITS4-B were used. To make a 10µl master mix for

amplification, 2.5µl of extracted DNA was added to 4.175µl PCR grade water,

1µl 10x buffer, 1µl DNTPs stock (containing premixed bases), 0.65µl MgCl2,

0.3µl forward primer, 0.3µl reverse primer and 0.075µl high fidelity Taq

polymerase to make a. Amplification was carried out in a G-Storm GS2

thermocycler (G-Storm Ltd., Surrey, UK) using the following 3 stage

amplification conditions: Stage1-denaturing at 94°C for 2 min, Stage 2 -35 cycles

271

of denaturation @ 94°C for 15 seconds, annealing @ 55°C for 30 seconds,

elongation @ 72°C for 60 seconds; Stage 3 -final elongation @ 72°C for 7 min

The resulting amplicon was visualized on a 2% Sybr-Safe agarose gel and

examined under ultraviolet light.

For RFLP analysis of the samples, the PCR fragments were digested with

the restriction endonucleases AluI and HinfI (Roche Applied Sciences, West

Sussex, U.K.). 10µl of amplified PCR product was incubated for 1.0 h with 1.5µl

of 10x buffer (supplied with restriction enzyme), 8.4µl PCR grade water and 0.1µl

restriction enzyme, at the specified temperature (AluI and HinfI 37ºC; Taq1 65ºC).

The resulting PCR-RFLP fragments were separated on a 2% Sybr-Safe agarose

gel with a negative control. The molecular sizes of the fragments were calculated

using the band analysis feature of the GeneTools software (Syngene UK,

Cambridge UK.) and recorded by measurement against a DNA molecular weight

marker (DNA molecular weight marker VIII, Roche Applied Sciences, West

Sussex, U.K.). This information was then entered into a database using the GERM

(Good Enough RFLP Matcher) excel tool (Dickie et al. 2003). Matches between

sporocarps and ectomycorrhizal tips were calculated using the recommended

settings for matches as advocated by Dickie et al. (2003); backward and forward

matches of the bands were accepted only if the molecular weights were within 25

base pairs (bp) of each other. If the sum of the differences between observed and

expected bands exceeds 100 bp, then the samples were not a match.

Sequencing of amplicons: PCR products were purified prior to sequencing

using the QiaQuick PCR purification kit (Qiagen, California, U.S.A). All

ectomycorrhizal samples that amplified were sent for sequencing to Eurofins

MWG Operon, London, U.K. The resulting sequences were entered into Bio-Edit

Software (Hall 1999) and all sequences were compared to identify matches with

the optimal global alignment option and their identity similarity index was

calculated according to the “IDENTIFY” matrix. Sequences with similarity values

greater than 0.98 were accepted as the same species. The identify matrix setting

was used because it has high penalties for mis-matches (BIOedit manual; Hall

2005) and is therefore useful for separating species within the same genus. The

un-edited sequence data was entered into the UNITE data base (Koljalg et al.

2005) and a Blast-n search was carried out using both the UNITE database and the

272

Genbank database. Sequences with identities above 98% were treated as species-

level matches, while identification at the genus level was based on Blast

consensus of over 90% (Barroetavena et al. 2010). The name suggested by

UNITE, a curated database for ECM fungi (Koljalg et al. 2005) was used

preferentially, and that of Genbank only if there was no entry in UNITE.

6.3.4 Statistical analysis

ANOVA with Tukey's post-hoc test was used to identify any significant

differences in the mean morphotype richness and mean morphotype abundance

between the forest types and individual forest plots.

Rarefaction and morphotype richness estimation

The length of roots sampled was used to compare the morphotype richness of the

different forest types at a similar sampling intensity, sample-based rarefaction

curves with 95% confidence intervals were calculated, using the computer

program EstimateS (Colwell 2004). The settings for the rarefaction analysis are

the same as those used in Section 4.3.4 Chapter 4, except that one thousand

permutations were carried out on the core data from the individual plots within a

specific forest type. Sample-based rarefaction was used instead of individual-

based rarefaction as fungal species have been shown to be non-randomly

distributed (Taylor 2002; Tedersoo et al. 2003).

Possible levels of morphotype richness were estimated using the Chao2,

Ace and Ice estimators. The settings for the richness estimation are the same as

those used in Section 4.3.4, Chapter 4. Rareness parameter was defined at 2; i.e.

any morphotype that was found in less than 2 cores was defined as rare.

Fungal species diversity was estimated using the species richness

estimators in EstimateS. The Simpson’s diversity and Simpson’s evenness

measure was calculated for each forest type. The settings used were similar to

those used in Section 4.3.5, Chapter 4.

6.3.5 Community similarity analysis

To assess relative ECM distributions within the forest type communities, rank

abundance plots (Whittaker 1965) were constructed following Magurran (2004).

The relative frequency of MorphotypeX in forest typen was calculated as:

273

=Number of appearances of MorphotypeX in all sub-samples in forest typen / Number of appearances of all Morphotypes in all sub-samples in forest typen. The fitting of log-normal distribution model was carried out as in Section 5.3.5,

Chapter 5.

To analyse the ECM community similarity between the different sites, the

abundance-based Jaccard similarity index (Chao et al. 2005) was calculated in

Estimate-S. Median values for the Jaccard index (JI) were calculated by pooling

the index values for individual sites based on tree type. Median JI values and the

minimum and maximum values were calculated within forest types and between

forest types.

To test if the fungal communities were similar above- and below-ground,

the data from the plots which were sampled for sporocarps and ECM roots was

analysed using a Mantel test. The above-ground species plot matrix consisted of

the sub-plot frequency (Section 5.3.5) data for 81 ECM species from 12 plots,

while the below-ground species plot matrix consisted of morphotype abundance

data for 56 morphotypes from the same plots. Similarity matrixes were created

using the Sørenson distance index and were compared using PC-Ord and

evaluated for significant using a Monte Carlo test with 9999 randomizations.

6.3.6 Multivariate community analysis

Nonmetric multidimensional scaling (NMS) was used to examine the relationship

between the different ECM assemblages. MRPP was used to assess whether the

ECM assemblages would group significantly according to dominant tree type. The

settings used for NMS in this chapter are identical to those used in Section 5.3.6,

Chapter 5. NMS was run on the morphotype abundance (% frequency) data from

all of the cores from each site thus producing a 51 x 12 species by plot matrix

using PC-Ord version 4.36 (MjM Software Design, Gleneden Beach, OR). A

statistical technique called Beals’ smoothing (Beals 1984) was carried out on the

ECM abundance matrix following the advice of McCune and Grace (2002) in

order to lessen the effect of zeros on the ordination. Beals’ smoothing is a

technique used to replace zero values in an ecological matrix with “probability of

occurrence” values calculated from the joint occurrences of the target species at a

274

given plot with the remaining species in the data table (Caceres and Legendre

2008).

275

6.4 Results

6.4.1 ECM morphotypes

In total 51 distinct morphotypes were collected from 12 forest plots (Table 6.4).

This included 16 morphotypes identified solely by morphotyping methods, 3

morphotype identified by RFLP, 17 morphotypes identified by sequencing and 15

morphotypes remaining unidentified. Twenty-one ECM types were found in oak,

20 in Scot’s pine and 18 in Sitka spruce.

Morphotyping of unknown ectomycorrhizal types

A total of 15 ECM types could not be identified to genus level and so were given

a code name. Both morphological and RFLP analysis allowed for the separation

of these types into their own morphotypes, but it was not possible to assign them

to genera or species either by RFLP-matching of sporocarps or sequencing of the

ITS region. The 15 types are described in the following section with

corresponding Figs. 1-36 available in Unidentified ECM folder, within Folder 4 in

attached disk in Appendix. The final three descriptions given in this section are of

identified ECM morphotypes that do not currently have descriptions in either

Agerer (1987-2002) or BCERN

Morphotype: 3a7 (Fig. 1; Appendix) Forest type: oak Morphology: White/beige colour, smooth texture with monopodial pinnate branching. Extramatrical hyphae inconspicuous; rhizomorphs absent.. Mantle anatomy: The mantle plectenchymatous with hyphae irregularly arranged; corresponding to mantle type B in Agerer’s types. No other microscopically distinguishing features were noted. Possible identification: Resembles Cantharellus cibarius morphotype. Morphotype: 3b7 (Fig. 2) Forest type: Oak Morphology: White weblike structure on a crème backround. Cottony texture with monopodial pinnate branching of the ectomycorrhiza. Mantle anatomy: The mantle was plectenchymatous with ring like arrangement of the hyphal bundles matching Agerer’s type A mantle. Possible identification: Possibly a Hebeloma species

276

Morphotype: 3c7 (Figs. 3, 4) Forest type: Oak Morphology: Light brown colour with a smooth texture. Short monopodial pinnate branching of the ectomycorrhiza give the morphotype a clumped structure. Mantle anatomy: Mantle was pseudoparencyhmatous made up of epidermoid cells creating a jig-saw like pattern. It corresponded to mantle type M according to Agerer’s types. Morphotype: 3c9 (Figs. 5, 6) Forest type: Sitka spruce Morphology: The ectomycorrhiza was grey in colour and had a felty appearance. It displayed monopodial pinnate branching and due to copious amounts of fine hyphal strands it held many soil particles in a hyphal mesh. Mantle anatomy: The mantle was plectenchymatous with the hyphae irregularly arranged and corresponded to mantle type B in Agerer’s types. No other microscopically distinguishing features were noted. Morphotype: 3e7 (Figs. 7, 8, 9) Forest type: Oak Morphology: This ectomycorrhiza had a golden colouring and due to abundant fine hyphal strands had a cottony appearance. It appeared to be dichotomously branched. Mantle anatomy: The inner mantle was plectenchymatous with the hyphae irregularly arranged and corresponded to mantle type B in Agerer’s types. The outer mantle was also plectenchymatous but with the hyphae arranged into ring like bundles corresponding to Agerer’s type A mantle. Morphotype: 3e9 (Figs. 10, 11, 12) Forest type: Sitka spruce Morphology: The morphotype colour was very light brown with a smooth texture. It showed monopodial pinnate branching with a twisted appearance to the individual branches. Mantle anatomy: The mantle was transitional between plectenchymatous and pseudoparenchymatous and matched Agerer’s type H mantle. Possible identification: Resembles a Laccaria species. Morphotype: 4c1 (Figs. 13, 14, 15) Forest type: Scot’s pine Morphology: The morphotype was dark brown in colour and had a smooth outer texture. It showed monopodial pinnate branching with the branches forming close together and giving the ectomycorrhiza a clustered appearance. . Mantle anatomy: Both the inner and outer mantles were pseudoparenchymatous with the former being composed of angular cells and the latter being composed of epidermoid cells. The two types corresponded to Agerer’s types L and M mantles. Possible identification: Possibly a Tomentella species.

277

Morphotype: 4c3 (Figs. 16, 17) Forest type: Scot’s pine Morphology: The morphotype had a light brown colour with whitish growing tips. It has a smooth texture and lacked emanating hyphae. It showed dichotomous branching pattern typical of many Scot’s pine ectomycorrhizas. Mantle anatomy: The mantle was plectenchymatous with the hyphae irregularly arranged and corresponded to mantle type B in Agerer’s types. Morphotype: 4e1 (Figs. 18, 19, 20) Forest type: Scot’s pine Morphology: The morphotype had a white frosty appearance on a crème background. The outer frosty layer gave the morphotype a rough appearance and the branching followed a monopodial pinnate structure. Mantle anatomy: The mantle had a pseudoparenchymatous structure composed of angular cells with circular cells found in clusters and corresponded to Agerer’s type K mantle. Morphotype: 4f9 (Figs 21, 22, 23) Forest type: Sitka spruce Morphology: The morphotype had a grey/silver colouration and a cottony texture. It followed an irregular pinnate branching pattern and had a bendy appearance. Mantle anatomy: The mantle was transitional between plectenchymatous and pseudoparenchymatous and matched Agerer’s type H mantle. Morphotype: 4h5 (Figs. 24, 25) Forest type: Scot’s pine Morphology: This morphotype had a light brown colour with white growing tips and a smooth texture. It had the typical Scot’s pine dichotomous branching system. Mantle anatomy: The mantle was transitional between plectenchymatous and pseudoparenchymatous and matched Agerer’s type H mantle. Morphotype: 4j7 (Figs. 26, 27, 28) Forest type: Oak Morphology: The morphotype had a distinctive orange to light brown colour and a smooth texture. It followed a monopodial pinnate branching system and had bendy segments. Mantle anatomy: The mantle was plectenchymatous with the hyphae irregularly arranged and corresponded to mantle type B in Agerer’s types. Possible identification: Possibly Paxillus involotus Morphotype: 5a7 (Figs. 29, 30) Forest type: Oak Morphology: This distinctive morphotype was composed of a white frosty mantle with long emanating hyphae which formed long rhizomorphs. It had an irregularly pinnate branching pattern. Mantle anatomy: The mantle was plectenchymatous with the hyphae irregularly arranged and corresponded to mantle type B in Agerer’s types Possible identification: Cortinarius species or Scleroderma citrinum or Boletus species.

278

Morphotype: 5c9 (Figs. 31, 32, 33, 34) Forest type: Scot’s pine Morphology: This very distinctive morphotype had a white web like outer layer on a brown inner mantle layer. The most distinctive characteristic was the abundant large white sclerotia formed by the type. Mantle anatomy: The mantle was plectenchymatous with the hyphae arranged into ring like bundles corresponding to Agerer’s type A mantle Possible identification: Cortinarius species Morphotype: 5e7 (Figs. 35, 36) Forest type: Oak Morphology: This morphotye was dark brown in colour and had a smooth appearance. It underwent monopodial pinnate branching with 3-5mm between branches. Mantle anatomy: The mantle was plectenchymatous with the hyphae arranged into ring like bundles corresponding to Agerer’s type A mantle. The other mantle was pseudoparencyhmatous made up of angular cells. It corresponded to mantle type L according to Agerer’s types. Morphotype: Inocybe cincinnata (Fig. InocybeCincinnata in Identified ECM folder, Folder4, Appendix) Forest type: Sitka spruce Morphology: This morphotype was grey with a hint of purple in colour and had a smooth appearance. It underwent monopodial pinnate branching with 1-2 mm between branches. Mantle anatomy: The mantle was plectenchymatous with the hyphae irregularly arranged and corresponded to mantle type B in Agerer’s types Morphotype: Russula puellaris (Fig. RussulaPuellaris and RussulaPuellarisM in Identified ECM folder, Folder4, Appendix) Forest type: Scot’s pine Morphology: This morphotype was crème to brown in colour and underwent typical pine dichotomous branching giving the ectomycorrhiza a clumped appearance. Mantle anatomy: The mantle formed resembled a pseudoparenchymatous type with epidermoid cells bearing a delicate hyphal net, matching type Q in Agerer’s types. Morphotype: Entoloma serrulatum (Fig. EntolomaSerrulatum and EntolomaSerrulatumM in Identified ECM folder, Folder4, Appendix) Forest type: Scot’s pine Morphology: The ectomycorrhiza formed a white frost on a light brown background. It branched dichotomously similar to many of the other pine ectomycorrhizas. Mantle anatomy: The mantle was pseudoparenchymatous with angular cells and matched Agerer’s type L mantle.

279

RFLP –matching

A total of 35 ECM tips and 43 suspected ECM fruitbodies from the plots were

analysed using the restriction enzymes HinfI and AluI. The resulting restriction

fragments were visualised under ultraviolet light (Fig. 6.1). The molecular

weights of the resulting bands were calculated with reference to the standard

molecular weight marker MW VIII (Table 6.1). Using the GERM spreadsheet

based program, ECM tips were matched to similar tips and also to fruitbodies

based on their RFLP fragment lengths (Table 6.1). Morphotypes which were not

matched to fruitbodies by RFLP analysis were re-amplified and amplicons were

sequenced and sequences checked against online databases (Table 6.2).

280

Fig. 6.1 RFLP patterns for a select number of macrofungal sporocarps and ECM morphotypes. Top lane; 1= ladder, 2= 3 b9, 3= 3 e3, 4= 3 f3, 5= 3 g3, 6= 3 g7, 7= 3 h9, 8= 3 i3, 9= (-) Control, 10= 4 e9, 11= Amanita fulva, 12= 5 i1, 13= 4 a7, 14= A. muscaria, 15= Cortinarius acutus, 16= C. cinnamomeus, 17= (-) Control , 18= C. evernius, 19= C. obtusus , 20= (-) Control ,21= C. scandens, 22= Ladder. Bottom lane; 1= ladder, 2= C. umbrinolens, 3= C. venetus, 4= C. flexipes, 5=Entoloma cetratum, 6= E. conferendum, 7= E. conferendum, 8=Inocybe

geophylla, 9= (-) Control, 10= I. lanuginosa, 11= Laccaria amethystina, 12= Lactarius camphoratus, 13= L. deterrimus, 14= (-) Control , 15= L. hepaticus, 16= Russula fragilis, 17= R. ochroleuca, 19= blank , 20= Ladder. The molecular weight bands are also labelled with their corresponding mass in kb.

281

Table 6.1 Restriction digests of the of PCR-ampliied ITS-rDNA from ECM morphotypes or sporocarps using the restriction enzymes Hinf1 and Alu1. Column 2 lists source of the sample material, either m= mycorrhiza or f= fruitbody. The numbers listed are the masses of the fragments in kb as measured against a standard molecular weight marker (mw VIII). Numbers in bold-face are faint bands. Examples of exact matches between the RFLP patterns of types listed in the table are indicated by matching superscript letters following the morphotype.

Sample T Hinf1 Alu1

3 a5 m 493 467 369 411 260

3 a9 m 485

3 b5 m 423 764

3 b9 m 408 372 675 574 96

3 c7 m 675

3 e3c m 459 316 566 368

3 f3 m 490 532

3 g3c m 460 565 368

3 g7 m 410 593

3 h5 m 498 485 589 213 160

3 h9 m 493 454 548 382

3 i3 m 424 521 307 181 87

3 j3 m 443 675

4 a3 m 469 330 566 490 382

4 b1 m 461 288 192 144 506

4 b3 m 414 313 388 245 205

4 b7 m 481 454 451 196

4 c3a m 492 414 538 211 85

4 c5 m 454 206 556 208

4 c9 m 420 463 177

4 d1 m 414 237 178 480 106

4 d5 m 482 446 533 326

4 e3 m 491 358 127 580 326

4 e9 m 419 804 725 628

4 f3b m 461 367 352 239 106

4 g3b m 465 352 239 104

4 h5 m 463 252 195 501 108

4 i3 m 491 469 556 308

4 i5 m 491 427 810 637 506 203

4 j5 m 299 175 459 349 133

5 a9a m 491 418 491 101

5 c7 m 439 313 216 146 549 388

5 c9 m 457 785

5 e9 m 492 418 549 223

5 i1 m 462 319 199 148 532 467 364 85

282

Amanita fulva f 433 529 181

Amanita

muscaria

f 465 418 149 500 486 324 109 86

Amanita

phalloides

f 461 146 847 95

Amanita

rubescens

f 496 423 581

Boletus badius f 489 427 580 512

Boletus

chrysenteron

f 496 446 136 949 728 245 96

Cortinarius

acutus

f 493 554 373

Cortinarius

anomalus

f 491 439 146 522 289 90

Cortinarius

bolaris

f 493 418 415 211

Cortinarius

cinnamomeus

f 491 428 801 697

Cortinarius

evernius

f 414 144 767 502 303 176 87

Cortinarius

flexipes var.

flabellus

f 411 683

Cortinarius

flexipes var.

inolens

f 491 522

Cortinarius

obtusus

f 268 826 720 606 236 103

Cortinarius

rubellus

f 495 427 560

Cortinarius

scandens

f 419 147 731 604 499

Cortinarius

stilatitious

f 428 302 218 601 522

Cortinarius

unbrinolens

f 411 692 499

Cortinarius

venetus

f 491 442 140 806 665 230 174 131

Entoloma

cetratum

f 347 280 147 395 320 254 198 131

Entoloma

clypeatum

f 435 755

Entoloma

conferrendum

f 344 276 147 391 300 250 131

Entoloma

conferrendum

var. pusillim

f 273 152 558 395 246

Hydnum

rufescens

f 451 588

283

Hygrophoropsis

aurantiaca

f 494 461 549 336 91

Inocybe

geophylla var.

lilacina

f 411 721 496 281 160 129

Inocybe

lanuginosa

f 472 419 630 143 130

Inocybe napipes f 409 745 495 447 343

Laccaria

amethystina

f 490 423 501 472 226 133

Lactarius

camphoratus

f 485 288 683 536 153 135

Lactarius

detterimus

f 492 400 273 220 566 455 353 130

Lactarius

hepaticus

f 492 353 133 529 455 362 129

Lactarius

quietus

f 491 354 577 325 65

Lactarius rufus f 491 350 145 769 574 288 208 157

Lactarius

subdulcis

f 494 357 146 560 207 150 90

Paxillus

involutus

f 499 395 139 1070 924 91

Russula illiota f 473 146 586 443 340

Russula fragilis f 490 451 501 450 358 130

Russula

ochroleuca

f 493 463 840 536 455 362

Russula

parazurea

f 485 975 856

Tricholoma

album

f 494 465 156 408 163 96

Tricholoma

atrosquamosum

f 498 451 377 177

There were some ECM morphotypes which differed in morphology but

had similar RFLP patterns. An example of this is the morphotypes 3e3 and 3g3.

These morphotypes were classed as being different based on morphology but the

same based on RFLP digests and later turned to be two Russula ochroleuca

ectomycorrhizas. In this case the morphotype 3e3 lacked the distinctive yellow

encrustations that are normally used to distinguish Russula ochroleuca

ectomycorrhizas. In this project the results of sequencing were taken as the most

definite, followed by RFLP matches and then by morphology methods. This

284

means if samples appeared to be different morphologically but matched in their

RFLPs then they were taken to be the same morphotype.

Sequencing and online database comparisons

Poor sequences with small fragment lengths (<500 bases), did not show high

similarity when checked against the online databases (Table 6.2). A failed read is

defined as “when the number of bases is less than 100 after quality clipping is

carried out“ (Eurofins MWG 2011) and is set at this level due to quality controls

defined for the particular method of sequencing. If a sequence has less than 100

bases, then each mismatched base or missing base will reduce the overall

similarity of the match by a percentage point. Thirteen morphotypes were

identified to species level, with a further nine identified to genera. A total of nine

morphotypes from this project have their sequences added to the Genbank

database. These are Russula ochroleuca (HQ703019), Inocybe cincinnata

(HQ703020), Cortinarius obtusus (HQ703021), Tomentella sublilacina

(HQ703022), Lactarius hepaticus (HQ703023), Russula puellaris (HQ703024),

Tomentella badia (HQ703025), Pseudotomentella griseopergamacea

(HQ703026) and Suillus variegatus (HQ703027).

285

Table 6.2 Samples which were sequenced using the primer ITS1-F. All samples were first checked for sufficient (>98%) matches against the UNITE database. If no matches were found then the Genbank database was also searched. In some cases where the accession was locked for editing, the accession number was not available at the time of checking, and the accession number column is therefore marked with na (not available). In cases where the match is not very similar, the accession number is not given. e= sequence from ectomycorrhizal morphotypes, f= sequences from fruitbodies

Sample name Sequence

length

Closest

accession

match

Accession name Similarity

4i5 e 97 UDB001476 Amanita citrina 95%

3c5 e 567 na Amphinema sp. >90% (518

bits, e-148)

3g7 e 591 na Amphinema sp. >90% (601

bits, e-172)

Camarophyllopsis

atropuncta f

615 Camarophyllopsis spp. 91%

3f3 e 664 UDB000127 Cortinarius obtusus 99%

Cortinarius

rubellus f

661 UDB002427 Cortinarius rubellus 99%

Cortinarius

scandens f

142 Cortinarius (Telamonia

sp.)

92%

4d1 e 596 UDB000592 Entoloma serrulatum 98%

Entoloma

conferendum var.

pusillum f

786 GQ397990 Entoloma sp. cf cetratum 94%

3a9 e 824 AM882850 Inocybe cincinnata 99%

4 e3 e 665 UDB000861 Lactarius hepaticus 100%

Cordyceps

forquignoni f

638 AJ786562 Ophiocordyceps

forquignoni

99%

4g3 e 652 UDB001617 Pseudotomentella

griseopergamacea

99%

4d5 e 75 UDB000353 Russula caerulea 91%

Russula fellea f 834 UDB000110 Russula fellea 100%

Russula foetans f 806 DQ422024 Russula illiota 99%

3 e3 e 802 AY254880 Russula ochroleuca 99%

3g3 e 802 AY254880 Russula ochroleuca 99%

5i1 e 810 AY254880 Russula ochroleuca 99%

4a3 e 802 AY061709 Russula puellaris 99%

Russula fragilis f 27 UDB001637 Russula sardonia 84%

4j5 e 670 UDB000664 Suillus variegatus 100%

4b1 e 772 UDB001656 Tomentella badia 99%

3j3 e 719 UDB003349 Tomentella sublilacina 98%

3b5 e 56 Tylospora sp. 95%

There were also mismatches found after sequencing between

morphological identification, RFLP identifications and sequence identification.

Morphotype 3f3 was found to be Cortinarius obtusus according to sequencing of

286

the ITS region but this morphotype did not match the published description of

Agerer (1987-2002). The morphotype 3f3 was brown with a woolly texture.

Cortinarius obtusus in Agerer’s descriptions is white with many rhizomorphs. It is

likely that the morphotype 3f3 is an older sample of Cortinarius obtusus and thus

it has lost its distinctive colour and texture. An example of a mismatch between

RFLP and sequencing is found in the morphotype 4e3. Morphotype 4e3 was not

found to be a match to Lactarius hepaticus (sporocarp) through RFLP analysis but

it matched this species following sequencing and database searches.

Troubleshooting PCR and other molecular methods used.

Common problems encountered when carrying out PCR included non-

amplification of products, streaky products and contamination of genomic DNA

with DNA from multiple species. In order to solve the problem of non-

amplification of products, all extracted genomic DNA was run for 15 mins on an

agarose gel and visualised to identify if genomic DNA was present. In order to

test if the primers were functioning, the annealing temperature was lowered to

55ºC and genomic DNA was amplified. Lowering the annealing temperature

decreases the specificity of the primers and allows amplification of non-specific

products (Mc Phearson and Moller 2000). In order to reduce the streaking of

products, different concentrations of MgCl2 were tested until a satisfactory band

pattern was achieved. The problem of amplification of multiple samples of

genomic DNA was not solved by any methods. The band-stab method (Bjourson

and Cooper 1992) was attempted, but to no avail. As multiple ectomycorrhizas

can occupy the same root segment (Bruns 1995); amplification of multiple

products, or no products at all can be a common result. Effective sorting and

removal of extra ECM types was the only effective solution found for this

problem.

6.4.2 ECM richness and abundance in the forest types

Overall, 51 ECM morphotypes were collected from 12 forest plots of three forest

types (Table 6.4). Twenty-one ECM types were found in oak, 20 in Scot’s pine

and 18 in Sitka spruce. Two species, Cenococcum geophilum and Russula

287

ochroleuca, were common to all forest types, two species and two morphotypes,

Laccaria amethystina, L. laccata, type 5e7 and Piceirhiza nigra, were common to

two of the forest types (the first three morphotypes were shared between oak and

Scot’s pine while the last morphotype was shared between Scot’s pine and Sitka

spruce). The remaining morphotypes were found in only one forest type. The five

most abundant types were Cenococcum geophilum, 5e7, Piceirhiza horti-inflata,

Pseudotomentella griseopergamacea, and 4h5, with mean % frequencies of 44,

13, 12, 8 and 8% across all forest types. Of the 9 morphotypes identified by

sequencing: Inocybe cincinnata, Russula puellaris and Entoloma serrulatum do

not have published descriptions in either Agerer (1987-2002) or BCERN and are

given above in Section 6.4.1.

The relative frequency of the different ECM types was calculated for each

forest type and rank-abundance curves were constructed. Visual examination of

the rank abundance curves suggests that the ECM communities from all of the

forest types follow a log-normal distribution. This indicates that these

communities have a few species which are very abundant, and a long tail of rarer

species (Figs. 6.2a, b). Kolmogorov-Smirnov two-sample tests (Magurran 2004)

found that the rank abundance curves for each forest type were not significantly

different from each other (Table 6.3).

Table 6.3 Comparison of rank-abundance distributions from each forest type using Kolmogorov-Smirnov two-sample test. Columns indicate which forest types were compared. n1= number of samples in forest type 1, n2= number of samples in forest type 2. D= largest unsigned difference between the two relative cumulative frequency distributions from the two forest types. D0.05= Critical significance value for two sample Kolmogorov- Smirnov test calculated for that number of samples. SP= Scot’s pine, SS= Sitka spruce. ns= not significantly different from each other.

OAK-SP OAK-SS SP-SS

n1 21 21 24 n2 24 18 18 D 47.6784 50.9166 36.504 D0.05 198 159 180 Significantly different ns ns ns

288

Table 6.4 ECM types present in the sites and their % frequency in each site. For site codes see Table 3.2, Chapter 3. The final two rows list the total morphotype richness and total morphotype abundance (appearance of morphotype in all 30 sub-samples) of ECM types in the sites. OAK SP SS

ABBEY

KILMAC

RAHEEN TOOMI

ANNAGH BNSHA BRITT TORC

BOHAT CHEVM CHEVY

DOOAR

3 a7 0 0 50 0 0 0 0 0 0 0 0 0

3 b7 0 0 10 0 0 0 0 0 0 0 0 0

3 c7 0 0 3.3 0 0 0 0 0 0 0 0 0

3 c9 0 0 0 0 0 0 0 0 0 0 13.3 0

3 e7 0 0 3.3 0 0 0 0 0 0 0 0 0

3 e9 0 0 0 0 0 0 0 0 0 0 16. 7 0

4 c1 0 0 0 0 0 0 10 0 0 0 0 0

4 c3 0 0 0 0 0 0 13.3 0 0 0 0 0

4 e1 0 0 0 0 0 0 40 0 0 0 0 0

4 f9 0 0 0 0 0 0 0 0 0 0 0 3.3

4 h5 0 0 0 0 0 46.7 0 46.7 0 0 0 0

4 j7 6.7 0 0 0 0 0 0 0 0 0 0 0

5 a7 0 0 0 3.3 0 0 0 0 0 0 0 0

5 c9 0 0 0 0 0 0 0 40 0 0 0 0

5 e7 13.3 0 43.3 30 0 26.7 40 0 0 0 0 0

5 i7 0 0 0 20 0 0 0 0 0 0 0 0

Amanita citrina 0 53.3 0 0 0 0 0 0 0 0 0 0

Amanita spp.2 0 0 0 0 0 0 0 0 0 0 20 0

Amanita spp.3 0 0 0 0 0 53.3 0 0 0 0 0 0 Amphinema spp. 1 0 0 0 0 0 0 0 0 3.3 0 0 0 Amphinema spp. 2 0 0 0 0 0 0 0 0 0 23.3 0 0

289

Cantharellus cibarius 0 0 0 0 0 0 0 0 0 0 33.3 0 Cantharellus tubaeformis 0 0 0 0 0 0 0 10 0 0 0 0 Cenococcum geophilum 10 76.7 83.3 60 0 73.3 50 76.7 10 0 80 3.3 Cortinarius evernius 0 0 0 0 0 0 0 0 16.7 0 0 0 Cortinarius obtusus 0 0 0 0 0 0 0 0 13.3 0 6.7 0 Cortinarius rubellus 0 0 0 0 0 0 0 0 20 0 0 0

Cortinarius spp.4 0 0 16.7 0 0 0 0 0 0 0 0 0 Elaphomyces granulatus 3.33 80 0 0 0 0 0 0 0 0 0 0 Entoloma serrulatum 0 0 0 0 0 0 10 0 0 0 0 0

Genea spp.1 16.7 0 0 0 0 0 0 0 0 0 0 0 Inocybe cincinnata 0 0 0 0 0 0 0 0 0 33.3 0 0 Laccaria amethystina 6.7 13.3 0 6.7 0 0 6.7 0 0 0 0 0

Laccaria laccata 0 0 0 33.3 26.7 0 0 0 0 0 0 0 Lactarius hepaticus 0 0 0 0 66.7 0 0 0 0 0 0 0

Lactarius spp.2 0 0 0 0 0 0 0 0 0 0 73.3 0

Paxillus involutus 0 0 0 10 0 0 0 0 0 0 0 0 Piceirhiza horti-inflata 0 0 0 0 0 0 0 0 0 43.3 3.3 93.3

Piceirhiza nigra 0 0 0 0 0 13. 3 0 0 0 20 10 0 Pseudotomentella griseopergamacea

0 0 0 0 80 0 0 20 0 0 0 0

Russula nobilis 0 10 0 0 0 0 0 0 0 0 0 0

290

Russula ochroleuca 0 10 0 0 0 3.3 0 0 23.3 0 0 0

Russula puellaris 0 0 0 0 3.3 0 30 0 0 0 0 0

Russula spp.4 0 0 0 0 0 0 0 20 0 0 0 0 Scleroderma citrinum 6.7 0 0 0 0 0 0 0 0 0 0 0

Suillus variegatus 0 0 0 0 0 23.3 23.3 0 0 0 0 0

Tomentella badia 0 0 0 0 0 0 10 0 0 0 0 0

Tomentella spp.3 0 33.3 0 0 0 0 0 0 0 0 0 0

Tomentella spp.4 0 0 6.7 0 0 0 0 0 0 0 0 0 Tomentella sublilacina 0 0 0 0 0 0 0 0 20 0 0 0 Tylospora spp. 0 0 0 0 0 0 0 0 36.7 0 0 0

Morphotype richness 7 7 8 7 4 7 10 6 8 4 9 3

Morphotype abundance 19 83 65 49 53 72 70 64 43 36 77 30

291

0 2 4 6 8 10 12 14 16 18 20 220.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

OakSPSS

Cenococcum geophilum5e7Elaphomyces granulatusAmanita citrina

3a7Cenococcum geophilum

Pseudotomentella griseopergamacea4h55e7Lactarius hepaticusPiceirhiza horti-inflata

Cenococcum geophilumLactarius spp.2Tylospora spp.2

Cantharellus cibarius

Rank

Rela

tive f

req

uen

cy

0 10 20 30 40 500.00

0.05

0.10

0.15

0.20

0.25

Cenococcum geophilum5e7Piceirhiza horti-inflataPseudotomentella griseopergamacea

4h5

Rank

Rela

tive f

req

uen

cy

(a)

(b)

Figs 6.2 (a & b). Rank abundance plots of the ECM morphotypes from each forest type along with the five most abundant morphotypes from each forest type. (a) Plots of each forest type investigated, Green squares and text= oak forest type, orange triangles and text= Scot’s pine (SP) forest type and brown inverted triangles and text= Sitka spruce (SS) forest type. (b) Rank abundance plot of the ECM morphotypes identified from all of the forest types.

292

293

Differences in morphotype richness and abundance between the different forest

types.

Using ANOVA and Tukey’s post-hoc test, there were no significant differences in

ECM morphotype richness between the forest types in (Fig. 6.3). Significant

differences were found in the morphotype abundance of between the different

forest types (Fig.6.4). Scot’s pine had a significantly higher (P<0.05) abundance

of ECM morphotypes per core than the Sitka spruce forest type (Fig. 6.4).

Fig. 6.3 Mean ECM morphotype richness per soil core from the different forest types. Mean, lower and upper quartile are displayed along with highest records (Circles). Means are not significantly different (P>0.05).

Fig. 6.4 Mean ECM morphotype abundance per soil core from the different forest types. Mean, lower and upper quartile are displayed. Box plots with the same letter have means which are not significantly different at P<0.05.

294

Fig. 6.5 Mean number of ECM types per soil core in the forest plots. The first four bars are oak sites, bars 5-8 are Scot’s pine sites and bars 9-12 are Sitka spruce sites. Error bars = standard deviation.

Fig. 6.6 Mean morphotype abundance per soil core in the forest plots. The first four bars are oak sites, bars 5-8 are Scot’s pine sites and bars 9-12 are Sitka spruce sites. Error bars = standard deviation.

295

ECM diversity

According to the Simpson’s diversity index, Scot’s pine was the most diverse in

terms of ECM diversity with an index values of 9.77, Sitka spruce was the second

most species-diverse with 9.67 and oak was the least diverse with values of 6.9

(Fig. 6.7).

OAK SP SS0.00

0.25

0.50

0.75

1.00

Simpsons 1/DE1/D

6

7

8

9

10

Forest type

Ind

ex v

alu

e

Fig. 6.7 Bar chart showing the values for the Simpson’s diversity and evenness index for the different forest types based on their ECM morphotype communities. SP= Scot’s pine, SS= Sitka spruce. E1/D= Simpson’s evenness index. Note break in y-axis from values 1-6.

Comparison of ECM richness and estimated richness at similar sampling

intensities

As the amount of sampling was not equal between the forest types (based on

length of roots sampled), sample based rarefaction was used to equate the

sampling intensity between the forest types. The effect of sample size on

morphotype richness can be seen from the morphotype accumulative curve (Fig.

6.8). Rarefaction was used to find the number of ECM types at a common sample

size. At 13,700 cm of root length (the lowest amount sampled for a tree type: oak)

the rarefied mean (and 95% lower and upper confidence bounds) number of ECM

types was 18 (±7), 15 (±6) and 16 (±7) for oak, Scot’s pine and Sitka spruce

forests (Fig. 6.9). This compares with the actual morphotype richness of the oak,

296

Scot’s pine and Sitka spruce forest types which was 21, 20 and 18 morphotypes

respectively. Visual examination of the curves and 95% confidence intervals

shows that the number of ECM morphotypes at a standard sample size is not

significantly different (P<0.05; Fig. 6.9).

To investigate if sufficient sampling had taken place to saturate the species

accumulation curve, species richness estimators were calculated for the different

forest types. The ACE, ICE and CHAO2 estimated morphotype richness curves

were calculated (Fig. 6.10 a,b,c). The Chao2 and the ICE estimators reached an

asymptote for all forest types before the ACE estimator did. The Chao2 and ICE

estimators reached a stable estimate after 5 cores (25% sampling). The ability of

an estimator to reach a stable estimate before 100% sampling is one of the main

traits looked for in a suitable estimator of species richness for a taxonomic group

(Magurran 2004). Taking these estimates as the lowest estimate of ECM richness,

the higher estimator (ICE) was used to estimate the possible morphotype richness

of the forest types. Oak is estimated to have 20 (±5) morphotypes, Scot’s pine

may have 25 (±10) ECM types while Sitka spruce is estimated to have 20 (±5)

morphotypes. In reality oak had 21 types, Scot’s pine had 20 types and Sitka

spruce had 18 types. This represents between 84-100%, 57-100%, and 71-100%

of the morphotype richness discovered in oak, Scot’s pine and Sitka spruce forests

respectively.

297

0 10 20 30 40 50 600

5

10

15

20

25

30

35

40

45

50

55

No. of Cores

EC

M m

orp

ho

typ

es

Fig. 6.8 Morphotype accumulation curves for all of the forests types. Black staggered line is the actual cumulative species numbers for the tree type as samples increased. Blue symbols are the sample based rarefaction curve for the cumulative species from all of the forest types combined with whiskers indicating the 95% upper and lower confidence bounds. Plots were sampled randomly with replacement using 1000 permutations.

298

1373

10

2500

5000

7500

1000

0

1250

0

1500

0

1750

0

2000

0

1373

1

0

5

10

15

20

25

OAKSPSS

Root lenght (cm)

No

. o

f m

orp

ho

typ

es

Fig. 6.9 Sample based rarefaction curves of mean (symbols) morphotypes with 95% upper and lower confidence intervals (whiskers) for oak (triangle symbols), SP= Scot’s pine (cross symbols) and SS= Sitka spruce (Circular symbols). Cores were sampled randomly with replacement using 500 permutations for each sample size.

299

OAK

0 5 10 15 200

5

10

15

20

25

30

35

40

45

EC

M m

orp

ho

typ

es

SP

0 5 10 15 200

10

20

30

40

EC

M m

orp

ho

typ

es

SS

0 5 10 15 200

10

20

30

40

ACEICECHAO2

Sample

EC

M m

orp

ho

typ

es

(a)

(b)

(c)

Fig. 6.10( a, b, c). Mean morphotype richness estimators (and standard deviations: dotted lines) for the number of ECM types in (a) oak, (b) Scot’s pine and (c) Sitka spruce forests. Triangle symbols = Abundance Based coverage Estimator ACE, cross symbols= Incidence Based coverage Estimator ICE and circle symbols = Chao2 estimator CHAO2. Cores were sampled randomly with replacement using 500 permutations for each sample size.

300

The species richness estimator curves above all show levelling off of the

species accumulation curve. This would indicate that sufficient sampling had

taken place to estimate the number of ECM types in the forest sites.

6.4.3 Similarity of below-ground ECM assemblage and above-ground ECM

sporocarp assemblage

Using a modified form of the Jaccard similarity index, the similarity of the above-

ground ECM sporocarps and the below-ground ECM morphotypes was examined

(Tables 6.5; 6.6; 6.7). The similarity between the above- and below-ground ECM

assemblages was low. The 4 oak plots had a mean similarity value of 12±9%, the

Scot’s pine sites 7±1%, and the Sitka spruce sites 9±14% similarity (Table 6.8).

Over the entire project there was less than 10% similarity found between the ECM

fruitbodies and the ECM morphotypes in the 12 sites examined.

Spearman’s rank correlation was used to test for significant relationships

between the below-ground richness and abundance values and the above-ground

sporocarp richness and abundance values. No significant relationships (P>0.05)

found between these variables from the data from all of the forest types.

The abundance of certain sporocarps was in some cases useful in deciding

what species the ECM root tip belonged to. For example, the white “frosty”

morphotype recorded in the Abbeyleix oak plot was similar to Scleroderma

ectomycorrhizas described by Agerer (1987-2002). As Scleroderma citrinum was

also found as a sporocarp in this site, the presence of the fungal fruitbody helped

to indicate possible identities for the ECM morphotype. In other cases however,

many of the species found as sporocarps were not found belowground, and vice

versa for belowground species which were not found fruiting as sporocarps.

301

Table 6.5 Species and ECM morphotype list from the oak sites. The second row lists the macrofungal species found above-ground which were not found below-ground. The third row lists the species which were found above and below-ground while the third row lists the morphotypes which were not found to have matching sporocarps above-ground. OAK Abbeyleix Kilmacrea Raheen Tomies

Above-

ground only

Boletus chrysenteron

Inocybe sericatum

Laccaria lacata

Lactarius quietus

Lactarius tabidus

Russula aeruginea

Russula atropurpurea

Russula foetens

Russula ochroleuca

Russula parazurea

Boletus badius

Cortinarius acutus

Cortinarius flexipes

Entoloma conferendum

Inocybe lanuginosa

Laccaria laccata

Lactarius pyrogalus

Lactarius quietus

Lactarius rufus

Lactarius vietus

Scleroderma areolatum

Scleroderma citrinum

Amanita citrina

Amanita phalloides

Amanita rubescens

Cantharellus cibarius

Cortinarius acutus

Cortinarius bolaris

Cortinarius flexipes

Cortinarius sanguineus

Cortinarius stillatitius

Entoloma conferendum

Laccaria amethystina

Laccaria laccata

Lactarius subdulcis

Russula aeruginea

Russula foetens

Russula nobilis

Tricholoma album

Tricholoma columbetta

Boletus badius

Cortinarius acutus

Cortinarius flexipes

Cortinarius flexipes var.

inolens

Cortinarius stillatitius

Cortinarius umbrinolens

Elaphomyces granulatus

Inocybe napipes

Lactarius camphoratus

Lactarius hepaticus

Lactarius quietus

Lactarius tabidus

Russula betulinum

Russula ochroleuca

Above and

below-ground

Elaphomyces granulatus,

Laccaria amethystina,

Scleroderma citrinum

Elaphomyces granulatus, Laccaria

amethystina, Russula nobilis, Russula

ochroleuca

Laccaria amethystina,

Laccaria laccata

Below-

ground only

Genea spp.1 5 e7 Cenococcum geophilum

4 j7

Amanita citrina

Cenococcum geophilum

Tomentella spp.3

3 a7 3 b7 3 c7 3 e7 5 e7 Cenococcum geophilum

Cortinarius spp.4 Tomentella spp.4

5 a7 5 e7 Russula nobilis

Cenococcum geophilum

Paxillus involutus

302

Table 6.6 Species and ECM morphotype list from the Scot’s pine sites. The second row lists the macrofungal species found above-ground which were not found below-ground. The third row lists the species which were found above and below-ground while the third row lists the morphotypes which were not found to have matching sporocarps above-ground. SP Annagh Bansha Brittas Torc

Above-ground only Amanita fulva

Cortinarius flexipes

Cortinarius umbrinolens

Hygrophoropsis aurantiaca

Inocybe lanuginosa

Lactarius quietus

Lactarius rufus

Lactarius tabidus

Paxillus involutus

Russula ochroleuca

Entoloma cetratum

Entoloma conferendum

Laccaria laccata

Laccaria amethystina

Lactarius quietus

Amanita rubescens

Cortinarius acutus

Entoloma conferendum

Hydnum repandum

Inocybe geophylla

Laccaria laccata

Lactarius camphoratus

Lactarius tabidus

Cantharellus cibarius

Cortinarius acutus

Cortinarius anomalous

Cortinarius semisanguineus

Cortinarius stillatitius

Elaphomyces granulatus

Hydnum repandum

Hydnum rufescens

Russula fellea

Russula fragilis

Above and below-

ground

Lactarius hepaticus Russula ochroleuca Laccaria amethystina Cantharellus tubaeformis

Below-ground only Laccaria laccata

Pseudotomentella-

griseopergamacea

Russula puellaris

4 b7 4 h5 5 e7 Amanita spp.3 Cenococcum geophilum

Suillus variegatus

4 c1 4 c3 4 e1 5 e7 Cenococcum geophilum

Entoloma serrulatum

Russula puellaris

Suillus variegatus

Tomentella badia

5 c9 5 e9 Cenococcum geophilum

Pseudotomentella

griseopergamacea

Russula spp.4

303

Table 6.7 Species and ECM morphotype list from the Sitka spruce sites. The second row lists the macrofungal species found above-ground which were not found below-ground. The third row lists the species which were found above and below-ground while the third row lists the morphotypes which were not found to have matching sporocarps above-ground. SS Bohatch Chevy chase M Chevy chase Y Dooary

Above-

ground

only

Cortinarius flexipes

Entoloma cetratum

Inocybe lanuginosa

Laccaria amethystina

Laccaria laccata

Russula nobilis

Boletus

subtomemtosus

Inocybe geophylla

Inocybe geophylla

var. lilacina

Inocybe rimosa

Laccaria amethystina

Laccaria laccata

Amanita fulva

Amanita rubescens

Cortinarius acutus

Cortinarius cinnamomeus

Cortinarius flexipes

Cortinarius scandens

Cortinarius umbrinolens

Elaphomyces granulatus

Laccaria amethystina

Laccaria laccata

Lactarius deliciosus

Lactarius tabidus

Russula fragilis

Russula nigricans

Russula ochroleuca

Cortinarius acutus

Cortinarius cinnamomeus

Cortinarius evernius

Cortinarius obtusus

Cortinarius venetus

Entoloma cetratum

Laccaria amethystina

Laccaria laccata

Lactarius deterrimus

Russula queletii

Above and

below-

ground

Cortinarius obtusus,

Cortinarius rubellus,

Cortinarius evernius, Russula

ochroleuca

Cantharellus cibarius,

Cortinarius obtusus

Below-

ground

only

Amphinema spp. 1 Cenococcum geophilum

Tomentella sublilacina

Tylospora spp.

Piceirhiza nigra

Amphinema spp. 2 Inocybe cincinnata

Piceirhiza horti-

inflata

Piceirhiza conspicua

3 c9 3 e9 Piceirhiza nigra

Amphinema spp. 2 Cenococcum geophilum

Piceirhiza horti-inflata

4 f9 Cenococcum geophilum

Piceirhiza horti-inflata

304

Table 6.8 Jaccard similarity (%) between above- and below-ground ECM communities in the different plots and forest types. Each plot is listed along with the similarity value (in parenthesis). The final column lists the mean (and standard deviation) Jaccard value for each forest type. The key to the plot abbreviations used is found in Table 3.2, Section 3.3.1.

Forest

type

Jaccard similarity (%) Mean

(±SD)

Oak Abbey (18) Kilmac(21) Raheen (0) Tomies (10) 12 (±9)

Scot’s pine Anna (7) Bansha (8) Britt (6) Torc (6) 7 (±1)

Sitka

spruce

Bohat (29) Chev M (0) ChevY (8) Dooar (0) 9 (±14)

A Mantel test was carried out to determine if the similarity matrices of the

above-ground and below-ground ECM assemblages (based on the Sørenson

index), were significantly related to one another. The Monte Carlo test showed

that the relationship between the two matrixes was not due to chance (P= 0.023).

The effect size r (according to the standardized Mantel statistic) was 0.48. This

result shows that the above-ground and below-ground ECM assemblages

differentiate the sites in a similar fashion which would indicate that there are

similar forest specific assemblages/communities based on above- and below-

ground aspects of the ECM community.

6.4.4 ECM community analysis

In order to test the hypothesis that ECM communities will be most similar

between plots of the same forest type, the Jaccard abundance-based index was

calculated for each pair of plots (Table 6.9). Mean and standard deviations of all

of the possible forest type pair combinations were calculated from pooled site data

(Table 6.10). It was found that plots of the same forest type were on average more

similar in community composition than plots of a different forest type (Table

6.10). Of all of the possible paired combinations of forest types analysed, the oak

plots had the highest average similarity (Table 6.10). Scot’s pine and Sitka spruce

plots had the lowest average similarity value (Table 6.10). Annagh and Chevy

Chase (Mature) had the lowest average similarity values overall. Both of these

sites shared very few species with any other forest type (Table 6.9).

305

Table 6.9 Jaccard estimated similarity index between the different plots (shaded cells) and the number of shared morphotypes in the clear cells. SP= Scot’s pine, SS= Sitka spruce.

Tree type OAK OAK OAK OAK SP SP SP SP SS SS SS SS

Site ABBEY KILMAC RAHEEN TOOMI ANNAGH BNSHA BRITT TORC BOHAT CHEVM CHEVY DOOAR

OAK ABBEY X 0.416 0.25 0.36 0 0.24 0.35 0.15 0.09 0 0.12 0.11

OAK KILMAC 3 X 0.13 0.21 0 0.19 0.18 0.16 0.21 0 0.12 0.11

OAK RAHEEN 2 1 X 0.26 0 0.24 0.26 0.18 0.1 0 0.13 0.11

OAK TOOMI 3 2 2 X 0.1 0.25 0.35 0.21 0.11 0 0.15 0.12

SP ANNAGH 0 0 0 1 X 0 0.06 0.07 0 0 0 0

SP BNSHA 2 2 2 2 0 X 0.33 0.37 0.23 0.08 0.19 0.12

SP BRITT 3 2 2 3 1 3 X 0.17 0.1 0 0.13 0.11

SP TORC 1 1 1 1 1 2 1 X 0.12 0 0.16 0.13

SS BOHAT 1 2 1 1 0 2 1 1 X 0 0.18 0.09

SS CHEVM 0 0 0 0 0 1 0 0 0 X 0.09 0.26

SS CHEVY 1 1 1 1 0 2 1 1 2 2 X 0.35

SS DOOAR 1 1 1 1 0 1 1 1 1 1 2 X

306

307

Table 6.10 Mean and standard deviations (in parenthesis) of JI index (shaded cells) between the ECM communities of the different forest types. The numbers of shared morphotypes between the different forest types are also given (clear cells).

Oak Scot’s pine Sitka spruce

Oak 0.27 (±0.10) 0.18 (±0.11) 0.09 (±0.06)

Scot’s pine 5 0.17 (±0.15) 0.08 (±0.08)

Sitka spruce 2 3 0.16 (±0.13)

Within-plot variation

Jaccard similarity was also calculated between the five individual cores collected

from each plot in order to get an estimate of the homogeneity of the ECM

assemblages in individual plots (Table 6.11). The Sitka spruce plots showed the

highest within-plot similarity (0.53) while having the lowest mean number of

ECM morphotypes per plot (6). Oak plots in contrast had the lowest within-plot

similarity (0.40), while having the highest mean number of ECM morphotypes per

plot (7.3). The corresponding values for Scot’s pine sites were 0.46 and 6.8.

These results suggest the distribution of ECM types in the forests types were

patchier in the oak plots than in the other forest types.

Table 6.11 Morphotype richness (sum of the 5 soil cores) and the within-plot similarity for each forest site. The within-plot similarity is the degree of similarity between the 5 sample cores taken from a plot in respect of the ECM types present. The values in column 3 are the mean Jaccard index values with the standard deviations in parenthesis.

Site Morphotype richness Within site JI

Oak

Abbeyleix 7 0.09 (0.17)

Kilmacrea 7 0.47 (0.24)

Raheen 8 0.62 (0.26)

Tomies 7 0.43 (0.21)

Scot’s pine

Annagh 4 0.66 (0.36)

Bansha 7 0.45 (0.16)

Brittas 10 0.36 (0.23)

Torc 6 0.37 (0.29)

Sitka spruce

Bohatch 8 0.15 (0.23)

Chevy chase M 4 0.5 (0.39)

Chevy chase Y 9 0.61 (0.27)

Dooary 3 0.87 (0.1)

308

NMS ordination of ECM communities

The NMS using the % frequency abundance data for each plot failed to produce a

statistically significant ordination. The large percentage of zero values in the data

set was a likely reason for the failed ordination, and so Beal’s smoothing was

carried out on the % frequency of ECM types matrix following the advice of

McCune and Grace (2002) in order to lessen the effect of zeros on the ordination.

Following this data modification, NMS found a three-dimensional solution (Figs.

6.11, 6.12, 6.13) with a final stress of 5.06. A Monte Carlo test showed that the

probability of a similar final stress being obtained by chance was 0.00001. Axis 1

represented 38.7% of the variance with a further 25.3% of the variance

represented by axis 2. Axis 3 represented the remaining 28.5% of the variance to

give the ordination a cumulative r2 value of 92.6%. Nine correlations were found

between the three ordination axes and 12 environmental variables (Table 6.12).

Table 6.12 Pearson’s correlations (rp) for ordination axes, environmental variables and species richness. %OM= organic matter, precipitation= mean annual precipitation, N total = total soil available nitrogen, NO3= total soil available nitrate, NH4= soil available ammonium, CWD total= total volume of coarse woody debris in m2/ha, ns= non-significant correlation.

Axis 1 Axis 2 Axis 3

pH -0.68 ns ns

Precipitation 0.63 ns ns

NO3 ns ns -0.486

NH4 ns ns -0.487

Calcium -0.45 ns 0.475

Potassium ns -0.7 ns

CWD total ns ns ns

Phosphorus ns -0.43 ns

N- Total ns ns -0.6

309

Fig. 6.11 NMS ordination of axis 1 and 2 (cumulative r2= 0.64) showing the plots identified by the tree type of the site and the ordinations of the ECM morphotypes. SP = Scot’s pine, SS= Sitka spruce.

310

311

Fig. 6.12 NMS ordination of axis 3 and 2 (cumulative r2= 0.54) showing the plots identified by the forest type of the site and a biplot of environmental and physical variables. SP = Scot’s pine, SSa= Sitka spruce young age group, SSb= Sitka spruce mature age group. N-Tot= available soil nitrogen, N-NO3= available soil nitrate nitrogen, N-NH4= available soil ammonium nitrogen, Ca= available soil calcium, K= available soil potassium.

312

313

Fig. 6.13 NMS ordination of axis 3 and 1 (cumulative r2= 0.67) showing the plots identified by the forest type of the plot and a biplot of environmental and physical variables. SP = Scot’s pine, SSa= Sitka spruce young age group, SSb= Sitka spruce mature age group. N-Tot= available soil nitrogen, N-NO3= available soil nitrate nitrogen, N-NH4= available soil ammonium nitrogen, Ca= available soil calcium, and precipitation= mean annual rainfall.

314

315

The NMS ordinations of the ECM communities separated the plots into

groups that corresponded well to forest type, particularly axis 2 and 3 of the

ordination. MRPP analysis of the ECM communities in each forest type found

that the ECM communities did differ significantly between plots based on the

main dominant type of the plot and on the vegetation communities present at the

sites (A= 0.13, P< 0.05). These results indicate that distinct ECM communities are

found in each forest type. The oak plots were ordinated along a gradient of

increasing soil nutrient status (nitrogen, potassium and phosphorus) from the

coniferous plots.

The distribution of the different ECM morphotypes amongst the plots was

not equal. With many morphotypes only present in single plots (Fig. 6.11) and

others present in very large and very low abundances over the plots, correlations

between individual species and ordination axes are difficult to explain. One

species which was present in a large proportion of the plots and samples was

Cenococcum geophilum. It showed a strong negative (rp-0.58) correlation to axis

3. Axis 3 was also negatively correlated with available nitrogen and positively

correlated with available calcium (Table 6.12). This could indicate that the

abundance of C. geophilum morphotype is positively related to the amounts of

available nitrogen in the plot and negatively related to the available calcium in the

plots.

316

6.5 Discussion

6.5.1 Ectomycorrhizal diversity

The number of ECM types found in this study (51) is comparable to other studies

utilising both morphological and molecular methods for identification. Peter et al.

(2001) collected 79 morphotypes in a three-year study of ECM communities of

Norway spruce in Switzerland. Taylor (2002) identified 37 ECM types in his

Scot’s pine plots in Sweden. It was found that oak forests yielded the most

morphotypes, followed by Scot’s pine and Sitka spruce, but the mean numbers of

morphotypes per core were not significantly different between the forest types.

Oak forests revealed the highest number of unidentified morphotypes, followed

by Scot’s pine and Sitka spruce. This unequal distribution of unidentified

morphotypes between the forest types, is most likely due to the inequality in the

numbers of published descriptions of mycorrhizas between the different forest

types examined. For example, Agerer and Rambold (2004-2011) give 554

descriptions, of which 121 are of Picea ectomycorrhizas, 100 are of Pinus

ectomycorrhizas and only 74 are from Quercus ectomycorrhizas.

The species richness estimation identified the Scot’s pine forest type as

being potentially the most morphotype rich. This result agrees with the results

from the sporocarp study (Chapter 4), as Scot’s pine plots were found to have the

greatest proportion of ECM sporocarps of all the forest types. Scot’s pine forests

are known to be very ECM morphotype rich, as studies of the effect of fire on

ECM communities in Sweden revealed 135 morphotypes from four forest stands

(Jonsson et al. 1999). Species richness estimation revealed that between 57 and

100% of the ECM morphotype richness was found in these forest types. These

results are broadly similar to those found by Tedersoo et al. (2006), while

examining the ECM community in a tree species rich wooded meadow in Estonia.

They found that Chao2 estimates revealed that a possible 52% of the morphotype

richness was recorded by their study. However, their estimates were much higher

than in this study (325 vs. max 35 in this study). The higher actual and estimated

species richness in Tedersoo et al. study is not surprising, as the positive

relationship between tree species richness and ECM richness (Ishida et al. 2007),

317

and the positive effect of multi-species neighbouring trees on ECM diversity

(Hubert and Gehring 2008) has been recognised.

The fact that sample-based rarefaction showed no significant differences

between the number of morphotypes from each forest type, shows that plantation

forests of Sitka spruce and Scot’s pine can have similar levels of ECM richness

below-ground as the native oak woods of Ireland. The species richness estimation

also showed that the estimated number of morphotypes in these forest types are

comparable. A similar result has been found for macrofungal species richness

according to the sporocarp survey (Fig. 4.4, Section 4.4.2, Chapter 4).

The overall community structure with regard to the abundance of species

is similar to other studies of ectomycorrhizas on the roots in forests soils. The

rank abundance curves follow a log-normal distribution, which agrees with

previous studies in coniferous (Krannabetter 2004; Luoma et al. 2006a; Cline et

al. 2005) and deciduous (Courty et al. 2008) forests. This distribution is due to the

samples being dominated by a few ECM species with very high abundances, and

having a long “tail” of rare species with low abundances. Cenococcum geophilum

was a ubiquitous member of the ECM communities of all forest types in this

study. Many other studies have found that it is often the most abundant

ectomycorrhizal species below-ground in temperate forests (Molina and Trappe

1982; Goodman and Trofymow 1998; Luoma et al. 2006a; Ishida et al. 2007).

NMS and MRPP analysis showed that each forest type had a reasonably

specific ECM community, thus agreeing with previous studies identifying

distinctive ECM communities according to forest type in temperate forests

(Tedersoo et al. 2006; Ishida et al. 2007). There was considerably less within-plot

than between-plot variation. Two (one oak and one Sitka spruce) of the plots had

relatively low within-plot similarity, and this may have been caused by a number

of factors, but most likely by either:

(1) The presence of different niches within the site

(2) The functional morphology of some ectomycorrhizas.

318

(1) Bruns (1995) was the first to describe how niche partitioning may be one of

the reasons for the large diversity of ECM types in forest soils. This

hypothesis was later tested and added to by Dickie et al. (2002), Tedersoo et

al. (2003) and Buee et al. (2007), as it was shown that certain

ectomycorrhizas are more common on certain substrates (e.g. Thelephora

spp. on coarse woody debris).

(2) The functional morphology of ectomycorrhizas may also cause large variation

between ECM communities within plots. Agerer (2001) noted that many

ECM species produce copious amounts of hyphal strands, which could aid in

the exclusion of other ectomycorrhizas from the surrounding root systems.

These competitive interactions of many ECM fungi have been reviewed by

Kennedy (2010), and the author stresses that these competitive interactions

are environmentally context-dependant, and so the results of these

interactions in nature can be highly stochastic. The study of ECM

morphology has been reviewed by Peay et al. (2011), and the current

hypothesis is that it is the amounts of roots which determine the success of

the different ectomycorrhizas depending on their exploration type

morphology. Morphotypes with long range exploration types (i.e. Cortinarius

spp. with long emanating hyphae) are more effective at colonizing roots in the

low density root zone by “piggy-backing” to new roots using their long range

hyphae and already colonized roots as an exploration platform.

6.5.2 The ECM communities of the forest types

The oak ECM community

In this study oak had the highest within-plot Jaccard similarity values and were

also the most undisturbed forest type investigated. The long history of all of the

oak sites as oak woodlands would allow time for the formation of a particular

assemblage of ECM species on the oak roots. It is known from ECM sporocarp

studies in coniferous forests of British Columbia, that ECM communities of

mature stands are relatively stable in comparison to young stands and change very

little as the forest ages (Kranabetter et al. 2005).

319

The most common ECM morphotype in the oak sites was Cenococcum

geophilum. It was found in all of the oak plots, with abundances ranging from 10

to 83% of the samples from the different oak sites. Richard et al. (2005) also

found that the ECM community of Holm oak Quercus ilex woods was dominated

by Cenococcum geophilum. Indeed, C. geophilum has been found to be the most

abundant ECM type in many Quercus forests e.g. Q. petraea in France (Buee et

al. 2007), Q. garryana in Oregon (Valentine et al. 2004), Q.rubra/Q. prinus in

North America and Germany (Walker et al. 2005; Gebhardt et al. 2007) and Q.

douglasii/Q. wislizeni forests in California (Morris et al. 2008).

The next most common ECM types in this study were types 5e7 and

Elaphomyces granulatus. Type 5 e7 resembles ectomycorrhizas of the Russula or

Tomentella type, as it lacks emanating hyphae and its main colours consist of dark

earthy shades. Ectomycorrhizas of these genera were very common in the study of

Q. petraea woods by Buee et al. (2007). Elaphomyces granulatus is probably very

common in Irish oak sites, as sporocarps of the fungus have been found in three of

the five oak sites examined in this study (see Chapter 4). Elaphomyces species

were also found in the oak woodland study by Courty et al. (2008), and this genus

along with other hypogeous genera are known to have spatially restricted, but

highly structured genet sizes (Kretzer et al. 2004). Low spore dispersal ability, but

high spore concentrations where they are found mean that hypogeous species of

fungi can compete effectively with other species of ectomycorrhizas that have

high dispersal rates (Grubisha et al. 2007). This explains the very high abundance

of Elaphomyces granulatus at one oak site (80% of samples from Kilmacrea oak

site), and the total lack of the species in another oak site (Raheen). The coupling

of low dispersal, high spore counts per unit area and longevity of the spores of

many hypogeous fungi (Bruns et al. 2009), could make the presence of hypogeous

species in Irish woodlands a good indicator of ancient woodlands on a site.

The Scot’s pine ECM community

The Scot’s pine sites, Bansha, Brittas and Torc, were similar regarding their ECM

assemblages, while the plot at Annagh had less in common with the other pine

plots regarding its ECM community. Annagh is a first rotation Scot’s pine site

planted on a site which was previously a bog. The past land use of the forest has

320

large effects on the ectomycorrhizal community present, as many ectomycorrhizal

species colonise newly planted forests from retained mature trees or from the

spore bank built up during the site over a number of years (Jones et al. 2003). The

morphotype Cenococcum geophilum was present in abundance in three of the

Scot’s pine plots, yet absent from Annagh; while the morphotype Suillus

variegatus was present in two of the Scot’s pine plots yet absent from Annagh.

Suillus variegatus was found to be much more common in mature Scot’s pine

forests in Spain (Bonnet et al. 2004), possibly indicating that this species fits into

the “late stage” grouping (sensu Deacon and Fleming 1992). The “early stage”

morphotype Laccaria lacatta was present in Annagh, yet absent from the rest of

the Scot’s pine plots. Laccaria lacatta is known to be a pioneer species in forests

(Deacon and Fleming 1992), and this is confirmed by its presence in the young

and absence from the old pine plots investigated.

Morphotype 5c9 (probably Cortinarius subgenus Sericeocybe cf.

violaceus) was very common (100% frequency in sub-samples) in two of the

cores in Torc where it formed large sclerotia. These can aid in the proliferation

and perennation of ectomycorrhizas, because they are relatively stable against

environmental stresses such as drought (Carlisle and Watkinson 1994).

Fruitbodies of this fungus were found to be very common in the Torc site over the

years 2007-2010 (Chapter 4). The ectomycorrhizas of Suillus variegatus was

found exclusively in pine sites. The genus Suillus, which is a member of the

Suilloid group, which also includes the genera Rhizopogon, Truncocolumella,

Gomphidius and Chroogomphus, are well known to exhibit strong host

preferences to pine species (Bruns et al. 2002).

The ectomycorrhizas of Cenococcum geophilum were found in all three of

the four Scot’s pine plots except Annagh. The unknown dispersal methods of C.

geophilum, which are probably relatively short-range (LoBuglio and Taylor

2002), would make it very difficult for this ECM species to enter the site of a new

forest unless it is already present on the roots of the seedlings from the nursery

stock. Cenococcum geophilum has been recorded on the roots of Pinus sylvestris

seedlings in Lithuania by Menkis et al. (2005), and therefore its presence in

nursery soils may aid in its dispersal. Cenococcum geophilum is thought to be

mainly dispersed by water or animals (Trappe 1969), and the lack of forested land

use near the Annagh plot may have impeded its dispersal into the plot.

321

In the native Scot’s pine forests of Scotland, Pickles et al. (2010) also

found Cenococcum geophilum and Suillus variegatus in their pine plots, but no

other similarities in ectomycorrhizal morphotypes were found between their study

and this study. However, five of the morphoptypes found in the pine plots of

Pickles et al., were found as sporocarps in the pine plots of this study (Chapter 4).

Heinonsalo et al. (2007) also found S. variegatus and Cenococcum geophilum

morphotypes in their study of ECM colonization after clear-cutting, and therefore

it is possible these two ECM species may not be adversely effected by the short

rotation cycles of Scot’s pine forests in Ireland.

The Sitka spruce ECM community

The Sitka spruce plots were the most variable of all the forest types with regard to

their ECM communities. This mirrored similar variability in respect of above-

ground sporcarp communities. There are many possible reasons for this high intra-

forest type variability in the Sitka spruce forest type. For example, It is known that

clear-cutting alters the ECM communities of forests (Jones et al. 2003), therefore

the Sitka spruce forests would rarely develop the stable fungal communities

characteristic of late stage forests. Differences in the soil types and abiotic

variables of the Sitka spruce plots (Chapter 3), or the ectomycorrhizal generalist

nature of Sitka spruce (Alexander and Watling 1987) could all be reasons for high

intra-forest type variability. The above-ground macrofungal communities of Sitka

spruce and oak were more similar to each other than to other forest types, but this

was not found for the below-ground ECM communities in these two forest types.

Many of the ECM morphotypes found in the spruce plots are often found on

nursery saplings e.g. Piceirhiza horti-inflata, Piceirhiza nigra, Tylospora spp.,

and were not found in the oak ECM communities. The frequent occurrence of

nursery ECM fungi would be likely if there was a long delay between clear-cut

and the re-forestation of the site. Other ECM species that produce long-lived

reproductive structures, such as the chlamydospores produced by Wilcoxina

ectomycorrhizas (previously called E-strain mycorrhizas) (Egger et al. 1991),

would also be favoured in these short rotation clear-cut forests, as was found in

Sitka spruce forests in the U.K. (Thomas et al. 1983).

322

Ectomycorrhizas of Tylospora cf. fibrillosa, Paxillus involutus and

Russula ochroleuca were common in Sitka spruce sites in this project and in Sitka

spruce forests in England (Palfner et al. 2005). Paxillus involutus has been

recorded on Sitka spruce trees in Ireland before (Heslin et al. 1992), along with 9

other types, including Cenococcum geophilum and Cortinarius obtusus, which

were also found with Sitka spruce in this project. The species Cortinarius croceus

and Cortinarius tubarius (listed as Dermocybe croceus and Dermocybe sphagneti)

were found in both of Heslin’s pure Sitka spruce plots, yet were lacking from this

study. It is possible that these species were locally common to the sites

investigated by Helsin et al., as sporocarps of these species are listed as

widespread but rarely reported (Legon and Henrici 2005). An investigation into

the ECM fungi on Sitka spruce under natural regeneration in Scotland (Flynn et

al. 1998), revealed 13 morphotypes of which Tylospora fibrillosa, Russula

ochroleuca, Laccaria sp., ITE2 and Cenococcum geophilum were common to this

study. The study by Thomas et al. (1983) of Welch Sitka spruce forests, also

found Cenococcum geophilum and Russula ochroleuca, and in addition they also

identified two Amanita species and four Lactarius species, of which some may

match the unidentified Amanita and Lactarius species from Sitka spruce in this

study.

6.5.3 Relationship between above-ground ECM sporocarps and below-

ground ECM morphotypes

The overall similarity between the fruitbodies above-ground and ectomycorrhizas

below-ground was low, on average, less than 10% of the species occurred as both

sporocarps and ectomycorrhizas on roots. It is common in studies which examine

the above- and below-ground ECM fungi to find disparities between the two

systems, in respect of species identities and richness and similar results have been

found in many different forest types from many different countries (Dahlberg et

al. 1997, Gardes and Bruns 1996 ; Peter et al. 2001; Richard et al. 2005; Gebhardt

et al. 2007). Nonetheless, the Mantel tests identified that the relationship between

ECM communities above-ground (sporocarps) and below-ground (ECM

morphotypes) was statistically much greater than expected by chance. In Norway

323

spruce forests in Switzerland, large differences between above-ground and below-

ground ECM richness were also found, but nonetheless, Peter et al. (2001) found

a similar relationship between the above-ground and below-ground communities

using Mantel tests, although their significance was much weaker (Mantel’s r=

0.11) than in this study.

Of those species represented above both above- and below-ground,

Russula ochroleuca was more common above- than below-ground. Its sporocarps

were found in great abundance in three of the oak, two of the Scot’s pine and two

of the Sitka spruce forests, whereas the easily-recognisable ectomycorrhizas of R.

ochroleuca were found only in one oak, one Scot’s pine and one Sitka spruce plot.

Palfner et al. (2005) found a similar trend in their Sitka spruce plots, where R.

ochroleuca was found as sporocarps at two sites and as ectomycorrhizas at only

one site. In fact, this species and other Russula and Cortinarius species have been

noted in many studies to account in relative terms, for much more of the

sporocarp biomass above-ground than ECM types below-ground (Peter et al.

2001; Dahlberg et al. 1997).

There were many species that were only found as sporocarps and never as

ECM morphotypes. From the sporocarp survey a total of 33, 29 and 22 species of

Cortinarius, Russula and Lactarius were found in the forest types. Conversely

only 4, 4 and 2 species from these genera were found as ECM morphotypes.

Differences in above- and below-ground species richness and community

composition are often due to the larger amount of sampling devoted to sporocarp

study than the ECM root examination. This study like many others before it

(Dahlberg et al. 1997; Luoma et al. 2006a), devoted more time (>1 sampling

seasons) to the study of the above-ground sporocarps than to the below-ground

ECM types (1 sampling season), which may further deepen the rift between the

above and below ground species lists. Although this is the norm in previous

studies examining the above- and below-ground fungal component of forest

ecosystems, Horton and Bruns (2001) question the larger resources and time

dedicated to the above-ground collection of sporocarps at the expense of the

below-ground examination of the roots. However, it was pointed out by Taylor

(2002) that to equally sample above- and below-ground communities on an area

basis, the number of roots tips to be examined would be overwhelming (between 7

and 72 x 104 root tips per m2 forest floor). In order to examine if there was a

324

relationship between the sample size and ECM types identified in forests, data

from published studies (Table 6.13) was used in a stepwise regression analysis

(PASW version 18 ; SPSS Inc., Chicago, Illinois), with the sample size and total

volume of soil sampled as the independent variables and ECM richness as the

response variable. A straight line relationship (F1,14 = 15.82, r2=0.53, P<0.01),

(Fig. 6.14) that explained over 50% of the variation was found. It was found that

the numbers of individual samples examined had the highest effect on the number

of morphotypes recovered. It is known that ectomycorrhizas can have clumped or

patchy distributions (Horton and Bruns 2001, Taylor 2002 and Tedersoo et al.

2003), and therefore more samples are likely to yield more distinct morphotypes.

The effect of increasing sampling was noted in this project, as it was found that

Sitka spruce had the lowest number of morphotypes overall, and this was shown

to be due to the lower length of Sitka spruce roots sampled. When the sampling

intensity was equalled between the forest types (according to rarefaction) it was

found that morphotype richness was equal amongst the forest types.

325

Table 6.13 Details of below ground ECM studies carried out on oak, Scot’s pine and Sitka spruce forests. Column 3 refers to diversity of ECM fungi based on morphotyping or molecular methods. Column 4 lists the number of samples taken (cores or blocks of soil) and the volume of the sample unit (in parenthesis).

Country Tree type ECM

richness

Sample size Total

volume

sampled

Study objectives Reference

Ireland Oak (Quercus petraea, Q.

robur) 21 20 soil cores

(1178cm3) 47,124cm3 This project

France Oak (Q. petraea) 36 40 soil samples (2500cm3)

100,000cm3 Test of niche effect on ECM communities

Buee et al. 2007

France Oak (Q. petraea) 75 90 soil cores (1964cm3)

176,714cm3 Test of temporal variation on ECM communities

Courty et al. 2008

Germany Oak (Q. rubra) 61 128 soil cores (295cm3)

37,699cm3 Test of host age on ECM communities

Gebhardt et al. 2007

U.S.A. Oak (Q. rubra) 33 96 soil cores (196cm3)

18,850cm3 ECM communities of urban and rural oak forests

Baxter et al. 1999

U.S.A. Oak (2 species Q.

douglasii, Q. wislizeni) 140 64 soil cores

(900cm3) 57,600cm3 ECM communities of

different oak habitats Morris et al. 2008

Ireland Scot’s pine (Pinus

sylvestris) 20 20 soil cores

(1178cm3) 47,124cm3 This project

Scotland Scot’s pine (P.Sylvestris) 24 217 soil cores (393cm3)

85,216cm3 Spatial and temporal variation of ECM fungi

Pickles et al. 2010

Sweeden Scot’s pine (P.Sylvestris) 135 120 soil cores (92cm3)

11,084cm3 Response of ECM communities to fire

Jonsson et al. 1999

Finland Scot’s pine (P.sylvestris) 34 60 seedlings N/a Effect of clear-cut practice on ECM communities

Heinonsalo et al. 2007

Sweeden Scot’s pine (P.sylvestris) 37 30 cores (92cm3)

2771cm3 Relationship between sampling effort and ECM

Taylor 2002

326

species accumulation

Spain Maritime pine (P.pinaster) 45 >90 seedlings N/a Effect of fire severity on ECM communities

Rincon and Pueyo 2010

Ireland Sitka spruce (Picea

sitchensis) 18 20 soil cores

(1178cm3) 47,124cm3 This project

Ireland Sitka spruce (P.sitchensis) 10 Not given N/a ECM communities on Sitka spruce

Heslin et al. 1992

U.K. Sitka spruce (P.sitchensis) 13 40 soil cores (200cm3) + 25 seedlings

>8000cm3 Chronosequence effect on ECM communities

Palfner et al. 2005

U.K. Sitka sruce (P.sitchensis) 13 30 seedlings 900cm3 ECM colonisation of Sitka spruce under natural regeneration

Flynn et al. 1998

U.K. Sitka sruce (P.sitchensis) 25 Not specified N/a ECM community changes following out-planting

Thomas et al. 1983

U.S.A.

(Alaska)

Sitka spruce (P.sitchensis) 17 17 (31cm3) 531cm3 ECM community changes across an ecotone

Wurzburger et al. 2004

Canada Hybrid spruce (P. glauca x

sitchensis) 28 27 seedlings N/a Effects of ECM

communities on seedling growth

Krannabetter 2004

Switzerla

nd

Norway spruce (P.abies) 79 106 soil cores (251cm3)

26,641cm3 ECM communities of Norway spruce forests

Peter et al. 2001

Canada White spruce (P.glauca) 52 32 seedlings N/a ECM communities at forest edges

Krannabetter et al. 1999

Canada Black spruce (P.mariana) 65 45 seedlings N/a Effect of a moisture gradient on ECM communities

Robertson et al. 2006

327

0 25 50 75 100 1250

25

50

75

100

125

150

Number of samples

EC

M t

yp

es

Fig. 6.14 Linear regression curve of the number of ectomycorrhizal morphotypes identified against the number of individual samples taken from the studies listed in table 6.13. Regression curve was found to be highly significant (F1,14 = 15.82, R2=0.53, P<0.01). The equation of the line is Y = 0.55Xnum samples + 15.15.

328

6.5 Conclusions

• The ECM communities of Irish forests showed a strong correspondence

with the dominant tree type of the forest.

• There were no significant differences between the total richness of ECM

types across the different forest types.

• Sitka spruce forests have an abundance of ECM morphotypes that are

normally found on nursery trees. The short rotation of Sitka spruce forests

in Ireland may favour these nursery ECM communities and also ECM

species that can form long-lived resistant propagules.

• The similarity of above- and below-ground ECM species was found to be

very low in this project with less than 10% of the ECM morphotypes

found as sporocarps.

329

Chapter 7: General discussion and

conclusions

330

331

7.1 Macrofungal diversity in Irish forest sites

Macrofungal diversity was assessed in this project from two different aspects: by

recording of sporocarps and identification of ectomycorrhizas on tree roots. A

large number (409) of macrofungal species were identified from sporocarps in the

three years of this project, including 48 species new to Ireland. These results

concur with the hypothesis of O’Hanlon and Harrington (in press), which suggests

that the majority of Ireland’s as yet undiscovered agaricomycete diversity may be

found in forests. Between 54% and 68% of the species, depending on forest type,

were only found on one plot visit and so would appear to sporulate irregularly.

Sitka spruce forests were found to contain the greatest species richness of

macrofungi of all the forest types, followed by oak, Scot’s pine and ash forests.

Sitka spruce forests in the U.K. have similarly been shown to support a high

species richness of macrofungi (Humphrey et al. 2003; C. Quine, unpub. data),

although not as high as the pedunculate oak forests also investigated in the U.K.

This project utilized sample-based rarefaction to overcome the methodological

issue of unequal numbers of plots (a similar non-orthogonal study design was

employed by Humphrey et al. 2003). Using sample-based rarefaction, it was

found that the species richness of the forest types was not significantly different

between oak, Scot’s pine and Sitka spruce forests. There was no significant

difference between the species richness of the young or mature Sitka spruce plots.

It is possible that the age of the mature site grouping was not old enough to allow

for the development of specific “late-successional” fungi as found in coniferous

forests in Canada (Krannabetter et al. 2005) and Spain (Bonnet et al. 2004).

Species-richness estimators (Chao2, Incidence based Coverage Estimator

[ICE], Abundance based Coverage Estimator [ACE]) produced the same results

for the ranking of the forest types based on species richness. Richness estimation

revealed that 45, 65, 52 and 77% of the species richness had been recorded in the

ash, oak, Scot’s pine and Sitka spruce forests types respectively. The use of

species-richness estimators in macrofungal studies is not common and the

reliability of such estimators in macrofungal studies has been questioned (Schmit

et al. 1999; Unterseher et al. 2008). However, Schmit et al. (1999) found that the

Chao and Jacknife estimators gave fairly accurate estimates of macrofungal

diversity in oak forest by estimating from data compiled in year 1 of the study, to

332

subsequent years. The richness estimate of their 1000m2 oak and maple plot in

Indiana using three year’s data was 178 species, which is comparable to estimates

for the 100m2 oak plots in this study. These close estimates aside, they concluded

that the estimators were not suited to their macrofungal data as they showed large

increases with successive sample season’s data.

This study rated the richness estimators according to the criterion set out in

Magurran (2004) for richness estimators, i.e. that the estimators reach a stable

asymptote at low sample sizes and preferably before 100% of the sampling effort

has been reached. On this basis, none of the species richness estimators would

qualify for use in macrofungal studies in the forest types examined in this study,

as all of the estimates continued to increase with increasing sampling. The Chao2

estimator could however be used as an indication of the lower bound of species

richness in these forests. According to the Chao2 estimator, ash forests were

estimated to have 124 species, oak 173 species, Scot’s pine 170 species and Sitka

spruce 186 macrofungal species.

All species richness estimators assume a degree of homogeneity across the

sample areas. Macrofungal species are not randomly distributed in forests, and the

community is often dominated by a few species, based on sporocarp observations.

These two points, (i) habitat heterogeneity and (ii) un-equal catch-ability, are two

factors that will diminish the power of species richness estimators when used with

macrofungal data. In this study, similar to the study by Schmit et al. (1999), the

presence/absence of a species in a number of contiguous sub-plots was used to

create a quantitative measure of abundance rather than counts of sporocarps. This

methodology would improve the performance of the estimators, by smoothing out

a degree of the habitat heterogeneity and also decrease the influence of common

species (species producing abundant sporocarps) on the estimation procedure. In

conclusion, it could be stated that the Chao2 and ICE species richness estimators

are promising methods of estimating total macrofungal species richness in forests

as long as (i) data from a number of years sampling (≥3) is used (ii) sporocarp

abundance is not used as a measure of dominance and (iii) sample plots are placed

in homogenous areas of the forest to reduce the effect on increasing niches on

species richness values.

333

7.2 Macrofungal functional group diversity

The proportions of the different functional group species in each forest type varied

significantly. Oak forests were found to have more wood-decay macrofungi than

the coniferous forests. A lack of coarse woody debris has been identified in the

past in Irish plantation forests (Sweeney et al. 2010a), and this was most likely the

cause for the inequalities in functional groups across the forest types. Previous

work in U.K. forests, has shown that the richness of wood-decay macrofungal

species was positively linked to the amount of coarse woody debris (Ferris et al.

2000a).

The Sitka spruce forest type was found to be the most species rich forest

type, with particularly high species richness in the litter-decay and

ectomycorrhizal functional groups. Sitka spruce forests are known to have high

species richness in some litter-decay genera, such as Mycena (Outerbridge 2002).

A reason for this may be the low levels of monoterpenes released from Sitka

spruce needles during decomposition (Ludley et al. 2008). Monoterpenes are

known to have anti-bacterial and anti-fungal properties, and their low levels in

Sitka spruce litter could allow for decomposition of the needles by many different

macrofungal species. Alexander and Watling (1987) offer the high tree species

diversity encountered by Sitka spruce in its home range, as a reason for the large

number of ectomycorrhizal species capable of forming ectomycorrhizas with

Sitka spruce. The ability to form ectomycorrhizas with more ectomycorrhizal

fungi would allow regenerating Sitka spruce seedlings to quickly tap into the

common mycorrhizal networks of the surrounding mature trees, and thus gain an

advantage over ectomycorrhizal selective tree species. A point that would agree

with Alexander and Watling’s hypothesis is that the most species rich genera of

ectomycorrhizal fungi in Sitka spruce forests were Cortinarius and Russula, both

of which were identified as ectomycorrhizal genera with low host specificity

(Molina et al. 1992).

When ectomycorrhizal species are excluded, forest type comparisons

including the ash forest type, show that Sitka spruce forests would be the most

species rich, followed by oak, ash and Scot’s pine. Although ash forest lack

ectomycorrhizal species, the old un-managed ash forests investigated in this

project had similar levels of rarefied species richness as many of the oak plots (21

species after four plot visits). Old ash forests can support very high species

334

richness of macrofungi (Hering 1966; Straatsma and Krisai-Greilhuber 2003),

with many species capable of sporulating on dry substrates (e.g. Xylaria

hypoxylon) selected for by the dry climate experienced in ash forests, particularly

after defoliation (Gibbs 1957).

The species richness of Scot’s pine forests may be adversely affected by

the lack of native Scot’s pine stands in Ireland. Distance to native pine woods was

positively related to the numbers of rare macrofungal species recorded in the U.K.

plots of Humphrey et al. (2000), and therefore it may be assumed that native

Scot’s pine forests provide a source of macrofungal inoculum for establishing

Scot’s pine forests. The retention of mature patches of forest stands after final

harvesting, as advocated by Peterken et al. (1992), would serve to provide a

source of macrofungal inoculum in the form of spores and mycelia, that could

then recolonise the newly planted forest and increase the functional diversity of

macrofungi in these forests.

7.3 Below-ground ectomycorrhizal diversity

Scot’s pine forests were found to have the highest proportional ectomycorrhizal

species richness, based on sporocarps, of all the forest types. The Scot’s pine

forest type was also the second most morphotype-rich forest type, according to

total morphotype richness and to the species richness estimators using below-

ground morphotype results. Pinus is known to associate with large numbers of

ectomycorrhizal species (Newton and Haigh 1998), few of which are restricted to

pine (~7%), in contrast to Betula (18%) and Quercus (13%), and this may help

increase the species richness of ectomycorrhizal morphotypes in Scot’s pine

forests.

The ectomycorrhizal community of Sitka spruce forests was found to be

composed of more “nursery type” ectomycorrhizas than oak forests. Nursery type

ectomycorrhizas are ectomycorrhizas that grow well in nursery conditions, with

relatively controlled environmental stresses and sufficient nutrients for tree

growth. The replacement of these ectomycorrhizas can begin soon after out-

planting (Palfner et al. 2005), as species found to colonise roots by mycelial

contact through emanating hyphae have been shown to be more common on roots

than species such as nursery-type ectomycorrhizas that colonise from spores, at

least in undisturbed mature forests (Pickles et al. 2010; Peter et al. 2001).

335

In this study as in others (Horton and Bruns 2001), the similarity between

below- and above-ground ECM communities was low. Some of the possible

explanations for the discrepancies between above- and below-ground abundance

are:

• Some species rely more on spores than mycelia contact to colonize new roots,

and therefore produce more sporocarps than species that rely on mycelial

contact in the rhizosphere (Peay et al. 2011). Peay et al. propose that species

which colonise readily from spores in lab experiments (e.g. Laccaria spp.,

Hebeloma spp.) often need to produce large amounts of fruitbodies in order to

ensure their survival in a habitat. Laccaria lacatta and L. amethystina were

two of the most common fruitbody producers in the three forest types in this

project, and these species are known to rely on spores for their continued

presence in the rhizosphere (Gherbi et al. 1999). Russula ochroleuca was also

very common as sporocarps in this project, and results of molecular marker

studies in pine forests in California have revealed that spore propagation

plays a large part in the colonization strategies of some Russula and Lactarius

species (Redecker et al. 2001). Colour plates 3, 4, 5 and 6 highlight the

prolific fruiting of some Russula and Lactarius species in this project, and

would lend support to the Peay et al. hypothesis about the relationship

between sporocarp production and ECM functional morphology.

• Insufficient below-ground sampling intensity (Taylor 2002). It is possible that

increased sampling would lessen the disparities between the above- and

below-ground species lists. It is unavoidable that some species that do not

form sporocarps will always be missed above-ground (e.g. Cenococcum

geophilum), but sample size, or more specifically the number of separate

samples taken, will increase the species richness discovered in the forest. The

difficulty with trying to give equal time and resources to the above- and

below-ground study of ectomycorrhizas has been commented upon by Taylor

(2002), and it is estimated that equal sampling of the above- and below-

ground communities is not feasible, at least not on an equal area basis.

7.4 The macrofungal communities of Irish forests

There were discontinuities in the distributions of macrofungi between the four

forest types, giving rise to distinct assemblages or communities. These

336

communities were composed of common species (e.g. Mycena leptocephala),

exclusive species (e.g. Daldinia concentrica with ash forests) and characteristic

species that were more common in one forest type than in other forest types (e.g.

Laccaria amethystina with oak forests). Wilkins et al. (1937) applied these

groupings to the macrofungi found in British deciduous forests, and noted that

they could be used to clearly define a macrofungal assemblage/community in

British oak and beech forests. Difficulties arise in using ephemeral and sporadic

sporocarps to distinguish macrofungal assemblages (Watling 1995). The unknown

relationship between sporocarp abundance and functional role in ecosystems, have

hampered the effectiveness of macrofungal sporocarps to describe functional

communities of fungi. This study utilized the presence of species in a number of

contiguous sub-plots as a measure of abundance, rather than using counts of

fruitbodies. Similar methods using sub-plots frequencies as quantitative measures

have been used by Bills et al. (1986) and Schmit et al. (1999), and their use is

more ecologically meaningful than that of sporocarp counts. Sporocarp production

is highly related to meteorological variables (Krebs et al. 2008), therefore, using

sporocarp abundance as a quantitative measure will produce data with large

seasonal fluctuations and may diminish the power of multi-variate statistical

techniques. However, the obvious pitfalls to using sporocarp presence as an

indicator of species presence have not reduced the significance of sporocarp based

studies; and these have successfully described distinctive macrofungal

communities in a wide range of forest types (Ferris et al. 2000a; Humphrey et al.

2000; Kranabetter et al. 2005; Buee et al. 2011). Information as to host

preference, species’ response to abiotic, biological and environmental variables,

and ectomycorrhizal richness response to forest management techniques, have all

been identified in the sporocarp only studies listed above.

This study revealed possible host preference for a number of macrofungal

species. Using indicator species analysis, seventeen species were identified that

showed some levels of preference for a certain forest type. The majority of the

species were decay species, with only four ectomycorrhizal species (oak:

Laccaria amethystina, Lactarius quietus; Scot’s pine: Lactarius hepaticus; Sitka

spruce: Russula emetica) showing preference for a specific forest type. The

resource specificity of certain decay macrofungi is well established through both

field-based (Frankland 1998; Unterseher and Tal 2006) and laboratory studies

337

(Boddy 2001; Ludley et al. 2008), so the fact that the majority of macrofungal

species found to strongly indicate tree species preference are wood- and litter-

decay species is not surprising. These species are more likely to discriminate

macrofungal communities, especially as levels of ectomycorrhizal specificity

world-wide (Molina et al. 1992) and in the U.K. (Newton and Haigh 1998) have

been noted as being low. The majority of species that showed restriction to a

single forest type in this study were decay fungi. Moreover, there are certain

species that can associate with several tree species, but often associate with one

tree species more than others. It is this quantitative aspect of macrofungal species

ecology that the community analysis of macrofungi should focus upon (Tyler

1992).

The results of this study found that the Scot’s pine and oak macrofungal

communities were the most distinctive. The plots of these two species formed

distinct groups in ordination and were also indicated strongly by certain wood-

decay species (e.g. Stereum hirsutum with oak and Trichaptum abietinum with

Scot’s pine). In a previous study of the macrofungal community of Scot’s pine

forests in the U.K. (Humphrey et al. 2003), it was found that both “old growth”

and plantation Scot’s pine forests formed distinct macrofungal assemblages. A

related study on plantation Scot’s pine forests in England (Ferris et al. 2000a),

found that the decay species Auriscalpium vulgare, Baeospora myosura,

Trichaptum abiteinum and ectomycorrhizal species Lactarius hepaticus were

common to all the Scot’s pine plots. These four species were also found to be

either exclusive or characteristic species of Scot’s pine forests in this study.

The finding that oak forests have a distinctive macrofungal community

appears to conflict with the existing information of oak forest macrofungal

communities (Tyler 1992; Wilkins et al. 1937; Watling 1974, 2005), which

identified oak communities as having a non-specific macrofungal community. It is

likely that differences in study design between this project and the previously

listed references are the reason for this conflict of results. The above studies all

investigated the macrofungal communities of oak plots and beech plots. When the

two are compared, the oak macrofungal community is much less distinctive than

that of a beech or birch forest. For example, Tyler (1992) noted 26 species as

showing some sort of affinity to beech forests, 24 with hornbeam, but only 12

species showed a significant relationship with oak forests in Sweden, based on the

338

relative frequency of species observations with a particular tree species. These

findings are in agreement with Watling (2005), that the Scottish oak wood mycota

may be an impoverished version of the English oak mycota, which in turn is an

impoverished form of the continental European mycota. Results from this study

and also from the meta-analysis by O’Hanlon and Harrington (in press), show that

the genera that are characteristically rich in British oak woods (e.g. Boletus), may

be nationally rare/extinct from Ireland (North and South). The historical removal

of Ireland’s oak forests may have caused extinction of oak associated species,

particularly ectomycorrhizal species, as these cannot survive when cut-off from a

host supply of carbohydrate.

The macrofungal community of Sitka spruce forests was found to share

many species similarities with Irish oak and Scot’s pine forests. The genus Picea

has been identified as an ectomycorrhizal generalist in the U.K. (Newton and

Haigh 1998), with Sitka spruce being especially noted as an ectomycorrhizal

generalist in its home range (Alexander and Waling 1987). Results of this study

would corroborate with Alexander and Watling’s hypothesis regarding the

ectomycorrhizal generalist nature of Sitka spruce. The results from the

Biodiversity in Britain’s planted forests (Humphrey et al. 2003; C. Quine, unpub.

data) would also indicate that in Western Europe, Sitka spruce is an

ectomycorrhizal generalist and forms ectomycorrhizas with many native

ectomycorrhizal fungi. Of the 111 ectomycorrhizal species found in British Sitka

spruce forests, 41 were also found in either oak or Scot’s pine forests and 31

species were common to oak, Scot’s pine and Sitka spruce forests. The

communities of other organisms in Sitka spruce forests in Ireland have also been

noted as being composed mainly of generalist species. Examinations of the

vascular species (French et al. 2007), spider communities (Oxbrough et al. 2006a)

and bird assemblages (Sweeney et al. 2010b) have shown that Sitka spruce

plantations are often dominated by generalist species. This stems from the fact

that Sitka spruce is not a native tree species to Ireland, and the species that

normally are found with it in its home range may not have reached Ireland. The

high species richness of native macrofungi in Sitka spruce forests in this study

therefore agree with the conclusions of Smith et al. (2006) for Ireland and Quine

and Humphreys (2010) for the U.K.: that Sitka spruce plantations in Ireland and

339

England are “complementary habitats” (Barlow et al. 2007) for the conservation

of native species groups.

Other researchers have also identified differences in abiotic variables

(Wilkins and Patrick 1939; Ruhling and Tyler 1990) and understory vegetation

(Wilkins et al. 1938; Baar 1996) as factors affecting the macrofungal communities

in forests. It is known that certain ectomycorrhizal species have preferences for

low or high pH (Hung and Trappe 1983), low or high soil nitrogen (Lilleskov et

al. 2002) and amounts of soil organic matter (Tyler 1985; Harrington and Mitchell

2005a). This research tested for correlations of the abiotic, understory cover and

soil organic matter, and the ordination axes derived from the multi-variate

analysis of the macrofungal communities. Although relationships between soil

abiotic variables and soil organic matter were found, it is believed that they were

not the main drivers of community structure in these forests types. This can be

stated because:

(i) Certain species were found to be just as common in all sites of a particular

forest type, despite differences in soil abiotic variables. The species Lactarius

hepaticus and L. rufus were of similar frequency in the Scot’s pine plots at

Annagh, Gortnagowna and Derryhogan. L. hepaticus is noted as a species that is

broadly tolerant of nitrogen and probably increasing in distribution due to

increasing levels of nitrogen deposition in European forests (Heilmann-Clausen et

al. 1998). Other species that are present in these three Scot’s pine plots are

Mycena leptocephala, Ricknella fibula, Trichaptum abietinum and Russula

ochroleuca. The frequencies of all these species are similar in the three

aforementioned Scot’s pine plots, despite differences in soil pH, available

nitrogen, calcium, magnesium and phosphorus. This finding would lend more

weight that macrofungal communities are more structured by the influence of the

dominant tree species than soil abiotic variables.

(ii) The data from Chapter 3 was used to classify the forest plots into their

woodland classifications according to Fossitt (2000). When an MRPP analysis

was run (as in Section 5.3.6, Chapter 5), two different grouping variables were

used to test if the macrofungal communities of the plots were not randomly

distributed between the different forest plots, but rather formed distinct

assemblages, according to the user-assigned grouping variable of the plots. In this

analysis, both the dominant tree species of the plots (ash, oak, Scot’s pine or Sitka

340

spruce) and the vegetation classification (five groupings: WN1, WN2, WN7,

WD1 and WD4) of the sites were used as grouping variables. Although both

grouping variables were significant at grouping the macrofungal communities, the

effect of dominant tree species was higher, according to a higher A statistic

(0.0687 for tree type vs. 0.0618 for vegetation type).

7.5 The value of plantation forests as habitats for macrofungal diversity

Other studies of plantation forests in Ireland have also found that plantation

forests could serve as habitats for native flora and fauna and so aid in the

conservation of native species. French et al. (2008) examined the vascular plant

communities of plantation Sitka spruce, larch and ash forests. They found that in

some plantation forests, low levels of plant diversity are found, due to competitive

exclusion by fast growing species, but that vascular species biodiversity can be

increased by increasing the CWD quantity and quality remaining on site and

through the promotion of chronologically heterogeneous forests. Oxbrough et al.

(2006a) found that the initial effects of afforestation on the spider communities of

grassland and bog habitats, was a reduction in species diversity due to the

replacement of habitat specialist species with more generalist species. However,

they did find that plantations supported a number of forest specialist species. In

related research, Oxbrough et al. (2006b) found that the inclusion of open spaces

(both glades and road/path verges) in the forest could be used to increase the

biodiversity of spiders found in plantation forests. These open spaces would

provide a different habitat type, to that of the forest canopy and forest floor, thus

increasing the range of habitats available in plantations. Research into the bird

communities of native and plantation forests in Ireland has also identified the

creation of a diversity of habitat types within plantation forests as important for

increasing bird species richness and diversity (Sweeney et al. 2010b). Sweeney et

al. compared the bird communities of oak, ash and plantation Sitka spruce forests,

and found that bird species richness was highest in oak, followed by ash and then

Sitka spruce forests. They related this decreasing species richness to the levels of

understory vegetation cover, and proposed that bird species richness and diversity

would increase, if management techniques that decreased canopy cover and

increase understory vegetation were applied to plantation forests.

341

The diversity and species richness of carabid beetles, in both first and

second rotation Sitka spruce plantations, was investigated by Oxbrough et al.

(2010). They found that beetle species richness and diversity increased as the

forest aged. This result is the opposite of that found for the spider assemblages in

the same sites, with spider species richness decreasing as the canopy closed and

the forest aged. They attribute this to the differing colonisation methods of beetles

from those of spiders. Beetles rely on flight to enter habitats, thus they take longer

to re-colonise a habitat after clearfelling of standing trees. Again, the creation of

chronological heterogeneity was identified as a management technique that would

lead to increased beetle richness and diversity in plantation forests. Arroyo et al.

(2010) examined the mesostigmetid mites of Irish Sitka spruce plantation forests.

They found that the mite species richness was lower than that recorded for Sitka

spruce in its home range of North America, but that there were species that were

restricted to specific niches (e.g. epiphytic moss habitat). Using Detrended

Correspondence Analysis they identified a number of different moss and bare

surface mite communities. Although not commented on in the research, it could

be hypothesised that by increasing canopy open-space and ground floor

vegetation, more micro-niches would be created and so increase the habitats for

more mite species.

The studies by French et al. (2008), Oxbrough et al. (2006b), Sweeney et

al. (2010b), Oxbrough et al. (2010) and Arroyo et al. (2010), all indicated that

management techniques that increase the number of different micro-habitats

within plantation forests would also increase biodiversity. The retention of mature

patches of forest stands after final harvesting, as advocated by Peterken et al.

(1992), Ferris et al. (2000a) and Humphrey (2005), would help in the creation of

many micro-habitats within the plantation forest. These retained patches would

also serve to provide a source of fungal inoculum in the form of spores and

mycelia, which could then recolonise the newly planted forest and increase the

functional diversity of macrofungi in these forests.

Biodiversity indicators are useful methods of estimating the biodiversity of

a forest site without having to examine the diversity of the individual taxonomic

groups (Noss 1999; Lindenmayer et al. 2000). It has been suggested that vascular

plant diversity in forests can be used to estimate macrofungal diversity (Schmit

and Mueller 2007). From previous studies that examined both vascular plant

342

species richness and macrofungal species richness, a 2:1 ratio for macrofungi to

vascular plants has been found in Canadian forests (Villeneuve et al. 1989); while

on the Hebrides, Scotland, a 1.5:1 ratio has been recorded (Dennis 1986). Results

from this study indicate that the macrofungi to vascular species richness was

similar to that of Villeneuve et al. (1989) (~2:1) for ash and Scot’s pine forests,

but higher for oak (~3:1) and Sitka spruce (~6:1) forests. There was also high

variability in the ratios between the individual sites in each forest type.

The estimate found for Sitka spruce forests using the macrofungal to plant

ratio, was much lower than the actual species richness of this forest type. Sitka

spruce plantations in Ireland constitute a “novel ecosystem” (sensu Hobbs et al.

2006). The ecosystem created by Sitka spruce plantations in Ireland is highly un-

natural, with a low species richness of vascular plants due to the closed canopy of

Sitka spruce monocultures. In an old-growth state in its home range, the Sitka

spruce forest type is characterised by a mixture of tree species, an abundance of

CWD (especially fallen mature trees) and a mixture of chronological stages of the

forest cycle (Peterson et al. 1997). These three factors have the effect of

increasing the diversity of vascular plant species in native Sitka spruce forests.

The application of estimation techniques created for natural ecosystems to these

novel ecosystems should be done with caution, as the new biotic assemblages of

these novel ecosystems are not fully understood. Changes in key interactions and

processes, such as plant–animal interactions, microbial communities and the

breaking down organic matter in soils are all important yet unknown factors in

these habitats. Young coniferous forests (e.g. Sitka spruce) can be poor in

vascular plants, but can have very high macrofungal species richness; while

deciduous forests (e.g. ash) can have high plant species richness, yet very low

macrofungal species richness. A greater understanding of these novel ecosystems

is needed before such broad-scale estimation is applied to native and plantation

forests.

7.6 Unanswered questions and future research directions

This study, the first systematic long-term study of macrofungal diversity in Irish

forests, will hopefully provide a basis for further investigations in the following

areas:

343

• This study primarily used sporocarp collection and identification as a means

of confirming a species presence in a plot. Recently, the focus is changing

from species which fruit in forests to species that are present as mycelium in

forests. The recent advances in next-generation sequencing techniques has

allowed for the identification of large levels (1000 operational taxonomic

units) of species richness from fungal DNA extracted directly from forest

soils in one year (Buee et al. 2009). Compare this to the ca. 900 species

identified from deciduous and coniferous plots in Switzerland over seven

years sampling by Straatsma and Krisai-Greilhuber (2003). In a pine forest in

the U.K. Orton (1987) identified 662 species of fungi over the course of 30

years. The large amount of species identified from a small sampling time

period is one of the main benefits of next-generation sequencing. However,

next generation sequencing methods are not without their disadvantages,

problems with the application of this relatively new method to fungal ecology

are discussed by Tedersoo et al. (2010), Ovaskainen et al. (2010) and Hibbett

et al. (2011). These authors unanimously agree that these methods will reveal

large amounts of important data on fungal distributions and community

composition, so long as the experimental procedures, data management and

sequence matching methods are standardised and a high quality sequence

database is used for identifications. This project has now set a baseline for

information for fungal species richness and community structure in these

plots and forest types. A follow up study using next-generation sequencing

methods should reveal links between the fungi that produce sporocarps, ECM

root fungi and the fungi present as mycelium in the soils of these forests.

• Many studies have shown that the cumulative species curve for macrofungi

can continue to increase after many years of sampling (Orton 1987; Tofts and

Orton 1998; Straatsma et al. 2003). Further periodic sampling in the plots

from this project will reveal even more information about the undiscovered

species and the fungal communities in these forests regarding the:

(1) Total species richness of macrofungi in these plots: including the species

where fruiting was missed, species with minute sporocarps and species

which only fruit on rare occasions.

344

(2) Validation of the species richness estimations that were generated for the

forest types and forest plots to be tested with real data.

(3) Information about the change in fungal community as the forest ages,

similar to studies by Deacon and Fleming (1992) and Kranabetter et al.

(2005).

• A logical follow up to this project would be to examine the macrofungal and

ectomycorrhizal communities of other tree species, such as beech and birch

woods. Other exotic species, i.e. lodgepole pine and Norway spruce, are also

widely planted in Ireland in commercial forests. It would be interesting to

investigate the macrofungal communities of beech and birch woods, as these

forests are known to have a rather specific macrofungal community, at least

in Scotland (Watling 2005). Following on from the investigation of the

macrofungal community of the other tree species widely planted in Ireland,

an investigation into the effect that mixed tree species forests have on

macrofungal diversity would reveal information relating the effects of tree

species composition on fungal biodiversity. It is known that mixing certain

tree species into forest plantations (e.g. birch trees mixed in with Spruce

woodlands in Scotland; Alexander and Watling 1987) can produce a

significant increase in fungal species richness of the forest, over

corresponding single species forests. Investigation into the differences in

macrofungal communities of very mature (>80 years old) stands would also

reveal interesting information about the late-stage fungal communities of

Sitka spruce forests in Ireland.

• The relationship of macrofungal diversity to that of other taxonomic groups

in these forest types needs to be investigated. This research was part of a

larger work-package which sought to examine the diversity of soil

decomposers and predatory and parasitic arthropods. The synthesis of the

work-package will draw together the results from fungi and soil-arthropods

and make conclusions about the relationship between the biodiversity of the

two groups and recommendations about ways to foster increases in the

biodiversity of both groups.

345

• A final recommendation that can be made from the results of this project is

that the knowledge revealed about Irish forest macrofungi in this project

should be built upon. In agreement with O’Hanlon and Harrington (in press),

two of the main points for future consideration should be the establishment of

a community of fungal recorders that conduct regular macrofungal surveying

in Irish forests and the promotion of a macrofungal recording database along

the lines of The National Biodiversity Data Centre Mapping System

(http://www.biodiversityireland.ie/). Once records of species are confirmed

by a trained taxonomist, their distribution and notes on ecology can be

entered into this database, which would serve as a valuable resource in

discovering the fungal part of Ireland’s biodiversity.

7.7 Final conclusions: Plantations as a habitat for native macrofungi

This project is foremost a follow-up to the investigation of the biodiversity of

other taxonomic groups in Irish plantation and native forests (Smith et al. 2005;

Iremonger et al. 2006; Smith et al. 2006). To date, the biodiversity of vascular

plants, bryophytes, hoverflies, spiders, birds, soil microarthropods, parasitoid

wasps, nematodes and fungi (macrofungi and ectomycorrhizal morphotypes) have

been investigated through COFORD projects. The results of this project are in

agreement with those investigating biodiversity of other taxonomic groups in

plantation forests (Humphrey et al. 2003; Smith et al. 2005); that plantation

forests can provide complementary habitats for macrofungi and so help in the

conservation of biodiversity in Ireland. This finding highlights the potential for

the use of plantation forests to conserve biodiversity, and help Ireland meet the

key condition of the new draft of the Conference of the Parties to the Convention

on Biological Diversity, which states that “by 2020, areas under agriculture,

aquaculture and forestry are managed sustainably, ensuring conservation of

biodiversity” (Anon. 2010).

346

7.8 Key findings of this project

• A total of 409 macrofungal species, 68 vascular plant species and 51 below-

ground ectomycorrhizal morphotypes were identified over the three years

sampling. Forty eight macrofungal species that are new to Ireland were

recorded.

• The four forest types formed distinctive communities of vascular plants,

macrofungi and ectomycorrhizal morphotypes, and this community formation

was driven by the dominant tree species of the forest.

• The macrofungal species richness of native oak and non-native Scot’s pine

and Sitka spruce forests was not significantly different.

• Significant differences were identified in the macrofungal functional group

richness between the different forest types.

• The below-ground ectomycorrhizal morphotype richness of the different

forest types was not significantly different.

347

348

References

349

Agarwala, M., Kumar, S., Treves, A., Naughton-Treves, L. (2010) Paying for wolves in Solapur, India and Wisconsin, USA: Comparing compensation rules and practice to understand the goals and politics of wolf conservation. Biological

Conservation, 143: 2945–2955. Agerer, R. (1987-2002). Colour Atlas of Ectomycorrhizae (1st-12th ed.). Schwäbisch Gmünd: Einhorn-Verlag. Agerer, R. and Rambolt, G. (2004-2011) DEEMY. [online], available: http://www.deemy.de/ [accessed 07/04/2011]. Agerer, R. (2001) Exploration types of ectomycorrhizae –A proposal to classify ectomycorrhizal mycelial systems according to their patterns of differentiation and putative ecological importance. Mycorrhiza, 11: 107-114. Alexander, I. and Watling, R. (1987) Macro fungi of Sitka spruce in Scotland. Proceedings of the Royal Society of Edinburgh, 93: 107-115. Allen, M.F. (1987) Restablishment of mycorrhizas on Mount St Helens: migration vectors. Transactions of the British Mycological Society, 88: 413-417. Allen, M.F. (1991) Ecology of Mycorrhizae. Cambridge: Cambridge University Press. Allen, M.F., Allen, E.B., Dahm, C.N. and Edwards, F.S. (1993) Preservation of biological diversity in mycorrhizal fungi: importance and human impacts, In: International symposium on human impacts on self recruiting populations, Sundnes, G. (ed.). Trondheim: Norwegian academy of sciences, pp 81-108. Allison, S.M. and Prosser, J.I. (1991) Urease activity in neutrophilic autotrophic ammonia-oxidizing bacteria isolated from acid soils. Soil Biology and

Biochemistry, 23: 45-51. Altschul, S. F., Madden, T. L., Schäffer, A. A., Zhang, J., Zhang, Z., Miller, W. and Lipman, D. J. (1997) Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Research, 25: 3389-3402. Amaranthus, M.P. (1992) Mycorrhizas, forest disturbance and regeneration in the pacific north west United States. In: Mycorrhizas in ecosystems, D.J. Read, D.H. Lewis, A.H. Fitter, and I.J. Alexander (eds.). Wallingford: CAB international. Amaranthus, M.P. and Perry, D.A. (1987) Effect of soil transfer on ectomycorrhizal formation and the survival and growth of conifer seedlings on old non-reforested clear-cuts. Canadian Journal of Forest Research, 17: 944-950.

350

Anderson, P., Brundrett, M., Grierson, P. and Robinson, R. (2010) Impact of severe forest dieback caused by Phytophthora cinnamomi on macrofungal diversity in the northern jarrah forest of Western Australia. Forest Ecology and

Management, 259: 1033–1040. Andersson, S., Jensen, P. and Soderstrom, B. (1996) Effects of mycorrhizal colonization by Paxillus involutus on uptake of Ca and P by Picea abies and Betula pendula grown in limed and unlimed peat. New Phytologist, 133: 695-704. Anon. (1996) Growing for the future. A strategic plan for the development of the

forestry sector in Ireland. Dublin: Department of Agriculture, Food and Forestry. Anon. (2000) Forest Biodiversity Guidelines. Department of Marine and Natural Resources. [online], available: http://www.agriculture.gov.ie/forestry/publications/biodiversity.pdf [accessed 08/11/2008]. Anon. (2005). Union Wood cSAC Site Code 638 Co. Sligo. National Parks and Wildlife service. [Online], available: http://www.npws.ie/en/media/Media,4666,en.pdf [accessed 03/09/10]. Anon. (2006) Sampling and soil analysis. [online], available: http://www.icp-forests.org/pdf/Chapt_3a_2006(2).pdf [accessed 08/11/2008]. Anon. (2009) Irish Standard for Forest Management Certification: DRAFT No. 4

for public consultation. Dublin: Irish Forestry Certification Initiative Ltd. Anon. (2010) Draft decisions for the tenth meeting of the conference of the parties

to the Convention on Biological Diversity, Nagoya, Japan. [online], available: http://www.cbd.int/cop10/doc/ [accessed 07/04/2011]. Arnolds, E. (1981) Ecology and coenology of macrofungi in grasslands and moist heathlands in Drenthe, the Netherlands. Bibliographica Mycology, 90: 1-501. Arnolds, E. (1991) Decline of ectomycorrhizal fungi in Europe. Agriculture

Ecosystems and Environment, 35: 209-244. Arnolds, E. (1992) The role of macrofungi in environmental conservation. Giornale Botanico Italiano, 126: 779-795. Arnolds, E. and de Vries, B. (1993) Conservation of fungi in Europe. In: Fungi of

Europe –investigation recording and mapping, Pegler, D.N., Boddy, L., Ing, I. and Kirk, P.M. (eds.). Kew: Royal Botanic Gardens, pp. 231–238. Arora, D. (1986) Mushrooms demystified. California: Ten speed press. Arroyo, J., Moraza, M.L. and Bolger, T. (2010) The mesostigmatid mite (Acari, Mesostigmata) community in canopies of Sitka spruce in Ireland and a comparison with ground moss habitats. Graellsia, 66: 29-37.

351

Ashkannejhad, S. and Horton, T.R. (2006) Ectomycorrhizal ecology under primary succession on the coastal sand dunes: interactions involving Pinus

contorta, suilloid fungi and deer. New Phytologist, 169: 345–354. Avis, P.G., McLaughlin, D.J., Dentinger, B.C. and Reich, P.B. (2003) Long-term increase in nitrogen supply alters above- and below-ground ectomycorrhizal communities and increases the dominance of Russula spp. in a temperate oak savanna. New Phytologist, 160: 239–253. Azul, A.M., Castro, P., Sousa, J.P. and Freitas, H. (2009) Diversity and fruiting patterns of ectomycorrhizal and saprobic fungi as indicators of land-use severity in managed woodlands dominated by Quercus suber--a case study from southern Portugal. Canadian Journal of Forest Research, 39: 2404-2417. Baar, J. (1996) The ectomycorrhizal flora of primary and secondary stands of Pinus sylvestris in relation to soil conditions and ectomycorrhizal succession. Journal of Vegetation Science, 7: 497-504. Baar, J. and Kuyper, T.W. (1993) Litter removal in forests and effect on mycorrhizal fungi. In: Fungi of Europe: Investigation, recording, and

conservation, D.N. Pegler, L. Boddy, B.Ing and P.M. Kirk (eds.) Kew: Royal Botanic Gardens, pp. 275-286. Baar, J. and ter Braak, C. J. F. (1996) Ectomycorrhizal sporocarp occurrence as affected by manipulation of litter and humus layers in Scot’s pine stands of different age. Applied Soil Ecology, 4: 61-73. Baar, J., Ozinga, W.A., Sweers, I.L. and Kuyper, T.W. (1994) Stimulatory and inhibitory effects of needle litter and grass extracts on the growth of some ectomycorrhizal fungi. Soil Biology and Biochemistry, 26: 1073-1079. Baath, E. (1981) Microfungi in a clear-cut pine forest soil in central Sweden. Canadian Journal of Botany, 59: 1331-1337. Bader, P., Jansson, S., Jonsson, B.G. (1995) Wood-inhabiting fungi and substrate decline in selectively logged boreal spruce forests. Biological Conservation, 72: 355-362. Baptista, P., Martins, A., Tavares, R.M. and Lino-Neto, T. (2010) Diversity and fruiting pattern of macrofungi associated with chestnut (Castanea sativa) in the Tras-os-Montes region (northeast Portugal). Fungal Ecology, 3: 9-19. Barkman, J. J. (1976) Terrestrische fungi in Jenever besstruwelen. Coolia, 19: 94-110.

352

Barlow, J., Gardner, T.A., Araujo, I.S., vila-Pires, T.C.A., Bonaldo, A.B., Costa, J.E., Esposito, M.C., Ferreira, L.V., Hawes, J., Hernandez, M.I.M., Hoogmoed, M.S., Leite, R.S., Lo-Man-Hung, N.F., Malcolm, J.R., Martins, M.B., Mestre, L.A.M., Miranda-Santos, R., Nunes-Gutjahr, A.L., Overal,W.L., Parry, L., Peters, S.L., Ribeiro-Junior, M.A., da Silva, M.N.F., da Silva Motta, C. and Peres, C.A. (2007) Quantifying the biodiversity value of tropical primary, secondary, and plantation forests. Proceedings of the National Academy of Sciences USA, 104: 18555–18560. Barreotavena, M.B., Pildain, M.B., Salgado, S. and Eberhart, J.L. (2010) Molecular identification of ectomycorrhizas associated with Ponderossa pine seedlings in Patagonian nurseries (Argentina). Canadian Journal of Forest

research, 40: 1940-1950. Bas, C., Kuyper, T.W., Noordeloos, M.E. and Vellinga, E.C. (1990) Flora

Agaricina Neerlandica, Vol 2. Holland: Balkema. Baxter, J.W. and Dighton, J. (2001) Ectomycorrhizal diversity alters growth and nutrient acquisition of grey birch (Betulla populifolia Marshall) seedlings in host-symbiont culture conditions. New Phytologist, 152: 139-149. Baxter, J.W., Steward, Pickett, S.T.A., Carreiro, M.M. and Dighton, J. (1999) Ectomycorrhizal diversity and community structure in oak forest stands exposed to contrasting anthropogenic impacts. Canadian Journal of Botany, 77: 771-782. Beiler, K.J., Durall, D.M., Simard, S.W., Maxwell, S.A. and Kretzer, A.M. (2009) Architecture of the wood-wide web: Rhizopogon spp. genets link multiple Douglas-fir cohorts. .New Phytologist, 185: 543–553. Berch, S.M., Baumbrough, B., Battigelli, J., Kroeger, P., Strub, N. and de Montigny, L. (2001) Preliminary Assessment of Selected Communities of Soil

Organisms under Different Conifer Species. Research report. Victoria, B.C: Ministry of Forests Research Program. Berch, S.M., Brockley, R.P., Battigelli, J.P., Hagerman, S. and Holl, B. (2006) Impacts of repeated fertilization on components of the soil biota under a young lodgepole pine stand in the interior of British Columbia. Canadian Journal of

Forest Research, 36: 1415-1426. Beschta, R.L., and W.J. Ripple (2010) Recovering riparian plant communities with wolves, in northern Yellowstone USA. Restoration Ecology, 18 (3): 380-389. Bever, J.D. (2003) Soil community feedback and the coexistance of competitors: conceptual frameworks and empiricle tests. New Phytologist, 157: 465-473. Bills, G., Holtzman, G., and Miller, O. (1986) Comparison of ectomycorrhizal- basidiomycete communities in red spruce versus northern hardwood forests of West Virginia. Canadian Journal of Botany, 64: 760-768.

353

Blasius, D. and Oberwinkler, F. (1989) Succession of Mycorrhizae: a matter of tree and stand age. Annals of Forest Science, 46: 758-761. Boddy, L. (2001) Fungal Community Ecology and Wood Decomposition Processes in Angiosperms: From Standing Tree to Complete Decay of Coarse Woody Debris. Ecological Bulletins, 49: 43-56. Boddy, L. and Rayner, A.D. (1983) Ecological roles of basidiomycetes forming decay communities in attached oak branches. New Phytologist, 93: 77-88. Boddy, L. and Watkinson, S.C. (1995) Wood decomposition, higher fungi, and their role in nutrient redistribution. Canadian Journal of Botany. 73(Suppl. 1): 1377-1383. Boddy, L. Bardsley, W. and Gibbon, O. (1987) Fungal communities in attached ash branches. New Phytologist, 107: 143-154. Bohus, G. (1984) Studies on pH requirement of soil inhabiting mushrooms: the R spectra of mushroom assemblages in deciduous forest communities. Annals of

Botany, 30: 155-171. Bolger, T. (2001). Functional value of species biodiversity. Biology and

Environment: Proceedings of the Royal Irish Academy, 101(3): 199-224. Bonet, J.A., Fisher, C.R. and Colinas, C. (2004) The relationship between forest age and aspect on the production of sporocarps of ectomycorrhizal fungi in Pinus

sylvestris forests of the central Pyrenees. Forest Ecology and Management, 203: 157–175. Borchers, S.L., and Perry, D.A. (1990) Growth and ectomycorrhizal formation of Douglas fir seedlings grown in soils collected at different distances from pioneering hardwoods in south west Oregon clear-cuts. Canadian Journal of

Forest Research, 20: 712-721. Boyle, G.M. and Farrell, E.P. (2005) Monitoring of Forest Ecosystems in Ireland, FOREM 9 project, Final Report. Forest Ecosystem Research Group Report Number 72 Department of Environmental Resource Management. Dublin: University College Dublin. Brady, N.C. and Weil, R.R. (2008) The Nature and Properties of Soils, 14th edition. New Jersey: Prentice Hall. Brandrud, T.E., Lindström, H., Marklund, H., Melot, J. and Muskos, S. (1990–1998) Cortinarius, Flora Photographica. Vols. 1 -4. Sweden: HB, Klövervägen, Matfors. Breda, N.J. (2003) Ground-based measurements of leaf area index: a review of methods, instruments and current controversies. Journal of Experimental Botany, 54: 2403-2417.

354

Breitenbach, J., and Känzlin, F. (1984-2005) Fungi of Switzerland. Vol 1-6. Lucerne: Verlag Mykologia Luzern Publishers. Brockerhoff, E.G., Jactel, H., Parrota, J.A., Quine, C.P. and Sayer, J. (2008) Plantation forests and biodiversity: oxymoron or opportunity? Biodiversity and

Conservation, 17: 925–951. Brokaw, N. V. and Lent, R.A. (1999) Vertical structure, In: Maintaining

biodiversity in forest ecosystems, M. L. Hunter Jr. (ed.). Cambridge: Cambridge University Press, pp. 373-399. Brown, N., Jennings, S., Wheeler, P., and Nabe-Nielson, J. (2000) An improved method for the rapid assessment of forest understorey light environments. Journal

of Applied Ecology, 37: 1044-1053. Brundrett, M.C. (2002) Coevolution of roots and Mycorrhizae of land plants. New

Phytologist, 154: 275-304. Bruns, T.D. (1995) Thoughts on the processes that maintain local species diversity of ectomycorrhizal fungi. Plant and Soil, 170: 63–73. Bruns, T.D., Bidartondo, M. and Taylor, D.L. (2002) Host specificity in ectomycorrhizal communities: what do the exceptions tell us?. Integrated and

Computational Biology, 42: 352–359. Bruns, T.D., Peay, K.G., Boynton, P.J., Grubisha, L.C., Hynson, N.A., Nguyen, N.H. and Rosenstock, N.P. (2009) Inoculum potential of Rhizopogon spores increases with time over the first 4 yr of a 99-yr spore burial experiment. New

Phytologist, 181: 463–470. Buee, M., Courty, P.E., Mignot, D., Garbaye, J. (2007) Soil niche effect on species diversity and catabolic activities in an ectomycorrhizal fungal community. Soil Biology and Biochemistry, 39: 1947-1955. Buee, M., Maurice, J.P., Zeller, B., Andrianarisoa, S., Ranger, J., Courtecuisse, R., Marçais, B. and Le Tacon, F. (2011) Influence of tree species on richness and diversity of epigeous fungal communities in a French temperate forest stand. Fungal Ecology, 4: 22-31. Burnham, K.P. and Overton, W.S. (1978) Estimation of the size of a closed population when capture probabilities vary among animals. Biometrika, 65: 623-633. Burton, R.F. (1998) Biology by Numbers. Cambridge: Cambridge university press. Carey, M., McCarthy, R. and Miller, H. (1988) More on nursing mixtures. Irish

Forestry, 45: 7-20.

355

Carlisle, M. J., and Watkinson, S. C. (1994) The fungi. London: Academic publishers. Carrier, P. (2003) Effects of water addition on biotic and abiotic components of a

dry boreal forest in the Yukon. Unpublished PhD. Thesis, Department of Zoology, University of British Columbia, Vancouver, B.C. Carter, G. (1998) The mycoflora of natural and semi-natural grasslands in

Ireland. Unpublished MSc. thesis, University College Dublin, Dublin. Chao, A. (1984) Nonparametric estimation of the number of classes in a population. Scandanavian Journal of Statistics, 11: 265-270. Chao, A. (1987) Estimating the population size for capture recapture data with unequal catchability. Biometrics, 43: 783-791. Chao, A. and Lee, S.M. (1992) Estimating the number of classes via sample coverage. Journal of the American Statistical Association, 87: 210-217. Chao, A., Chazdon, R.L., Colwell, R.K. and Shen, T. (2005) A new statistical approach for assessing similarity of species composition with incidence and abundance data. Ecology Letters, 8: 148–159. Chazdon, R.L., Colwell, R.K., Denslow, J.S., and Guariguata, M.K. (1998) Statistical methods for estimating species richness of woody regeneration in primary and secondary rain forests of northeastern Costa Rica. In: Forest

Biodiversity Research Monitoring, and Modeling: Conceptual Background and

Old World Case Studies, Dallmeier, F. and Comisky, J.A. (eds.). Paris: Parthenon Publishing. Christensen, M. and Emborg,J. (1996) Biodiversity in natural versus managed forests in Denmark. Forest Ecology and Management, 85: 47-51. Clapp, J.P., Young, J.P.W., Merryweather, J.W., and Fitter, A.H. (1995) Diversity of fungal symbionts in arbuscular mycorrhizas from a natural community. New

Phytologist, 130: 259-265. Clarke, D.A. and Clarke, D.B. (1992) Life history diversity of canopy and emergent trees in a neotropical rain forest. Ecological Monographs, 62: 315-344. Cline, E. T., Ammirati, J. F. and Edmonds, R.L. (2005) Does proximity to mature trees influence ectomycorrhizal fungus communities of Douglas-fir seedlings?. New Phytologist, 166: 993–1009. Colgan, W., Carey, A., Trappe, J., Molina, R., and Thysell, D. (1999) Diversity and productivity of hypogeous fungal sporocarps in variably thinned douglas fir forests. Canadian Journal of Forest Research, 29: 1259-1268.

356

Coll, L., Messier, C., Delagrange, S., and Berninger, F. (2007) Growth, allocation and leaf gas exchanges of hybrid poplar plants in the establishment phase on previously forested sites: effect of different vegetation management techniques. Annuls of Forest Science, 64: 1-11. Coll, M.T. and Bolger, T. (2007) Biodiversity and species composition of caribidae in irish coniferous forests: additional insight from the use of paored sites in comparison with the fauna of open habitats. Biology and Environment: Proceedings of the Royal Irish Academy, 107B: 1-11. Colwell, R.K. (2004) EstimateS: statistical estimate of species richness and shared species from samples, Version 7. [Online], available: http://viceroy.eeb.uconn.edu/estimates [accessed: 23/03/2011]. Colwell, R. K., and Coddington, J. A. (1994) Estimating terrestrial biodiversity through extrapolation. Philosophical Transactions of the Royal Society, 345B: 101-118. Colwell, R.K. (2009) Estimate S, users guide. [Online], available: http://viceroy.eeb.uconn.edu/EstimateSPages/EstSUsersGuide/EstimateSUsersGuide.htm [accessed 06/04/2011]. Colwell, R.K., Mao, C.X. and Chang, J. (2004) Interpolating, extrapolating, and comparing incidence-based species accumulation curves. Ecology, 85: 2717–2727. Coote, L., Smith, G.F., Kelly, D.L., O’Donoghue, S., Dowding, P., Iremonger, S. and Mitchell, F.J.G. (2008) Epiphytes of Sitka spruce (Picea sitchensis) plantations in Ireland and the effects of open spaces. Biodiversity and

Conservation, 17: 953–968. Courty, P., Franc, A., Pierrat, J. and Garbaye, J. (2008) Temporal Changes in the Ectomycorrhizal Community in Two Soil Horizons of a Temperate Oak Forest. Applied and Environmental Microbiology, 74: 5792–5801. Cripps, C. (2004) Ectomycorrhizal fungi above and below ground in a small Aspen stand: a simple system reveals fungal fruiting strategies and an edge effect. In: Fungi in Forest Ecosystems: Systematics, Diversity and Ecology, Cripps, C. (ed.). New York: New York Botanical Gardens. Crites, S. and Dale, M.R.T. (1998) Diversity and abundance of bryophytes, lichens, and fungi in relation to woody substrate and successional stage in aspen mixed wood boreal forests. Canadian Journal of Botany, 76: 641-651. Cross, J.R. (1998) An outline and map of the potential natural vegetation of Ireland. Applied Vegetation Science, 1: 241-245.

357

Cross, J.R. (2006) The potential natural vegetation of Ireland. Biology and

Environment: Proceedings of the Royal Irish Academy, 106B: 65-116. Cumming, J., R. (1996) Phosphate-limitation physiology in ectomycorrhizal pitch pine (Pinus rigida) seedlings. Tree Physiology, 16: 977-983. Curtis, T. G. F & McGough, N. (1988) The Irish Red Data Book: 1 Vascular

Plants. Wildlife Service Ireland, Dublin: The Stationery Office. Dahlberg A., Genny, D.R., Heilmann-Clausen, J. (2009) Developing a comprehensive strategy for fungal conservation in Europe: current status and future needs. Fungal Ecology, doi:10.1016/j.funeco.2009.10.004. Dahlberg, A., Jonssen, L. and Nylund, J (1997) Species diversity and distribution of biomass above and below ground among ectomycorrhizal fungi in an old-growth Norway spruce forest in south Sweden. Canadian Journal of Botany, 75: 1323- 1335. Danielson, R.M. (1984) Ectomycorrhizal associations in Jack pine stands in North eastern Alberta. Canadian Journal of Botany, 62: 932-939. Day, R. W. and G. P. Quinn. (1989) Comparisons of treatments after an analysis of variance in ecology. Ecological Monographs, 59: 433-463. De vries, B., Jansen, E., and Kuyper, T. (1995) Partial restoration of fungal and plant species diversity by removal of litter and humus layers in stands of Scot’s pine in the netherlands. Biodiversity and Conservation, 4: 156-164. Deacon, J.W. and Fleming, L.V. (1992) Interactions of ectomycorrhizal fungi. In: Mycorrhizal functioning: An integrative plant fungal process, Allen, M.F. (ed.). New York: Chapman and Hall. Dennis, R. W. G. (1986) Fungi of the Hebrides. Kew: Royal Botanic Gardens. Dickie, I.A, Bing, X., Koide, R. (2002) Vertical niche differentiation of ectomycorrhizal hyphae in soil as shown by T-RFLP analysis. New Phytologist, 156: 527–535. Dickie, I.A. and Reich, P.B. (2005) Ectomycorrhizal fungal communities at forest edges. Journal of Ecology, 93(2): 244-255. Dickie, I.A., Avis, P.G., McLaughlin, D.J. and Reich, P.B. (2003) Good-enough RFLP matcher (GERM) program. Mycorrhiza, 13: 171–172. Diekmann, M. (1994) Deciduous forest vegetation boreonemoral Scandinavia. Acta Phytogeographica Suecica, 80: 1-112. Diekmann, M. (1999) Southern deciduous forests. Acta Phytogeographica

Suecica, 84: 33-53.

358

Dighton, J., Poskitt, J.M., and Howard, D.M. (1986) Changes in occurrence of basidiomycete fruit bodies during forest stand development: with specific reference to mycorrhizal species. Transactions of the British Mycological Society, 87: 163-171. Dixon, K.W., Kell, S.P., Barrett, R.L. and Cribb, P.J. (2003) Orchid conservation. Kota Kinahalu: Natural History Publications. Duchesne, L.C., Peterson, R.L., and Ellis, B.E. (1998) Pine root exudates stimulate the synthesis of antifungal compounds by the ectomycorrhizal fungus Paxillus involutus. New Phytologist, 111: 693-698. Dufrene, M. and Legendre, P. (1997) Species assemblages and indicator species: the need for a flexible asymmetrical approach. Ecological Monographs, 67: 345-366. Durall, D.M., Gamiet, S., Simard, S.W., Kudrna, L. and Sakakibara, S.M. (2006) Effects of clearcut logging and tree species composition on the diversity and community composition of epigeous fruitbodies formed by ectomycorrhizal fungi. Canadian Journal of Botany, 84: 966-980. Edmonds, R. and Lebo, D. (1998) Diversity, production, and nutrient dynamics of fungal sporocarps on logs in an old-growth temperate rain forest, Olympic National Park, Washington. Canadian Journal of Forest Research, 28: 665-673. EEA (2010) Assessing biodiversity in Europe - the 2010 report. Luxemburg: European Environment Agency. Egger, K.N., Danielson, R.M., and Fortin, J.A. (1991) Taxonomy and population structure of E-strain mycorrhizal fungi inferred from ribosomal and mitochondria DNA polymorphisms. Mycological Research, 95: 866-872. Engelmark, O. and Hytteborn, H. (1999) Coniferous forests. Acta

Phytogeographica Suecica, 84: 33-55. English Nature (1998) The conservation of wild mushrooms. The wild mushroom

pickers code of conduct. Peterborough, U.K.: English Nature. Enttema, C.H. and Wardle, D.A. (2002). Spatial soil ecology. Trends in Ecology

and Evolution, 17: 177-183. Erland, S. and Taylor, A.F.S. (2002) Diversity of ectomycorrhizal communities in relation to the abiotic environment. In: The Ecology of Mycorrhizas, Ecological Studies Series. Volume. 157, M. Van der Heijden and . Sanders, I. (eds.). Germany: Springer-Verlag, pp163-200.

359

Esseen, P. A., Ehnstrdm, B., Ericson, L. and Sjoberg, K. (1992) Boreal forests - the focal habitats of Fennoscandia. In Ecological principles of Nature

conservation, Applications in temperate and boreal environments, L. Hansson (ed.). London : Elsevier Applied Science, pp. 252-325. EUROPA (2009). Annexed habitats, Atlantic region. [Online], available from: http://ec.europa.eu/environment/nature/natura2000/sites_hab/biogeog_regions/docs/ref_list_atlantic.pdf [accessed 11/05/2011]. Evans, S.E., Ing, B., Henrici, A. and Rotheroe, M. (2003) A Provisional Red Data

List of British Fungi. [Online], available: http://www.britmycolsoc.org.uk/files/RDL_of_Threatened_British_Fungi_Version_1.0.pdf [accessed 25/02/10]. Fahy, O. and Gormally, M. (1998) A comparison of plant and carabid beetle communities in an Irish oak woodland with a nearby conifer plantation and clearfelled site. Forest Ecology and Management, 110: 263-273. Felton, A., Lindbladh, M., Brunet, J. and Fritz, O. (2010) Replacing coniferous monocultures with mixed-species production stands: an assessment of the potential benefits for forest biodiversity in northern Europe. Forest Ecology and

Management, 260: 939–947. Fernandez, L.M., Agreda, T. and Olano, J.M. (2006) Stand age and sampling year effect on the fungal fruit body community in Pinus pinaster forests in central Spain. Canadian Journal of Botany, 84: 1249-1258. Ferris, R. Humphrey, J.W. (1999) A review of potential biodiversity indicators for application in British forests. Forestry, 72: 313-328. Ferris, R., Peace, A.J., Humphrey, J.W., Broome, A.C. (2000b) Relationships between vegetation, site type and stand structure in coniferous plantations in Britain. Forest Ecology and Management, 136: 35-51. Ferris, R., Peace, A.J., and Newton, A.C. (2000a) Macro fungal communities of low land Scot’s pine and Norway spruce plantations in England: relationships with site factors and stand structure. Forest Ecology and Management, 131: 255-267. Ferris-Kaan, R., Peace, A.J., and Humphreys, J.W. (1998) Assessing forest structure in managed forests. In: Assessment of forest biodiversity for improved

forest management, P. Bachmann, M. Kohl and R. Paivainenn (eds.). Dordrecht: Kluwer, pp 331-342. Fossitt, J.A. (2000) A guide to habitats in Ireland. [Online], available: http://www.heritagecouncil.ie/publications/habitats/Guide_to_Habitats.pdf [accessed 03/01/2009].

360

FRA (2010) Global Forest Resource Assessment 2010 Main report. Rome: Food and Agriculture Organisation. Franklin, J.F. and DeBell, D.S. (1973) Effects of various harvesting methods on forest regeneration. In: Proceedings of the Symposium on Even-Age Management,

School of Forestry, Oregon State University, Corvallis, 1 August 1972, Hermann, R.K., Lavender, D.P. (Eds.), pp. 29–57. Franklin, J.F., Shugart, H.H. and Harmon, M.E. (1987) Tree death as an ecological process. The causes, consequences, and variability of tree mortality. Bioscience, 37, 550-556. FRDBI (2010) Fungal records database of Britain and Ireland. [Online], available: http://www.fieldmycology.net/FRDBI/FRDBI.asp [accessed 23/03/2011]. French, L.J., Smith, G.F., Kelly, D.L., Mitchell, F.J.G., O’Donoghue, S., Iremonger, S.F., and McKee, A. (2008) Ground flora communities in temperate oceanic plantation forests and the influence of silvicultural, geographic and edaphic factors. Forest Ecology and Management, 255: 476-494. Friend, D.T. (1961) A simple method of measuring integrated light values in the field. Ecology, 42: 577-580. Gabel, A.C. and Gabel, M.L. (2007) Comparison of Diversity of Macrofungi and Vascular Plants at Seven Sites in the Black Hills of South Dakota. American

Midlands Naturalist, 157: 258-296. Gardener, M.J. and Radford, T. (1980) Soil Associations of Ireland and Their

Land Use Potential. Dublin, Ireland: An Foras Talúntais. Gardes, M., and Bruns, T.D. (1993) ITS primers with enhanced specificity for basidiomycetes - application to the identification of mycorrhizae and rusts. Molecular Ecology, 2: 113-118. Gardes, M., Bruns, T.D. (1996) Community structure of ectomycorrhizal fungi in a Pinus muricata forest: above- and below-ground views. Canadian Journal of

Botany, 74: 1572–1583. Garnica, S., Weis, M., Walther, G., and Oberwinkler, F. (2007) A framework for a phylogenetic classification in the genus Cortinarius (Basidiomycota, Agaricales) derived from morphological and molecular data. Canadian Journal of Botany, 83: 1457-1477. Garrison, G.A. (1949) Uses and modifcations for the Moosehorn crown closure estimator. Journal of Forestry, 47: 733-735. Gaston, K.J. (1994) Rarity. London: Chapman and Hall.

361

Gaston, K.J. and Spicer, J.I. (2004) Biodiversity: an introduction. Oxford: Blackwell publishing. Gebhardt, S., Neubert, K., Wöllecke, J., Münzenberger, B. and Hüttl, R.F. (2007) Ectomycorrhiza communities of red oak (Quercus rubra L.) of different age in the Lusatian lignite mining district, East Germany. Mycorrhiza, 17: 279–290. Gherbi, H., Delaruelle, C., Selosse, M.A. and Martin, F. (1999) High genetic diversity in a population of the ectomycorrhizal basidiomycete Laccaria

amethystina in a 150-year-old beech forest. Molecular Ecology, 8: 2003–2013. Gibbs, R.D. (1957) Patterns in the seasonal water content of trees. In: The

Physiology of Forest Trees, K.V. Thimann, W.B. Critchfield and M.H. Zimmerman (Eds.). New York: Ronald Press Co. Gill, G.E., Pearson, A.A., Dennis, R.W.G., Allen, C.W. and O’Connor, P. (1948) The fungal flora of county Kerry. Irish Naturalists’ Journal, 9: 181-185. Goodman, D.M. and Trofymow, J.A. (1998) Comparison of communities of ectomycorrhizal fungi in old growth and mature stands of Douglas fir at two sites on southern Vancouver Island. Canadian Journal of Forest Research, 28: 574-581. Goodman, D.M., Durall, D.M., Trofymow, J.A., and Berch, S.M. (1997) Concise

Descriptions of North American Ectomycorrhizae . Victoria, B.C.: Mycologue Publications. [Online], available: http://www.cfp.scf.rncan.gc.ca/biodiversity/bcern/manual/index_e.html [accessed 11/12/2008]. Grayston, S.J., Addison, J.A., Berch, S.M., DeMontigny, L.E., Durall, D.M., Egger, K., N., Jones, M.D., Modesto, R., Mohn, W.W., Panesar, T.S., Prescott, C.E. and Srivastava, D.S. (2006) Green tree retention: A tool to maintain soil function after harvest. BC Journal of Ecosystems and Management, 7: 9-12. Grebenc, T., Christensen, M., Vilhar, U., Cater, M., Martín, M.P., Simoncic, P. and Kraigher, H. (2009) Response of ectomycorrhizal community structure to gap opening in natural and managed temperate beech-dominated forests . Canadian

Journal of Forest Research, 39: 1375-1386. Griffith, G.W., Bratton, J.H. and Easton, G. (2004) Charismatic megafauna, the

conservation of waxcap grasslands. Britain: British wildlife. Grubisha, L.C., Bergemann, S.E. and Bruns, T.D. (2007) Host islands within the California northern Channel Islands create fine-scale genetic structure in two sympatric species of the symbiotic ectomycorrhizal fungus Rhizopogon. Molecular Ecology, 16: 1811–1822.

362

Gulden, G., Hoiland, K., Bendiksen, K., Brandrud, T., Foss, B., Jenssen, H., and Laber, D. (1992) Macromycetes and air pollution. Bibliotheca Mycologia. Berlin: Cramer, pp.144. Gunnar-Jonsson, B. and Jonsell, M. (1999) Exploring potential biodiversity indicators in boreal forests. Biodiversity and Conservation, 8: 1417-1433. Hall, T.A. (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series, 41: 95–98. Hall, T.A. (2005) BioEdit v7.0.5. [Online], available: http://www.mbio.ncsu.edu/bioedit/page1.html [accessed 07/04/2011]. Halme, P., Kotiaho, J., Ylisirniö, A.-L., Hottola, J., Junninen, J., Kouki, J., Lindgren, M., Mönkkönen, M., Penttilä, R., Renvall, P., Siitonen, J. and Similä, M. (2008). Perennial polypores as indicators of annual and red-listed polypores. Ecological Indicators, 9: 256–266. Hangs, R.D., Greer, K.J. and Sulewski, C.A. (2004) The effect of interspecific competition on conifer seedling growth and nitrogen availability measured using ion-exchange membranes. Canadian Journal of Forest Research, 34: 754-761. Hanson, P.A. (1988) Prediction of macrofungai occurrence in Swedish beech forests from soil and litter variable models. Vegetation, 78: 31-44. Harley, J.L. and Harley, E.L. (1987) A check-list of mycorrhiza in the British flora. New Phytologist, 105: 1–102. Harrington, T.J (1994) New Irish fungal records. Irish Naturalists’ Journal, 24: 413-414. Harrington, T.J. (1996) Observations on the macrofungi of Dryas communities in the Burren, Western Ireland. Irish Naturalists’ Journal, 25: 157-192. Harrington, T.J. (2003) Relationship between macrofungi and vegetation in the Burren. Biology and Environment: Proceedings of the Royal Irish Academy, 103B: 147-159. Harrington, T.J. and Mitchell, D.T. (2002) Characterization of Dryas octopetala ectomycorrhizas from limestone karst vegetation, Western Ireland. Canadian

Journal of Botany, 80: 970-982. Harrington, T.J. and Mitchell, D.T. (2005a) Ectomycorrhizas associated with a relict population of Dryas octopetala in the Burren, western Ireland I. Distribution of ectomycorrhizas in relation to vegetation and soil characteristics. Mycorrhiza, 15: 425-433.

363

Harrington, T.J. and Mitchell, D.T. (2005b) Ectomycorrhizas associated with a relict population of Dryas octopetala in the Burren, western Ireland II. Composition, structure and temporal variation in the ectomycorrhizal community. Mycorrhiza, 15: 435-445. Hartley, J., Cairney, J.W.G. and Meharg, A.A. (1997) Do ectomycorrhizal fungi exhibit adaptive tolerance to potentially toxic metals in the environment. Plant

Soil, 189: 303-319. Hassell, F.C. (1953) The fungi of North Bull Island, Co Dublin. Irish Naturalists’

Journal, 2: 3-7. Hawkins, D. (2009) Biomeasurement : a student's guide to biological statistics. Oxford: Oxford University Press. Hawksworth, D. L. (1991) The fungal dimension of biodiversity: magnitude, significance, and conservation. Mycological Research, 95: 641–655. Hawksworth, D.L. and Mueller, G.M. (2005) Fungal communities: their diversity and distribution. In: The fungal community: its organization and role in the

ecosystem, Dighton. J., White, J.F. and Oudemans, P. (eds) 3rd edn. Boca Raton: Taylor & Francis. Hawksworth, D.L. and Rossman, A.Y. (1997) Where are all the undescribed fungi. Phytopathology, 87: 888-891. Heilmann-Clausen, J. and Christensen, M. (2005) Wood inhabiting fungi in Danish beech forests conflicting diversity patterns and their implications in a conservation perspective. Biological Conservation, 122: 633-642. Heilmann-Clausen, J., Christensen, M. (2003) Fungal diversity on decaying beech logs – implications for sustainable forestry. Biodiversity and Conservation, 12: 953–973. Heilmann-Clausen, J., Verbeken, A. and Vesterholt, J. (1998) The genus

Lactarius. Fungi of Northern Europe, Vol. 3. Denmark: Svampetryk. Heinonsalo, J., Koskiahde, I. and Sen, R. (2007) Scot’s pine bait seedling performance and root colonizing ectomycorrhizal fungal community dynamics before and during the 4 years after forest clear-cut logging, Canadian Journal of

Forest Research, 37: 415–429. Heinze, E., Boch, S., Fischer, M., Hessenmoller, D., Klenk, B., Muller, J., Prati, D., Detlef Schulze, E., Seele, C., Socher, S., Halle, S. (2011) Habitat use of large ungulates in northeastern Germany in relation to forest management. Forest

Ecology and Management, 261: 288-296.

364

Hermy, M., Honnay, O., Firbank, L., Grashof-Bokdam, C.J. & Lawesson, J.-E. (1999) An ecological comparison between ancient and other forest plant species of Europe and the implications for forest conservation. Biological Conservation, 91: 9-22. Hering, T. (1966) The terricolous higher fungi of the four Lake District woodlands. Transactions of the British Mycological Society, 49: 369-383. Heslin, M.C., Blasius, D., McElhinney, C. and Mitchell, D.T. (1992) Mycorrhizal and associated fungi of Sitka spruce in Irish forest mixed stands. European

Journal of Forest Pathology, 22: 46-57. Hibbett, D.S., Ohman, A., Glotzer, D., Nuhn, M., Kirk, P. and Nilsson, R.H. (2011) Progress in molecular and morphological taxon discover in Fungi and options for formal classification of environmental sequences. Fungal Biology

Reviews, 25: 38-47. Hickey, K.R. (2000) A geographical perspective on the decline and extermination of the Irish wolf Canis lupus— an initial ssessment. Irish Geography, 33: 185-198. Hill, M.O. (1979) The development of a flora in even-aged plantations. In: The

Ecology of Even-Aged Forest Plantations, E.D. Ford, D.C. Malcolm and J. Atterson (eds). Cambridge: Institute of Terrestrial Ecology Hill, M.O. (1986) Ground flora and succession in commercial forests. In: Trees

and Wildlife in the Scottish Uplands, Jenkins, D. (Ed.). Cambridge: Institute of Terrestrial Ecology. Hobbie, E.A. (2006) Carbon allocation to ectomycorrhizal fungi correlates with total belowground allocation in culture studies. Ecology, 87: 563–569. Hobbs, R.J., Arico S. Aronson J., Jill S. Baron, J.S., Bridgewater, P., Cramer, V.A., Epstein, P.R., Ewel, J.J., Klink, C.A., Lugo, A.E., Norton, D., Ojima, D., Richardson, D., Sanderson, E.W., Valladares, F., Vilà, M., Zamora, R. and Zobel, M. (2006) Novel ecosystems: theoretical and management aspects of the new ecological world order. Global Ecology and Biogeography, 15: 1–7. Hoiland, K., and Bendiksen, E. (1996) Biodiversity of wood inhabiting fungi in a boreal coniferous forest in Sor-Trondelag county, Central Norway. Nordic

Journal of Botany, 16: 643-659. Horgan, T., Keane, M., McCarthy, R., Lally, M. and Thompson, D. (2004) A

guide to forest tree species selection and silviculture in Ireland. Dublin: COFORD. Horton, T.R. and Bruns, T.D. (2001) The molecular revolution in ectomycorrhizal ecology: Peeking into the black-box. Molecular Ecology, 10: 1855-1871.

365

Huang, W.Z. and Schoenau, J.J. (1996) Forms, amounts and distribution of carbon, nitrogen, phosphorus, and sulfur in a boreal aspen forest soil. Canadian

Journal of Soil Science, 76: 373-385. Hubert, N.A. and Gehring, C.A. (2008) Neighboring trees affect ectomycorrhizal fungal community composition in a woodland-forest ecotone. Mycorrhiza, 18: 363-374. Huhta, V., Karppinen, E. Nurminen, M. and Valpas, A. (1967) Effect of silvicultural practices upon arthropod, annelid and nematode populations in coniferous forest soil. Annales Zoologici Fennici, 4: 87-143. Huhta, V., Nurminen, M. and Valpas, A. (1969). Further notes on the effect of silvicultural practices upon the fauna of coniferous forest soil. Annales Zoologici

Fennici, 6: 327-334. Huhta, V. (1971). Succession in the spider communities of the forest floor after clear-cutting and prescribed burning. Annales Zoologici Fennici, 8: 483-542. Huhta, V. (1976) Effects of clear-cutting on numbers, biomass and community respiration of soil invertebrates. Annales Zoologici Fennici, 13: 63-80. Humphrey, J.W. (2005) Benefits to biodiversity from developing old-growth conditions in British upland spruce plantations: a review and recommendations. Forestry, 78: 33-53. Humphrey, J.W. and Coombs, E.L. (1997) Effects of forest management on understorey vegetation in a Pinus sylvestris L. Plantation in NE Scotland. Botanical Journal of Scotland, 49: 479- 488. Humphrey, J.W., Ferris, R. and Quine, C. (2003) Biodiversity in Britain’s planted

forests. Edinburgh: Forestry commission. Humphrey, J.W., Holl, K., Broome, A.C., (1998) Birch in Spruce Plantations:

Management for Biodiversity. Forestry Commission Technical Paper 26. Edinburgh: Forestry Commission. Humphrey, J.W., Newton, A.C., Peace, A.J., and Holden, E. (2000) The importance of conifer plantations in northern Britain as a habitat for native fungi. Biological Conservation, 96: 241-252. Ing, B. (1992) A provisional red data list of British fungi. Mycologist, 6: 124-128. Ingleby, K., Mason, P.A., Last, F.T. and Fleming L.V. (1990) Identification of

Ectomycorrhizas. Institute of terrestrial ecology. Research publication no 5. London: Her Majesty’s Stationary Office. Iremonger, S. (1999) Guidelines for Forestry and Biodiversity. Dublin: Forest Service, Department of the Marine and Natural Resources.

366

Iremonger, S., Gittings, T., Smith, G.F., Wilson, M.W., Oxbrough, A., Coote, L., Pithon, J., O'Donoghue, S., McKee, A., O'Halloran, J., Kelly, D.L., Giller, P.S., O'Sullivan, A., Neville, P., Mitchell, F.J.G., O'Donnell, V., Kelly, T.C. and Dowding, P. (2006) Investigation of experimental methods to enhance biodiversity in plantation forests. BIOFOREST Project 3.1.3 Final Report to COFORD. [Online], available: http://bioforest.ucc.ie/pages/project_three.htm [Accessed 09/03/09]. Ishida, T.A., Nara, K. and Hogetsu, T. (2007) Host effects on ectomycorrhizal fungal communities: insight from eight host species in mixed conifer–broadleaf Forests. New Phytologist, 174: 430–440. Jennings, S.B., Brown, N.D. and Sheil, D. (1999) Assessing forest canopies and understory illumination: canopy closure, canopy cover and other methods. Forestry, 72: 59-73. Jiang, Q.Q. and Bakken, L.R. (1999) Comparison of Nitrosospira strains isolated from terrestrial environments. FEMS Microbial Ecology, 30: 171-186. Johnson, D.W., Hungate, B.A., Dijkstra, P., Hymus, G. and Drake, B. (2001) Landscape and Watershed Processes: Effects of Elevated Carbon Dioxide on Soils in a Florida Scrub Oak Ecosystem. Journal of Environmental Quality, 30: 501-507. Jones, M.D, Durall, D.M., and Cairney, J. (2003) Ectomycorrhizal fungal communities in young forest stands regenerating after clearcut logging. New

Phytologist, 157: 399-422. Jongmans, J.G., Van Breeman, N., Lundstrom, U., Van Hees, P.A., Finlay, R.D., Srinivasan, M., Unestam, T., Giesler, R., Melkerud, P.A. and Olsson, M. (1997) Rock eating fungi. Nature, 389: 682-683. Jonsson, L., Dahlberg, A., Nilsson, M.C., Zackrisson, O., and Karen, O. (1999) Ectomycorrhizal fungal communities in late successional Swedish boreal forests and their composition following wildfire. Molecular Ecology, 8: 205-215. Jowkin, V. and Schoenau, J.J. (1998) Impact of tillage and landscape position on nitrogen availability and yield of spring wheat in the brown soil zone in southwestern Saskatchewan. Canadian Journal of Soil Science, 78: 563-572. Joyce, P. and O’Carroll, N. (2002) Sitka spruce in Ireland. Dublin: COFORD. Jumpponen, A., Trappe, J.A., Cazares, E. (2002) Occurrence of ectomycorrhizal fungi on the forefront of retreating Lyman Glacier (Washington, USA) in relation to time since deglaciation. Mycorrhiza, 12: 43-49. Kaldorf, M., Renker, C., Fladung, M. and Buscot, F. (2004) Characterization and spatial distribution of ectomycorrhizas colonizing aspen clones released in an experimental field. Mycorrhiza, 14: 295-306.

367

Kappes, H., Catalano, C., Topp, W. (2007) Coarse woody debris ameliorates chemical and biotic soil parameters of acidified broad-leaved forests. Applied Soil

Ecology, 36: 190–198. Keil, P., Dziock, F., and Storch, D. (2008) Geographical patterns of hoverfl y (Diptera, Syrphidae) functional groups in Europe: inconsistency in environmental correlates and latitudinal trends. Ecological Entomology. 33:748-757 Keil, P., Dziock, F., and Storch, D. (2008) Geographical patterns of hoverfly (Diptera, Syrphidae) functional groups in Europe: inconsistency in environmental correlates and latitudinal trends. Ecological Entomology, 33: 748-757. Kelly, D.L. and Iremonger, S.F. (1997) Irish wetland woods: the plant communities and their ecology. Biology and Environment: Proceedings of the

Royal Irish Academy 97B: 1-32. Kelly, D.L. and Kirby, E.N. (1982). Irish native woodlands over limestone. In: Studies on Irish Vegetation (ed. J. White). Dublin: Royal Dublin Society, pp. 181-198. Kelly, D.L. and Moore, J.J. (1975) A preliminary sketch of the Irish acidophilous oakwoods. In: Colloques Phytosociologiques III: La Vegetation des Forets

Caducifoliees Acidophiles (ed. J. M. Gehu). Vaduz: J. Cramer, pp. 375-387. Kelly, D.L. (2002) The regeneration of Quercus petraea. (Sessile oak) in southwest Ireland: a 25-year experimental study. Forest Ecology and

Management, 166: 207-226. Kelly, D.L. (2005) Woodland on the Western Fringe: Irish Oakwood Diversity and the Challenges of Conservation. Botanical Journal of Scotland, 57: 21-40. Kelly, D.L. and Kirby, E.N. (1982) Irish native woodlands over limestone. Journal of Life Science Royal Dublin Society, 3: 181–198. Kennedy, P. (2010) Ectomycorrhizal fungi and interspecific competition: species interactions, community structure, coexistence mechanisms, and future research directions. New Phytologist, 187: 895-910. Kennedy, P.G. and Bruns, T.D. (2005) Priority effects determine the outcome of ectomycorrhizal competition between two Rhizopogon species colonizing Pinus

muricata seedlings. New Phytologist, 166: 631-638. Kent, M. and Coker, P. (1992) Vegetation description and analysis: a practical

approach. London: Belhaven Press. Kernaghan, G. and Harper, K. A. (2001) Community structure of ectomycorrhizal fungi across an alpine/subalpine ecotone. Ecography, 24: 181–188.

368

Kilbride, C.A., Byrne, K.A. and Gardiner, J.J. (1999) Carbon sequestration and

Irish forests. Dublin: COFORD. Kirby, K.J., Reid, C.M., Thomas, R.C. and Goldsmith, F.B. (1998) Preliminary Estimates of Fallen Dead Wood and Standing Dead Trees in Managed and Unmanaged Forests in Britain. Journal of Applied Ecology, 35: 148-155. Kirk, P., Cannon, P. and Stalpers, J. (2008) Dictionary of the Fungi, 10th ed. Wallingford: CABI. Kjoller, R. and Clemmensen, K.E. (2009) Belowground ectomycorrhizal fungal communities respond to liming in three southern Swedish coniferous forest stands. Forest Ecology and Management, 257: 2217-2225. Klironomos, J.N., and Hart, M.M. (2001) Food web dynamics- animal nitrogen swap for plant carbon. Nature, 410: 651-652. Kohl, M. (1996) Assessing and monitoring forest biodiversity in Switzerland and Germany. In: Assessment of Biodiversity for Improved Forest Management. EFl Proceedings No 6. Joensuu: European Forest Institute, pp. 95-104. Kohm, K.A., and Franklin, J.F. (1997) Creating a forestry for the 21st century. Washington, D.C.: Island Press. Kõljalg, U., Larsson, K-H,, Abarenkov, K., Nilsson, R.H., Alexander, I.J., Eberhardt, U., Erland, S., Høiland, K., Kjøller, R., Larsson, E. Pennanen, T., Sen, R., Taylor, A.F.S, Tedersoo, L., Vrålstad, T. and Ursing, B.M. (2005) UNITE: a database providing web-based methods for the molecular identification of ectomycorrhizal fungi. New Phytologist, 166: 1063–1068. Korhonen, K., Capretti, P., Karjalainen, R., Stenlid, J. (1998) Distribution of Heterobasidion annosum intersterility groups in Europe. In: Heterobasidion

annosum. Biology, Ecology, Impact and Control. Woodward, S.; Stenlid, J.; Huttermann, A.; Karjalainen and R. Oxon (eds). New York: CAB International. Kranabetter , J.M. and Wylie, T. (1998) Ectomycorrhizal community structure across forest openings on naturally regenerated western hemlock seedlings. Canadian Journal of Botany, 76: 189-196. Kranabetter, J.M. (1999) The effect of refuge trees on a paper birch ectomycorrhiza community. Canadian Journal of Botany, 77: 1523–1528. Kranabetter, J.M. (2004) Ectomycorrhizal community effects on hybrid spruce seedling growth and nutrition in clearcuts. Canadian Journal of Botany, 82: 983-991.

369

Kranabetter, J.M., Friesen, J., Gamiet, S., Kroeger, P. (2005) Ectomycorrhizal mushroom distribution by stand age in western hemlock lodgepole pine forests of northwestern British Columbia. Canadian Journal of Forest Research, 35: 1527–1539. Krannabetter, J.M., Hayden, S., and Wright, E.F. (1999) A comparison of ectomycorrhiza communities from three conifer species planted. Canadian

Journal of Botany, 77: 1193-1198. Krebs, C.J., Carrier, P., Boutin, S., Boonstra, R., and Hofer, E. (2008) Mushroom crops in relation to the weather in southwestern Yukon. Botany, 86: 1497-1502. Kretzer, A.M., Dunham, S., Molina, R. and Spatafora, J.W. (2004) Microsatellite markers reveal the below ground distribution of genets in two species of Rhizopogon forming tuberculate ectomycorrhizas on Douglas fir. New

Phytologist, 161: 313–320. Kruys, N., Fries, C., Gunnar Jonsson, B., Lamas, T. and Stahl, G. (1999) Wood-inhabiting cryptogams on dead Norway spruce (Picea abies) trees in managed Swedish boreal forests. Canadian Journal of Forest Research, 29: 178–186. Lagana, A., Loppi, S. and De Dominicis, V. (1999) Relationship between environmental factors and the proportions of fungal trophic groups in forest ecosystems of the central Mediterranean area. Forest Ecology and Management, 124: 145-151. Landeweert, R., Hoffland, E., Finlay, R.D., Kuyper, T.W. and van Breemen, N. (2001) Linking plants to rocks: ectomycorrhizal fungi mobilize nutrients from minerals. Trends in Ecology and Evolution, 16: 248-254. Landy, E.T. (2001) A study of the fungi present on temperate salt marsh at North

Bull Island, Co. Dublin. Unpublished PhD thesis, University College Dublin. Lang, R. (1915). Versuch einer exakten Klassifikation der Boden in klimatischer und geologischer Hinsicht. Internationalen Mitteilungen fur Bodenkunde 5, 312. Last, F.T., Dighton. J. and Mason. P.A. (1987). Successions of sheathing Mycorrhizal fungi. Trends in Ecology and Evolution, 2: 157-161. Leake, J.R., and Read, D.J. (1997) Mycorrhizal fungi in terrestrial habitats. In: The Mycota, IV, Environmental and microbial relationships. Berlin: Springer-Verlag. Leake, J.R., Johnson, D., Donnelly, D., Muckle, G.E., Boddy, L. and Read D.J. (2004) Networks of power and influence: The role of mycorrhizal mycelium in controlling plant communities and agro-ecosystem functioning. Canadian Journal

of Botany, 82: 1016-1045.

370

Lee, K.J. and Lee, H. (2004) Fungal diversity and fruitbody production in relation to vegetation structure, topography and soil properties in a Quercus mongolica forest in central Korea. In: Fungi in forest ecosystems: Systematics, diversity and

ecology, Cripps, C. (ed). New York: The New York Botanical Gardens. Legon N.W. and Henrici A. (2005) Checklist of the British and Irish

Basidiomycota. Kew: Royal Botanic Gardens. Lehto, T. and Zwiazek, J.J. (2011) Ectomycorrhizas and water relations of trees: a review. Mycorrhiza, 21: 71–90. Liang, J. and Schoenau, J.J. (1995) Development of resin membranes as a sensitive indicator of heavy metal toxicity in the soil environment. International

Journal of Environmental and Analtical Chemistry, 59: 265-275. Lilleskov, E.A., Fahey, T.J., Horton, T.R. and Lovett, G.M. (2002) Belowground ectomycorrhizal fungal community change over a nitrogen deposition gradient in Alaska. Ecology, 83: 104-115. Lindenmayer, D. B., Fischer, J, Felton, A., Crane, M., Michael, R., Macgregor, C., Montague-Drake, R., Manning, A. and Hobbs, R.J. (2008) Novel ecosystems resulting from landscape transformation create dilemmas for modern conservation practice. Conservation Letters, 1: 129–135. Lindenmayer, D.B., Margules, C.R. and Botkin, D. (2000) Indicators of Biodiversity for Ecologically Sustainable Forest Management. Conservation

Biology, 14: 941–950. LoBuglio, K.L. and Taylor, J.W. (2002) Recombination and genetic differentiation in the mycorrhizal fungus Cenococcum geophilum Fr. Mycologia, 94: 772-780. Longino, J.T., Colwell, J. and Coddington, R.K. (2002) The ant fauna of a tropical rainforest: estimating species richness three different ways. Ecology, 83: 689-702. Loreau, M., Naeem, S. and Inchausti, P. (2007) Biodiversity and ecosystem

functioning. Oxford: Oxford University Press. Ludley, K.E., Robinson, C.H., Jickell, S., Chamberlain, P.M. and Whitaker, J. (2008) Differential response of ectomycorrhizal and saprotrophic fungal mycelium from coniferous forest soils to selected monoterpenes. Soil Biology and

Biochemistry, 40: 669–678. Lugo, A.E. (1997) The Apparent Paradox of Re-establishing Species Richness on Degraded Lands with Tree Monocultures. Forest Ecology and Management, 99: 9-19. Luoma, D.L., Eberhart, J.L., Molina, R., and Amaranthus, M.P. (2004) Response of ectomycorrhizal fungus sporocarp production to varying levels and patterns of green-tree retention. Forest Ecology and Management, 202: 337–354.

371

Luoma, D.L., Eberhart, J.L., Molina, R., Stockdale, C.A. (2006a) The spatial influence of Psuedotsuga menziessi retention trees on ectomycorrhizal diversity. Canadian Journal of Forest Research, 36: 2651-2573. Luoma, D.L., Eberhart, J.L., Abbott, R., Moore, A., Amaranthus, M.P., and Pilz, D. (2006b) Effects of mushroom harvest technique on subsequent American matsutake production. Forest Ecology and Management, 236: 65–75. MacAurthur, R. and Wilson, O. (1969) The Theory of Island Biogeography. Princeton: Princeton University Press. MacAurthur, R.H. (1964) Environmental factors affecting bird species diversity. American Nature, 98: 387-397. Magurran, A.E. (2004) Measuring Biological diversity. Oxford, UK: Wiley Publishing. Manachère, G., (1980) Conditions essential for controlled fruiting of macromycetes—a review. Transactions of the British Mycological Society, 75: 255–270. Mao, C.X., Colwell, R.K. and Chang, J. (2005) Estimating species accumulation curves using mixtures. Biometrics, 61: 433–441. Markkola, A.M., Ohtonen, A., Ahonen-Jonnarth, U. and Ohtonen, R. (1996) Scot’s pine responses to CO2 enrichment. I. Ectomycorrhizal fungi and soil fauna. Environmental Pollution, 94(3): 309-316. Mason, P.A., Last, F.T., Wilson, J. Deacon, J.W., Fleming. L.V. and Fox. F.M. (1987) Fruiting and succession of ectomycorrhizal fungi. In: Fungal infection of

plants, Pegg, G.F. and Ayres, P.G. (eds.). Cambridge: Cambridge University Press, pp. 253-268. Massicotte, H.B., Molina, R., Tackaberry, L.E., Smith, J.E. and Amaranthus, M.P. (1999) Diversity and host specificity of ectomycorrhizal fungi retrieved from three adjacent forest sites by five host species. Canadian Journal of Botany, 77: 1053-1076. McCarthy, B.C. and Bailey, D.R. (1996) Composition, structure, and disturbance history of Crabtree Woods: an old-growth forest of western Maryland. Bulletin of

the Torrey Botany Club, 123: 350-365. McComb, W. C. and Lindenmayer, D. (1999) Dying, Dead, and Down Trees. In: Maintaining Biodiversity in Forest Ecosystems, Hunter, M. L., Jr., (ed). Cambridge: Cambridge University Press. McCullagh, P. and Nelder, J. (1989) Generalized Linear Models, Second Edition. Boca Raton: Chapman and Hall/CRC.

372

McCune, B. and Grace, J.B. (2002) Analysis of ecological communities. Oregon: MjM Software Design. McHugh, R., Mitchel, D., Wright, M. and Anderson, R. (2001) The Fungi of Irish grasslands and their value for nature conservation. Biology and Environment:

Proceedings of the Royal Irish Academy, 101B: 225–242. Menkis, A., Vasiliauskas, R., Taylor, A.F., Stenlid, J. and Finlay, R. (2005) Fungal communities in mycorrhizal roots of conifer seedlings in forest nurseries under different cultivation systems, assessed by morphotyping, direct sequencing and mycelial isolation. Mycorrhiza, 16: 33–41. Millennium Ecosystem Assessment (2005) Ecosystems and Human Well-being:

Biodiversity Synthesis. Washington, DC: World Resources Institute. Mitchel, D. (2006) Survey of the Grassland Fungi of County Clare. Dublin: The Heritage Council. Mitchel, D. (2007) Survey of the Grassland Fungi of the Vice County of West

Cork. Dublin: The Heritage Council. Mitchel, D. (2008) Survey of the Grassland Fungi of the Vice County of West

Mayo. Dublin: The Heritage Council. Mitchel, D. (2009) Survey of the Grassland Fungi of the Vice County of West

Donegal. Dublin: The Heritage Council. Mitchell, F.J.G. (1988) The vegetational history of the Killarney Oakwoods, S. W. Ireland: evidence from fine spatial resolution pollen analysis. Journal of

Ecology, 76: 415-436. Mitchell, F.J.G. (1995) The dynamics of Irish post-glacial forests. In: Wood’s, Trees and Forests in Ireland, Pilcher, J.R. and Mac an tSoir, S. (eds.). Royal Irish Academy, Dublin. Molina, R. and Trappe, J. (1982) Lack of mycorrhizal specificity by the ericaceous host Abrutus menziesii and Arctostaphylus uva-ursi. New Phytologist, 90: 495-509. Molina, R., Trappe, J.M. (1982) Pattern of ectomycorrhizal host specificity and potential amoung pacific northwest conifors and fungi. Forest Science, 28: 423-458. Molina, R.J., Massicotte, H.B., Trappe, J.M., (1992) Specificity phenomena in mycorrhizal symbioses: Community-ecological consequences and practical implications. In: Mycorrhizal Functioning, an Integrative Plant-fungal Process, Allen, M.F. (Ed.). New York: Routledge, Chapman and Hall, pp. 357–423.

373

Moore, D., Nauta, M., Evans, S., and Rotheroe, M. (2001) Fungal conservation-

Issues and solutions, Cambridge: Cambridge University Press. Morris, M.H., Smith, M.E., Rizzo, D.M., Rejmanek, M. and Bledsoe, C.S. (2008) Contrasting ectomycorrhizal fungal communites on the roots of co-occuring oaks (Quercus spp.) in a California woodland. New Phytologist, 178: 167–176. Moser, M. (1983) Keys to Agarics and Boleti. 4th ed. London: Roger Phillips. Mueller, M.G., J.P. Schmit, P.R. Leacock, B. Buyck,. Cifuentes, J., Desjardin, D.E., Halling, R.E., Hjortstam, K., Iturriaga, T., Larsson, k., Lodge, J., May, T.W., Minter, D., Rajchenberg, M., Redhead, S.A., Ryvarden, L., Trappe, J.M., Watling, R. and Wu, Q. (2007) Global diversity and distribution of macrofungi. Biodiversity and Conservation, 16: 37-48. Muller, M. and Hallaksela, A. (1998) Diversity of Norway spruce endophytes in various mixed and pure Norway spruce stands. Mycological Research, 102: 1183-1189. Muñoz, J.A. (2005) Boletus: Fungi Europaei. Vol. 1. Alassio: Edizioni Candusso. Muskett, A.E. and Malone, J.P. (1978) Catalogue of Irish fungi – I. Gasteromycetes. Proceeding of the Royal Irish Academy, B78, 1-11. Muskett, A.E. and Malone, J.P. (1980) Catalogue of Irish fungi – II. Hymenomycetes. Proceeding of the Royal Irish Academy, B80: 197-276. Naeem, S. (1999) Power behind Nature’s throne. Nature, 401: 653–654. Nantel, P. and Neumann, P. (1992) Ecology of ectomycorrhizal-basidiomycete communities on a local vegetation gradient. Ecology, 73: 99–117. National Forest Inventory (2006) [online], available: http://www.agriculture.gov.ie/index.jsp?file=forestry/NFI/NFI_results.xml [accessed 16 Feb. 2009]. NBDC (2010). Data Centre Mapping System. [online], available: http://maps.biodiversityireland.ie/ [accessed 22/03/2011]. Newton, A. C. and Pigott, C. D. (1991) Mineral nutrition and mycorrhizal infection of seedling oak and birch I. Nutrient uptake and the development of mycorrhizal infection during seedling establishment. New Phytologist, 117: 37–44. Newton, A.C. (2007) Forest ecology and conservation. Oxford: Oxford University Press.

374

Newton, A.C. and Haigh, J. (1998) Diversity of ectomycorrhizal fungi in the UK: a test of the species-area relationship and the role of host preference. New

Phytologist, 138: 619-627. Newton, A.C., Holden, E., Davy, L.M., Ward, S.D., Fleming, L.V. and Watling, R. (2002) Status and distribution of stipitate hydnoid fungi in Scottish coniferous forests. Biological Conservation, 107: 181-192. Nihlgard, B., and Lindgreen, L. (1977) Plant biomass, primary production and bioelements of three mature beech forests in southern Sweden. Oikos, 28: 95-104. Noordeloos, M.E. (1992). Entolomataceae. In Knudsen & Hansen (eds.) Nordic

Macromycetes, vol. 2. Copenhagen: Nordsvamp, pp.341-359. Noordeloos, M.E. (2005) General Part. In: Flora Agaricina Neerlandica,

Noordeloos, M.E., Kuyper, T.W. and Vellinga , E.C. (eds.) vol. 6. Rotterdam: Balkema publishers, pp. 1-19. Norden, B., Gotmark, F., Ryberg, M., Paltto, H., Allmer, J. (2008) Partial cutting reduces species richness of fungi on woody debris in oak-rich forests. Canadian

Journal of Forest Research, 38: 1807–1816. Norden, B., Gotmark, F., Tonnberg, M., and Ryberg, M. (2004) Dead wood in semi-natural temperate broadleaved woodland: contribution of coarse and fine dead wood, attached dead wood and stumps. Forest Ecology and Management, 194: 235–248. Norden, B., Paltto, H., Gotmark, F. and Wallin, K. (2007) Indicators of biodiversity, what do they indicate? - Lessons for conservation of cryptogams in oak-rich forest. Biological Conservation, 135: 369-379. Noss, R.F. (1999). Assessing and monitoring forest biodiversity: a suggested framework and indicators. Forest Ecology and Management, 115: 135–146. Nuñez, M.A., Horton, T.R. and Simberloff, D. (2009) Lack of belowground mutualisms hinders Pinaceae invasions. Ecology, 90: 2352–2359. O’Carroll, M. (1978) The nursing of Sitka spruce I. Japanese larch. Irish Forestry, 35: 60-65. O’ Connell, T. and Bolger, T. (1997) Stability, ephemerality and dispersal ability: microarthropod assemblages on fungal sporophores. Biological Journal of the

Linnean Society, 62: 111–131. O’ Connell, T. and Bolger, T. (2007) Fungal fruiting bodies and the structure of the fungus micro-arthropod assemblages. Biology and Environment: Proceedings

of the Royal Irish Academy 97B: 249–262.

375

O’Dell, T.E., Ammirati, J.F., and Schreiner, E.G. (1999) Species richness and abundance of ectomycorrhizal basidiomycete sporocarps on a moisture gradient in the Tsuga heterophylla zone. Canadian Journal of Botany, 77: 1699-1711. O’Hanlon, R. and Harrington, T.J. (2011; Appendix 2 this volume) Diversity and distribution of mushroom forming fungi (Agaricomycetes) in Ireland. Biology and

Environment: Proceedings of the Royal Irish Academy

O’Toole, L., Fielding, A.H. and Haworth, P.F. (2002) Re-introduction of the golden eagle into the Republic of Ireland. Biological Conservation, 103: 303–312. Odor, P., Heilmann-Clausen, J., Christensen, M., Aude, E., van Dort, K.W., Piltaver, A., Siller, I., Veerkamp, M.T., Walleyn, R., Standovar, T., van Hees, A.F.M., Kosec, J., Matocec, N., Kraigher, H. and Grebenc, T. (2006) Diversity of dead wood inhabiting fungi and bryophytes in semi-natural beech forests in Europe. Biological Conservation, 131: 58–71. Ohenoja, E. (1993) Effect of weather conditions on the larger fungi at different

forest sites in northern Finland in 1976-1988. Unpublished Ph.D. thesis, Department of Botany, Oulu University, Oulu, Finland. Okland, B., Gotmark, F., and Norden, B. (2008) Oak woodland restoration: testing the effects on biodiversity of mycetophilids in Sweden. Biodiversity and

Conservation, 17: 2599-2616. Okland, B., Gotmark, F., Norden, B., Franc, N., Kurina, O., and Polevoi, A. (2005). Regional diversity of Mycetophilids (Diptera: Sciaroidea) in Scandinavian oak-dominated forests. Biological Conservation, 121: 9–20. Ortega , A. and Lorite, J. (2007) Macrofungi diversity in cork-oak and holm-oak forests in Andalusia (southern Spain); an efficient parameter for establishing priorities for its evaluation and conservation. Central European Journal of

Botany, 2: 276–296. Orton, P.D. (1987) Fungi of northern Pine and Birch woods. Bulletin of the British

Mycological Society, 20: 130-145. Orton, P.D. (1994) Some comments on ‘A provisional red data list for British Fungi’. Mycologist, 8: 66-67. Osono, T., Fukasawa, Y. and Takeda, H. (2003). Roles of diverse fungi in larch needle-litter decomposition. Mycologia, 95: 820-826. Ostry, M.E., and Laflamme, G. (2009) Fungi and diseases- natural components of healthy forests. Botany, 87: 22-25.

376

Outerbridge, R.A., and Trofymow, J. (2004) Diversity of ectomycorrhizae on experimentally planted Douglas-fir seedlings. Canadian Journal of Botany, 82: 1671-1681. Outerbridge, R.A. (2002) Macrofungus ecology and diversity under different

conifer monocultures on southern Vancouver Island. Unpublished PhD Dissertation, University of Victoria, BC. Ovaskainen, O., Nokso-koivisto, J., Hottola, J., Rajala, T., Pennanen, T., Ali-kovero, H., Miettinen, O., Oinonen, P., Auvinen, P., Paulin, L., Larsson, K. and Makip, R. (2010) Identifying wood-inhabiting fungi with 454 sequencing –what is the probability that BLAST gives the correct species?. Fungal Ecology, 3: 274–283. Oxbrough, A.G., Gittings, T., O’Halloran, J., Giller, P.S. and Smith, G.F. (2005) Structural indicators of spider communities across the forest plantation cycle. Forest Ecology and Management, 21: 171-183. Oxbrough, A.G., Gittings, T., O’Halloran, J., Giller, P.S., and Kelly, T.C. (2006b) The influence of open space on ground-dwelling spider assemblages within plantation forests. Forest Ecology and Management, 237: 404-417. Oxbrough, A.G., Gittings, T., O’Halloran, J., Giller, P.S., and Kelly, T.C. (2006a) The initial effects of afforestation on the ground-dwelling spider fauna of Irish peatlands and grasslands. Forest Ecology and Management, 237: 478-491. Oxbrough, A.G., Irwin, S., Kelly, T.C. and O’Halloran, J. (2010) Ground-dwelling invertebrates in reforested conifer plantations. Forest Ecology and

Management, 259: 2111–2121. Paletto, A. and Tosi, V. (2009) Forest canopy cover and canopy closure: comparison of assessment techniques. European Journal of Forest Research, 128: 265–272. Palfner, G., Angélica Casanova-Katny, M. and Read, D.J. (2005) The mycorrhizal community in a forest chronosequence of Sitka spruce [Picea sitchensis (Bong.) Carr.] in Northern England. Mycorrhiza, 15: 571-579. Paul, E.A. (2007) Soil Microbiology, Ecology and Biochemistry. Massachusetts: Academic press. Peay, K.G., Bruns, T.D., Kennedy, P.G., Bergemann, S.E. and Garbelotto, M. (2007). A strong species–area relationship for eukaryotic soil microbes: island size matters for ectomycorrhizal fungi. Ecology Letters, 10: 470–480. Peay, K.G., Kennedy, P.G., Bruns, T.D. (2011) Rethinking ectomycorrhizal succession: are root density and hyphal exploration types drivers of spatial and temporal zonation?. Fungal Ecology, 4: 233-240.

377

Peck, J.E. (2003) Using PC-ORD for Multivariate Data Analysis. Minneapolis, MN, USA: PC-ORD Workshop, pp. 11. Pegler, D.N., Roberts, P.J. and Spooner, B.M. (1997) British Chanterelles and Tooth Fungi. Kew, Richmond: Royal Botanic Gardens. Perez-Moreno, J. and Read, D.J. (2001) Nutrient transfer from nematodes to plants: a direct pathway provided by the mycorrhizal mycelial network. Plant,

Cell and Environment, 24: 1219-1226. Perini, C., Barluzzi, C. and De Dominicis, V. (1989) Mycocoenological research on evergreen oak wood in the hills adjacent to the Maremma coastline (NW of Grosseto, Italy). Phytocoenologia, 17: 289-306. Perrin, P., Martin, J., Barron, S., O’Neill, F., McNutt, K. and Delaney, A. (2008a) National survey of Native Woodlands 2003–2008: Volume I main report. Dublin: National Parks & Wildlife Service. Perrin, P., Martin, J., Barron, S., O’Neill, F., McNutt, K. and Delaney, A. (2008b) National survey of Native Woodlands 2003–2008: volume II woodland classification. Dublin: National Parks & Wildlife Service. Perrin, P., Kelly, D.L. and Mitchell, F.J.G. (2006b) Long-term deer exclusion in yew-wood and oakwood habitats in southwest Ireland: Natural regeneration and stand dynamics. Forest Ecology and Management, 236: 356–367. Perrin, P.M., Barron, S.J. & Martin, J.R. (2006a) National Survey of Native

Woodlands in Ireland: Second Phase Report. Dublin: National Parks & Wildlife Service. Perry, D. A., Oren, R. & Hart, S. (2008) Forest ecosystems. Baltimore, MD: John Hopkins University Press. Perry, D.A., Borchers, J.G., Borchers, S.L. and Amaranthus, M.P. (1990) Species migrations and ecosystem stability during climate change: the belowground connection. Conservation Biology, 4: 266-274. Peter, M., Ayer, F., Egli, S. and Honegger, R. (2001) Above and belowground community structure of ectomycorrhizal fungi in three Norway spruce (Picea

abies) stands in Switzerland. Canadian Journal of Botany, 79: 1134–1151. Peterken, G.F. and Jones, E.W. (1989) Forty years of change in Lady ParkWood: the young-growth stands. Journal of Ecology, 77: 401–429. Peterken, G.F., Ausherman, D., Buchanan, M. and Forman, R.T.T. (1992) Old-growth conservation within British upland conifer plantations. Forestry, 65: 127-144.

378

Peterson, E. B., Peterson, N. M., Weetman, G. F. and Martin, P. J. (1997) Ecology

and management of Sitka spruce, emphasizing its natural range in British

Columbia. Vancouver, B.C.: University of British Colmbia. Peterson, R.H. and Cifuentes, J. (1994) Notes on mating systems of Auriscalpium

vulgare and A. villipes. Mycological Research, 98: 1427-1430. Phillips, R. (2006) Mushrooms. UK: Pan Macmillian. Pickles, B.J., Genney, D.R., Potts, J.P., Lennon, J.J., Anderson, I.C. and Alexander, I.J. (2010) Spatial and temporal ecology of Scot’s pine ectomycorrhizas. New Phytologist, 186: 755–768. Pilz, D. and Perry, D. (1984) Impact of clearcutting and slash burning on ectomycorrhizal associations of Douglas fir seedlings. Canadian Journal of

Forest Research, 14: 94-100. Pim, G. (1898) The fungi of counties Dublin and Wicklow. Irish Naturalists’

Journal, 2: 173-185. Pinna, S., Gévry, M.F., Côté, M. and Sirois, L. (2010) Factors influencing fructification phenology of edible mushrooms in a boreal mixed forest of Eastern Canad. Forest Ecology and Management, 260: 294–301. Qian, P., and Schoenau, J.J. (2002) Practical applications of ion exchange resins in agriculture and environmental soil research. Canadian Journal of Soil Science 82: 9–21. Quine, C.P. and Humphrey, J.W. (2010) Plantations of exotic tree species in Britain: irrelevant for biodiversity or novel habitat for native species?. Biodiversity and Conservation, 19: 1503–1512. Raaijmakers, J.G. (1987) Statistical analysis of the Michaelis–Menten equation. Biometrics, 43: 793–803. Rackham, O. (1995) Looking for ancient woodland in Ireland. In: Wood’s, Trees

and Forests in Ireland, Pilcher, J.R. and Mac an tSoir, S. (Eds.). Dublin: Royal Irish Academy. Ramsbottom, J. (1936) The Killarney foray: 20–26 September, 1936. Transactions of the British Mycological Society, 22: 5-11. Redecker, D., T. M. Szaro, J. R. Bowman, and T. D. Bruns. (2001) Small genets of Lactarius xanthogalactus, Russula cremoricolor, and Amanita francheti in late stage ectomycorrhizal successions. Molecular Ecology, 10: 1025–1034. Renvall, P. (1995) Community structure and dynamics of woodrotting Basidiomycetes on decomposing conifer trunks in northern Finland. Karstenia, 35: 1-51.

379

Richard, F., Millot, S., Gardes, M. and Selosse, M.A. (2005) Diversity and specificity of ectomycorrhizal fungi retrieved from an old-growth Mediterranean forest dominated by Quercus ilex. New Phytologist, 166: 1011–1023. Richardson, M. J. (1970) Studies on Russula emetica and other agarics in a Scot’s pine plantation. Transactions of the British Mycological Society 55: 217-229. Ries, L., Robert, J., Fletcher, J., Battin, J. and Sisk, T.D. (2004) Ecological responses to habitat edges: mechanisms, models, and variability explained. Annual Review of Ecology Evololution and Systematics, 35: 491–522. Rincón, A. and Pueyo, J.J. (2010) Effect of fire severity and site slope on diversity and structure of the ectomycorrhizal fungal community associated with post-fire regenerated Pinus pinaster Ait. seedlings. Forest Ecology and Management, 260: 361–369. Riva, A. (2003) Tricholoma, Fungi Europaei 3. Alassio: Edizioni Candusso. Roberts, C., Ceska, O., Kroeger, P. and Kendrick, B. (2004) Macrofungi from six habitats over five years in Clayoquot Sound, Vancouver Island. Canadian Journal

of Botany, 82: 1518–1538. Robertson, S.J., Tackaberry, L.E., Egger, K.N. and Massicotte, H.B. (2006) Ectomycorrhizal fungal communities of black spruce differ between wetland and upland forests. Canadian Journal of Forest Research, 36: 972–985. Roche, J.R., Mitchell, F.J.G. and Waldren, S. (2009) Plant community ecology of Pinus sylvestris, an extirpated species reintroduced to Ireland. Biodiversity and

Conservation, 18: 2185-2203. Rodin, L.E., and Basilevic, N.I. (1968) World distribution of plant biomass. In: Functioning of terrestrial ecosystems at the primary production level. F.E. Eckardt (ed.). Paris: UNESCO. pp. 45 -52. Rodwell, J.S. and Cooper, E.A. (1995) Scottish pinewoods in a European context. In: Our pinewood heritage, Aldhous, J.R. (ed). Inverness: Forestry Commission, The Royal Society for the Protection of Birds, Scottish Natural Heritage. Rogerson, T.L., (1976) Soil water deficits under forested and clearcut areas in northern Arkansas. Soil Science Society of America, 40: 802–805. Rosenwald, R. and Lõhmus, A. (2008) For what, when and where is green-tree retention better than clear-cutting? A review of the biodiversity aspects. Forest

Ecology and Management, 255: 1–15. Rosling, A., Landeweert, R., Lindahl, B.D., Larsson, K-H., Kuyper, T.W., Taylor, A.F.S. and Finlay, R.D. (2003) Vertical distribution of ectomycorrhizal fungal taxa in a podzol profile determined by morphotyping and genetic verification. New Phytologist, 159: 775–783.

380

Ruhling, A., and Tyler, G. (1990) Soil factors influencing the distribution of macrofungi in oak forests of sounthern Sweden. Holartic Ecology, 13: 11-18. Sakakibara, S.M., Jones, M.D., Gillespie, M., Hagerman, S.M., Forrest, M.E., Simard, M.E., Durall, D.M. (2002) A comparison of ectomycorrhiza identification based on morphotyping and PCR-RFLP analysis. Mycological Research, 106: 868–878. Sakura, T., Gimingham, C.G. and Miller, C.S. (1985). Effect of tree density on ground vegetation in a Japanese larch plantation. Scottish Forests, 39: 191-198. Salo, K. (1993) The composition and structure of macrofungus communities in boreal upland type forests and peatlands in north Karelia, Finland. Karstenia, 33: 61-99. Sarrionandia, E., Rodriguez, N. and Salcedo, I. (2009) A study of the macrofungal community in the beech forests of Altube (Basque Country, Northern Spain). Cryptogamie Mycologie, 30: 67-83. Såstad, S. (1995) Fungi-vegetation relationships in a Pinus sylvestris forest in central Norway. Canadian Journal of Botany, 73: 807-816. Scherer-Lorenzen, M., Korner, C., and Schulze, E. (2005) Forest Diversity and

Function. Berlin: Springer-Verlag. Schild, D.E., Kennedy, A. and Stuart, M.R. (1988) Isolation of symbiont and associated fungi from ectomycorrhizas of Sitka spruce. European Journal of

Forest Pathology, 18: 51-61. Schmit, J. P. and G. M. Mueller (2007) An estimate of the lower limit of global fungal diversity. Biodiversity and Conservation, 16: 99–111. Schmit, J.P., Mueller, G.M., Leacock, P.R., Mata, J.L., Wu, Q. and Huang, Y. (2005) Assessment of tree species richness as a surrogate for macrofungal species richness. Biological Conservation, 121: 99–110. Schmit, J.P., Murphy, J.F., and Mueller, G.M., (1999) Macrofungal diversity of a temperate oak forest: a test of species richness estimators. Canadian Journal of

Botany, 77: 1014– 1027. Schweiger, O., Musche, M., Bailey, D., Billeter, R., Diekotter, T., Hendrickx, F., Herzog, F., Liira, J., Malfait, J., Speelmans, J. and Dziock, F. (2007) Functional richness of local hoverfly communities (Diptera, Syrphidae) in response to land use across temperate Europe. Oikos, 116: 461-472.

381

Senn-Irle, B., Heilmann-Clausen, J., Genney, D., Dahlberg, A. (2007) Guidance

for the conservation of Fungi. Prepared for the Convention on the Conservation

of European and natural habitats. European Council. Document T-PVS (2007) 13 (rev). Strasbourg 17 Oct 2007. [Online], available: www.coe.int/t/dg4/cultureheritage/conventions/bern/T-PVS/sc27_tpvs13_en.pdf. [accessed 10/06/09]. Senn-Irlet, B. and Bieri, G. (1999) Sporocarp succession of soil-inhabiting macrofungi in an autochothonous subalpine Norway spruce forest of Switzerland. Forest Ecology and Management, 124: 169-175. Shannon, C.E. and Weaver, W. (1949) The Mathematical Theory of

Communication. Urbana: University of Illinois Press. Siitonen, P., Lehtinen, A. and Siitonen, M. (2005) Effects of Forest Edges on the Distribution, Abundance, and Regional Persistence of Wood-Rotting Fungi. Conservation Biology, 19: 250-260. Simard, S.W. and Durall, D.M. (2004) Mycorrhizal networks: a review of their extent, function, and importance. Canadian Journal of Botany, 82: 1140-1165. Simard, S.W., Perry, D.A., Jones, M.D., Myrold, D.P., Durall, D.M., and Molina, R., (1997) Net transfer of carbon between ectomycorrhizal trees species in the field. Nature, 388: 579-582. Simard, S.W., Perry, D.A., Jones, M.D., Smith, J.E. and Molina, R., (1997b) Shared compatibility of ectomycorrhizae on Psuedotsuga menziesii and Betula

papyrifera seedlings grown in mixture in soils from southern British Columbia. Canadian Journal of Forest Research, 27: 331-342. Simpson, E.H. (1949) Measurements of diversity. Nature, 163: 668. Sippola, A., Renvall, P. (1999) Wood-decomposing fungi and seed-tree cutting: A 40-year perspective. Forest Ecology and Management, 115: 183-201. Smith, E.P. and Belle, G.V. (1984) Nonparametric estimation of species richness. Biometrics, 40: 119–129. Smith, G.F., Iremonger, S., Kelly, D.L., O’Donoghue, S. and Mitchell, F.J.G. (2007) Enhancing vegetation diversity in glades, rides and roads in plantation forests Biological Conservation, 136: 283-294. Smith, G.F., Gittings, T., Wilson, M.W., Oxbrough, A., Iremonger, S., O'Donoghue, S., McKee, A.M., O'Halloran, J., Kelly, D.L., Pithon, J., O'Sullivan, A., Neville, P., Mitchell, F.J.G., Giller, P.S., O'Donnell, V. and Kelly, T.C. (2006). Biodiversity Assessment of Afforestation Sites. BIOFOREST Project 3.1.1 Final Report to COFORD. [Online], available: http://www.coford.ie/iopen24/pub/pub/Reports/BIOFOREST312text.PDF. [accessed 06/08/08].

382

Smith, G.F., Gittings, T., Wilson, M.W., French, L., Oxbrough, A., O’Donoghue, S., Pithon,. O’Donnell, V., McKee, A.M., Iremonger, S., O’Halloran, J., Kelly, D.L., Mitchell, F.J.G., Giller, P.S. and Kelly, T. (2005) Assessment of biodiversity

at different stages of the forest cycle. BIOFOREST Project 3.1.2 Final Report to COFORD. [Online], available: http://bioforest.ucc.ie/pages/project_two.htm [accessed 09/03/09]. Smith, J.E., Molina, R., and Perry, D.A. (1995) Occurance of ectomycorrhizas on ericaceous and coniferous seedlings grown in soils from the Oregon coast range. New Phytologist, 129: 73-81. Smith, J.E., Molina, R., Huso, M.M.P., Luoma, D.L., McKay, D., Castellano, M.A., Lebel, T., Valachovic, Y., (2002) Species richness, abundance, and composition of hypogeous and epigeous ectomycorrhizal fungal sporocarps in young, rotation-age, and old-growth stands of Douglas-fir (Pseudotsuga

menziesii) in the Cascade Range of Oregon USA. Canadian Journal of Botany, 80: 186–204. Smith, S.E. and Read, D.J. (2008) Mycorrhizal symbiosis, 3rd edition. London: Acedemic Press. Sokal, R.R. and Rohlf, F.J. (1995) Biometry, 3rd edition. New York: W.H. Freeman and Co. Stark, N.M. (1982) Soil fertility after logging in the northern Rocky Mountains. Canadian Journal of Forest Research, 12: 679–686. Stendell, E.R., Horton, T.R. and Bruns, T.D. (1999) Early effects of prescribed fire on the structure of the ectomycorrhizal fungus community in a Sierra Nevada ponderosa pine forest. Mycological Research, 103:1353–1359. Straatsma, G. and Krisai-Greilhuber, I. (2003) Assemblage structure, species richness, abundance, and distribution of fungal fruiting bodies in a seven year plot based survey near Vienna. Mycological Research, 107: 632-640. Straatsma, G., Ayer, F. and Egli, S. (2001) Species richness, abundance and phylogeny of fungal fruit bodies over 21 years in a Swiss forest plot. Mycological

Research, 105: 512-523. Strong, D.R. and Levin, D.A. (1975) Species Richness of the Parasitic Fungi of British Trees. Proceedings of the National Academy of Sciences, 72: 2116-2119. Stubblefield, S.P. and Taylor, T.N. (1988) Recent advances in paleomycology. New Phytologist, 108: 3-25. Su, Q., Maclean, D. and Needham, T. (1996) The influence of hardwood content on balsam fir defoliation by spruce budworm. Canadian Journal of Forest

Research, 26: 1620-1628.

383

Sweeney, O.F., Martin, R.D., Irwin, S., Kelly, T.C., O’Halloran, J., Wilson, M.W. and McEvoy, P.M. (2010a). A lack of large-diameter logs and snags characterises dead wood patterns in Irish forests. Forest Ecology and Management, 259: 2056-2064. Sweeney, O.F., Wilson, M.W., Irwin, S., Kelly, T.C. and O’Halloran, J. (2010b) Are bird density, species richness and community structure similar between native woodlands and non-native plantations in an area with a generalist bird fauna?. Biodiversity and Conservation, 19: 2329–2342. Tarvainen, O., Markkola, A.M. and Strömmer, R. (2003) Diversity of macrofungi and plants in Scot’s pine forests along an urban pollution gradient. Basic and

Applied Ecology, 4: 547–556. Taylor, A.F.S. (2002) Fungal diversity in ectomycorrhizal communities: sampling effort and species detection. Plant and Soil, 244: 19-28. Taylor, A.F.S. and Alexander, I.J. (1989) Demography and population dynamics of ectomycorrhizas of Sitka spruce fertilized with N. Agriculture, Ecosystems and

Environment, 28: 493–496. Taylor, D.L. and Bruns, T.D. (1999) Community structure of ectomycorrhizal fungi in a Pinus muricata forest: minimal overlap between the mature forest and resistant propagule communities. Molecular Ecology, 8: 1837–1850. Teagasc (2005) Irish Agriculture and Food Development Authority. [Online], available: www.teagasc.ie [Accessed 11/05/11]. Tedersoo, L., Koljalg, U., Hallenberg, N. and Larsson, K. (2003) Fine scale distribution of ectomycorrhizal fungi and roots across substrate layers including coarse woody debris in a mixed forest. New Phytologist, 159: 153–165. Tedersoo, L., May, T.W. and Smith, M.E. (2010) Ectomycorrhizal lifestyle in fungi: global diversity, distribution, and evolution of phylogenetic lineages. Mycorrhiza, 20: 217-263. Tedersoo, L., Nilsson, R.H., Abarenkov, K., Jairus, T., Sadam, A., Saar, I., Bahram, M., Bechem, E., Chuyong, G. and Kõljalg, U. (2010) 454 Pyrosequencing and Sanger sequencing of tropical mycorrhizal fungi provide similar results but reveal substantial methodological biases. New Phytologist, 188: 291–301. Tedersoo, L., Suvi, T., Larsson, E., and Koljalg, U. (2006) Diversity and community structure of ectomycorrhizal fungi in a wooded meadow. Mycological

Research, 110:734-748. Tennant, D. (1975). A test of a modified line intersect method of estimating root length. Journal of Ecology, 63: 995-1001.

384

Termorshuizen, A.J. (1991) Succession of Mycorrhizal fungi in stands of Pinus

sylvestris in the Netherlands. Journal of Vegetation Science, 2: 555–564. Thomas, G.W., Rogers, D. and Jackson, R. (1983) Changes in the mycorrhizal status of Sitka spruce following outplanting. Plant and Soil, 71: 319-323. Tofts, R.J. and Orton, P. D. (1998) The species accumulation curve for agarics and boleti from a Caledonian pine wood. Mycologist, 12: 98-102. Tomppo, E. (1996) Biodiversity monitoring in Finnish forest inventories. In: Assessment of Biodiversity for Improved Forest Management. EFI Proceedings No 6. Joensuu: European Forest Institute, pp. 87-94. Topp, W., Kappes, H., Kulfan, J., Zach, P. (2006) Distribution pattern of woodlice (Isopoda) and millipedes (Diplopoda) in four primeval forests of the Western Carpathians (central Slovakia). Soil Biology and Biochemistry, 38: 43–50. Trappe, J.M. (1969) Studies on Cenococcum graniforme I. An efficient method for isolation from sclerotia. Canadian Journal of Botany, 47: 1389–1390. Twieg, B.D., Durall, D.M., Simard, S.W. and Jones, M.D. (2009) Influence of soil nutrients on ectomycorrhizal communities in a chronosequence of mixed temperate forests. Mycorrhiza, 19: 305–316. Tyler, G. (1985) Macrofungal flora of swedish beech forest related to soil organic mattes and acidity characteristics. Forest Ecology and Management, 10: 13-29. Tyler, G. (1989) Edaphical distribution patterns of macrofungal species in deciduous forest in South Sweden. Acta Ecology, 10: 309-326. Tyler, G. (1992) Tree species affinity of decomposer and ectomycorrhizal macrofungi in beech (Fagus sylvatica L.), oak (Quercus robur L.) and hornbeam (Carpinus betulus L.) forests. Forest Ecology and Management, 47: 269–284. UNECE (2003) State of Europe’s Forests 2003: The MCPFE’s Report on

Sustainable Forest Management in Europe. 4th Ministerial conference on the protection of forests in Europe. Vienna: MCPFE Liason Unit. Unterseher, M., and Tal, O. (2006) Influence of small scale conditions on the diversity of wood decay fungi in a temperate, mixed deciduous forest canopy. Mycological Research, 110: 169-178. Unterseher, M., Otto, P. and Morawetz, W. (2005) Species richness and substrate specificity of lignicolous fungi in the canopy of a temperate, mixed deciduous forest. Mycological Progress, 4: 117–132. Unterseher, M., Schnittler, M., Dormann, C. and Sickert, A. (2008) Application of species richness estimators for the assessment of fungal diversity. FEMS

Microbiology Letters, 282: 205–213.

385

Valentine, L.L., Fiedler, T.L., Hart, A.N., Petersen, C.A., Berninghausen, H.K. and Southworth D. (2004) Biodiversity of ectomycorrhizal fungi associated with Quercus garryana. Canadian Journal of Botany, 82: 123–135. Vare, H., Ohenoja, E., Ohtonen, R. (1996) Macrofungi of oligotrophic Scot’s pine forests in northern Finland. Karstenia, 36:, 1–18. Vesterdal, L., Dalsgaard, M., Felby, C., Raulund-Rasmussen, K. and Jùrgensen, B.B., (1995) Effects of thinning and soil properties on accumulation of carbon, nitrogen and phosphorus in the forest floor of Norway spruce stands. Forest

Ecology and Management, 77: 1-10. Villeneuve, N., Grandtner, M.M. and Fortin, J.A. (1989) Frequency and diversity of ectomycorrhizal and saprophytic macrofungi in the Laurentide Mountains of Quebec. Canadian Journal of Botany, 67: 2616-2629. Waksman, S.A. (1916) Do fungi live and produce mycelium in the soil?. Science, 44: 320-322. Walker, J.F., Miller, O.K. and Horton, J.L. (2005) Hyperdiversity of ectomycorrhizal fungus assemblages on oak seedlings in mixed forests in the southern Appalachian Mountains. Molecular Ecology, 14: 829–838. Wallace, H.L., Good, J.E.G. and Williams, T.G., (1992). The effects of afforestation on upland plant communities: an application of the British National Vegetation Classification. Journal of Applied Ecology, 29: 180-194. Wang, B. and Qiu, Y.L. (2006) Phylogenetic distribution and evolution of mycorrhizas in land plants. Mycorrhiza, 16: 299–363. Wang, H., Dai, L.M., Yang, B.S., Lang, Q.L. and Gu, H.Y. (2005). Occurrence and culture of Mycorrhizal fungi associated with oaks in Dandong Region, Limning Province. Pedosphere, 15(2): 232-237. Wang, Y. and Hall, I.R. (2004) Edible ectomycorrhizal mushrooms: challenges and achievements. Canadian Journal of Botany, 82: 1063–1073. Warren W.G. and Olsen P.E. (1964) A line intersect technique for assessing logging waste. Forest Science, 10: 267-276. Waters, J., McKelvey, K., Zabel, C., and Oliver, W. (1994) The effects of thinning and broadcast burning on sporocarp production of hypogeous fungi. Canadian Journal of Forest Research, 24: 1516-1522. Watling, R, (1995) Assessment of fungal diversity: macromycetes, the problems. Canadian Journal of Botany (Suppl. 1) 73: S15-S24. Watling, R. (1974) Macrofungi in oak woods in Britain. In: The British Oak, M.G. Morris & F.H. Perring(eds.). Berkshire, UK: BSBI. pp. 235-249.

386

Watling, R. (1997). Biodiversity of lichenised and non-lichenised fungi in Scotland. In: Biodiversity in Scotland: Status, Trends and Initiatives, Fleming, L.V., Newton, A.C., Vickery, J.A., Usher, M.B. (Eds.). Scottish Natural Heritage, The University of Edinburgh. Edinburgh: The Stationery Office. pp 77-88. Watling, R. (2005) The fungi of Scottish Western oakwoods. Botanical Journal of

Scotland, 57: 155-165. Watson, K. (2007) Soils illustrated: Field descriptions. British Columbia: International remote sensing surveys ltd. Webber, J.F., Mullett, M. and Brasier, C.M. (2010) Dieback and mortality of plantation Japanese larch (Larix kaempferi) associated with infection by Phytophthora ramorum. New Disease Reports, 22: 19. Whalley, A.J.S. and Watling, R. (1982) Distribution of Daldinia concentrica in the British Isles. Transactions of the British Mycological Society, 78: 47–53. Whipps, J.M. (2004) Prospects and limitations for mycorrhizas in biocontrol of root pathogens. Canadian Journal of Botany, 82: 1198–1227. White tailed eagle project (2010) [Online], available: http://www.goldeneagle.ie/portal.php?z=9 [accessed 22/03/2011]. White, J. & Doyle, G. (1982). The vegetation of Ireland: a catalogue raisonne In: Studies on Irish vegetation, J. White (ed.),. Dublin: Royal Dublin Society, pp. 289-368. White, T.J., Bruns, T., Lee, S., Taylor, J.W. (1990) Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: PCR Protocols:

A Guide to Methods and Applications, Innis MA, Gelfand DH, Sninsky JJ, White TJ (eds.). New York: Academic Press Inc. Whittaker, R. H. (1965) Dominance and diversity in plant communities. Science, 147, 250–260. Wilkins, W.H., Ellis, E.M. and Harley, J.L. (1937) The Ecology of the Larger Fungi 1. Constancy and frequency of fungal species in relation to certain vegetation communities, particularly oak and beech. Annals of Applied Biology, 24: 703-732. Williams, S.E. and Hero, J.M. (2001) Multiple determinants of Australian tropical frog biodiversity. Biological Conservation, 98: 1-10. Winterhoff, W. (1984). Analyse der Pilze in Pflanzengesellschaften, insbesondere der Makrornyzeten. In: Sampling Methods and Taxon Analysis in Vegetation

Science, R. Knapp(ed.). Amsterdam: Hague, pp. 227-248.

387

Witt, C. (2010) Characteristics of aspen infected with heartrot: Implications for cavity-nesting birds. Forest Ecology and Management, 260: 1010–1016. Wurzburger, N., Hartshorn, A.S. and Hendrick, R.L. (2004) Ectomycorrhizal fungal community structure across a bog-forest ecotone in southeastern Alaska. Mycorrhiza, 14: 383–389. Zak, J.C. and Willig, M.R. (2004) Fungal Biodiversity patterns. In: Biodviersity of

fungi: inventory and monitoring methods, Mueller, G.M., Bills, G.F., Foster, M.S. (eds.). New york: Academic press. Zinko, U., Seibert, J., Dynesius, M. and Nilsson, C. (2005) Plant species numbers predicted by a topography based groundwater flow index. Ecosystems, 8: 430-441.