patrones de diversidad de especies leÑosas en los …oa.upm.es/9873/1/emilia_martin_queller.pdf ·...

195
UNIVERSIDAD POLITÉCNICA DE MADRID ESCUELA TÉCNICA SUPERIOR DE INGENIEROS DE MONTES PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS BOSQUES ESPAÑOLES A DIFERENTES ESCALAS ESPACIALES: EFECTOS DE LAS PERTURBACIONES SELVÍCOLAS Y OTROS FACTORES AMBIENTALES Y DE PAISAJE TESIS DOCTORAL EMILIA MARTÍN QUELLER Licenciada en Ciencias Ambientales 2011

Upload: others

Post on 08-Jan-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

UNIVERSIDAD POLITÉCNICA DE MADRID

ESCUELA TÉCNICA SUPERIOR DE INGENIEROS DE MONTES

PATRONES DE DIVERSIDAD DE ESPECIES

LEÑOSAS EN LOS BOSQUES ESPAÑOLES A

DIFERENTES ESCALAS ESPACIALES: EFECTOS

DE LAS PERTURBACIONES SELVÍCOLAS Y

OTROS FACTORES AMBIENTALES Y DE

PAISAJE

TESIS DOCTORAL

EMILIA MARTÍN QUELLER

Licenciada en Ciencias Ambientales

2011

Page 2: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 3: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INVESTIGACIÓN FORESTAL AVANZADA

ESCUELA TÉCNICA SUPERIOR DE INGENIEROS DE MONTES

PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS

EN LOS BOSQUES ESPAÑOLES A DIFERENTES

ESCALAS ESPACIALES: EFECTOS DE LAS

PERTURBACIONES SELVÍCOLAS Y OTROS FACTORES

AMBIENTALES Y DE PAISAJE

Diversity patterns of woody plant species in Spanish forests at

different spatial scales: effects of silvicultural disturbances and

other environmental and landscape factors

EMILIA MARTÍN QUELLER

Licenciada en Ciencias Ambientales

Director:

SANTIAGO SAURA MARTÍNEZ DE TODA

2011

Page 4: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 5: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

iii

Tribunal nombrado por el Sr. Rector Magnífico de la Universidad Politécnica de Madrid, el día de de 20 Presidente: _____________________________________________________________ Vocal: _________________________________________________________________ Vocal: _________________________________________________________________ Vocal: _________________________________________________________________ Secretario: _____________________________________________________________ Suplente: ______________________________________________________________ Suplente: ______________________________________________________________ Realizado el acto de defensa y lectura de la Tesis el día de de 20 en la E.T.S.I. / Facultad EL PRESIDENTE LOS VOCALES

EL SECRETARIO

Page 6: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 7: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

v

A mi madre

Page 8: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 9: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

vii

INDEX

ABSTRACT ................................................................................................................... ix

RESUMEN ..................................................................................................................... xi

INTRODUCTION .......................................................................................................... 1

1. BIODIVERSITY..................................................................................................... 3

1.1 Biodiversity: A complex concept ..................................................................... 3

1.2. Richness and diversity of woody plant species as indicators of forest

biodiversity .............................................................................................................. 4

1.3. Why biodiversity? The effect of biodiversity on ecosystem functioning and

services ..................................................................................................................... 4

2. FACTORS EXPLAINING PATTERNS OF BIODIVERSITY ......................... 6

2.1. A landscape approach ..................................................................................... 6

2.1.1 The spatial scale........................................................................................... 6

2.1.2. Patterns and processes from a landscape ecology perspective ................... 7

2.1.3.The Spanish National Forest Inventory....................................................... 8

2.2. Drivers of species richness and diversity across scales ................................ 8

2.2.1. Factors related with contemporary processes............................................. 9

2.2.1.1. Climate/Productivity............................................................................ 9

2.2.1.2. Environmental heterogeneity............................................................. 10

2.2.1.3. Biotic interactions ............................................................................. 11

2.2.1.4. Elevation ............................................................................................ 11

2.2.1.5. Landscape structure .......................................................................... 12

2.2.2. Historical factors ...................................................................................... 14

2.2.2.1. Palaeobiogeography.......................................................................... 15

2.2.2.2. Human history ................................................................................... 15

Page 10: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

viii

3. SILVICULTURAL DISTURBANCES AND BIODIVERSITY ...................... 16

3.1. Disturbances as part of the natural dynamics of ecosystems .................... 16

3.2. Human-induced disturbances in the Mediterranean and in the Iberian

Peninsula ............................................................................................................... 17

3.3. Forestry practices in Spain ........................................................................... 18

3.4. Ecophysiological consequences of silviculture on plant communities ...... 20

3.5. Silviculture and vascular plant diversity..................................................... 21

MOTIVATION OF THE THESIS.............................................................................. 23

OBJECTIVES AND OUTLINE OF THE THESIS .................................................. 25

CHAPTER 1.................................................................................................................. 31

CHAPTER 2.................................................................................................................. 53

CHAPTER 3.................................................................................................................. 75

CHAPTER 4................................................................................................................ 101

DISCUSSION.............................................................................................................. 123

CONCLUSIONS......................................................................................................... 137

CONCLUSIONES ...................................................................................................... 141

ACKNOWLEDGEMENTS ....................................................................................... 143

Agradecimientos ......................................................................................................... 143

REFERENCES ........................................................................................................... 147

APPENDICES………………………………………………………………………………………...….......179

Page 11: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

ix

ABSTRACT

Biodiversity provides many services to humans and influences many ecosystem functions.

Among forest biodiversity indicators, species richness is the most frequently used, although

all components of biodiversity are important. Patterns of species richness and their drivers

are scale-dependent. In Spain most studies analyzing the factors explaining plant species

richness at the landscape scale have focused on relatively large grain and extent spatial

scales, or on smaller grain sizes but only in particular and relatively small Spanish regions.

Other studies have shown that silvicultural disturbances may influence species richness of

vascular plants, but most of them have been developed in boreal or temperate forests; our

understanding of such influences in Spanish woodlands still remains much limited. Effects

of management on Mediterranean forests are expected to be different due to a long lasting

human disturbance regime, and to differences in the ecophysiological responses in the

understorey to canopy opening as related to the prevalent climatic conditions in

Mediterranean areas.

In this PhD we evaluate the factors, and potential underlying ecological processes, that

explain forest biodiversity patterns in Spain, mainly focusing on the species richness of

woody plants in central Spain. In this context, we place special emphasis on the assessment

of the role of silvicultural treatments as a potential factor explaining those patterns. The

spatial scale is a key transversal element throughout the work, since different processes

operate at different scales and potentially influence each other. Intermediate and local grain

sizes were analysed covering relatively large regional extents in order to intend to provide

broad insights that could be of interest beyond a particular stand type, forest land, or

mountainous system.

Species richness of woody plants was better explained than other structural indicators of

forest biodiversity at the habitat level. In general, environmental factors were more strongly

associated to tree and shrub species richness than landscape structure or management

factors. Variables representing biotic interactions had also relatively important effects in

some cases.

While clearcutting treatments had a negative impact on species richness of woody plants

both at local and landscape scales, selection cuttings or improvement treatments benefited

species richness in some cases. Particularly, landscapes in central Spain were richer in tree

Page 12: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

x

and shrub species richness with increasing proportion of forests managed through these

treatments. However, above a threshold of 40-60% the tendency was reversed.

Forest stands selectively cut had also in general more tree species than unmanaged stands.

Although this result was maintained in coniferous and broadleaved forests (either deciduous

or sclerophyllous), the response to selection cuttings varied in relation to coarse climatic

patterns. Indeed, the increase in tree species after selection harvesting was not observed in

most xeric Mediterranean regions. Under summer stressful conditions, canopy facilitation

may prevail to competitive exclusion.

Species richness of trees also depended on ecological processes operating at intermediate

spatial scales. Management practices such as those of dehesas or plantations seemed to be

associated to a reduced tree species in neighbouring forest stands. The number of species in

a managed stand also depended on the proportion of managed forests in the surrounding

landscape. Positive effects on local tree species richness were found when riparian or

unmanaged forests were increasingly represented in a landscape. Actually, we showed that

an increased diversity of potential seed fluxes colonizing a stand was linked to a

proportional increase in the number of tree species in that stand, highlighting the

importance of species pools and connectivity patterns in the forest landscape.

In general, the results obtained in this PhD point out the necessity to widen the spatial scale

at which the forest management plans are usually conducted in order to better achieve the

goals of conserving forest biodiversity and related ecosystem services.

Page 13: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

xi

RESUMEN

La biodiversidad proporciona numerosos servicios al hombre e influye sobre muchas

funciones de los ecosistemas. Entre los indicadores de biodiversidad forestal, la riqueza de

especies es el más utilizado, si bien todos los componentes de la biodiversidad son

importantes. Los patrones de riqueza de especies y los factores que los explican varían en

función de la escala espacial. En España, la mayoría de los estudios que han analizado los

factores que explican la riqueza de especies vegetales han utilizado escalas espaciales

(grano y extensión) relativamente grandes, o bien tamaños de grano menores pero

restringiéndose a regiones concretas relativamente pequeñas. Por otra parte, algunos

estudios han demostrado que las perturbaciones selvícolas pueden influir sobre la riqueza de

especies vegetales. Sin embargo, muchos de estos estudios se han desarrollado en bosques

boreales o templados y por tanto nuestro conocimiento sobre dicha influencia en el caso de

los montes españoles es todavía bastante limitado. Los efectos de la gestión en los bosques

mediterráneos son presumiblemente diferentes debido a la larga historia de perturbaciones

antrópicas a la que han sido sometidos, y a las diferencias en las respuestas ecofisiológicas

del sotobosque a la apertura del dosel arbóreo bajo climas mediterráneos.

En esta tesis se evalúan los factores, y potenciales procesos subyacentes, que explican los

patrones de biodiversidad forestal en España, centrándonos principalmente en la riqueza de

especies de plantas leñosas en el centro de España. En este contexto, se hace especial

hincapié en evaluar cuál es el papel de los tratamientos selvícolas en la explicación de

dichos patrones. La escala espacial es un elemento clave y trasversal a lo largo de este

trabajo, dado que diferentes procesos operan a diferentes escalas y potencialmente

interactúan unos con otros. En este trabajo se utilizaron tamaños de grano intermedios y

locales cubriendo extensiones relativamente grandes con el fin de intentar obtener

conclusiones amplias que pudieran ser de interés más allá de un tipo de rodal o monte

concreto.

La riqueza de especies leñosas fue explicada en mayor medida que otros indicadores

estructurales de biodiversidad forestal a nivel de hábitat. Además, en general, los factores

ambientales se asociaron más a la riqueza de árboles y arbustos que los factores de

estructura de paisaje o de gestión. Por otra parte, las interacciones biológicas también

tuvieron un efecto relativamente importante en algunos casos.

Mientras que las cortas a hecho tuvieron un impacto negativo en la riqueza de leñosas tanto

a escala local como de paisaje, la entresaca o los tratamientos de mejora resultaron

Page 14: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

xii

beneficiosos en algunos casos. En concreto, los paisajes del centro de España fueron más

ricos en especies arbóreas y arbustivas a medida que aumentó la proporción de bosques

gestionados mediante dichos tratamientos. Sin embargo, a partir de un umbral de 40-60%

dicha tendencia se invirtió.

También a escala local los rodales entresacados tuvieron en general más especies arbóreas

que los no gestionados. Este resultado se observó en bosques de coníferas y frondosas

(caducifolias o esclerófilas), pero varió en función de las condiciones climáticas regionales.

De hecho, no se observó un aumento de especies arbóreas tras la entresaca en las regiones

mediterráneas más xéricas. Es posible que bajo las condiciones de estrés hídrico estival en

estas regiones, las interacciones de facilitación con el estrato arbóreo superen a las de

exclusión competitiva.

La riqueza de especies arbóreas también dependió de procesos ecológicos que operan a

escalas espaciales intermedias. Por ejemplo, la mayor presencia en el paisaje de las

prácticas de gestión asociadas a las dehesas o las plantaciones se tradujeron en una

reducción en la riqueza arbórea en rodales cercanos. El número de especies de árboles en

los rodales gestionados también estuvo influido por la proporción de bosques gestionados a

su alrededor. Asimismo se encontró un efecto positivo del aumento en el paisaje de bosques

de ribera o no gestionados sobre la riqueza local. De hecho, los resultados demostraron que

un aumento en la diversidad de flujos de semillas que pueden colonizar un rodal está ligado

a un aumento proporcional en el número de especies arbóreas en dicho rodal, lo que subraya

la importancia de la riqueza regional de especies y de los patrones de conectividad en el

paisaje forestal para explicar estos procesos.

En general, los resultados de esta tesis señalan la necesidad de ampliar la escala espacial de

monte individual a la que generalmente se plantean los proyectos de ordenación y planes de

gestión forestal con el fin de contribuir a una mejor consecución de los objetivos de

conservación y fomento de la biodiversidad forestal y de los servicios ecosistémicos

asociados.

Page 15: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

Page 16: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 17: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

3

INTRODUCTION

1. BIODIVERSITY

1.1 Biodiversity: A complex concept

Biodiversity is the diversity of life in all its forms and all its levels of organization (Hunter,

1990), including a broad spectrum of biotic scales, from genetic variation within species to

biome distribution on the planet (Gaston, 1996). Given the complexity encompassed by the

term biodiversity, indicators are used to assess each of its aspects. Three main components

of biodiversity have been identified (Noss, 1990): compositional, structural and functional.

These components have been used to develop indicators for biodiversity (Spanos et al.,

2009).

Compositional indicators are related to the biological component; the identity and variety of

elements (Spanos et al., 2009). The most commonly used are, by far, species richness (the

number of species present in a system) and species diversity (the number and evenness of

abundance of the species present in a system). Other compositional indicators are

presence/absence of key species (e.g. endemism, endangered or sensitive species), and the

composition or identity of the assemblages.

Structural indicators are related to the physical organization of the elements (Spanos et al.,

2009), and can drive ecosystem processes and thus shape biological diversity (Spies, 1998).

For example, in the case of forest ecosystems, accepted structural indicators are stand

complexity, deadwood or forest age (COST Action E43, 2004-2008). Complex stands (i.e.

with high vertical and horizontal heterogeneity) normally harbor higher diversity of biota

(e.g. MacArthur and MacArthur, 1961; Brokaw and Lent, 1999; Gil-Tena et al., 2007).

Increased amount of deadwood promotes diversity of insectivores (particularly birds) and

cavity-nesting species, by providing key niches (see Vandekerkhove et al., 2009). Finally,

the specific ecological conditions of old forests are required by some species, called

ancient- or true-forest species, to persist (e.g. Honnay et al., 1999). Also in old-growth

forests, their high diversity of resources, the presence of old trees, or the indirect effect of

increased stand complexity and amount of deadwood benefit species richness of forests

birds and arboreal mammals (e.g. Díaz, 2006; Loyn and Kennedy, 2009).

Finally, functional indicators are those related to ecological and evolutional processes.

Some examples are cited in section 1.3.

Page 18: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

4

1.2. Richness and diversity of woody plant species as indicators of forest biodiversity

Woody plants are characterized by their large size and longevity compared to non-woody

specimens, and by an important difference in size between individuals (stratification). These

attributes make woody plants ecosystem engineers (Wright and Jones, 2006) and landscape

modulators (Shachak et al., 2008). This biological group, and especially trees, regulates the

ecological conditions (microclimatic, edaphic and structural) of the forest system

dynamically building resource niches that ultimately determine the composition of forest

communities. Tree species richness has often been linked to increased diversity of other

organisms such as butterflies (e.g. Vesseby et al., 2002), birds (e.g. Gil-Tena et al., 2007;

Kissling et al., 2008) or arthropods (e.g. Sobek et al., 2009). Nonetheless, negative

interactions can sometimes lead to opposed diversity patterns in other forest layer

components, such as herbs or shrubs (e.g. Lloret et al., 2009).

1.3. Why biodiversity? The effect of biodiversity on ecosystem functioning and services

There is an increasing public and government interest in the conservation of biodiversity

(Rands et al., 2010). Biodiversity provides numerous essential services to society. Recent

estimates claim that the economic value of benefits from biodiverse ecosystems may be 10

to 100 times the cost of maintaining them (TEEB, 2009). Services to society include

material goods (for example, food, timber, medicines, and fiber) and nonmaterial benefits

such as a positive impact on human mental health (Millennium Ecosystem Assessment,

2005). Nonmaterial aspects, including ethical and aesthetical, have also direct economical

implications (e.g. ecotourism). There is also increasing concern about the consequences of

biodiversity loss for ecosystem functioning. A long history of ecological experimentation

and theory supports the postulate that ecosystem goods and services, and the ecosystem

properties from which they are derived, depend on biodiversity, broadly defined (Hooper et

al., 2005).

The effects of species richness or diversity, as indicators of biodiversity, on different

ecosystem properties are more controversial (Hooper et al., 2005). Experimental and

observational studies agree in the patterns of relationships between species

diversity/richness and some ecosystem functional properties (see references in Hooper et

al., 2005). For instance, in species-rich communities, complementarity in the pattern of

resource use by certain combinations of species results in increased rates of productivity

and nutrient retention. Also, higher species richness within a particular site has been shown

to reduce invasion by exotic species. However, results of the influence of species diversity

Page 19: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

5

on other ecosystem functions (e.g. decomposition, nutrient cycling) are more variable.

Instead, functional diversity or even functional traits (e.g. leaf chemistry, phenology) of the

dominant plant species, community composition, trophic structure or disturbance regime,

seem to play a major role in determining these ecosystem properties (Hooper et al., 2005

and references therein).

Species richness can also affect stability of ecosystem process rates (note than this is not

necessarily the same than stability of community composition); theory and several

mathematical models and manipulative studies support this (see Hooper et al., 2005).

“Stability” in biotic communities can refer to a large number of potential phenomena,

including, but not limited to, resistance and resilience to disturbance (see references in

Hooper et al., 2005). Redundancy and compensation mechanisms are hypothesized to

explain this pattern. Ecosystem properties are expected to be buffered against perturbations

when there is a redundancy of types of functional effects among species in a community

(i.e., several species have a similar influence on a particular ecosystem property), and at the

same time, diversity of functional-response types favors compensation among species

(Hooper et al., 2005). For example, a mix of species with two different functional responses

to fire disturbances, seeders and sprouters, makes communities more resilient (Lavorel,

1999). Higher diversity of species is generally associated to an increase in diversity of

functional traits (Schmid et al., 2002). However, the evolutionary history of a community

might have favored the selection of species adapted to a particular disturbance regime or

environmental conditions, i.e., with a relatively restricted range of functional characteristics.

An increase in species richness, in this context, might result in a higher proportion of

species unable to tolerate the new conditions when rare disturbance events occur (e.g. Lloret

el al., 2007; see also below the section “Disturbances as part of the natural dynamics of

ecosystems”).

The effect of diversity on ecosystem properties can be confounded with variation in

environmental conditions, an important control of species diversity. The difficulty of

finding diversity gradients in homogeneous environmental conditions in natural systems

explains, in part, the predominance of the local-scale, theoretical and experimental studies

in this context, mainly in grasslands. Out of these particular and restrictive conditions, one

pattern that is often observed is a unimodal relationship in which species richness is greatest

at intermediate levels of resource availability, stress, productivity, or disturbance (see

references in Hooper et al., 2005). However, this macroecological pattern is seen in

response to gradients of abiotic factors that influence productivity (Hooper et al., 2005); that

is, it reflects the effect of productivity on species richness, rather than the opposite

Page 20: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

6

relationship. Similarly, studies focusing on more complex systems, such as forests, and

consequently relatively large spatial scales, have also found a positive or at least neutral

association between tree diversity and functions related to productivity and soil parameters

(decomposition, nutrient cycling). However, confounding environmental factors were

generally not controlled (see Nadrowski et al., 2010), with few exceptions (e.g. Vilà et al.,

2007).

One of the most convincing arguments for preserving biodiversity may be provided by the

insurance hypothesis (see Bengtsson et al., 2000). The major long-term importance of

diversity may be as a source of species capable of performing desired ecosystem functions

if the present species should disappear. While global extinction of a species is clearly an

important conservation concern, local extinctions are more common than global extinctions

and can be very difficult to reverse (Hooper et al., 2005). Ecosystem management should be

based in the precautionary principle: “sacrificing those aspects of ecosystems that are

difficult or impossible to reconstruct, such as diversity, simply because we are not yet

certain about the extent and mechanisms by which they affect ecosystem properties, will

restrict future management options even further” (Hooper et al., 2005).

2. FACTORS EXPLAINING PATTERNS OF BIODIVERSITY

2.1. A landscape approach

2.1.1 The spatial scale

An important, general effort in ecology is to understand how deterministic and stochastic

forces combine to shape ecological processes across spatial and temporal scales (Diez and

Pulliam, 2007) and create the observed ecosystem properties. The crucial necessity of

conserving and enhancing biodiversity has propelled extensive research on the ecological

processes underlying observed patterns of species richness, as an indicator of biodiversity,

for different taxa, at varying spatial scales and in contrasted bioclimatic regions (see

references throughout this introduction). However, patterns of species richness or diversity

appear to vary as a function of spatial scale (Ricklefs, 1987; Levin, 1992; Rosenzweig,

1995; Gaston, 2000; Rahbek, 2005; Kallimanis et al., 2008). Additionally, it has been

shown that the predictive power, as well as the rank order of explanatory variables in

models of species diversity, or even the form of the relationship, is function of spatial scale

(Auerbac and Shmida, 1987; Whittaker et al., 2001; Field et al., 2009). The definition and

choice of the grain and extent of scale can directly affect the results of any given analysis

Page 21: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

7

and the comparability of results between similar studies. Indeed, variation in grain (unit of

analysis) size has been shown to significantly influence the relative importance of factors

determining large scale patterns of species richness (Rahbek and Graves, 2001; Whittaker et

al., 2001). As for extent, which captures inter-unit variation, it is expected to be important

because it defines the portion of the planet’s surface that was sampled. The range in

heterogeneity sampled determines whether the sample is representative and sufficient for

describing and analyzing relationships (Field et al., 2009). For instance, while processes

larger than the extent of observation appear as trends or constants in the observation set;

processes smaller than the grain size of observation become noise in the data, since

homogeneity is assumed within each unit (Wu et al., 2006).

So, which scale should be used in the study of diversity patterns? Given that each species

observes environment on its own unique suit of spatial and temporal scales; and that for a

particular species this scales vary among life history stages (e.g. local scale for adult plants,

and landscape-scale for seeds; see fragmentation, connectivity and metapopulation

dynamics below in the section “Landscape structure”) (see Levin, 1992); considering also

that each natural phenomenon has its own distinctive scale (or ranges of scales) (Wu et al.,

2006), there is no single ‘correct’ scale on which to describe communities or ecosystems.

“The problem is not to choose the correct scale of description, but rather to recognize that

change is taking place on many scales at the same time, and that it is the interaction among

phenomena on different scales that must occupy our attention” (Levin, 1992).

2.1.2. Patterns and processes from a landscape ecology perspective

Our ability to evaluate hypotheses about causal mechanisms from observed data, i.e. to

match patterns and processes, is constrained by the intermingled nature of natural systems,

especially over broader areas (Zavala, 2004). First, patterns in biodiversity at a particular

scale are likely to be generated by several contributory mechanisms, rather than by a single

one (Gaston, 2000), and by their interactions. But different patterns can result from

variation in intensity of the same process; and the same pattern can be formed by different

mechanisms (Fortin and Dale, 2005). Furthermore, patterns and processes are intertwined in

space and in time, and feedback effects between them are difficult to distinguish (Fortin and

Dale, 2005). Finally, the historical destruction by humans of many of the natural patterns

hinders even more this task (Nogués-Bravo et al., 2008). Large-scale phenomena cannot be

studied experimentally, for ethical and practical reasons. We must limit to observational,

theoretical and correlational studies, and mathematical and statistical models, finding a

Page 22: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

8

balance between mechanistic understanding and complexity. Although ecological and

historical processes cannot be deduced directly from observed patterns of biodiversity, this

type of studies provide a basis for generating hypothesis about underlying processes.

2.1.3. The Spanish National Forest Inventory

Large datasets accurately representing large study areas are not easily obtained. In the last

years there has been a shift in society demands towards a multifunctional sustainable forest

management. This has motivated the adaptation of the Spanish National Forest Inventory

(SNFI) to this demand in its third cycle (3SNFI; Ministerio de Medio Ambiente, 1997-

2007). Particularly, information about Spanish forest biodiversity has been incorporated

through a set of compositional and structural indicators. The 3SNFI provides information

about woody species composition, number of snags, large trees or structural complexity at

the plot level for the entire Spanish territory (Alberdi et al., 2005). Plots in the 3SNFI are

located systematically in the intersections of the 1 km x 1 km UTM grid that fall inside

forests and other woodlands (defined as lands with tree canopy cover above 5%). Its

systematic sampling design and reasonable spatial resolution make the 3SNFI a valuable

tool for the analysis of forest biodiversity patterns across large spatial extents (Bravo et al.,

2002). Actually, there has been recently an increased number of studies that benefited from

SNFI data for landscape ecology studies (e.g. Lloret et al., 2007; Vilà et al., 2007; García

López and Allué Camacho, 2008; Montoya et al., 2008; Lloret et al., 2009; Gil-Tena et al.,

2009; Torras et al., 2009; Montoya et al., 2010).

2.2. Drivers of species richness and diversity across scales

As mentioned before, the drivers of diversity and/or their relative influence may differ with

spatial scale. Field et al. (2009) performed a meta-analysis to determine the group of factors

that most strongly correlated with species richness at different scales of grain and extent.

Plants and animals, land and water habitats, and insular and connected systems were

represented in the analysis. Their study strongly supports the previously stated claim that

species richness shows strong correlations with climate at large grains and extents

(continental and regional scale), but less so at smaller scales. Influences on species richness

that vary at finer scales tend to average out at larger grain sizes, and are found to be

important at meso- or local scales. Although the relative importance of other factors at

smaller scales is more variable across taxa and scales, in general, environmental

Page 23: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

9

heterogeneity appear as the main factor at mesoscales, and disturbances and biotic

interactions become more relevant at local scales (see references in next sections).

There is a dispute about history and geography versus contemporary processes as the main

modelers of current diversity patterns (Ricklefs, 2004). These two approaches are normally

applied separately. However, it is most plausible that spatial species-richness patterns may

have both contemporary ecological and historical evolutionary origins, which are not

mutually exclusive (see Rahbek, 2005).

Indirect factors (e.g. latitude, altitude) are those that in themselves have no direct impact on

organisms or resources availability but may be correlated with one or more primary factors.

Because the correlation between indirect variables and more direct variables is location

specific and need not to be linear, there is not theoretical expectation regarding the shape of

species responses to indirect variables (Austin, 2007). Many authors have outlined the need

to focus on biological meaningful variables rather than on indirect, in order to understand

the mechanisms controlling species richness (Gaston, 2000; Lobo et al., 2001; Pausas et al.,

2003; Field et al., 2009; Vetaas and Ferrer-Castan, 2008).

Finally, the factors that explain the diversity of woody plants may differ from those of non

woody species because of their larger size, their stratification, and their longevity (e.g.

Gonzalez et al., 2009). The study of woody plant diversity needs to account for these

specificities.

2.2.1. Factors related with contemporary processes

2.2.1.1. Climate/Productivity

Climate affects diversity patterns either by direct physiological effects or by determining

resources availability. Most hypotheses about the mechanisms underlying climate-diversity

relationships are based on the effects of climate on productivity (Currie and Paquin, 1987).

Increased productivity is sometimes associated to wider ranges of available resources,

promoting higher species richness, although it can influence the amount of resources rather

than their variability (Begon et al., 2006). Higher species richness could also be originated

in the increase of individuals with productivity (energy-richness hypothesis), in the larger

availability of species adapted to higher productivity conditions (physiological tolerance

hypothesis), or in the increase in speciation rates with productivity (see Currie et al. (2004)

for a discussion of these and other hypotheses).

Page 24: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

10

The climatic attributes influencing productivity are energy, water availability, and their

combination. These attributes have been quantified through a large set of variables (e.g.

precipitation, temperature, potential or actual evapotranspiration, Thornthwaite Index,

Priestly-Taylor’s α Index), thus evaluating different properties of the ecosystem (e.g.

growing conditions, available environmental energy, plant productivity, moisture deficit).

The use of such different indices, considering or no the interaction energy-water, as well as

the variability in spatial scales and taxa studied, might be the cause of inconsistent and

highly variable relationships between water- or energy-related variables and richness (e.g.

Currie and Paquin, 1987; O’Brien, 1993; O’Brien, 1998; Pausas et al., 2003; Field et al.,

2005; Hawkins et al., 2007; Moreno-Saiz and Lobo, 2008; Vetaas and Ferrer-Castan, 2008).

While direct measurements of net primary productivity are hard to attain, especially at

coarser scales, the assumption of treating most climate variables as directly substitutable for

productivity is problematic (Whittaker and Heegaard, 2003). Keeping in mind this possible

bias, the relationship between productivity and species richness tends to be hump-shaped at

local and intermediate scales and more monotonically positive at regional scales

(Mittelbach et al., 2001; Whittaker and Heegaard, 2003). This is consistent with the thesis

that at small scales, competitive controls can act strongly and limit richness in the most

productive areas. At larger scales, competition forms a less efficient control on diversity,

and the positive effects of resource availability are more evident (e.g. Sarr et al., 2005).

2.2.1.2. Environmental heterogeneity

After accounting for the effect of gross climate variation on species pools, environmental

heterogeneity is typically the best correlate of mesoscale richness (Sarr et al., 2005; Field et

al., 2009). The increase in environmental heterogeneity implies an increase in niches,

allowing more species to coexist via competition-colonization trade-offs (Gardner and

Engelhardt, 2008). Spatial and temporal heterogeneity are also a powerful force for

maintaining genetic diversity within natural populations by providing a background of

continuously changing selection pressures to which the species must respond (Jump et al.,

2009).

Heterogeneity of variable environmental aspects has been analyzed as a diversity driver:

edaphic (e.g. Heikkinen and Neuvonen, 1997), lithological (e.g. Aparicio et al., 2008),

landscape structure (e.g. Miller et al., 1997), etc. However, topography is certainly the most

common source of environmental heterogeneity found to correlate with species diversity

(e.g. Wohlgemuth, 1998; O’Brien et al., 2000; Rey-Benayas and Scheiner, 2002; Pausas et

Page 25: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

11

al., 2003; Vetaas and Ferrer-Castan, 2008). According to Vetaas and Ferrer-Castan (2008),

the explanatory power of topographic heterogeneity might reflect (1) the higher surface area

found in a grid cell with mountain relief; (2) climatic variability due to aspect variability

and orographic rainfall; (3) lithological variability; (4) the role of mountains as refugia

during Pleistocene contraction/expansion cycles.

2.2.1.3. Biotic interactions

Processes related with biotic interactions (e.g. competition, predation, mutualism, disease)

explain species richness patterns mainly at local and intermediate scales (Sarr et al., 2005;

Field et al., 2009). Some examples of variables used as indicators of biotic interactions are

shading, grazing intensity, woody litter, number of old trees, etc. (Field et al., 2009). Few

studies have examined the role of biotic interactions at larger scales (Field et al., 2009)

probably due to the difficulty of separating climatic and biotic influences upon changing

ecosystem patterns (Levin, 1992; Field et al., 2009). In this sense, biotic interactions have

been identified as secondary factors at the landscape scale (e.g. Whittaker et al., 2001;

Begon et al., 2006). However, it has been suggested that macro-scale range boundaries of

species are not determined by physiological limitations directly determined by climate, but

by competition with superior competitors, i.e. a suit of species with faster growth rates

(Whittaker et al., 2001). Similarly, in the case of heterotrophs, an association between

species richness and climatic gradients could be interpreted as reflecting the underlying

strong relationship between plant species richness and climate, by a plant-to-animal trophic

exchange (e.g. Andrews and O’Brien, 2000). Finally, biotic interactions occurring at local

scales may vary depending on the environmental conditions. For example, facilitation

mechanisms seem to be more common in stressful environments (e.g. Callaway et al.,

2002).

2.2.1.4. Elevation

Studies of altitudinal gradients in species richness are numerous (see Rahbek (2005) for a

meta-analysis). Altitudinal gradients represent surrogates for several environmental

gradients such as area, net primary productivity and geometric constraints (Nogués-Bravo et

al., 2008). Because it is not a direct causal factor, interpretation of elevation-species

richness patterns in term of ecological hypotheses can be problematic (Willig et al., 2003).

Also, a general pattern of the relationship has not been found, given the variability in coarse

Page 26: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

12

climate conditions, organisms, spatial scales and history of human intervention across the

world (Rahbek, 2005; Nogués-Bravo et al., 2008).

Different mechanisms have been suggested to interpret elevation-species richness

relationships (see Rahbek, 2005). Altitude has traditionally been viewed as a good surrogate

for productivity, although both a decrease of productivity from sea-level to high altitudes,

and a peak at intermediate elevations have been observed. Another possible explaining

mechanism is the reduction in size of climatically equivalent altitudinal bands with altitude,

which imposes significant constraints on population viability at maximum elevations.

Additionally, it is important to note that human activities have generally affected worldwide

the lower and upper slopes more than the mid-altitudinal habitats (Nogués-Bravo et al.,

2008).

Rahbek (2005), in a quantitative analysis of altitudinal species richness gradients including

204 data sets demonstrated that about 50% of the pattern distributions were hump-shaped.

Nogués-Bravo et al. (2008) showed that diminishing the extent of the study with the same

data set (i.e. the altitudinal gradient surveyed) resulted in change of the pattern from hump-

shaped to a linear decrease. On the contrary, using large extents that encompass different

regional climates can result in misleading patterns (Austin et al., 1984).

2.2.1.5. Landscape structure

Landscape structure has been shown to influence many ecological processes and ultimately

species richness patterns (Turner, 1989). The basic properties of landscape structure,

composition and configuration, are a result of environmental and anthropogenic forces, and

their interaction (Forman and Godron, 1986). The relative importance of such influences is

not well known yet. Ortega et al. (2008) found that the composition of Spanish rural

landscapes depends more on environment than the landscape configuration. In this study

landscape structure was also significantly explained by geographical variables, representing

historical controls of diversity.

Shape irregularity of the patches in a landscape is one aspect of landscape configuration that

has been found to correlate with species richness in various study areas and scales of

analysis (e.g. Honnay et al., 1999; Moser et al., 2002; Honnay et al., 2003; Saura and

Carballal, 2004; Baessler and Klotz, 2006; Saura et al., 2008; Torras et al., 2009). This

association has been explained by edge processes, associated to increased edge length in

more complex or elongated patches. Forest edges receive more radiation, and have higher

Page 27: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

13

temperature and lower relative humidity than the inner core habitat. These microclimatic

conditions favour the occurrence of pioneer species (e.g. Hill and Curran, 2005). Forest

edges are also more exposed to disturbances from adjacent ecosystems, such as human

impact, which hinders the existence of vulnerable species, but enhances occurrence of

generalist species, that tolerate frequent disturbances and contrasting environmental

conditions (Forman, 1995). Longer boundaries also promote interception of animals, which

may increase the proportion of animal-dispersed plant species (see Hill and Curran, 2005),

but also may foster predator pressure on plant recruitment. Finally, some authors have

pointed out that shape irregularity may be an indicator to the level of human energy input in

a system (Moser et al., 2002; Saura and Carballal, 2004).

Probably, the most remarkable process related with landscape structure influencing

biodiversity is habitat loss and fragmentation. The process of habitat loss is defined as the

reduction in the proportion of landscape composed of suitable habitat for a focal species

(Smith et al., 2009), while fragmentation per se is a change in the arrangement or

configuration of the remaining habitat (Smith et al., 2009). Habitat loss and fragmentation

have the following consequences on landscape structure (Fahrig, 2003): reduction in habitat

amount, increased number of habitat patches, reduced sizes of habitat patches, and a higher

isolation of patches. All these aspects are interconnected and their effect on biodiversity is

difficult to assess separately (Fahrig, 2003; Smith et al., 2009). For instance, disappearance

of a habitat patch removes biodiversity held in the lost habitat, but also results in increased

potential isolation of the habitat in the surrounding landscape. According to metapopulation

theory (Hanski, 1999), populations in the remaining habitat patches are maintained by a

dynamic balance of local extinctions and colonizations, and therefore habitat reduction in

the landscape can end in biotic collapse and species extinction in the landscape (extinction

threshold in habitat amount, Fahrig, 2001). Although the strong negative effect of habitat

loss on species richness is widely accepted (Bascompte and Rodríguez, 2001; Fahrig, 2003;

Montoya et al., 2008; Montoya et al., 2010), contrasted results on the effects of

fragmentation per se on biodiversity have been observed (see Fahrig, 2003; Smith et al.,

2009). It has been suggested that fragmentation per se aggravates the effects of habitat loss

through an accelerated decrease of population size across the gradient of shrinking habitat,

and a sooner appearance in this gradient of the extinction threshold (Fahrig, 2002).

Nonetheless experimental results do not clearly confirm theoretical expectations (e.g.

Montoya et al., 2010). The reduction of the size of a patch also implies a threshold response

in the persistence of populations inhabiting it (see Fahrig, 2003), although this threshold

depends on the relative importance of abundance-dependent local extinction versus

Page 28: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

14

colonization from regional pool (Honnay et al., 1999). The consequences of reduced patch

size may even lead to a cascade of species loss, by the decline of frugivorous vertebrates,

recruitment of vertebrate-dispersed shrubs, and eventually breaking up of facilitative plant

interactions in some ecosystems (Alados et al., 2010). Fragmentation per se can also

influence species richness through the above-mentioned edge related processes, induced by

size reduction of habitat patches (e.g. Fahrig, 2003; Montoya et al., 2010). Finally, it should

be noted that, given the difficulty to evaluate fragmentation as a process (Fahrig, 2003),

most studies consider fragmentation as a state, comparing structure among different

landscapes instead of evaluating the same landscape across a temporal gradient. This fact

explains that many studies found that several small patches rather than a single large one

enhances biodiversity, given that they tend to cover more heterogeneous abiotic

environments (e.g. Honnay et al., 1999; Bascompte and Rodríguez, 2001; Honnay et al.,

2003).

But fragmentation is not an antonym of functional connectivity, since it does not necessarily

imply isolation if the increased distance can still be covered by a particular dispersal flux

(Saura, 2010). Connectivity is defined as the degree to which the landscape facilitates the

movement of species and other ecological flows (Taylor et al., 1993). Connectivity

promotes exchange of diaspores and individual animals among populations, assuring local

and landscape species persistence through metapopulation dynamics (Wilson and Willis,

1975; Noss, 1991; Nathan and Muller-Landau, 2000; Leibold et al., 2004). Increased animal

flux could also raise seed predation, benefiting species otherwise excluded by seedling

competition (e.g. Orrock et al., 2003). Connectivity also facilitates the exchange of pollen,

increasing genetic diversity (Neel, 2008) and promoting within patch recruitment of plants

(Orrock et al., 2003). These mechanisms might explain the enrichment in species in a site

when its connectivity is improved, either as a short-term response (e.g. Malanson and

Armstrong, 1996; Gilbert et al., 1998; Gonzalez et al., 1998; Damschen et al., 2006), or as a

delayed response (Lindborg and Eriksson, 2004).

2.2.2. Historical factors

The dynamics of plant species has been constrained by processes acting at different

temporal scales. On the one hand, palaeoclimatic changes have determined extinction

processes and range dynamics at continental scales and over extended periods. On the other

hand, stochastic processes such as historical contingencies and metacommunity assembly

dynamics seem to play a more relevant role at intermediate or more local scales (see

Page 29: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

15

Rodríguez-Sánchez et al., 2010). However, variables representing historical factors and

their influence on current patterns of species richness at coarse scales not only are difficult

to measure, but also are likely correlated with climate (Ricklefs, 2004). In general, this type

of data is rarely available for incorporation into a GIS. In their meta-analysis, Field et al.

(2009) observed that few studies tested historical controls on plant diversity, especially at

coarse scales. According to the authors, these constraints might explain the few cases where

history appeared as the best correlate of taxa richness. Historical issues may also correlate

with geographical variables, as in Lobo et al. (2001) where the distance to the Pyrenees was

used to assess palaeobiogeographical controls. The distance to the Pyrenees hardly

explained however Iberian plant species richness at the grain of 50x50 km.

2.2.2.1. Palaeobiogeography

The relatively rapid adjustment of the range boundaries of species such as trees to

palaeoclimatic changes (e.g. Benito et al., 2007) is consistent with current macroscale

patterns of species richness as function of energetic (climatic) controls. However, at smaller

scales, past climatic changes may have higher explanatory power of contemporary species

ranges. For example, some boreo-alpine conifers and temperate broadleaved species better

adapted to the harsh glacial conditions during the Pleistocene, have survived in the

mountain ranges under the Mid-Holocene and current warmer climate (Benito et al., 2007).

Indeed, far back to the Mid-Holocene, tree species richness was predicted to be higher in

Iberian mountains than in the plateaus (Benito et al., 2007). Palaeobiogeographical

processes may have also had important imprints on spatial genetic structure. In the case of

the Iberian Peninsula, important geographical barriers such as the Pyrenees, palaeoclimatic

changes and constrains, edaphic controls and limitation on seed dispersal, have resulted in a

remarkably clear-cut divide between lineages from southeastern and western regions in

populations of deciduous oak species (Quercus spp.), Quercus ilex, Pinus pinaster and, to a

lesser extent, Quercus suber species (Rodríguez-Sanchez et al., 2010).

2.2.2.2. Human history

Although the possible effect of historical human influence on contemporary plant

communities has been overlooked by plant ecologists (Briggs et al., 2006), land-use legacy

has been shown to affect current plant communities (see Hermy and Verheyen, 2007 for a

review; Rozas et al., 2009). In a temperate forest in central Europe, Dupouey et al. (2002)

Page 30: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

16

demonstrated that two centuries of farming during Roman times induced gradients in soil

nutrient availability and plant composition that are still measurable almost 2000 years later.

It has been shown, for instance, that former agricultural land use results in a lower number

and abundance in recent forests (i.e. afforested patches after agricultural abandonment) of

herbaceous species typically confined to ancient forests (the so called ancient forest

species), although this is not necessarily true for the total number of plant species (Hermy

and Verheyen, 2007). In particular, competitive inferiority under higher phosphorus

concentrations and other changes in soil properties, the dispersal limitation of ancient forest

species towards recent forests, and conservation of soil quality by biochemical cycling

induced by cultivation, seem to explain this effect. The stronger manipulation by men of

woody plants limits conclusions with respect to past land use influences on this layer.

Ancient forests refer to those which already existed before a certain threshold date, not

necessarily primary forests (those that have never been cleared). Primary forests are scarce,

especially in the Mediterranean Basin (Blondel and Aronson, 1999).

3. SILVICULTURAL DISTURBANCES AND BIODIVERSITY

3.1. Disturbances as part of the natural dynamics of ecosystems

Based on the niche theory (Hutchinson, 1957; Chase and Leibold, 2003), it has long been

emphasized the existence of an equilibrium state of ecosystems in terms of community

composition, defined by the ranking in the competitive ability of species in the system (see

de Angelis and Waterhouse, 1987; Rivas Martínez, 1987; Allué-Andrade, 1990).

Nevertheless, there is wide recognition that the expected composition according to

environmental properties is being continuously modified by biotic interactions and

disturbances. Dynamic- or non-equilibrium paradigm points up disturbances as part of

natural dynamics of ecosystems (e.g. Shea et al., 2004). Disturbance occurrence alters niche

opportunities and results in a disturbance-succession cycle where reorganization of the

system takes place in the time interval between successive disturbances (Huston, 1979).

However, this dynamic equilibrium can also be disrupted when a threshold of intensity,

frequency, duration or extent of the disturbance is exceeded. If this happens, disturbances

can carry ecosystems into a different stability domain (e.g. Bengtsson et al., 2000; Arroyo et

al., 2004). Then, how can we predict the resilience of an ecosystem to different types of

disturbances? Ecosystems may be resilient to the disturbance regime under which they have

evolved (e.g. Sousa, 1984; Bengtsson et al., 2000; Decocq et al., 2004). This statement is

the basis of current management strategies which intend to emulate natural disturbance

Page 31: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

17

regimes (Crow and Perera, 2004; Drever et al., 2006; Kuuluvainen, 2009; Long, 2009) or

the historical range and variation of ecological conditions across multiple time and space

scales (Keane et al., 2009). The second approach acknowledges that human-induced

disturbances may have shaped current plant communities and recognizes the importance to

consider ecological complexity induced by successional dynamics in plant communities,

and not accounted for when just focusing on post-disturbance ecosystem responses.

3.2. Human-induced disturbances in the Mediterranean and in the Iberian Peninsula

The first significant impact of humans was their probable role in the extinction of a number

of large mammals back in the Palaeolithic. The disappearance of certain megaherbivores

had important consequences for the structure of landscapes, notably their role in keeping

forests open and in maintaining a mosaic vegetation structure (see Bengtsson et al., 2000;

Blondel, 2006). Nevertheless, it is during the Neolithic revolution (ca. 6000 yr BP in Spain),

with the beginning of human farming practices, that human influence on forests became

remarkable. Man use of fire since then has been determinant in vegetation dynamics in

Spain (Arroyo et al., 2004). Up until the industrial revolution in the nineteenth century, the

social structure in Spain was basically founded upon a fine balance between agricultural and

pastoral activities, with forests and woodlands being cleared and exploited as needed

through wood-cutting, coppicing and controlled burning. Forests were also managed to

obtain firewood, charcoal and timber for housing, shipbuilding and other forms of

construction (Blondel and Aronson, 1999). The proportion and intensity of these agents of

deforestation have varied in parallel with the social and economic history. In Spain, relevant

periods contributing to the transformation and deforestation of landscapes were (Valbuena-

Carabaña et al., 2010) the colonization of Mediterranean eastern cultures: Phoenicians,

Greeks and Romans (ca. 9 yr BC - 5 yr AD); the strong power of sheep shepherds (La

Mesta) and their transhumance activities across hundreds of kilometers during five

centuries, but especially in the 16th-17th centuries; Spain as a naval power (15th-17th

centuries); the privatization (disentailment) of 36% of the territory in about a century (19th-

20th centuries); and the industrial revolution (19th century), requiring wood for glasswork,

metallurgy and other modern industries. Mining and wars have had also an important role.

Recovery of forests occurred between many of these periods, and also after the substitution

of wood by fossil fuels as a source of energy (20th century). Today, Spanish forest

landscapes are mainly modeled by reforestation-oriented forest policies, natural

afforestation of abandoned agricultural lands, more sustainable and multifunctional

Page 32: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

18

management practices, and increased frequency and extent of forest fires (see a more

detailed discussion about theses processes in the next section).

Overall, artificially increased fire frequency, grazing, deforestation and the consequent

desiccation of the Mediterranean Basin (Blondel, 2006), have historically favored evergreen

sclerophyllous oaks (mainly Quercus ilex), for their resprouting capacity, adaptation to dry

conditions and lower palatability to livestock (Blondel and Aronson, 1999; Blondel, 2006;

Urbieta et al., 2008; Valbuena-Carabaña et al., 2010), at the expense of deciduous oaks as

Quercus pubescens. Pines presenting mechanisms of fire adaptation (e.g. serotiny) such as

Pinus halepensis, and to a lesser extent, Pinus pinaster, has been favored at the expense of

pines as Pinus nigra and Pinus sylvestris (Tapias et al., 2004; Valbuena-Carabaña et al.,

2010).

The strong, continual and long-lasting human pressure in the Mediterranean may have

caused co-evolution of plant species with man over perhaps a period of thousands of years

(Di Castri, 1981); that is, organisms may have evolved life-history traits as a response to

human-induced habitat changes (Blondel and Aronson, 1999; Blondel, 2006). Actually,

Mediterranean-type ecosystems tend to have a high intrinsic resilience to disturbances such

as fire and grazing when these have been part of the evolutionary history (see Lavorel,

1999). There is debate, however, regarding whether Mediterranean ecosystems should be

considered as degraded or as landscapes adapted to disturbance (e.g. Bengtsson et al., 2000;

Hall, 2000; Fabbio et al., 2003; Blondel, 2006; Vogiatzakis et al., 2006).

3.3. Forestry practices in Spain

Forest legislation in Spain is scarce before 1748 and silvicultural management as a set of

well-defined rules and cultivation schemes originated in the late 19th century and became

widely applied in the 20th century (Valbuena-Carabaña et al., 2010). The first School of

Forest Engineers was founded in Spain in 1846. The corps of forest engineers, originating

in that period, focused mainly on the defense of public forests, given the historical context

of disentailment process, promoted by two different laws in 1837 and 1855. (Bauer-

Manderscheid, 1991; Madrigal, 2008; Saura, 2011) The first widely applied forestry

planning instructions in Spain date to 1890, although previous policies exist and the first

management plan dates back to 1882 (Bauer-Manderscheid, 1991; Saura, 2011). These

instructions were characterized by their rigidness and literal adoption of the practices

traditionally applied in Central Europe. Only the normal forest model with all age classes

covering the same area and managed through even-age silvicultural systems was

Page 33: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

19

considered, with the exception of selection cutting where protection objectives prevailed.

The particularities of Mediterranean forests (heterogeneity, instability and low productivity)

have lead to the increasing recognition of the necessity to relax the traditional, long-term

and rigid, planning systems, towards a Mediterranean silviculture and planning (Madrigal,

2008). The evolution of forestry legislation in Spain since 1890 is characterized therefore

by an increasing flexibility in the permitted planning methods and silviculture systems. This

implied relaxing the target of the normal forest with predefined and fixed regeneration

periods, tailoring locally adapted management strategies, and focusing mainly on the short-

term planning while considering long-term plans as a recommendation only (Montero and

Madrigal, 1999; Saura, 2011).

An important part of the efforts in forestry practices during the 20th century in Spain was

also directed towards reforestation. From 1901-39 many reforestations were promoted for

hydrological purposes, and most of these plantations have become nowadays mature forests

(Valbuena-Carabaña et al., 2010). In the period 1940-1983 11.34% of the Spanish forest

surface had been reforested with protective aims, using chiefly native pine forests

(Valbuena-Carabaña et al., 2010). These plantations were expected to facilitate

establishment of late-successional species such as oaks, and naturalize with time. However,

the lack of posterior silvicultural treatments has lead to a densification and excessive

canopy cover, hindering the growth of oak seedlings in the understory, increasing the risk of

wildfire ignition and spread (e.g. Pausas et al., 2004), and limiting bird species diversity

(Gil-Tena et al., 2007).

In the last decades Spain has undergone deep social and administrative changes. The

transfer of authority on forest policy and management to the regional governments and the

entry of Spain to the European Union conditioned wildlife conservation and forestry

practices. Rural depopulation, replacement of traditional energy sources and building

materials, and low economical profitability of wood products have caused the abandonment

of less productive and accessible agricultural fields, and the reduction in silvopastoral

practices (e.g. Fabbio et al., 2003; Poyatos et al., 2003; Moreira and Russo, 2007). As a

consequence, spread of forests in Spain through secondary succession in abandoned fields is

taking place, and at the same time preexisting forests are increasing in age (maturation) and

biomass. Excessive densification of forests results in relatively uniform stand structures

(Honnay et al., 1999), low spatial and temporal heterogeneity in light conditions (Marañón

et al., 2004), and predominance of shaded conditions and water deficit in the understory

(Valladares et al., 2004). These attributes might be associated to a decline in biodiversity.

Although expansion of forests might mitigate negative effects of climate change, improve

Page 34: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

20

habitat connectivity and foster diversity of forest dwelling species (e.g. Suárez-Seoane et

al., 2002; Moreira and Russo, 2007; Gil-Tena et al., 2009), it is not clear whether the

decline in the traditional patterns in Mediterranean landscapes and in open habitats will

have negative effects on species diversity of organisms such as reptiles or birds, especially

for Mediterranean species (e.g. Moreira and Russo, 2007; Sluiter and de Jong, 2007). One

of the greatest impacts of rural abandonment is probably related to fire disturbances.

Landscape homogenization and connectivity and fuel accumulation increase the probability

of occurrence and the extent of fires (Vega-García and Chuvieco, 2006). Large fires may

create even more fire-prone landscapes covered by scrublands (Loepfe et al., 2010), and the

subsequent increase in fire frequency could compromise autosuccession capacity of

Mediterranean ecosystems (Lloret, 2004; Pausas et al., 2004). From 1961 to 2005 a forest

surface representing 93% of the area reforested between 1940 and 1983 in Spain was

burned (Ministerio de Medio Ambiente, 2006); however, an overall steady state or slight

increase in forest land is observed even in frequently disturbed regions as eastern Spain

(Pausas et al., 2004). This is due to post-fire regeneration and old-field colonization.

3.4. Ecophysiological consequences of silviculture on plant communities

Silviculture involves a set of practices that have direct or indirect impacts on different

aspects of forest ecosystems. For instance, mechanical site preparation, when applied,

modifies microtopography, reducing the occurrence of pits and mounds, and thus inducing

microclimatic homogenization (e.g. Ramovs and Roberts, 2003). It also produces shifts in

the relative proportions of substrates in the soil (e.g. mineral soil, undisturbed litter, woody

debris), increases nutrient availability and may result in soil compaction (Bock and Van

Rees, 2002). Biotic conditions following direct plant removal together with the new

environmental characteristics will condition recruitment of new species. Abundance and

growth of residual sources of vegetation, that is, seed and diaspore seed bank, advance

regeneration, and vegetative reproductive organs, also decline after mechanical site

preparation (Newmaster et al., 2007). Timber skidding can have similar impacts on forest

ecosystems (e.g. Burke et al., 2008).

Changes in overstory composition can also influence seedling colonization and

establishment by modifying litter composition (and indirectly microclimate and seedbed

characteristics) and light conditions in the understory (determined for instance by the

deciduous/coniferous proportion in the overstory) (Ramovs and Roberts, 2003). Another

consequence of some types of management, determined by the frequency of the rotation

Page 35: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

21

period, is a decline in elements characteristic of old-growth forests, such as deadwood (e.g.

Atlegrim and Sjöberg, 2004; Torras and Saura, 2008, old and large trees (e.g. Torras and

Saura, 2008), among other features (see Bengtsson et al., 2000; Rowland et al., 2005).

Some of the most remarkable changes in forest ecosystems are those derived from canopy

removal (Valverde and Silvertown, 1997). Canopy-gap opening results in a direct increase

of light and summer temperature in the understorey. The increase in potential

evapotranspiration entails a decline in summer water availability, which reduces

photosynthetic capacity of plants in the understory and may compromise seedling viability

(Aussenac, 2000), especially in the Mediterranean (Montero et al., 2003; Valladares et al.,

2004a; Serrada et al., 2008); too strong radiation can also induce photodamage (Valladares

et al., 2004b). But an increase in the amount of radiation also promotes photosynthetic

activity in the understory. Furthermore, reduction in rain interception and transpiration by

the overstory leads to increased water availability (Aussenac, 2000; Montero et al., 2003;

Serrada et al., 2008). Additionally, the increase in temperature can foster decomposition of

organic matter, weathering of soil minerals, and nitrogen fixation (Kozlowski, 2002).

Therefore, establishment of new species in the understory can be favoured by a release in

resources after harvesting. Nonetheless, the increase in water availability might be temporal

due to water interception by the herbaceous layer, and thus competition with dominant

species might constraint survival of some plant species (Aussenac, 2000; Serrada et al.,

2008; Wagner et al., 2011). Canopy removal also leads to higher daily and seasonal

temperature variability. This increases the risk of late frosts (Aussenac, 2000; Wagner et al.,

2011), particularly in Mediterranean mountains. Finally, in areas with strong winds, higher

evapotranspiration reduces water availability because of the increased wind velocity

(Aussenac, 2000; Serrada et al., 2008). Summarizing, the responses of a plant community to

canopy removal will depend on a trade-off between all these processes, which at the same

time depend on the degree of canopy opening, and the prevailing climatic and resource

conditions of the site.

3.5. Silviculture and vascular plant diversity

Ecophysiological changes induced by silviculture disturbances alter plant community

composition and triggers succession dynamics. Consequently, silviculture practices may

modify plant species diversity in a community. Evaluation of the consequences of

silviculture on plant diversity has been the focus of substantial research in the last years,

mainly in North America and Scandinavia. These studies generally compare species

Page 36: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

INTRODUCTION

22

richness between managed and unmanaged stands, but there is a high variability in the

degree of harvest intensity analyzed, the time since disturbance, and the bioclimatic region.

Overall, an increase in plant species richness in the understory within 1-5 years after

disturbance has been observed in North America (e.g. Gilliam et al., 1995; Roberts and Zhu,

2002; Schumann et al., 2003; Zenner et al., 2006; Burke et al., 2008), Northern and Central

Europe (e.g. Degen et al., 2005; Götmark et al., 2005), and the Mediterranean Region (e.g.

González-Alday et al., 2009, 2010; Navarro et al., 2010). This short-term enrichment was

more marked with increasing harvest intensity, although in the extreme case in which forest

is clearcut, the comparison is done between mature forests and early successional stages,

which are very different regardless of whether or not they are managed (Paillet et al.,

2010a). The abovementioned studies also showed in general that the increased species

richness was due to colonization by early successional species and occurred at the expense

of forest specialist species.

Diversity patterns observed across successional gradients or, in other words, with varying

time after disturbance are less clear and may be system-specific (Roberts and Gilliam,

1995). Although some studies have monitored stands for relatively long periods since

harvest (e.g. Halpern and Spies, 1995; Brashears et al., 2004), most works use

chronosequences or compare harvested stands with old-growth, control stands. Some

authors have reported higher species richness of vascular plants in the understory (Scheller

and Mladenoff, 2002), or of woody plants (Onaindia et al., 2004), in stands clearcut 60-80

years before than in old-growth stands. Similarly, stands than had been partially harvested 9

or 20 years before had also more species of vascular plants in the understory than

unmanaged stands (Kraft et al., 2004; McDonald et al., 2008). Long-lasting decline (50

years after harvest) in herbaceous species richness has also been observed after clearcutting

(Meier et al., 1995; Duffy and Meier, 1992). Nonetheless, in many studies, the number of

vascular plant species in the understory recovered from initial changes after harvest or was

maintained, compared to old-growth stands, 30-60 years after clearcutting (Albert and

Barnes, 1987; Metzger and Shultz, 1984; Halpern and Spies, 1995; Okland et al., 2003;

Krzic et al., 2003; Kern et al., 2006), or 9-50 years after partial harvest (Albert and Barnes,

1987; Kern et al., 2006). All the mentioned studies concern North America, except Okland

et al. (2003), which study area is located in Norway; and Onaindia et al. (2004), in the

Basque Country (Spain). Few studies exist on the intermediate to long-term consequences

of silvicultural regeneration treatments on plant species diversity in the Mediterranean.

Initial research points to the enrichment in woody plant species some decades after partial

harvesting (Atauri et al., 2004; Torras and Saura, 2008).

Page 37: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

MOTIVATION OF THE THESIS

23

MOTIVATION OF THE THESIS

As we have explained, biodiversity is key for ecosystems’ functioning and services.

Understanding the ecological processes that drive biodiversity is a necessary step to protect,

manage and enhance it. Biodiversity is driven by multiple interacting factors acting at

multiple spatio-temporal scales. Therefore, landscape ecology represents an indispensable

approach to complement local studies on biodiversity. Other components of biodiversity

besides the most frequently studied species richness should also be considered in this task

(chapter 1). The Mediterranean Basin is one of the 25 biodiversity hotspots of the world

(Myers et al., 2000). However, this region is underrepresented within the scientific literature

on the factors explaining landscape patterns of species richness (Field et al., 2009). In

Spain, studies of plant species covering the whole Iberian Peninsula applied a grain size of

50x50 km (Lobo et al., 2001; Moreno-Saiz and Lobo, 2008; Vetaas and Ferrer-Castán,

2008), used regions between 100 and 5000 km2, or focused on local communities (Rey-

Benayas and Scheiner, 2002). Other studies have used smaller grain sizes, although they

focused on more restricted geographical ranges (spatial extent), being more difficult to

extrapolate. Some examples are Pausas et al. (2003) and Torras et al. (2009), who analyzed

plant, or tree, species richness, respectively, in Catalonia (NE Spain) in 10x10 km cells; or

Aparicio et al. (2008) and Torras and Saura (2008), where patterns of different biodiversity

indicators at local scale were analyzed at the extent of Catalonia. The understanding of the

causal mechanisms shaping current patterns of plant species richness requires the widening

of the range of spatial scales (grain or extent) analyzed until present in Spain (chapters 1, 2,

3, 4).

Ecosystem organisms seem to be adapted, and therefore be resilient, to the disturbance

regime to which they have been historically subjected. Silvicultural disturbances have been

shown to potentially influence plant species richness, and can affect ecosystem resilience.

We know very little about the consequences of silvicultural practices on biodiversity in the

Mediterranean. However, we can expect a different ecological response of Mediterranean

plant communities to silviculture than that of temperate or boreal communities. Indeed, the

Mediterranean region has a long-lasting history of human influence and, therefore, a

different historical regime of disturbances. Additionally, differences in light and water

availability among Mediterranean and temperate or boreal forests can be very marked.

Microclimatic changes after canopy opening will probably have a different impact on

Mediterranean plant communities, being water, rather than light, the limiting resource.

Page 38: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

MOTIVATION OF THE THESIS

24

Besides, abandonment of silvicultural practices is influencing current vegetation dynamics

and the regime of the main natural disturbance type in the Mediterranean, wildfire. In this

context, it urges the evaluation of the impacts of silvicultural practices on Mediterranean

forest biodiversity (Scarascia-Mugnozza et al., 2000), and particularly on species richness

of woody plants as an indicator of biodiversity (chapters 2, 3, 4). Deepening in this issue

will provide appropriate decision tools for forest, and landscape, planning and management.

Additionally, a landscape approach based on statistical modeling has scarcely been applied

in this context. However, this approach will allow assessing the role of human disturbances,

and consequent biotic interactions, in the context of other driving factors of plant species

richness more frequently explored in the literature. Previous research has reported the main

explanatory variables of species richness patterns at relatively coarse grain sizes in Spain.

An important question now is whether disturbances play a significant role modelling

landscape (gamma) species richness when working with intermediate grain sizes (chapter

2). But a landscape approach can also provide insightful information about processes

influencing local species richness. Ecophysiological consequences of canopy removal

depend on the main abiotic conditions. Therefore, evaluating the response of species

richness to silviculture along a wide gradient of climatic conditions will improve

understanding on the underlying ecophysiological mechanisms (chapters 2, 3). Finally,

given that communities are not isolated, their species richness and the reorganization after

disturbances will depend on local extinction-colonization dynamics with other communities

in the landscape (i.e. metacommunity dynamics). Understanding the consequences of

silviculture on biodiversity may not be possible without enlarging the focus to the landscape

surrounding the target community (chapters 3, 4).

Page 39: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

OBJECTIVES AND OUTLINE

25

OBJECTIVES AND OUTLINE OF THE

THESIS

This doctoral dissertation has been structured into four chapters that cover different aspects

and stages of the research. The two main objectives are:

- to study the factors (and underlying processes) shaping patterns of forest

biodiversity in Spain.

- to evaluate the impact of forest management practices on forest biodiversity from

a landscape ecology perspective.

The specific objectives of the thesis have been grouped according to the grain (scale) of the

analysed biodiversity patterns (landscape or local scale biodiversity):

Landscape (gamma) biodiversity

1) Which factors, and underlying processes, better explain biodiversity indicators

measured at the landscape scale? We considered different explanatory variables:

a. Environmental variables (climatic and topographic). Chapters 1, 2

b. Landscape structure variables (fragmentation and shape irregularity).

Chapter 1

c. Biotic variables. Chapter 2

2) How different indicators of biodiversity relate to the same explanatory

(environmental or landscape structure) variable? Chapter 1

3) Which are the effects of forest management practices on landscape (gamma)

richness of woody species? Chapter 2

4) Which are the spatial patterns of the analysed relationships and their possible

underlying ecological processes? Chapter 2

Page 40: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

OBJECTIVES AND OUTLINE

26

Local (alpha) biodiversity

5) Does the response of local tree species richness to silviculture depend on factors

and/or processes operating at multiple, larger scales, following a hierarchical

structure? Chapter 3

6) How is the richness of tree species in a community influenced by management

practices in the landscape surrounding it? Chapter 3

7) Is the diversity of potential seeds colonizing a locality from the surroundings

related with the number of tree species observed in that locality? Chapter 4

8) To what extent the response of species richness in a community to silvicultural

disturbances is affected by metacommunity processes from outside the

community? Chapter 4

The main characteristics of each chapter in terms of the spatial scale and biodiversity

indicators analyzed is shown in Table 1.

First of all, because biodiversity is a complex concept, we characterized forest

biodiversity by analysing a set of five structural and compositional indicators (chapter

1). Particularly, we assessed which environmental or landscape structure factors better

explained the patterns of each of these biodiversity indicators. In this first approach the

units of analysis were forest habitats; by removing variability due to forest types, we

intended to control for other ecological, historical or management factors not considered

in this first study.

In order to address the second main goal, in chapter 2 we focused on two compositional,

biodiversity indicators and used 25 km2 UTM cells as grain size. We wanted to test

whether forest management practices (agrosilvopastoral practices, forest regeneration

and improvement treatments) had a significant impact on biodiversity patterns at

intermediate spatial scales, and to discuss the observed relationships. In that purpose,

other environmental and biotic variables ought to be included in the analysis, and the

observed relationships were also discussed and compared with results in previous

literature. Furthermore, we evaluated spatial variation of the relationships by varying

the scale of extent to help interpreting the results.

Page 41: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

OBJECTIVES AND OUTLINE

27

Subsequently, we wanted to understand local biodiversity patterns in a large-extent

study area, by adopting a hierarchical approach of factors acting at multiple scales

(chapter 3). In this approach, the response of biodiversity to local silvicultural

treatments was influenced by factors operating at larger spatial extents (environmental

region) or organizational levels (functional group). Additionally, both local species

richness and its response to silvicultural disturbances were expected to depend on the

type of management of woodlands in the surrounding landscape.

As a further step, biodiversity responses to surrounding landscape were more deeply

analysed, by incorporating information about species composition and connectivity,

using a graph-theory approach (chapter 4). We assessed the potential role of

metacommunity dynamics, and their interaction with local silvicultural disturbances,

shaping patterns of local tree species richness.

Finally, we discussed the results obtained within a wider theoretical framework,

provided some guidelines for sustainable forest management and summed up into some

key conclusions.

Page 42: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1 CHAPTER 2 CHAPTER 3 CHAPTER 4

Patterns of biodiversity indicators

Biodiversity indicator

Tree species richness

Tree species diversity

% uneven-aged stands

Snag abundance

Shrub species richness

Tree species richness

Shrub species richness Tree species richness Tree species richness

GRAIN SCALE forest habitat UTM cell (25 km2) local stand (plot) local stand (plot)

EXTENT SCALE

Spatial Area 150000 km2 150000 km

2 150000 km

2 88000 km

2

Max. covered distance 1000 km 650 km 650 km 490 km

Max latitudinal distance (N-S) 650 km 570 km 570 km 440 km

Environmental Annual precipitation range 350 - 1940 mm 300 - 1900 mm 60 - 1860 mm 250 - 1500 mm

Annual temperature range 5.7 - 15.9 ºC 5.5 - 16.5 ºC 5.2 - 16.9 ºC 8.4 - 15.4 ºC

Elevation range 120 - 1950 m 300 - 1900 m 300 - 1990 m 450 - 1700 m

Table 1. Biodiversity indicators and spatial scale analyzed in each chapter

Page 43: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

Page 44: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 45: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

31

CHAPTER 1

Large-scale determinants of biodiversity across Spanish

forest habitats: accounting for model uncertainty in

compositional and structural indicators *

In this chapter we explore the influence of large-scale explanatory factors (climate, topography,

and landscape configuration) on the distribution of six forest biodiversity indicators across 213

Spanish habitats, taking into account the inherent model uncertainty when dealing with a complex

and large set of variables.

* This chapter has been published as: Martín-Queller, E, Torras, O., A lberdi , I., Solana, J., Saura, S. 2011.

Large-scale determinants of biodiversity across Spanish forest habitats: accounting for model uncertainty in

compositional and structural indicators. Forest Systems 20: 151-164. The affiliations of the other coauthors of this

article are: Universitat de Lleida (O. Torras), Instituto Nacional de Investigación y Tecnología Agraria y

Alimentaria (I. Alberdi) and Universidad Politécnica de Madrid (J. Solana)

Page 46: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

32

ABSTRACT

An integral understanding of forest biodiversity requires the exploration of the many

aspects it comprises and of the numerous potential determinants of their distribution. The

landscape ecological approach provides a necessary complement to conventional local

studies that focus on individual plots or forest ownerships. However, most previous

landscape studies used equally-sized cells as units of analysis to identify the factors

affecting forest biodiversity distribution. Stratification of the analysis by habitats with a

relatively homogeneous forest composition might be more adequate to capture the

underlying patterns associated to the formation and development of a particular ensemble of

interacting forest species. Here we used a landscape perspective in order to improve our

understanding on the influence of large-scale explanatory factors on forest biodiversity

indicators in Spanish habitats, covering a wide latitudinal and altitudinal range. We

considered six forest biodiversity indicators estimated from more than 30,000 field plots in

the Spanish national forest inventory, distributed in 213 forest habitats over 16 Spanish

provinces. We explored biodiversity response to various environmental (climate and

topography) and landscape configuration (fragmentation and shape complexity) variables

through multiple linear regression models (built and assessed through the Akaike

Information Criterion). In particular, we took into account the inherent model uncertainty

when dealing with a complex and large set of variables, and considered different plausible

models and their probability of being the best candidate for the observed data. Our results

showed that compositional indicators (species richness and diversity) were mostly

explained by environmental factors. Models for structural indicators (standing deadwood

and stand complexity) had the worst fits and selection uncertainties, but did show

significant associations with some configuration metrics. In general, biodiversity increased

in habitats covering wider topographic ranges and comprising forest patches with more

complex shapes. Patterns in other relationships varied between indicators (e.g. species

richness vs. diversity), or even were opposed (trees vs. shrubs). Our study (1) allowed

deepening the understanding of biodiversity patterns in a large set of Spanish forest habitats

and (2) highlighted the increasing complexity of identifying common landscape conditions

favouring forest biodiversity as the range of analysed biodiversity aspects is widened

beyond the more commonly assessed species richness indicators.

Page 47: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

33

1. INTRODUCTION

The effects of biodiversity on key ecological processes, such as productivity, stability or

nutrient cycling (Loreau et al., 2001; Hooper et al., 2005), have stimulated an increasing

effort to comprehend and characterize it (Huston, 1994; Begon et al., 2006). However,

given the difficulty of thoroughly measuring and quantifying biodiversity even in a small

area, suitable indicators have to be found (Duelli and Obrist, 2003). In the last years,

multiple international agreements such as the Ministerial Conference on Protection of

Forests in Europe (MCPFE, 2003), or the COST Action E43 (2004-2008) have resulted in a

general consensus of the scientific community about some common indicators of forest

biodiversity. Biodiversity indicators are frequently classified into three categories (Noss,

1990): compositional, structural, and functional. The more aspects are considered through

these types of indicators, the better the characterisation and monitoring of biodiversity.

Species richness and diversity are the most common compositional indicators. Examples of

structural indicators, less frequently used, are age evenness, which may indicate the degree

of naturalness and complexity in a forest; or the amount of deadwood, which provides

nesting, refugia and foraging sites for a variety of species.

Landscape ecology approaches provide a necessary complement to the conventional local

approaches in the study of forest biodiversity distribution. Indeed, metrics related to

landscape structure can be used as effective, indirect biodiversity indicators (Lindenmayer,

et al. 2000; Dauber et al., 2003; MCPFE, 2003; Lafortezza et al., 2010). Aspects such as

configuration are necessary for a comprehensive assessment of temporal changes in the

forest landscape (Ortega et al., 2008). See the general introduction for more details on the

association between aspects of landscape configuration such as fragmentation or shape

irregularity and species richness. Landscape biodiversity patterns are strongly influenced by

multiple factors, both biotic and abiotic (see general introduction). Therefore, underlying

environmental drivers need to be considered when analysing forest landscape biodiversity

patterns.

National Forest Inventories (NFI) are a very useful information source for the study of large

scale biodiversity patterns. However, a complex issue is how to scale up the biodiversity

data from the plot level (at which the NFI data are gathered) to wider spatial scales where

many of the drivers of forest biodiversity operate. Many landscape studies use squared

UTM cells as the unit of analysis (e.g. Lobo et al., 2001; Torras et al., 2009; Vetaas and

Ferrer-Castan, 2008). UTM cells are expected to represent relatively homogeneous values

of factors with spatial trends at intermediate to large (regional) scales or extents, such as

Page 48: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

34

climate. However, these units may comprise a variety of forest types (with considerably

different biota, management intensity, disturbance history, etc.) and correspond to artificial

boundaries that do not match with the physical and ecological boundaries actually

influencing the composition of forest communities and that artificially introduce a higher

spatial variability in some of the analyzed variables (such as topography, spatial

configuration, etc.).

For these reasons, here, instead of focusing of equally-sized cells, we used the less

frequently explored habitat level to characterize and analyse the distribution of six forest

biodiversity indicators over a large study area of about 150,000 km2 comprising different

Spanish provinces and forest types. We analysed large scale patterns of biodiversity once

removed the variability due to those forest types. The average of some of the explanatory

factors would be, in that case, more representative of actual causal factors at the habitat

level than at the cell level. Specifically, our objectives were (1) to explore which

environmental (climatic and topographic) and landscape variables (fragmentation and shape

irregularity) better explained the patterns of biodiversity, and (2) to compare the responses

of five different biodiversity indicators (compositional and structural) derived from national

forest inventory data. We aimed at an improved understanding of the relationships between

a large set of biodiversity indicators (larger than those typically considered in previous

studies, which have mainly focused on species richness) and the abovementioned

explanatory factors, which have less been explored in Mediterranean or Spanish landscapes

compared to other regions of the world. Finally, we improved the assessment of these

relationships using the information-theoretic approach by Burnham and Anderson (2002).

Most studies evaluate the strength and significance of the relationships with biodiversity

according to the full model, i.e. that single best model including all the explanatory

variables eventually selected according to statistical criteria. However, regression

coefficient estimates based on a single best model do not take into account model selection

uncertainty and therefore ignore the fact that the ‘best’ model is often highly variable

depending on the sample data set. The misestimation of the true explanatory power of a

particular variable that may result from the widespread use of a single statistical model is

thus avoided in this study (Stephens et al., 2005).

Page 49: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

35

2. METHODS

2.1. Study area and forest habitats

Sixteen Spanish provinces -belonging to the regions of Galicia, Asturias, Cantabria,

Navarra, Catalonia, La Rioja, Madrid, Extremadura and Murcia- were selected as the study

area (Figure 1). These were the provinces where the 3SNFI was already fully completed

when the study was started. The study area covered a large part of the Spanish latitudinal

and altitudinal range. While most of the study area falls mainly within the Mediterranean

region, approximately a third of the region belongs to the Atlantic bioclimatic region

(North-West), and the higher mountains correspond to the Alpine region.

Figure 1. Spanish provinces considered in the analysis (shown in black). See table 1 for the full province

names.

In order to avoid the potentially confounding factor of habitat heterogeneity, we stratified

data by forest habitats. A total of 238 different forest habitat types were initially

differentiated in the study area from the information in the Spanish Forest Map at a

1:50,000 scale (SFM). The SFM was developed within the Third Spanish National Forest

Inventory (3SNFI) (Ministerio de Medio Ambiente, 1997-2007). The SFM has a vector data

Page 50: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

36

structure and a minimum mapping unit of 2.25 ha. It has been developed from the

interpretation of aerial photographs combined with pre-existing maps and field inventory

data. Forest habitats were defined attending to the dominant tree species and its abundance,

the stand development stage and the forest canopy cover. Thereby, variability within these

units of analysis was reduced for these variables. Other environmental factors, such as

climate or topography, can still show a wide range of variation depending on the ecological

requirements of the dominant tree species, but still will correspond to those conditions that

are suitable for the development of a particular forest habitat. In addition, the division of

habitats according to the province in which they were present, as required by the design of

the SNFI and SFM, allowed a decrease of environmental variability within habitats (i.e.

Quercus ilex forests in the province of Cantabria and Q. ilex forests in the province of

Murcia were considered as two different habitat types). The definition of forest here

includes all areas with forest tree canopy cover ranging from 5 to 100%, as defined in the

SFM.

Riparian forests were excluded from this analysis, as their configuration characteristics are

strongly governed by hydrogeomorphic processes not considered in this study (Rex and

Malanson, 1990). We also excluded dehesas, which are scattered tree open woodlands

typical of Spanish extensive farmlands. The spatial and biodiversity patterns of these

agroforestry systems are determined by a particularly high degree of human influence.

Processes occurring in dehesas are outside the scope and the set of variables considered in

this study. Thus, a final set of 213 forest habitat types was considered for subsequent

analyses (Table 1). Forest habitats with dominance of Quercus ilex, Pinus halepensis, P.

pinaster, P. sylvestris, Eucaliptus globulus or Fagus sylvatica are the most abundant in the

study area (Table 1). The range of environmental conditions represented by the analysed

forest habitats is considerably wide, with mean altitude ranging from 122 to 1951 m, mean

annual precipitation from 353 to 1936 mm, and mean annual temperature from 5.7 to 15.9

ºC.

2.2. Forest biodiversity indicators

Forest biodiversity indicators for each of the analysed habitats were obtained from a total of

30,929 inventory plots in the 3SNFI (Ministerio de Medio Ambiente, 1997-2007). Plots in

the 3SNFI are located systematically in the intersections of the 1 km x 1 km UTM grid that

fall inside forests and other woodlands (defined as lands with tree canopy cover above 5%).

The average sampling intensity is of one plot per 1 km2 of land. Plots were circular and the

Page 51: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

37

inventory of tree stems depended on their diameter at breast height (DBH) and distance to

the plot centre, which ranged from 5 m for trees with DBH from 7.5 cm to 12.5 cm up to a

maximum radius of 25 m for trees with DBH of at least 42.5 cm.

PROVINCE DOMINANT FOREST TREE SPECIES

Asturias (A) Castanea sativa, Fagus sylvatica, Eucalyptus globulus

Cantabria (CA) Eucalyptus globulus, Fagus sylvatica, Quercus petraea, Quercus robur

Barcelona (B) Pinus halepensis, Pinus sylvestris, Quercus ilex

Girona (GI) Quercus ilex, Quercus suber, Pinus sylvestris, Pinus halepensis

Lleida (L) Pinus sylvestris, Quercus ilex, Pinus nigra, Pinus uncinata

Tarragona (T) Pinus halepensis

Badajoz (BA) Quercus ilex, Quecus suber

Cáceres (CC) Quercus ilex, Quercus suber, Quercus pyrenaica

A Coruña (C) Eucalyptus globulus, Pinus pinaster

Lugo (LU) Eucalyptus globulus, Pinus radiata, Quercus robur, Pinus pinaster

Ourense (OU) Pinus pinaster, Quercus pyrenaica

Pontevedra (PO) Eucalyptus globulus, Pinus pinaster

La Rioja (LO) Quercus pyrenaica, Fagus sylvatica, Pinus sylvestris

Madrid (M) Quercus ilex, Pinus sylvestris, Quercus pyrenaica

Murcia (MU) Pinus halepensis

Navarra (NA) Fagus sylvatica, Pinus sylvestris, Pinus halepensis

Table 1. Dominant tree species in the most abundant forest habitats in each province (corresponding to

those occupying nearly half of the forest area in the province).

We considered four compositional and two structural biodiversity indicators from the

information gathered in the 3SNFI: tree species richness, tree species diversity (quantified

through both Margalef and Shannon indices), shrub species richness, percentage of uneven-

aged stands and snag abundance (Table 2). Plant or tree species composition, forest (age)

structure and deadwood are accepted indicators of forest biodiversity (MCPFE, 2003;

COST E43, 2004-2008). Other important indicators, such as additional deadwood types

(apart from standing dead stems) and decay classes, were not inventoried in the 3SNFI.

Although for simplicity we will refer to shrub species richness throughout the manuscript, it

should be noted that this indicator, as measured in the 3SNFI, is not a strict estimator of the

total number of shrub species because it is based on a predefined list of 169 taxa (125

Page 52: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

38

species, 42 genera and 2 subfamilies). The inclusion of a certain plant in the list at the level

of species or genus depends on its abundance in the Iberian Peninsula: most frequent

species that can be easily identified are determined at the species level, whereas the rest of

shrubs are grouped at the genus level.

2.3. Landscape and environmental variables

Initially we considered 22 explanatory variables as potential candidates to explain the

distribution of the forest biodiversity indicators at the habitat level. The following fourteen

(computed through the SFM) were landscape structure variables: total forest area, number

of forest patches, mean size of the forest patches, maximum size of the forest patches,

percentage of core area at 100 and 300 m from forest edge, patch cohesion index, mean

distance to the nearest neighbour habitat patch, edge length of the forest patches, mean

shape index, area weighted mean shape index, density of shape characteristic points,

elongation index, and minimum circumscribing circle index. The remaining eight

independent variables were related to climate or topography: mean, maximum, minimum,

range and standard deviation of elevation, mean annual precipitation, mean annual radiation

and mean annual temperature.

To avoid multicollinearity problems, when the Pearson correlation coefficient between two

of the above variables was higher or equal than 0.6 one of them was discarded for

subsequent analyses. The selection of the final explanatory variables took into account that

each of the following categories of landscape or environmental factors should be

represented by at least one variable: fragmentation, shape irregularity, water-energy

availability (productivity), altitudinal gradient, and heterogeneity resulting from

topography. Each of these categories has, according to literature, a distinctive contribution

to relevant processes that affect biodiversity (see general introduction). In addition, in the

case that two highly correlated metrics fell in the same category, selection was based on

metric simplicity and therefore easier interpretation of the results. The eight independent

variables finally selected for subsequent analyses are described in Table 2.

The four landscape configuration metrics were computed for all forest habitat types using

the SFM. We considered that the spatial resolution of this map was enough for the goals of

our study (Diaz-Varela et al., 2009). The information source for the topographic variables

was the official Spanish Digital Elevation Model (DEM) at a resolution of 25 m (Ministerio

de Fomento, 1999). Climatic data were obtained from the Climatic Atlas of the Iberian

Peninsula at a resolution of 200 m (Ninyerola et al., 2005).

Page 53: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

39

Category Abbreviation Description

Biodiversity indicators

Forest

composition

Tree species

richness

Total number of different tree species found in all the plots in the

habitat, included the regeneration strata.

Forest

composition

Tree species

diversity

Calculated through the Margalef and Shannon diversity indices,

based on the proportion of basal area (m2/ha) of each species with

respect to the total basal area in the plot, and averaged at the

habitat level.

Forest structure % uneven-

aged stands

Percentage of plots with uneven-aged stands (as a measure of stand

structure complexity). Uneven-aged stands are defined in the SNFI

as those in which stems are distributed in at least three different

age classes (with at least 10% of the stems in each of them).

Forest structure Snag

abundance

Amount of standing dead wood (number of stems per ha).

Forest

composition

Shrub species

richness

Total number of different shrub species found in all the habitat plots

(based on a taxon list, see methods).

Explanatory variables

Configuration

Fragmentation Mean-SIZE Mean size (area) of the forest patches in the habitat.

Fragmentation MNND Mean nearest neighbour distance (arithmetic mean of the distance

between each habitat patch and the nearest patch belonging to the

same forest habitat).

Shape

irregularity

DSCP Density of shape characteristic points. The total number of shape

characteristic points is based in the minimum number of points

necessary to describe a patch boundary and computed on vector

data as the number of vertices of the polygons with a minimum

vertex angle of 160º (Moser et al. 2002). The density results from

dividing the sum of NSCP in the habitat by the total perimeter of the

forest patches (Saura and Carballal 2004).

Shape

irregularity

MCCI Minimum circumscribing circle index, based on the ratio between

the area of the patch and the area of the minimum circumscribing

circle around the patch. This index attains a minimum value (MCC =

0) for circular patches and increases for more elongated and narrow

patches, up to a maximum value of MCC = 1 (Saura and Carballal

2004).

Environmental

Altitudinal

gradient

Mean-ELEV Mean elevation of the forest habitat.

Heterogeneity Sd-ELEV Standard deviation of elevation of the forest habitat.

Water

availability

Mean-PREC Mean total annual precipitation in the forest habitat.

Energy

availability

Mean-RAD Mean annual solar radiation in the forest habitat.

Table 2. Description of the analysed forest biodiversity indicators and the explanatory variables finally

considered in the analysis. Categories represented by each explanatory variable are shown in italics.

Page 54: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

40

2.4. Data analysis

We performed multiple linear regression analyses, taking biodiversity indicators as

dependent variables and configuration and environmental data as explanatory variables.

Different sampling efforts are expected to influence the value of the indicators in each

forest habitat. To control this, simple linear regressions of each biodiversity indicator

against the total number of plots per habitat were performed. The residuals resulting from

these regressions were used as the final dependent variables in the models. By controlling

the sampling effort, the effect of the amount of habitat on each biodiversity indicator can be

assumed to be removed, given the intensive and systematic sampling design of the SNFI.

Dependent variables were previously transformed (√x, lnx and x2) to meet parametric

assumptions.

The selection of the best regression model (or set of models) for each biodiversity indicator

was based on the Akaike Information Criterion, correcting for sample size (AICc)

(Burnham and Anderson, 2002). We estimated the AICc for all the possible combinations of

the eight explanatory variables: 255 possible models for each indicator (28 minus the model

with an intercept only). The best model had the lowest AICc score, and each model was

compared to it in terms of the difference in the AICc value (∆i). Models for which ∆i < 2

were selected, as they were considered to have substantial support and to be reasonable

models for the data (Burnham and Anderson, 2002). Akaike weights (wi) provide a relative

importance of evidence for each model, and can be interpreted as the probability that a

particular model is the best for the observed data, given the candidate set of models

(Burnham and Anderson, 2002).

The same analysis, based on the AICc values and Akaike weights, was performed

considering only the four landscape explanatory variables (landscape models), and

considering only the four environmental explanatory variables (environmental models).

Thus, two sets of 15 (24-1) candidate landscape and environmental models, respectively,

were evaluated for each biodiversity indicator.

The relative importance of each explanatory variable was assessed using all the set of

models considered, instead of a single model, as proposed by Burnham and Anderson

(2002). For each biodiversity indicator, the sum of Akaike weights across all the models

where the considered explanatory variable occurred was calculated; the larger this sum is,

the more important that variable is, compared to the other variables. Regression parameters

were obtained from the average of their values on the subset of best models (those models

with ∆i<2 for each indicator) weighted by the Akaike weights (wi).

Page 55: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

41

3. RESULTS

3.1. Models performance

3.1.1 Global models

The ranking of the resulting best global models (comprising both environmental and

landscape explanatory variables) for the six biodiversity indicators was almost similar

according to R2 and Akaike weights criteria (Table 3). The highest amount of variance was

explained for the shrub species richness, with only two models with substantial empirical

support to be the best ones (∆i < 2) (Table 3). Tree species richness and diversity quantified

through the Margalef index had also relatively little uncertainty in model selection, with wi

of the best model above 0.2 and 40% and 33% of total explained variance, respectively

(Table 3). The uncertainty of the models increased in the case of Shannon diversity of tree

species, existing seven to nine models with substantial empirical support to be the best ones,

the best one explaining about 30% of the variance (Table 3). Finally, the worst fits and

greatest model uncertainties were found for the percentage of uneven-aged stands and snag

abundance models (Table 3).

GLOBAL MODEL ENVIRONMENTAL

MODEL LANDSCAPE MODEL

R2 wi

Models

∆i<2 R

2 wi

Models

∆i<2 R

2 wi

Models

∆i<2

Tree species richness 0.32 0.20 3 0.19 0.24 6 0.09 0.28 4

Tree species diversity

(Margalef index) 0.40 0.32 3 0.33 0.57 2 0.008 0.20 4

Tree species diversity

(Shannon index) 0.32 0.11 9 0.30 0.47 2 0.04 0.25 4

% uneven-aged stands 0.19 0.04 24 0.006 0.15 8 0.16 0.48 2

Snag abundance 0.08 0.05 13 0.021 0.19 5 0.06 0.25 4

Shrub species richness 0.50 0.45 2 0.46 0.87 1 0.17 0.38 2

Table 3. Corrected coefficient of determination (R2) and Akaike weight (wi) of the best global model

(considering both environmental and landscape explanatory variables), environmental model and

landscape model for the six biodiversity indicators studied. The number of models with ∆i<2 is also

indicated.

Page 56: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

42

3.1.2 Environmental and landscape models

In general, environmental models performed better than those using only landscape

variables, both in terms of explained variance and model uncertainty (Table 3). Even so, the

contrary occurred for the uneven-aged stands and snag abundance. It is noteworthy the case

of shrub species richness, where a single environmental model had a very high probability

of being the best (wi=0.87), with a R2=0.46 (Table 3).

3.2. The role of the explanatory variables

The density of shape characteristic points (DSCP) was highly associated to all the

compositional indicators except the Shannon index, and to the snag abundance (Table 4).

According to these indicators, higher biodiversity can be found in forest habitats with more

complex patches’ boundaries (Table 5). The minimum circumscribing circle index (MCCI)

and mean patch size (Mean-SIZE) were also positively associated to the percentage of

uneven-aged stands (Tables 4 and 5).

The standard deviation of elevation (Sd-ELEV) was one of the most remarkable

explanatory variables for all the indicators related to species richness and diversity (Table

4), with positive relationships (Table 5). Mean total annual precipitation (Mean-PREC) was

an important factor for the same indicators, except for the case of tree species diversity as

quantified through the Shannon index (Table 4). However, the effect of Mean-PREC on tree

species indicators was opposite in sign to that for the shrub species indicator (Table 5).

Forest habitats in more humid regions tended to have higher levels of tree species richness

and Margalef diversity but lower shrub species richness.

Mean elevation (Mean-ELEV) was a relevant factor for the shrub species richness, and the

tree species Shannon diversity (Table 4). Both biodiversity indicators tended to increase in

habitat forests distributed mainly in lowlands (Table 5). Mean radiation had a minor role in

the models compared to other explanatory variables (Table 4).

Mean-SIZE was especially relevant in the shrub model, where a greater size of forest

patches favoured shrub species richness (Tables 4 and 5). As for the mean distance to the

nearest habitat patch, it was a relative important correlate for tree species diversity

(Margalef) and snag abundance (Table 4), in all the cases showing a positive relationship

(Table 5).

Page 57: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

43

Tree species richness Tree species diversity

(Margalef index)

Tree species diversity

(Shannon index)

DSCP (1.00) Sd-ELEV (1.00) Sd-ELEV (1.00)

Sd-ELEV (1.00) Mean-PREC (1.00) Mean-ELEV (0.98)

Mean-PREC (1.00) DSCP (1.00) MNND (0.90)

MPS (0.87) MNND (0.97) Mean-PREC (0.86)

Mean-ELEV (0.77) Mean-RAD (0.93) Mean-RAD (0.74)

MNND (0.74) MPS (0.77) MPS (0.40)

MCCI (0.64) MCCI (0.68) DSCP (0.39)

Mean-RAD (0.39) Mean-ELEV (0.61) MCCI (0.38)

% uneven-aged stands Snag abundance Shrub species richness

MCCI (1.00) MNND (0.95) Mean-PREC (1.00)

MPS (1.00) DSCP (0.94) Mean-ELEV (1.00)

Sd-ELEV (0.65) Sd-ELEV (0.70) Sd-ELEV (1.00)

Mean-RAD (0.53) MCCI (0.56) MPS (0.99)

Mean-PREC (0.53) MPS (0.44) DSCP (0.97)

Mean-ELEV (0.44) Mean-RAD (0.44) Mean-RAD (0.92)

MNND (0.41) Mean-PREC (0.36) MCCI (0.32)

DSCP (0.39) Mean-ELEV (0.36) MNND (0.26)

Table 4. Ranking of the explanatory variables according to their relative importance as predictors for the

six biodiversity indicators considered. The relative importance was estimated considering the weight

evidence (wi) of the models where the considered explanatory variable appeared.

Page 58: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

44

Tree

species

richness

Tree

species

diversity

(Margalef

index)

Tree

species

diversity

(Shannon

index)

Uneven-

aged

stands (%)

Snag

abundance

Shrub

species

richness

MPS 0.16 0.13 -0.07 0.31 0.09 0.18

MNND 0.12 0.17 0.16 0.08 0.19 -

DSCP 0.44 0.28 0.09 0.09 0.21 0.18

MCCI 0.12 0.12 0.07 0.36 -0.13 0.04

Mean-ELEV -0.15 -0.13 -0.25 -0.12 0.06 -0.42

Sd-ELEV 0.37 0.38 0.50 -0.13 0.15 0.30

Mean-PREC 0.36 0.47 0.18 0.13 0.08 -0.45

Mean-RAD -0.08 -0.18 -0.14 -0.13 -0.09 0.16

R2 0.328 0.403 0.316 0.187 0.077 0.495

AVERAGE R2 0.326 0.398 0.316 0.178 0.076 0.495

K 7 8 5 6 4 6

Table 5. Average partial regression coefficients based on the subsets of models with ∆i<2 for each

indicator and on the Akaike weights. R2 of the best model and average R2 for the subset of models with

Ai<2 are reported. K is the number of independent variables appearing in the best model for each

biodiversity indicator.

4. DISCUSSION

4.1. Can environmental and landscape variables explain the distribution of the

biodiversity indicators in the Spanish forest habitats?

The amount of variation of the biodiversity indicators explained by the global models was

moderate, with a maximum of 49.5% for shrub species richness (Table 3). These

percentages of explained variance can be considered, however, relatively high for some

indicators given the wide scale and range of environmental conditions considered in this

analysis. For example, in similar studies the maximum percentages of explained variation in

species richness estimated was 65% (Lobo et al., 2001), 62% (Torras et al., 2009), or 24%

(Rey-Benayas and Scheiner, 2002), although caution in the comparison of these percentages

is needed given the different scales of study and data stratification.

The two groups of indicators, compositional and structural, presented clearly different

tendencies. In the case of the compositional indicators, despite the stratification of the data

Page 59: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

45

by forest habitats rather than cells, environmental models still explained higher variance and

had more acceptable selection uncertainty. The global models for the forest structure

variables (snag abundance and percentage of uneven-aged stands) were unsatisfactory.

These indicators are probably affected by factors acting at more local scales that vary within

the habitat (e.g. forestry practices applied at the stand level). Interestingly, our results at the

habitat level emphasize the role of configuration indices for structural indicators (especially

for the percentage of uneven-aged stands), rather than the environmental variables.

The two diversity indices used in this study for the tree stratum (Margalef and Shannon

indices) performed quite differently. Global and environmental models performed better for

the Margalef index than for the Shannon index. Actually, Margalef index followed the

tendency of the rest of indicators related to vegetation composition, both tree and shrub

species richness. By contrast, important factors found for these biodiversity indicators, such

as the Mean-PREC or the shape complexity (DSCP), were not relevant in Shannon diversity

models. The reason may be the higher bias of the Margalef index towards species richness

(i.e. it is more affected by rare species), while the Shannon index is biased towards species

dominance (i.e. it is more affected by changes in the abundance of the most common

species) (Magurran, 1989). This is supported by the Pearson correlation coefficients (r)

between biodiversity indicators, as the Margalef index was more closely related to tree

species richness (with r=0.870) than to the Shannon index (r=0.640). The results of this

study suggest that evenness on tree species abundance is more difficultly modelled, and is

barely affected by landscape configuration metrics related to shape complexity or

fragmentation, or by environmental variables such as Mean-PREC that did have a

prominent effect on tree species richness.

4.2. Responses of forest biodiversity indicators to environmental variables

Since forest habitats were defined here by the realised ecological niche of a particular

dominant tree species, higher ecological plasticity (e.g. tree species adapted to wide

altitudinal ranges) resulted in an increased heterogeneity within the habitat. According to

results, in general those habitats with higher topographic complexity were remarkably

associated to higher woody plant diversity. This factor has been reported to remarkably

influence diversity in other studies carried out in the Iberian Peninsula (Lobo et al., 2001;

Rey-Benayas and Scheiner, 2002; Pausas et al., 2003; Moreno-Saiz and Lobo, 2008; Vetaas

and Ferrer-Castán, 2008). In our study, high standard deviations of the elevation in a

particular type of forest habitat indicate two characteristics on its distribution (1) it can be

Page 60: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

46

found in mountainous, and thus topographically complex, regions; and/or (2) it extends in a

wide altitudinal range, including elevations from the see level up to 2000 m. In both cases,

the inclusion in the models of heterogeneity induced by topography corrected for the

unrealistic assumption that average climate values were uniform in the large grid cells or

habitat types (O’Brien et al., 2000). In the first case, mountains would also favour diversity

through a higher surface area, bedrocks heterogeneity and availability of refugia during the

last glacial period (Vetaas and Ferrer-Castán, 2008). The second characteristic would

influence indirectly biodiversity through heterogeneity of other environmental, biologic or

human variables not considered here. It should be noted that especially shrub species

richness, but also tree Shannon diversity, were found to decrease with mean habitat

elevation. This suggests that habitats covering a larger variety of environmental conditions

but not restricted to the upper mountainous regions had more shrub species and diversity of

trees. These results contrast with previous research in the Iberian Peninsula where the

higher number of vascular plant or fern species was found in mountainous areas (Castro-

Parga et al., 1996; Lobo et al., 2001; Moreno-Saiz and Lobo, 2008); this can probably be

attributed to several reasons. One could be the focus on different groups of plant species.

Most species in upper mountain areas are non woody phanerogams species, which are not

considered in this study. Pausas (1994) also showed a negative relationship of understorey

woody species richness with altitude in Pyrenean coniferous forests. To add to such

discussion, more than a third of the study area is located within the Atlantic and Alpine

bioclimatic regions, whereas in the studies cited above the Mediterranean region is

comparatively more represented. This implies that productivity limitation induced by hydric

stress at lower altitudes, one of the arguments given in previous studies for the increased

species richness at higher elevations (Lobo et al., 2001), might not be so applicable here.

Lobo et al. (2001) found that coastal cells with lower mean altitude presented an increase,

although less pronounced, in plant species richness; ocean influence might determine the

presence of species not found in other areas of the habitat range.

Water and energy variables have been shown to play a determinant role explaining

biogeographical patterns of plant species richness (O’Brien, 1993; O’Brien, 1998; Field et

al., 2005), agreeing with our results. However, weak correlations were found between

water-related variables and plant species richness in previous analysis in the Iberian

Peninsula (Lobo et al., 2001; Vetaas and Ferrer-Castán, 2008). Again, it should be noted

that rigorous comparisons between studies in the literature are difficult, given the wide

variety of variables used as surrogates of water availability (e.g. summer precipitation,

annual evapotranspiration, soil moisture). For example, Pausas (1994) found no significant

Page 61: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

47

relationship between plant species richness and Mean-PREC in Catalonia (cells of 10x10

km), whereas the Thornthwaite moisture index and aspect, also characterising water

availability, showed a positive relationship. Here we found that Mean-PREC was a

remarkable explanatory variable of almost all biodiversity indicators characterising

vegetation composition (except Shannon diversity index), while radiation had a negligible

role in most of the models. This fits with Whittaker et al. (2007), who concluded that plant

richness approximately south of 46º N (which covers all the study area here analysed)

should be more related to water variables than to energy. Higher tree species richness and

diversity (Margalef index) is expected in forest habitats within the Atlantic bioclimatic

region of the study area, where Mean-PREC is greater than in other Mediterranean regions.

This phenomenon has also been found for tree species in other regions (Currie and Paquin,

1987; Leathwick et al., 1998). Interestingly, the contrary occurred in the shrub layer.

Indeed, the number of shrub species had a strong negative correlation with water

availability, assessed by Mean-PREC in the forest habitat. This may be interpreted as a

consequence of tree biomass accumulation in humid forest habitats. The increase in Mean-

PREC determines greater productivity leading to canopy closure. As a consequence of the

decline in light penetration in the forest understorey, the lower availability of resources may

limit the number of shrub species that can tolerate these conditions.

4.3 Responses of forest biodiversity indicators to landscape configuration variables

Although environmental variables had a more relevant role in the models, some

configuration metrics had a remarkable association with the analysed biodiversity

indicators. The shape irregularity metrics resulted much more relevant than the

fragmentation ones for explaining the distribution of the indicators of forest biodiversity,

agreeing with Saura et al. (2008). Among these shape metrics, the DSCP clearly stood out

as a good correlate of almost all the indicators considered in the analysis. In some cases the

DSCP was even more relevant for explaining forest biodiversity than other environmental

variables. More complex shapes of forest patches at the landscape scale harboured greater

biodiversity in all cases. Similar results were observed in other studies for different groups

of plants, and in a moderate variety of scales, regions and analytical units (Moser et al.,

2002; Saura and Carballal, 2004; Baessler and Klotz, 2006; Saura et al., 2008; Torras et al.,

2009). It has been suggested that simpler shapes of forest patches indicate higher degree of

land use intensity and hence potentially less biodiverse habitats (Moser et al., 2002; Saura

and Carballal, 2004).

Page 62: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

48

The other shape metric considered in the analysis, MCCI, differentiates another

characteristic of patch shape: elongation (Saura and Carballal, 2004; Saura et al., 2008).

Forest habitats with more elongated patches tended to have more structurally complex

stands (as measured by the percentage of uneven-aged stands). This agrees with the results

of Saura and Carballal (2004) who found that in Galicia the MCCI was the index that better

discriminated between native and exotic forest patterns (typically even-aged plantation

forests). This shape elongation metric was significantly correlated with tree species richness

in a landscape analysis in Catalonia (NE Spain) by Torras et al. (2009), but this was not

verified here. Stand structure complexity was also positively associated with patch size in

the habitat; more homogeneous and simpler forest structures resulting from forest

management and other human-induced disturbances are actually characterised by a higher

fragmentation, accessibility and spatial mixture with other agricultural cover types.

The Mean-SIZE is not generally independent from habitat amount, as habitats that cover

larger areas tend to present patches of bigger sizes (e.g. Fernández-Juricic, 2000). If this

was not statistically controlled, the role of fragmentation in the models might have been

overestimated, given that the habitat loss is the primary driver of the loss of biodiversity and

should be evaluated independently from fragmentation (Fahrig, 2003). According to our

results, the relative importance of habitat fragmentation per se, despite a few exceptions,

was lower in most cases than shape complexity, precipitation or the standard deviation of

slopes.

Finally, Honnay et al. (1999) found that many small forest patches contain more plant

species than one large patch of the same total area. The authors concluded that this finding

could be the result of the probability of higher inter-patch diversity in more dispersed

habitat patterns, which would enhance biodiversity. Our analysis also suggests that the

potential negative effects of isolation (higher distances to the nearest patch) in terms of tree

species diversity are less important than the benefits of covering as much as possible

different physical environments. In the case of the snags, the processes that explain their

positive association with MNND are less evident and further studies are needed.

4.4. Limitations and final remarks

The present paper deepens the understanding of the patterns of forest biodiversity in a wide

range of Spanish habitats and provides useful considerations about the indicators included

in the 3SNFI. The stratification of the analysis by forest habitat types still highlighted the

importance of climatic and topographic factors as determinants of forest biodiversity

Page 63: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 1

49

distribution. Nonetheless, this approach conveys a different perspective, focused on

functional aspects and community composition rather than on space, which complements

and provides additional insights compared to other studies using a cell grid approach.

Finally, when widening the range of aspects of biodiversity considered (from the more

widely explored species richness indicators), identification of the landscape conditions

favouring forest biodiversity becomes increasingly complex. In our study, some habitat

conditions improved some particular aspects of biodiversity but not others (e.g. species

richness vs. diversity), or even produced different responses according to the indicator

considered (e.g. trees vs. shrubs, or compositional vs. structural). In the recently started

forth SNFI the number and quality of the assessed biodiversity indicators has already been

increased compared to the 3NFI; for example, indicators related to the herb layer and other

deadwood types (apart from standing dead stems) and decay classes are now considered.

This will allow improving and widening these analyses to a broader set of indicators within

a few years, as well as evaluating if more consistent trends on the determinants of

biodiversity in Spanish forest landscapes emerge as a result of the improved detail and

quality of the related data gathering in the filed plots.

We also showed that deriving conclusions on biodiversity determinants from a single best

model might produce misleading conclusions particularly when large and complex data sets

are explored. By basing our results in a set of plausible models, we were able to take into

account uncertainty in the selection of the best combination of explanatory variables and

provide a potentially less biased picture of the analysed relationships. We are however

aware that even the information-theoretic approach here applied assumes a priori selection

of suitable predictor variables to build a set of candidate models. Other variables not

included in this study might also be significant predictors of forest biodiversity patterns in

the Iberian Peninsula (e.g. forest management, continentality) and further studies

considering them may provide a more comprehensive assessment at a variety of spatial

scales and forestry contexts.

Page 64: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 65: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

Page 66: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 67: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

53

CHAPTER 2

Species richness of woody plants in the landscapes of

central Spain: the role of management disturbances,

environment and non-stationarity*

In chapter 1 we have shown that when widening the range of aspects of biodiversity

considered, identification of processes driving biodiversity becomes increasingly complex,

with some indicators being hardly predictable in a consistent manner. While we recognize

the need to consider the maximum possible number of biodiversity components for an

appropriate comprehension of ecosystem functioning, in this chapter we concentrated in a

more detailed analysis and understanding of the factors determining the species richness of

trees and of shrubs. These are two of the most relevant and widely used forest biodiversity

indicators, and they were shown in the previous chapter to present lower uncertainties in the

models attempting to describe their driving mechanisms. In addition, these indicators are

part of the basic information gathered in any forest inventory designed to orient a forest

management plan. They are also more homogeneously measured than other indicators

throughout the different Spanish provinces in the Third Spanish National Forest Inventory.

This is a particularly important issue to keep the indicators more trustfully comparable

when analysing large study areas (as is the case of this chapter). In this chapter we

particularly focused on the role of silvicultural practices as human-made disturbances that

could play a key role in the distribution of Mediterranean forest biodiversity even at wide

spatial scales, which remained untackled in chapter 1. In addition, we now used equally-

sized UTM cells as units of analysis, which are less ambiguously interpreted and can more

easily be compared with other studies.

* This chapter has been published as: Martín-Queller, E., Gil-Tena, A., Saura, S. 2011. Species richness

of woody plants in the landscapes of central Spain: the role of management disturbances, environment

and non-stationarity. Journal of Vegetation Science 22: 238-250. The affiliations of the other coauthor of

this article (A. Gil Tena) are Universitat de Lleida and Centre Tecnològic Forestal de Catalunya.

Page 68: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

54

ABSTRACT

This study addresses the following questions: How important is the effect of management

disturbances on gamma species richness of woody plants at intermediate landscape scales?

2. How is species richness related to other climatic and biotic factors in the study area? 3.

How does the assumption of spatial stationarity affect the assessment of relationships

among species richness and explanatory variables (e.g. management, biotic, and climatic

factors) across extensive study areas? In that purpose, a study area of 150,000 km2-extent

and 25 km2-grain size including the regions of Castilla y León, Madrid and Castilla-La

Mancha was analyzed. Information from 21,064 plots from the third Spanish national forest

inventory was used to evaluate richness of tree and shrub species at intermediate landscape

scales. In addition to variables that are well-known to explain biodiversity, including

environmental and biotic factors, the effect of management treatments was evaluated by

assessing clearcutting, selection cutting, stand improvement treatments, and

agrosilvopastoral systems (dehesas). Results from Geographically Weighted Regression

(GWR) techniques, which allow regression coefficients to vary spatially, were compared

with those from ordinary least squares regression (OLS). Patterns of gamma species

richness of woody plants, although strongly affected by both environmental and biotic

variables, were also significantly modified by management factors. Species richness

increased with the percentage of selection cutting stands and improvement treatments but

decreased with the percentage of clearcutting stands. A decline in species richness of woody

plants was associated with agrosilvopastoral practices. Species richness for trees was most

closely related to basal area, annual precipitation, and topographic complexity and species

richness for shrubs was most closely related to topographic complexity and

agrosilvopastoral systems. Most of the relationships between species richness and

environmental or biotic factors were non-stationary. Relationships between species richness

and management effects tended to be stationary, with a few exceptions. Landscape models

of biodiversity in Central Spain were more informative when they accounted for effects of

management practices at least at intermediate scales. In the context of the current rural

abandonment, our results showed that silvicultural disturbances of intermediate intensity

increased gamma species richness of woody plants. The exclusion of factors such as

agrosilvopastoral systems from our models could have led to spurious relationships with

other spatially covarying factors (e.g. summer precipitation). The patterns of spatial

variation in the relationships, provided by GWR models, allowed formulating hypotheses

about the potential ecological processes underlying them, beyond the generalization

resulting from global (OLS) models.

Page 69: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

55

1. INTRODUCTION

According to the ecological equilibrium paradigm, the internal regulation and stability of

ecosystems has traditionally been emphasised but the importance of disturbances for the

maintenance of vegetation diversity has increasingly been accepted (see Odion and Sarr,

2007). Some types of anthropogenic disturbances could be regarded as part of normal forest

functioning rather than as an external disturbance (Christensen et al., 1996; Decocq et al.,

2004), although this issue is controversial (Hall, 2000; Fabbio et al., 2003). In European

forests, the current flora might be adapted to regimes of secularly imposed disturbance

(Niemelä, 1999), such as forest management, particularly in the Mediterranean basin

(Blondel and Aronson, 1999). An extreme example of the close coevolution of species

composition and human activities are the dehesas. These agrosilvopastoral systems are

typical from the southwestern regions of the Iberian Peninsula (San Miguel-Ayanz, 1994).

Alpha (local) biodiversity tends to be higher in dehesas than in other natural habitats

(including Mediterranean forests from which they were created) for many functional groups

(Díaz et al., 2003; Pérez-Soba et al., 2007). This biodiversity depends on the maintenance of

traditional low-intensity farming.

Effects of forest management on biodiversity have been assessed in most studies at the

stand scale (e.g. Peltzer et al., 2000; Fabbio et al., 2003; Atauri et al., 2004; Torras and

Saura, 2008). These studies reveal that effects of forest management depend on the study

region and type of management practices. Because forestry practices are mainly designed

and applied at the stand or forest level, the most prominent impact of forest management

may be found at these local scales. At broader scales (i.e. wider extents and coarser grain

sizes), silvicultural factors (treatment type and intensity) might vary within the base unit of

analysis and large effects on indicators of biodiversity might not be easily observed (e.g.

Torras et al., 2009; Gil-Tena et al., 2010). In Spain, a focus on a broader scale for forest

management planning (from the classical stand and forest level to the landscape and

regional scales) has begun to be implemented only recently (Saura, 2010). Few empirical

studies have addressed the direct effect of forest management on biodiversity at a landscape

scale, particularly for vegetation indicators. More insights into the effects of silvicultural

disturbances and associated ecological processes on gamma (landscape) diversity are

necessary to provide guidelines for a sustainable forest landscape management.

In the assessment of the broad-scale effect of forest management on biodiversity,

underlying environmental drivers should be also considered because their relationship to

global patterns of biodiversity are widely recognised, as in the case of climate (O’Brien,

Page 70: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

56

1998; Field et al., 2005) (see the general introduction for more details on the drivers of

species richness at the landscape scale).

In most of studies, the relationships between indicators of biodiversity and explanatory

factors are assumed not to vary spatially, even in study areas of thousands of square

kilometres; i.e. the spatial stationarity of the underlying processes in landscape forest

biodiversity are assumed. However, spatial variation in the relationships can be expected

due to (1) random sampling variations, (2) intrinsic differences across space, (3) relevant

factors omitted from the model, or (4) an incorrect functional form of the relationships -

linear, unimodal, etc- (Fotheringham et al., 2002). If the relationships are non-stationary in

a given study area (i.e. they vary from one location to another), descriptive and predictive

utility of global models (considering the entire area) may be compromised.

The main objectives of this study are (1) to characterize effects of forest management

practices (agrosilvopastoral practices, forest regeneration and improvement treatments) on

gamma species richness of woody plants at an intermediate grain scale (5x5 km), (2) to

evaluate the relationships among species richness and other environmental and biotic

variables and to compare them with those obtained in previous literature, (3) to interpret the

observed spatial patterns of the relationships and possible underlying ecological processes.

The study was carried out in a region in central Spain which covers about 150,000 km2,

using the data from the Third Spanish National Forest Inventory and other thematic sources.

Local regression techniques were applied to evaluate the potential spatial non-stationarity of

the analysed relationships, particularly, geographically weighted regression techniques,

which are being increasingly used in ecological models (e.g. Osborne et al., 2007; Guo et

al., 2008).

2. METHODS

2.1. Study area and the Spanish national forest inventory

The present study was carried out in the Spanish regions of Castilla y León, Castilla-La

Mancha and Madrid (Figure 1). A total of 3996 UTM cells (5x5 km) were analysed. Cells

in Zone 29N (European Datum 1950) were excluded to avoid effects of differences in cell

size derived from the use of a common projection during data processing. We considered

only cells that included some forest area according to the Third Spanish National Forest

Inventory (3SNFI, Ministerio de Medio Ambiente, 1997-2007). The scale of 5x5 km

allowed (1) a sufficient number of SNFI plots to be captured to characterise the forests

Page 71: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

57

within each cell (see details below for sampling intensity) and (2) detection of potential

spatial variability by yielding a sufficiently large number of cells throughout the study area.

Figure 1. Map of the study area and corresponding administrative provinces. Environmental Zones

according to the European stratification by Metzger et al. (2005) are represented. ALS: Alpine South,

ATC: Atlantic Central, LUS: Lusitanian, MDM: Mediterranean Mountains, MDN: Mediterranean North,

MDS: Mediterranean South.

According to the stratification of Europe performed by Metzger et al. (2005), each of the

main environmental zones into which the European Mediterranean basin can be divided

were represented in our study area (Figure 1). South Alpine and Central Atlantic Zones also

appeared in some of the northern sectors of the study area, contrasting with the dominant

Mediterranean Zones. Mean elevation ranges from 300 to 1900 m, mean annual

precipitation from 300 to 1500 mm, and mean annual temperature from 5.5 to 16.5 ºC. The

most frequent tree and shrub species (taxa) in the study area are presented in Table 1, and

distribution patterns of the main dominant tree species in the study area can be consulted in

Appendix A.

Page 72: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

58

Environmental Zone Tree species Frequency (%) Shrub taxon Frequency (%)

Quercus pyrenaica 36 Cytisus spp. 32

Fagus sylvatica 31 Erica arborea 30

ALS (36 tree species, 44 shrub taxa)

Pinus sylvestris 31 Rosa spp. 29

Quercus ilex 44 Genista spp. 44

Pinus pinaster 41 Arctostaphylos uva-ursi 42

ATC (24 tree species, 49 shrub taxa)

Quercus faginea 34 Rosa spp. 37

Pinus sylvestris 30 Thymus spp. 42

Quercus pyrenaica 27 Rosa spp. 33

MDM (68 tree species, 108 shrub taxa)

Pinus nigra 24 Lavandula latifolia 16

Quercus ilex 43 Thymus spp. 60

Pinus pinaster 30 Rosmarinus officinalis 21

MDN (66 tree species, 116 shrub taxa)

Quercus faginea 14 Lavandula stoechas 18

Quercus ilex 75 Cistus ladanifer 30

Pinus pinea 13 Thymus spp. 21

MDS (57 tree species, 86 shrub taxa)

Pinus pinaster 11 Rosmarinus officinalis 19

Table 1. Frequencies of the three most common species (or taxa) of woody plants in each of the six

environmental zones in the sampling area according to Metzger et al. (2005).

The number of 3SNFI plots in the study area were 21,064. Plots were sampled from 2000 to

2004. See section 2.2 in chapter 1 for more details about the sampling design of the 3SNFI.

2.2. Biodiversity indicators and explanatory variables

For each 5x5 km UTM cell, two indicators of biodiversity were computed: gamma tree and

shrub species richness. Gamma species richness was estimated as the sum of the number of

species of tally trees (DBH ≥ 7.5 cm) or shrubs surveyed in the 3SNFI plots belonging to

one cell. To control variation in species richness of woody plants that was explained by

differences in the sampling effort (number of observations) between cells, simple linear

regressions of each indicator was computed against the number of 3SNFI plots per cell

(indicating the real sampled area). We used the residuals resulting from those regressions as

the final dependent variables in subsequent analyses.

Eight explanatory variables were included in the analysis (Table 2). Four variables were

related to forest management. We distinguished two different methods of silvicultural

regeneration treatments: the percentage of plots per cell treated with clearcutting

(CLEARCUT) or selection cutting (SELECT). The variable CLEARCUT included plots

with shelterwood treatments, although the shelterwood practices represented less than 15%

of the total number of plots managed through these two silvicultural systems that are typical

of even-aged stands. Stand improvement treatments were also considered (IMPR). These

Page 73: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

59

three variables were estimated from the specific information available in the 3SNFI plots.

Silvicultural practices were applied to 30% of the plots in the study area. Regeneration

treatments were applied to 15% of plots (selective cutting: 9%, clearcutting: 4% and

shelterwood: 1%). Improvement treatments were applied to 20% of plots. The forth

management variable, agrosilvopastoral practices (SILVOPAST), covered 6% of the total

forest area and was concentrated in the southwestern region (Figure 2). The area occupied

by these agrosilvopastoral systems was estimated using the Spanish Forest Map (SFM). The

SFM, developed within the 3SNFI, has a vector data structure, a scale of 1:50 000 and a

minimum mapping unit of 2.25 ha. It has been developed from the interpretation of aerial

photographs (dated 1997-1998 in the study area) combined with pre-existing maps and field

surveys.

CLEARCUT Percentage of the 3SNFI plots managed with clearcutting or shelterwood treatments

SELECT Percentage of the 3SNFI plots managed with selection-cutting treatments

IMPR Percentage of the 3SNFI plots presenting stand improvement treatments (cleaning,

precommercial thinning, thinning or pruning)

SILVOPAST Land area occupied by agrosilvopastoral practices (m2)

Ann-PREC Average total annual precipitation (mm)

Jul-TEM Average temperature in July (º C)

sd-SLOPE Standard deviation of slopes (º)

BA Average of the basal area in the 3SNFI plots (m2·ha

-1)

Table 2. Explanatory variables computed in each sampling unit (a 5x5 km cell) and considered in the

final analysis.

Environmental and biotic variables were selected according to the major relationships

identified in previous research (see the general introduction). Accordingly, we included

variables representing climate (water and energy), spatial heterogeneity, and biotic

interactions. However, the final set of variables was constrained by multicollinearity. The

selected variables should not have Spearman correlation coefficients (r) higher than 0.5

between them. Thus, final climatic variables were mean annual precipitation (Ann-PREC)

and mean July temperature (Jul-TEM). Climatic data were obtained from the Climatic Atlas

of the Iberian Peninsula at a resolution of 200 m (Ninyerola et al., 2005). Topographic

complexity was represented by the standard deviation of slopes in each cell (sd-SLOPE),

quantified from the official Spanish Digital Elevation Model at a resolution of 25 m

(Ministerio de Fomento, 1999). Mean basal area (BA) was estimated from the tally trees of

the 3SNFI plots to account for the effects of stand stocking and maturation on species

Page 74: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

60

richness of woody plants not covered by processes related to productivity (climate) and

management. Additionally, BA could be reflecting biotic interactions between the tree

stratum and other vegetation layers, such as shrubs. Spatial patterns of the explanatory

variables are represented in Figure 2. Each variable was standardised to zero mean and unit

standard deviation to eliminate the effect of differences in the measurement scale and to

allow a comparison of the regression coefficients between models (particularly in the case

of GWR).

2.3. Statistical analyses

The relationships between the explanatory variables and each biodiversity indicator were

evaluated using both global Ordinary Least-Squares regression (OLS) and local

Geographically Weighted Regression (GWR) models; GWR models were computed using

GWR 3.0 software (Charlton et al., 2003). OLS models but not GWR models were based on

the assumption of spatial stationarity of the relationships.

A stepwise OLS model selection was performed based on the Akaike Information Criterion

(AIC) to select the final OLS models. Each of the eight explanatory variables (Table 2) was

included in the GWR model, in which variable selection was not possible and in the initial

OLS model for the stepwise selection. The most complex potential OLS model included

each quadratic term and the interaction between energy (Jul-TEM) and water (Ann-PREC).

The simplest model included only the intercept. We used the step function in R statistical

software (www.r-project.org).

GWR provides local estimates of the parameters at each regression point –here the centroids

of the 3996 UTM cells- using a weighted calibration in which the influence of the other

observations decreases inversely to their distance to the location under consideration

through a spatial kernel function (Fotheringham et al., 2002). The bandwidth determines the

rate at which the weighting of an observation declines and reflects an approximation of the

extent of the zone considered around each local regression model.

Page 75: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

61

Figure 2. Geographical patterns of the explanatory variables in the study area. Mean elevation (mean-

ELEV) of the 5x5 km UTM cells is also represented in order to show the distribution of the main

mountainous chains across the study area, although this was not included as an explanatory variable in the

final models (see methods).

We applied a fixed-kernel Gaussian function. Bandwidth selection was based on the

analysis of the spatial autocorrelation of the residuals of the global (OLS) model.

Autocorrelation of residuals can invalidate inferences from a model (Legendre and

Legendre, 1998); allowing geographically varying relationships can reduce this problem

Page 76: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

62

(Fotheringham et al., 2002). We built correlograms with a lag distance of 5 km for each

biodiversity indicator using Rookcase Excel add-in (Sawada, 1999). We calculated the

distance at which the value of Moran’s I crossed the expected value for the absence of

spatial autocorrelation, indicating the spatial range of the pattern (Fortin and Dale, 2005). In

order to use the same extent and allow for the comparability of shrub and tree GWR

models, the maximum spatial range of both correlograms was used as a common

bandwidth. Remaining spatial dependence in OLS and GWR residuals was evaluated with

global Moran’s I coefficients using the corresponding bandwidth as the lag distance. The

significance of the autocorrelation coefficient was assessed through a Monte Carlo

randomisation test with 999 runs with the software Rookcase. In order to assess the

improvement in GWR models with respect to OLS models, differences in AIC values were

estimated.

Simultaneously testing statistical significance of all the local parameters in the study area

implies an increased Type I error (Fotheringham et al., 2002). The choice of applying

Bonferroni corrections is a conservative approach when considerably large sample sizes are

analysed, as is the case in this study. Alternative approaches have been suggested (e.g.

Benjamini and Yekutieli, 2001). Here we adjusted the family-wise error rate (i.e. 0.05)

dividing it by the effective number of parameters (ENP), which in initial experiments

suggests yielding similar results as in the cited approach (Dr. Martin Charlton, personal

communication). A Monte Carlo significance testing procedure was applied to examine if

variability in local coefficients was due to random variation or reflected a true geographical

trend (Fotheringham et al., 2002).

3. RESULTS

OLS models had higher AIC values than the GWR models (Table 3), indicating a better

performance for the latter. Correlograms from OLS model residuals revealed spatial ranges

of 60 and 135 km for tree and shrub species richness, respectively, and consequently the

selected bandwidth for both GWR models was set to 135 km. For the tree species, Moran’s

I values at distances above 60 km oscillated around the expected value (i.e. there was an

absence of significant spatial autocorrelation in the error term). GWR analyses accounted

for most of the spatial variation of the indicators. Residual autocorrelation, as measured by

global Moran’s I (Table 3), was much smaller for a GWR model than for the comparable

OLS model.

Page 77: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

63

OLS GWR

AIC Moran’s I p-value AIC Moran’s I p-value

Tree species richness 10,758 0.0034 0.001 10,664 0.0007 0.059

Shrub species richness 10,116 0.0493 0.001 9,859 0.0192 0.001

Table 3. Results of the analysis of spatial autocorrelation in residuals from ordinary least squares (OLS)

and geographically weighted regression (GWR) models. The bandwidth (135 km) was used as the lag

distance in the estimation of the Moran’s I coefficients. AIC values for both types of models are also

shown.

3.1. Relationships between explanatory variables and species richness of woody plants

in global (OLS) models

Each of the two indicators of biodiversity was highly correlated with environmental, biotic

and/or agrosilvopastoral factors. Comparatively lower coefficients were found for

silvicultural treatments (Figure 3). The main explanatory variables for tree species richness

were basal area and mean annual precipitation and for shrub species richness, standard

deviation of slopes and agrosilvopastoral practices. Each management variable was

significantly related to each of the two biodiversity indicators (Figure 3). Most relationships

were curvilinear except for mean July temperature, standard deviation of slopes and

agrosilvopastoral practices in both models (Figure 3). Linear regression coefficients for

agrosilvopastoral practices were negative for each biodiversity indicator. The coefficient

was especially large for shrub species richness (β=-0.18 versus -0.06 for trees).

Improvement and selection treatments had a weak, positive coefficient for models of both

trees and shrubs when less than approximately 40-60% of the plots in the cells received the

treatment. The association became negative beyond these thresholds (Figure 3). Only about

2 and 10% of the cells in the study area were beyond these thresholds. Clearcutting

association with both indicators was significantly negative in cells with percentages below

60-70, but became positive above these thresholds, only occurring in 1% of the cells.

The increase in annual precipitation was in general associated with an increase in tree

species richness but a decline of shrub species richness (Figures 3 and 4). For shrub species

richness, the effect of the interaction between the two climatic variables was significant

(Figure 4). Annual precipitation levels above 820 mm, which occurred mainly in the Alpine

South and Atlantic Central Zones, were negatively associated with tree species richness

(Figure 3). The highest and lowest levels of mean annual precipitation and July

Page 78: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

64

temperature, respectively, occurred in these zones, which coincided with the lowest number

of shrub species estimated by our model in those climatic conditions (Figure 4). Values of

standard deviation of slopes increased linearly with richness of tree and shrub species,

particularly for the latter. Basal area had a strong positive regression coefficient with the

number of tree species for the most common values of basal area (0-30 m2ha-1), and

negative for higher values. The quadratic term of basal area was significant in both models

but within the positive range of basal area values, its relationship with shrub species was

always negative (Figure 3).

Figure 3. Relationships, according to OLS results, among species richness for trees and shrubs and each

of the explanatory variables while maintaining constant the remaining variables (mean value). The X-axis

represents the entire range of variation of the explanatory variable in the study area. The vertical dashed

lines delimit the 90% X central distribution. The Y-axis range is constant in each plot and represents the

60% Y (trees or shrubs) central distribution. Partial regression coefficients in the formula are estimated

with both dependent and explanatory variables standardized. In order to facilitate plot interpretation, the

original X values are shown. n.s: non significant.

Page 79: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

65

3.2. Relationships between explanatory variables and species richness of woody plants

in local (GWR) models

The testwise error rate (i.e. 0.05/ENP) used to assess the significance of the GWR

regression coefficients was 0.0024. Almost all environmental and biotic factors had a

spatially non-stationary relationship with the analysed biodiversity indicators (Figure 5).

Monte Carlo tests did not show significant non-stationarity for most management variables.

For the stationary factors (mainly management factors) the relationships found in the OLS

models are assumed to be applicable to the whole study area and they were not mapped in

Figure 5. As shown in Figure 5, only the intensity of the non-stationary relationships varied

geographically (including an absence of a significant relationship in some localized areas),

whereas spatial changes on sign were not observed. The spatial patterns of the intensity of

the relationships between environmental or biotic factors and tree species richness were

opposed to those for shrub species.

Figure 4. Response surface of shrub species richness to the interaction Jul-TEM x Ann-PREC (β=0.05,

p<0.001). The rest of explanatory variables were kept constant (mean value). The displayed ranges of

both explanatory variables correspond to those found in the study area.

Page 80: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

66

Figure 5. Spatial variation of the local regression coefficients of GWR models for species richness of

trees (top) and shrubs (bottom). Only significant coefficients after multiple testing corrections (see

Methods) are drawn. Note that other explanatory variables may also be significant according to OLS

results, but are not shown here when their relationships with the biodiversity indicators did not

significantly vary spatially; significance level in Monte Carlo tests for the variables shown was always

less than 0.001. The approximate extent of the area encompassed by each local model according to the

size of the bandwidth (135 km) is shown by a circle.

4. DISCUSSION

4.1. Forest management factors

Patterns of biodiversity at biogeographical scales may not be affected by disturbances such

as those due to forest management, which might be a consequence of an observed

equilibrium linked to the enlargement of the scale (de Angelis and Waterhouse, 1987).

However, although at the meso-scale here analysed (i.e. at a grain size of 25 km2), patterns

of species richness of woody plants were strongly determined by environmental and biotic

variables, these patterns also were modified by ecological processes induced by silvicultural

treatments. This concurred with a recent study in Catalonia (NE Spain) by Torras et al.

(2009), who found that management intensity had a small, positive effect on tree species

richness and diversity at a grain size of 100 km2, but not on shrub species richness.

Page 81: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

67

Notably, we found that impacts on woody plant species varied among silvicultural

treatments. Intense disturbances resulting from clearcuts were associated with lower species

richness. In contrast, less intensive treatments such as selection cuttings and improvement

treatments were associated with higher species richness, at least below a proportion of

managed forest area. This proportion is not frequently exceeded in central Spain. Indeed, as

stated above, most plots in the study area are currently unmanaged (70%). Agricultural land

abandonment together with forest non-intervention is occurring in the last decades in some

areas of the Mediterranean Basin (e.g. Gil-Tena et al., 2010). The decline in forest

management derives from the replacement of traditional forest products –with a current low

profitability- with new fuel sources and building materials. In this context, some types of

management practices could be a source of structural heterogeneity within forest patches

that ultimately increases gamma biodiversity. In the framework of the non-equilibrium

paradigm, the increase in biodiversity with moderate disturbance fits with the intermediate

disturbance hypothesis. According to this hypothesis, the number of species is maximized if

the intensity of ecological disturbances is moderate. At moderate levels, competitive

exclusion is controlled, heterogeneity is favoured, and a balance with the consequent

environmental stress is achieved (Connell, 1978). Both selection cutting and improvement

treatments, as currently applied in the study area at the grain of 25 km2, seemed to match

with the intermediate disturbance intensity, frequency and gap sizes found to be the most

favourable for species diversity. Our results were consistent with previous findings at the

stand scale (e.g. Roberts and Gilliam, 1995; Atauri et al., 2004; Torras and Saura, 2008).

However, authors have suggested that at a coarser grain size, the heterogeneity that results

from the spatial age mosaic of even-aged clearcut stands (plots) might favour gamma

diversity rather than a continuous uneven-aged forest with a uniform application of

selection cuttings throughout the landscape (Decocq et al., 2004). The latter treatment may

increase woody plant diversity locally (alpha diversity, at the plot level) while

impoverishing the variety of species at wider spatial scales. The loss of gamma species

richness of woody plants that may result from homogenisation by selection cuttings was

found in our study at intermediate scales only when more than 60-70% of the forest

landscape was managed, which is rarely the case in central Spain. For the same reasons, the

positive response of species richness to the clearcutting of more than 60-70% of the total

forested landscape should be interpreted cautiously because such a response corresponded

only to a few outlier cells that are not representative of the typical current practices in the

study area (Figure 3).

Page 82: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

68

Although silvicultural impacts on species richness of woody plants were in general

stationary across the study area, the relationship between selection cuttings and shrub

species richness varied spatially. The dynamic equilibrium model (Huston, 1979) states that

diversity depends not only on the intensity, periodicity and extent of the disturbance, but

also on the growth rate. According to the model, the characteristics of the disturbances that

maintain maximum diversity will vary depending on site productivity because different

processes (competitive exclusion or environmental stress) will prevail (Odion and Sarr,

2007). Competitive control between shrub species might dominate those areas in which

selection cutting was found to favour species richness of shrubs (Figure 5). In these areas,

lower water availability reduces tree density or even favours shrubby land formations under

poorer soil conditions.

Species richness of both trees and shrubs declined with an increase in the extent of

traditional grazing systems (dehesas). The decline was particularly prominent in the western

part of the study region (Figure 5). The decline could have been anticipated: management

practices that are typical of traditional grazing systems (tree clearing and pruning, and

periodical ploughing) encourage scattered formations of Quercus ilex or Q. suber with less

than 5-10% of shrub cover (Plieninger, 2006; Ramírez and Díaz, 2008). This type of

management, combined with the grazing pressure, inhibits the regeneration and

establishment of other woody species and facilitates livestock grazing and crop cultivation.

Dehesas are highly diverse because of the herbaceous understorey and the variety of

animals that benefit from the coexistence of microhabitats typical of grassland and forest

formations (Díaz et al., 2003; Ramírez and Díaz, 2008). For instance, in the study by Peco

et al. (2006) more than 90% of the understorey plant species in these systems were annuals.

Therefore, despite their low species richness of woody plants, dehesas have a remarkable

intrinsic value and the need to conserve them is acknowledged (see Plieninger, 2006; Pérez-

Soba et al., 2007).

4.2. Environmental and biotic factors

In the Mediterranean Zones, higher annual rainfall was associated with higher tree and

shrub species richness in those regions with most extreme temperatures in summer (Figs. 3

and 4) (such as the Mediterranean South Zone; Figure 1) but with lower species richness of

shrubs under milder conditions (Figure 4). The positive association of woody species

richness with precipitation might be due to a decline in productivity associated with hydric

stress, as pointed out by Lobo et al. (2001). Hydric stress would be caused by scarce water

Page 83: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

69

and high energy inputs in the Mediterranean. Our results suggested that this stress has a

greater limiting effect on tree species than on shrub species under moderate summer

temperatures. These results could be also explained by a higher tree basal area expected in

most productive areas according to climate patterns.

Because temperature decreases with increasing altitude, effects of altitude could be

mediated by climate at the local scale (i.e. as captured by GWR models). Especially in

mountain regions in the edge of the study area, where correlation coefficients above |0.9|

were estimated in some local models in the Alpine Zone. This was not necessarily the same

for the whole study area (OLS models), where r for annual precipitation and July

temperature were, respectively, 0.55 and -0.65. This means that, on the one hand, the

different extent considered by GWR and OLS techniques could determine a focus on

different processes underlying the observed relationships. On the other hand, local

multicollinearity in the explanatory variables on GWR results (Wheeler and Tiefelsdorf,

2005) might have influenced the observed patterns in the relationships (but see Smith et al.,

2009). According to this, global pattern of annual precipitation might reflect a poorer

richness in tree species in the Alpine Zone when compared to other Mediterranean areas

(OLS results; Figure 3). Within the more reduced pool of species in the smaller extent of the

Alpine Zone (GWR results; Figure 5), a positive altitudinal gradient of tree species richness

might exist, linked to the decrease in July temperature with elevation. However, this did not

agree with Nogués-Bravo et al. (2008) who found that, at the same grain and extent (0-2000

m), species richness for vascular plants, lichens, and bryophytes decreased with increasing

altitude in the Pyrenees. Therefore, it seems that along the Iberian Peninsula, the patterns of

gamma species richness of woody plants with altitude at a landscape scale could vary

because of different underlying processes related to resource availability and productivity,

ultimately determined by larger scale climate patterns.

Species richness of woody plants increased with topographic complexity in our study. It

was a particularly important factor in the shrub species model. This factor would favour

species richness through an increase in surface area and a greater variety of microclimate

conditions and bedrocks (see Vetaas and Ferrer-Castán (2008) and references therein),

especially in the most xeric areas for shrubs, as shown by the GWR models. Additionally,

these results could reflect shrub encroachment and colonisation processes in those less

productive and less accessible areas that are more likely to be abandoned.

Landscape patterns of forest basal area that were not explained by silviculture or climate

patterns were a key factor positively associated with tree species richness. More mature

Page 84: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

70

forests with greater tree biomass were associated with an increase in the number of tree

species at the landscape scale. This could be explained by the association between site

productivity and higher basal areas and the significant relationship between productivity

and tree species richness. Causality could exist in the opposite sense: higher tree species

richness might allow more productivity and thus higher basal area (Vilà et al., 2007).

Additionally, forests with currently higher basal area might have suffered from lower

anthropogenic degradation in the past. However, cells with a mean basal area greater than

30 m2ha-1 had fewer tree species. Although this occurred in less than 2% of the studied cells,

it might have been due to a homogeneous forest structure and age within the 25 km2 area of

the cell, which probably leaded to the impoverishment of gamma diversity at this grain size.

The increase in landscape basal area was however associated with a poorer shrub stratum

independently of the range of BA. This was particularly true in the Atlantic Zone according

to GWR results, probably because the density of the tree canopy was greater, penetration of

light was reduced, and the growth of understory plants was limited.

5. CONCLUSIONS

Management factors are not frequently evaluated in biodiversity models at the landscape

scale. However, the impact of management on gamma species richness might be significant

at an intermediate grain size, as shown in our results. Furthermore, higher levels of species

richness of woody plants were found in landscapes with moderate intensity management

disturbances in Central Spain. The exclusion of factors such as agrosilvopastoral systems

from the models in a study of this kind, given their characteristic geographical distribution,

could have led to spurious relationships with other spatially covarying factors (e.g. summer

precipitation) and to potentially misleading conclusions about the biodiversity drivers in the

study area.

Many relationships showed non-stationarity across the study area. Thus, global trends that

were observed in vast study areas should be interpreted carefully (Fortin and Dale, 2005).

These trends could result from a counteraction of different strengths of the relationships

across the study area. Spatial variation can be discerned via GWR and hypotheses then can

be formulated to explain spatial variation, complementing OLS results. However, the

observed patterns of the relationships in OLS and GWR might not be comparable because

differences in the extent could also determine a focus on different ecological processes in

each analysis.

Page 85: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 2

71

We acknowledge that species richness of woody plants should not be viewed as a surrogate

of biodiversity. To have a more holistic view of the interactions between environment,

management and forest ecosystem functioning, the impact of forestry practices on other

functional groups and components of biodiversity (such as ecosystem and genetic diversity)

should be considered in further studies. Many studies have reported the negative effects of

some types of forest management on other taxa (e.g. birds, invertebrates, fungi) due to a

decline in the amount of deadwood and old trees among other factors (e.g. Loyn and

Kennedy, 2009; Vandekerkhove et al., 2009). However, other studies have suggested that

moderate-intensity forest practices favour forest biodiversity or have a scarce effect

comparing to other large-scale processes such as widespread maturation after large-scale

rural abandonment (see Gil-Tena et al. (2010) and references therein). Population isolation

and artificial regeneration should also be avoided in order to prevent the potential negative

effects of harvesting treatments on genetic diversity (Finkeldey and Ziehe, 2004). While we

here took into account different extents and related ecological processes (as captured by

GWR and OLS), the grain size was kept constant in all our analyses. We recognize the

importance of extending the study to other grain sizes different from the 25 km2 here

considered, since this scale component largely determines the relative influence of large

climate trends versus more local management factors on biodiversity patterns. This is part

of our ongoing research. Finally, the variation of species diversity with scale and

disturbance regime here reported may differ from the one existing in other regions with

different forestry practices and productivity contexts.

Page 86: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 87: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

Page 88: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 89: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

75

CHAPTER 3

Effects of silviculture on tree species richness:

interactions between management, landscape context

and regional climate

Spatial patterns of variation in the relationships observed in chapter 2 were hypothesized to

be explained by coarse patterns of variation in climate. In this chapter we specifically

evaluated the role of this interaction across scales in the processes influencing local richness

of tree species (including management). Although silvicultural impacts on landscape

species richness were in general constant across the study area (chapter 2), we hypothesize

that spatial non-stationarity will be significant when working at the stand level, the scale at

which we conducted the analysis in this third chapter. In addition, we further refined the

analysis by considering the potential effect of different forest and management types in the

landscape context on woody plant diversity in the analysed stands.

* This chapter has not yet been published. It will be submitted in the near future as an article for a

scientific journal with the following authors: Martín-Queller, E.; Diez J.M.; Ibáñez, I; Saura, S. The

affiliation of the other coauthors of this article (J.M. Diez and I. Ibáñez) is University of Michigan, USA

Page 90: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

76

ABSTRACT

Patterns of forest biodiversity are shaped by a complex set of processes operating over

different spatial scales. While climate may largely determine diversity at regional

scales, numerous biotic interactions and disturbances affect diversity more locally.

However, the likely interactions between these processes across different scales

complicates efforts to understand their relative importance. For example, the response

of biodiversity to a local disturbance may depend on regional productivity and on

species pools available at intermediate scales. In this study we quantified the interacting

effects of silvicultural disturbances and regional climate on tree species richness in an

area of 152,000 km2 in central Spain. We used data from the Spanish forest inventory,

together with hierarchical Bayesian models, considering different management types

and intensities both in the focal stands and in the surrounding landscape. Our study

supports the hierarchical structure of processes influencing species richness patterns; coarse

climatic patterns determined different responses of species richness to disturbances and

abiotic conditions. We showed that, in general, partial harvesting may allow the coexistence

of a higher number of tree species both in coniferous or broadleaved, Mediterranean forests.

However, this was not always true in our study. We hypothesize that in semi-arid regions

with water stress conditions, canopy facilitation benefits surpass those of resource release in

canopy gaps. In addition, forestry operations other than canopy removal can limit potential

positive effects of canopy gaps in selection cutting. Finally, intermediate degrees of canopy

removal might not have a positive impact on tree species richness without a minimum

amount of unmanaged forest in the landscape. Our results suggest that a sufficient pool of

species in the landscape is required to guarantee the seed recolonization of canopy gaps. In

general, metacommunity processes seemed to be key drivers of local species richness, since

increased species richness was found in forest stands surrounded by species-rich, riparian

forests, while prevalence of tree species poor dehesas and plantations in the landscape was

associated to decreased richness in surrounding forest stands. A common ecological history

probably accounts also in part for these patterns.

Page 91: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

77

1. INTRODUCTION

Determining which are the drivers of species diversity is one of the fundamental goals in

ecology. It is accepted that these drivers and/or their relative importance differ with spatial

scale (Auerbac and Shmida, 1987; Whittaker et al., 2001; Field et al., 2009). Particularly,

climate has been found to be the main driver of large, regional scale patterns of plant

diversity, while processes affecting biotic interactions, such as disturbances, have been

suggested to explain patterns at finer scales (Sarr et al., 2005; Kallimanis et al., 2007).

Not only natural disturbances but also human-caused disturbances can modify species

diversity in local communities (e.g. Roberts and Gilliam, 1995; Niemelä, 1999). For

example, the effect of silvicultural regeneration treatments on local forest biodiversity has

been the focus of substantial research in the last decades, especially in North America and

Scandinavia where intensive forest management is generally applied to large areas and has a

long history of use (Rowland et al., 2005; Kuuluvainen, 2009). Few studies exist, however,

on their effects in the Mediterranean region (but see for instance Torras and Saura, 2008;

González-Alday et al., 2009; Navarro et al., 2010). Additionally, many of the studies on the

effect of regeneration treatments on plant diversity are experimental, and therefore, short-

term (e.g. Hannerz and Hanell, 1997; Kraft et al., 2004; Götmark et al., 2005; Zenner et al.,

2006; Newmaster et al., 2007; Burke et al., 2008); thus, they mainly evaluate changes

occurring in the first successional stages after disturbances, and the majority focus on the

herbaceous community. But harvest practices have the potential to shape tree species

diversity compared to unmanaged stands (e.g. Halpern and Spies, 1995; Onaindia et al.,

2004; Torras and Saura, 2008). By influencing understorey dynamics during the first stages

after disturbance, the degree of canopy removal determines the competitive ability of tree

seedlings, the most sensitive life history stage on trees, and ultimately drives the structure

and successional path of the future stand (Wagner et al., 2011).

Scale and hierarchical structure of ecosystems (in space, time or ecological organization)

are fundamental to understand ecological processes influencing species diversity (Levin,

1992; McMahon and Diez, 2007). Indeed, species richness patterns might derive from the

interaction of factors influencing it at varying scales (e.g. Sarr et al., 2005). However, most

studies analyzing patterns of species richness or diversity do not account for spatial scale in

the analysis, focusing on a single scale, and generally use multivariate regression techniques

assuming that explanatory variables are simply additive.

One example of this interaction across scales might be the influence of large-scale

environmental constraints on the response of ecological communities to local disturbances.

Page 92: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

78

Actually, from a disturbance-theory perspective, the impact on a community of a particular

type -frequency, intensity, extent and duration- of disturbance is expected to vary depending

on the productivity of the site (Kondoh, 2001). The release of resources derived from the

disturbance will affect species interactions depending on the overall levels of resources in

the system. Similarly, the stress resulting from disturbances will be greater where growth is

already constrained by abiotic stresses from resource limitations and climatic extremes. But

also the attributes of silvicultural disturbances themselves depend on the productivity

context. Foresters adapt to the prevailing climatic (productivity) conditions by tailoring

local management rules (Fabbio et al., 2003). Furthermore, species perception of stress (e.g.

shade tolerance, drought and late-frost sensitivity) is usually associated to evolutionary

forces related with the historical regime of disturbances and coarse climatic conditions (e.g.

Lloret et al., 2007). Therefore, not only large scale patterns of species diversity are affected

by climate, as a surrogate of productivity (Currie and Paquin, 1987; Field et al., 2009), but

also the consequences of disturbances on local plant diversity might depend on the

prevailing large scale climatic conditions. Historical ecology (Legendre and Legendre,

1998) and metapopulation processes (Gardner and Engelhardt, 2008) occurring at

intermediate spatial scales can also modify the impact of forestry disturbances on local

species diversity. The long-term impact of a disturbance event will be negative when

recolonization is hampered by the intensity of the disturbance (e.g. when the seed bank is

depleted), or by the spatial isolation of the populations (Gardner and Engelhardt, 2008).

In Spain, thousands of years of human management have modelled current forests (Blondel

and Aronson, 1999) and still do nowadays. In particular, agrosilvopastoral practices in

South-western Spain have given rise to a traditional system, dehesa, characterized by

extensive semi-forested areas with evergreen oaks (Quercus ilex and Quercus suber)

scattered over grasslands or cereal crops (e.g. Díaz et al., 2003). Other forest type common

in Spanish landscapes are pine plantations; the policy of afforestation prevailing since the

beginning of the 20th century has promoted millions of hectares of planted forests in Spain

(Valbuena-Carabaña et al., 2010). As for silvicultural practices, their application in Spanish

forests, and in the European Mediterranean in general (Fabbio et al., 2003), has decreased in

the last decades, as a consequence of current low economical profitability of traditional

forest products. This phenomenon is resulting in the densification and biomass increase of

these forests. All these human interventions (or the lack of them), together with cropland

abandonment (e.g. Moreira and Russo, 2007) and their interaction with a fire-driven

disturbance regime (Pausas et al., 2004; Loepfe et al., 2010) are fundamental to understand

current forest dynamics in Spain.

Page 93: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

79

In this study, we use a hierarchical Bayesian model to account of the different factors, and

the scales they operate under, affecting the species composition of forests in a

Mediterranean region of central Spain (152,000 km2). In particular, we evaluated the impact

of different silvicultural practices on forest composition while accounting for the effects of

climate at the regional level, and the effects of the neighborhood landscape. Our main

questions were (1) how different types of silviculture regeneration treatments affect tree

species richness of forest stands, once accounting for processes operating at larger scales?

(2) How do the effects of silvicultural disturbances on biodiversity depend on forestry

practices in the surrounding landscape? (3) is the number of tree species in a forest stand

influenced by the proportion of other woodland types, such as dehesas or plantations, in the

surrounding landscape? and (4) a more general goal of this study was to provide hypotheses

about the ecological and historical mechanisms that might explain the estimated

relationships between tree species richness and other environmental and stand structure

factors.

2. METHODS

2.1. Study area

Figure 1. Localization of the study area. Environmental zones according to the European stratification by

Metzger et al. (2005) are presented. MDM, Mediterranean Mountain; MDN, Mediterranean North; MDS,

Mediterranean South.

Page 94: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

80

The study area covers an extension of 152,000 km2 in central Spain, including the regions

of Castilla-La Mancha, Madrid, and most of Castilla León (Figure 1). According to the

European environmental stratification by Metzger et al. (2005), each of the main

environmental zones into which the European Mediterranean basin can be divided was

represented in our study area (Figure 1). Metzger et al. (2005) performed a statistical

stratification of the European environment using climatic, geomorphological, oceanicity

and, northing-related variables, at 1 km2 resolution. Using Principal Component Analysis

(PCA), they came out with 84 strata that were aggregated into 13 Environmental Zones

(EnZ). This aggregation was based on the mean first component of the classification

variables, which fundamentally expressed the temperature gradient across Europe. Three

Mediterranean regions resulted: Mediterranean Mountain (MDM), Mediterranean North

(MDN) and Mediterranean South (MDS); Mediterranean mountains included all strata with

altitude above 1000 m. Monthly variation of temperature and precipitation for these regions

in our study area is presented in Figure 2. Seasonal patterns of precipitation and temperature

are characteristic of Mediterranean climate and very similar in all three regions. There is a

gradient MDM-MDN-MDS of increasing temperatures and decreasing precipitation.

Mediterranean mountains are characterized by a higher spatial variability in monthly

temperatures and, especially, precipitation levels. The southern region is characterized by a

particularly marked water deficit in summer. Table 1 shows information about the

characteristic vegetation composition of each EnZ (see more detailed spatial patterns of

some environmental variables across the study area in chapter 2 and of the forest species

and management types in appendices A and B).

Figure 2. Climate diagrams for each environmental zone (MDM, Mediterranean Mountain; MDN,

Mediterranean North; MDS, Mediterranean South) in the study area. Data source: Ninyerola et al. (2005)

Page 95: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

81

2.2. Data

The Third Spanish National Forest Inventory (3SNFI, Ministerio de Medio Ambiente,

1997-2007) was used as a data source. The 3SNFI has a systematic sampling design; plots

were located at the intersections of a 1x1-km UTM grid that fall inside forests and other

woodlands. A total of 24498 inventory plots were surveyed in the study area from the year

2000 to 2004. From these plots, patterns of tree species richness were only assessed for

those where tree overstorey had a minimum Forest Canopy Cover (FCC) of 5%, excluding

plantations, dehesas, riparian and burnt forests. 14,306 inventory plots met these criteria,

called hereafter simply ‘forests’. The rest of the plots in the study area were used to

calculate land use composition in the landscape surrounding these forest plots (see below).

EnZ FUNC Clearcutting Selection cutting No management

n plots Composition n plots Composition n plots Composition

MDM BD 21 90% Quercus pyrenaica 49 71% Quercus pyrenaica 1321 78% Quercus pyrenaica

18% Quercus faginea 15% Quercus faginea

SE 0 - 26 100% Quercus ilex 793 96% Quercus ilex

CO 179 90% Pinus sylvestris 523 44% Pinus pinaster 2649 30% Pinus nigra

28% Pinus nigra 30% Pinus sylvestris

22% Pinus sylvestris 17% Juniperus thurifera

14% Pinus pinaster

MDN BD 0 - 37 43% Quercus pyrenaica 846 59% Quercus faginea

43% Quercus faginea 29% Quercus pyrenaica

SE 0 - 64 97% Quercus ilex 2260

CO 427 90% Pinus pinaster 772 44% Pinus pinaster 2974 39% Pinus pinaster

33% Pinus pinea 21% Pinus halepensis

10% Pinus halepensis 14% Juniperus thurifera

8% Pinus nigra 11% Pinus pinea

MDS BD 0 - 0 - 69 30% Quercus faginea

28% Quercus pyrenaica

SE 0 - 42 98% Quercus ilex 939 96% Quercus ilex

CO 0 - 74 73% Pinus pinea 214 51% Pinus pinea

20% Pinus pinaster 22% Pinus pinaster

13% Juniperus oxycedrus

TOTAL 627 1587 12065

Table 1. Sample size and proportion of dominant tree species within the plots for each combination of

environmental zone (EnZ), functional group (FUNC) and regeneration treatment. MDM, Mediterranean

Mountain; MDN, Mediterranean North; MDS, Mediterranean South. BD, broadleaved deciduous; SE,

sclerophyllous evergreen; CO, coniferous.

Page 96: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

82

A functional group (FUNC) variable was defined, based on the functional identity of the

dominant tree species of each plot. The following categories were distinguished:

broadleaved deciduous (BD), sclerophyllous evergreen (SE), and coniferous forests (CO).

Each plot was assigned the functional group of its dominant tree species. Mixed-forest

plots, those with two co-dominant species of different functional groups were not

considered in the analysis of tree species richness (<1%).

Four different types of silvicultural regeneration practices (REG) were assigned to each plot

during field survey in the 3SNFI. These were shelterwood, Clearcutting (CL), Selection

cutting (SEL) and No-Management (NOM). Due to the small sample size (79 plots), species

richness were not assessed in shelterwood plots. Sample size and dominant tree species of

forest plots is presented in Table 1 for each combination of EnZ, FUNC and REG

categories in the study area.

Information about the landscape surrounding each forest plot was also obtained from the

3SNFI plots. For this purpose, not only forest plots, as defined in this study, but all

inventory plots were considered. Also, those outside the boundary of the study area,

including forest plots where tree species richness was not assessed, were considered in

order to avoid edge effects. Woodland plots from the 3SNFI were classified into the

following categories: forests, plantations, dehesas and riparian forests. Dehesas were

defined in the 3SNFI as woodlands with a minimum FCC of 5% with scattered trees

(normally with a FCC lower than 20%) and crop-lands or pastures in the understory. The

forest category was sub-classified according to the regeneration treatment observed in the

plot into three categories: even-aged managed forest (including both clearcut and

shelterwood plots), selection-cutting forest, and unmanaged forest. Thereby the percentage

of managed forests through each treatment was assessed in the landscape surrounding each

forest plot. Given the systematic sampling design of the 3SNFI, based on a grid of 1x1 km-

cells, a fixed number of plots was expected for a particular circle size. Thus, the difference

between the expected and the observed number of plots was used in order to estimate the

percent cover of the category ‘others’, which encompasses agricultural, urban and other

non-forest land uses. Thus, the percent cover of each category was calculated within a

radius of 3.5 and 5.5 km, respectively, around each forest plot (these were median distances

according to a negative exponential function in which maximal dispersal distance is equal to

10.5 and 16.5 km, respectively (see Cain et al., 2000; Jordano et al., 2007). The radius that

yielded the best fit for the model was selected (see below).

Page 97: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

83

Seven environment and forest structure variables that could explain the distribution of tree

species richness were assessed at the local (plot) scale: total precipitation from September to

May (SepMayPREC), maximum temperature in July (JulTEM), minimum temperature in

January (JanTEM), elevation (ELEV), slope, basal area (BA) and tree FCC. Water-energy

variables were selected to incorporate ecophysiological processed related to productivity,

reflecting community ecophysiological responses to water availability, summer stress and

frosts (e.g. Currie and Paquin, 1987; Mittelbach et al., 2001; but see Whittaker and

Heegaard, 2003). Elevation, although an indirect factor and therefore difficult to interpret

(Rahbek, 2005), was intended to account for as much environmentally or human driven

variation as possible (Rahbek, 2005; Nogués-Bravo et al., 2008). Basal area and FCC

accounted, to some extent, for differences in successional stage and biotic interactions,

although they might reflect also differences in forest types across the study area. The

selection of these variables was based on exploratory analyses and on the literature (see

chapter 2). Climate variables were obtained from the Climatic Atlas of the Iberian Peninsula

at a resolution of 200 m (Ninyerola et al., 2005). Elevation was calculated from the official

Spanish Digital Elevation Model at a resolution of 25 m (Ministerio de Fomento, 1999).

Slope and FCC were measured in the field in the 3SNFI. The basal area of each forest plot

(m2ha-1) was calculated as the sum of the basal area at the breast height of all the tally trees

(DBH≥7.5 cm) per unit area.

2.3. Model development

We included different factors at varying spatial and ecological scales and their interactions

to explain local patterns of tree species richness (see Figure 3). Tree species richness was

defined as the number of species of trees surveyed in a particular 3SNFI forest plot i (with

no minimum DBH threshold). Tree species richness was estimated as a function of locally

applied silvicultural treatments (disturbances), with a varying response depending on the

main life history traits of the overstorey (functional group), and the environmental region

where they were applied. The effect of climate, topography and forest structure at the site

also varied by regional environmental zone (EnZ). Finally, the effect of land use in the

surrounding landscape was also considered. We constructed a Hierarchical Bayesian model

to represent this structure. Bayesian statistics differ from conventional statistics by

weighting the likelihood values by the prior probabilities, to obtain posterior probabilities of

the model parameters (Ellison, 2004). Hierarchical Bayes treats the group-level model as

prior information in estimating subgroup- or individual-level coefficients (Gelman and Hill,

Page 98: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

84

2007). Additionally, hierarchical Bayes includes the potential to treat high dimensional

problems with full accommodation of different components of stochasticity (uncertainty in

the unknown parameters vs. variability not explained by deterministic processes in the

model), rather than ignoring stochasticity by treating it as error or noise (Clark, 2005).

Figure 3. Conceptual diagram of the hierarchical structure of factors influencing tree species richness.

Bayesian models were fit using OpenBUGS 3.1.0 (Thomas et al., 2006). The final model

was run for 150,000 iterations, and convergence of the posterior means of the estimated

parameters was assessed from two chains. Pre-convergence “burn-in” iterations (~50,000)

were discarded in the calculation of the parameters. Although we built these models based

on exploratory analyses of the data and ecological theory to better appreciate ecological

processes and answer the main questions of the study (McMahon and Díez, 2007), we also

scored models for selection. We estimated the Deviance Information Criteria (DIC) for each

model, from the simplest structure to the most complex, final model (see all models in

Table 2), in order to evaluate their fit of the data (Spiegelhalter and Best, 2000).

Page 99: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

85

Model DIC ∆DIC

1 ( ) ( )

2( ) ( ) ( ) ( )

1 1

log( )i i

M N

i m m n n

m n

λ ϕ= =

= + +∑ ∑β X θ X 53080 0

2 ( ) ( )

2( ) ( ) ( ) ( )

1 1

log( )i i

M N

i e m m n n

m n

λ ϕ= =

= + +∑ ∑β X θ X 52950 130

3 ( ) ( )

2( ) ( ) ( ) ( )

1 1

log( )i i

M N

i fe m m n n

m n

λ ϕ= =

= + +∑ ∑β X θ X 52630 320

4 ( ) ( )

2( ) ( ) ( ) ( )

1 1

log( )i i

M N

i rfe m m n n

m n

λ ϕ= =

= + +∑ ∑β X θ X 52420 210

5 ( ) ( )

2( ) ( ) ( ) ( )

1 1

log( )i i

M N

i rfe m e m e n e n e

m n

λ ϕ= =

= + +∑ ∑β X θ X 52250 170

6 ( ) ( ) ( ) ( )

2( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )

1 1 1 1

log( )i i i i

M N K H

i rfe m e m e n e n e k k h h

m n k h

λ ϕ= = = =

= + + + +∑ ∑ ∑ ∑β X θ X ω L γ R

(r=3.5 km)

51870 380

7 ( ) ( ) ( ) ( )

2( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )

1 1 1 1

log( )i i i i

M N K H

i rfe m e m e n e n e k k h h

m n k h

λ ϕ= = = =

= + + + +∑ ∑ ∑ ∑β X θ X ω L γ R

(r=5.5 km)

51790 80

8 ( ) ( ) ( ) ( )

2( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )

1 1 1 1

log( )i i i i

M N K H

i rfe m e m e n e n e k k h r h r

m n k h

λ ϕ= = = =

= + + + +∑ ∑ ∑ ∑β X θ X ω L γ R

51750 40

Table 2. DIC for models of increasing complexity. ∆DIC refers to the previous model with decreasing

complexity, not to the ‘best model’ (i.e. that with lowest DIC values).

The number of tree species iY was modeled with a Poisson distribution and log-link

function.

~ ( )i iY Poisson λ

( ) ( ) ( ) ( )

2( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )

1 1 1 1

log( )i i i i

M N K H

i rfe m e m e n e n e k k h r h r

m n k h

λ ϕ= = = =

= + + + +∑ ∑ ∑ ∑β X θ X ω L γ R

Page 100: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

86

2.3.1. Higher spatial and organizational levels

We included intercept parameters,ϕ , for each regeneration treatment, functional group and

environmental zone combination, rfeϕ . Each of these parameters was estimated in a

hierarchical fashion. Intercepts were drawn from prior normal distributions with parameter

values, mean and variances, representing all the functional group and environmental zone

combinations (Figure 4):

ϕrfe ~ Normal (α fe,σ fe2 )

The 32=9 hyper-parameters for the mean defining these prior distributions ( efα ) were

themselves grouped with common prior distributions for each EnZ (Figure 4), accounting

for variability of species richness between FUNC for each EnZ (sensu Clark, 2005). Finally,

hyper-parameters for the three EnZ ( eµ ) were assigned a common non-informative prior

distribution (all plots inform each other in the model) (Figure 4).

α fe ~ Normal(µe,σ e2 )

µe ~ Normal (0,10000)

The variances ( 2σ ) of the described distributions were also given non-informative prior

distributions ( 2 1~ (0.01,0.01);Gammaτ σ

τ= ), where τ is the precision, following

Gelman and Hill (2007).

In order to assess whether managed plots had higher or lower species richness with regard

to unmanaged plots, for each combination of EnZ (e) and FUNC (f), differences between

intercepts for clearcutting, or selection cutting, and unmanaged plots were calculated within

the model. The resulting posterior distributions of these differences were used to calculate

whether treatments had significant effects on species richness. For example, the difference

between clearcutting and no management was calculated as ϕ [clearcutting] –

ϕ [nomanagement] = D. The probability p that clearcutting resulted in lower species

richness than no management, or p(D<0), was then calculated as the cumulative density of

D up to zero. These differences are therefore used as a Bayesian form of a ‘test statistic’

from which the significance of a treatment is calculated:

, , , ,

, , , ,

differences between clearcutting and no management :

differences between selection cutting and no management :

e f clearcutting e f nomanagement

e f selectioncutting e f nomanagement

ϕ ϕϕ ϕ

−−

Page 101: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

87

Figure 4. Conceptual map of the model structure. X : vector of environmental and forest structure

covariates calculated at plot level; L and R : vector of covariates measuring the proportion of forest

land-uses and regeneration treatments, respectively, in the landscape surrounding each plot. Boxes

represent levels of ecological organisation; arrows represent parameter organization (i.e., prior probability

distributions for model parameters and hyperparameters, defined by other hyper-parameters). Note that

the landscape context accounts for an intermediate spatial scale but is calculated at the plot scale, and

does not take part of the parameter hierarchical structure. REG, Regeneration treatment: CL, clearcutting;

SEL, selection cutting; NOM, no management.

2.3.2. Site-level attributes: Landscape context

We included coefficients for the percent cover of each land-use category. Coefficients for

categories of land-use other than forest, ( )kω , (i.e. plantation, dehesa and riparian forest)

were not estimated hierarchically, and were therefore assigned non-informative priors. We

considered unnecessary adding more complexity to the model since this was far from the

main questions of the study.

( ) ~ (0,10000)k Normalω

In order to evaluate the interacting effect of forestry practices applied in the stand and the

percentage of those applied in the neighbouring landscape, we included coefficient

parameters, ( )h rγ , for each combination of local and landscape management treatments.

Page 102: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

88

Each of these coefficients were estimated hierarchically; they were drawn from prior

normal distributions with parameter values, mean and variances, representing each category

of regeneration treatment in the landscape (for instance, the percentage of even-aged

managed forests surrounding all forest plots in the study area informed each other,

independently of the type of management of the central plot), see Figure 4. Hyper-

parameters, mean and variance, for each landscape management treatment were given non-

informative prior distributions:

( )

2

~

~ (0,10000)

1~ (0.01,0.01);

h r h h

h N

Gamma

κ εκ

τ στ

+

=

γ

In order to explore coarse differences in the species richness of each woodland category

(forests, plantations, dehesas and riparian forests), we also estimated a modified version of

the simplest model 1 in Table 2 using a different intercept for each category of plots.

2.3.3. Site-level attributes: Environmental and forest structure variables

Seven environmental and forest structure explanatory variables, ( )mX , were included at the

site level: SepMayPREC, JulTEM, JanTEM, ELEV, SLOPE, BA and FCC. For those

variables showing a unimodal relationship with tree species richness in exploratory analyses

(PREC, ELEV, BA and FCC), the quadratic term 2( )nX was included in the model. In order

to incorporate the hierarchical structure of ecological processes, for each variable, we

included coefficient parameters, ( )m eβ and ( )n eθ , for each environmental zone (EnZ).

Coefficients for each EnZ were drawn from a common distribution, with non-informative

prior distributions (see Figure 4):

( )

( )

~ (0,10000)

~ (0,10000)m e

n e

Normal

Normal

β

θ

Page 103: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

89

3. RESULTS

Increasing complexity of the model resulted in a reduction in DIC over 100 with respect to

the previous less complex model in most cases (Table 2). The most significant reduction,

considering the number of variables added in each step, was due to the inclusion of the

interaction between two variables: FUNC and EnZ (model 3, table 2). Adding the six

landscape variables also had a considerable decrease in DIC values, especially when using

the 5.5 km-radius, rather than the 3.5 km-radius around each plot to define the landscape

(models 6 and 7). Although the stratification of coefficients for landscape silvicultural

treatments according to management practices on the central forest plot resulted in a small

decrease in the DIC values (∆DIC=40, model 8), their inclusion allowed us to respond to

the second main question of this study (see introduction).

The probability that clearcut forests had lower tree species richness than non-managed

forests was considerably higher in coniferous forests (probability ≥ 0.9, using the approach

to assess significance described in the methods), and lower in broadleaved deciduous forests

(probability = 0.7), in all environmental regions (Table 3). In Mediterranean mountain and

northern regions, selection cutting treatments resulted in an increase of species richness

with probabilities above 0.95 for sclerophyllous and coniferous forests (the latter only in

northern Mediterranean region), and 0.81 for broadleaved deciduous forests. On the

contrary, tree species richness was lower on selection-cutting coniferous forests in

Mediterranean mountains (p=0.83, Table 3). In the southern Mediterranean region, selection

cutting seemed not to affect tree species richness in coniferous forests (p=0.52), and a

slightly negative tendency was observed in sclerophyllous forests (p=0.66).

According to exploratory analysis of coarse differences in species richness, plantations and,

particularly, dehesas had significantly fewer species than forests, while riparian forests were

considerably more rich in species than the rest of woodland types (Figure 5) The

neighboring landscape had a particularly important negative impact on tree species richness

of local stands when dehesas or, especially, clearcut forests were abundant, regardless of

the management that these stands had received (Figure 6). Larger landscape proportions of

planted forests resulted also in lower number of species in the central forest plot, while

riparian forests in the landscape tended to increase it (p=0.9). Forest stands, independently

of the regeneration treatment received, had fewer species when located in a landscape with

abundant clearcut forests (Figure 6). Similarly, species richness decreased in selection

cutting forests with increasing selection cutting in their neighbourhood. Finally, landscapes

Page 104: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

90

with a large proportion of unmanaged forests influenced positively species richness of

central, unmanaged forest plots (Figure 6).

Clearcutting Selection cutting

EnZ Functional type Effect Prob Effect Prob

MDM Broadleaved deciduous - 0.70 + 0.81

Sclerophyllous evergreen no data no data + 0.99

Coniferous - 0.90 - 0.83

MDN Broadleaved deciduous no data no data + 0.81

Sclerophyllous evergreen no data no data + 0.96

Coniferous - 1.00 + 0.96

MDS Broadleaved deciduous no data no data no data no data

Sclerophyllous evergreen no data no data - 0.66

Coniferous no data no data - 0.52

Table 3. Effect of regeneration treatments on tree species richness in the forest plots where they were

applied, compared with unmanaged forests of the same functional type and in the same EnZ region. For

each combination of EnZ (e) and FUNC (f), the difference between intercepts ( efrϕ ) for clearcutting, or

selection cutting, and unmanaged plots was calculated, and a probability measure of the significance of

the differences is provided (see methods).

The predicted richness of tree species for each EnZ against each of the environmental and

forest structure variables is shown in Figure 7. A significant part of the range observed for

most variables overlapped in the three EnZs; in the case of JulTEM, JanTEM and ELEV,

however, a tendency of increasing temperature and decreasing mean elevation MDM-

MDN-MDS can be appreciated, as expected according to Figure 2. Climatic variables were

the most relevant drivers of tree species diversity, especially JulTEM in Mediterranean

mountains, JanTEM in the northern Mediterranean region, and SepMayPREC in the

southern Mediterranean region (Table 4 and Figure 7). Higher species richness was

associated with increased SepMayPREC, JulTEM, Slope, BA and FCC, and lower JanTEM

in all three regions. Nonetheless, extreme values (within the range observed in each EnZ) of

precipitation, basal area or FCC were not beneficial for species richness in some regions

(Figure 7). The relationship with elevation was unimodal, except in the southern

Mediterranean region where the negative association was weak (p=0.72, Table 4). In

general, the gradient MDM-MDN-MDS was linked to increased importance of

SepMayPREC, SLOPE and FCC, and decreased importance of JulTEM and BA, as controls

Page 105: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

91

of richness of tree species (Figure 7 and Table 4). BA had in general the lowest importance

(Table 4).

Figure 5. Posterior mean parameter estimates and the 95% credible intervals for the intercepts

representing tree species richness on each category of woodland when the other environmental and forest

structure variables are equal to zero (i.e. to their mean value in the study area, since they were previously

standardized). The simplest model in Table 2 was used to explore coarse differences on tree species

richness of these categories for the whole study area.

Figure 6. Posterior mean parameter estimates and the 95% credible intervals for the coefficients

representing effects of landscape cover variables on local tree species richness. Different coefficients for

the effect of silvicultural treatments across the landscape were distinguished depending on the

regeneration treatment applied in the central forest plot, where species richness was assessed. Those

variables with intervals that do not overlap the horizontal zero line may be considered significantly

different from zero at the 95% level.

Page 106: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

92

Figure 7. Predicted tree species richness for the 95% central range of each explanatory variable in each

environmental region, while maintaining constant the remaining variables (mean values). The overall

intercept for each EnZ was used, and landscape-context variables were set to zero. Note that uncertainty

of the parameter estimates is not represented for clarity reasons. MDM, Mediterranean Mountain; MDN,

Mediterranean North; MDS, Mediterranean South.

MDM MDN MDS

β x p θ x2 p β x p θ x

2 p β x p θ x

2 p

JulTEM 0.41 1.00 - - 0.21 1.00 - - 0.18 0.99 - -

JanTEM -0.27 1.00 - - -0.32 1.00 - - -0.09 0.85 - -

Prec 0.17 1.00 -0.14 1.00 0.26 1.00 -0.19 1.00 0.43 1.00 -0.18 0.99

ELEV -0.03 0.74 -0.21 1.00 -0.08 0.98 -0.48 1.00 -0.09 0.72 -0.03 0.60

SLOPE 0.14 1.00 - - 0.16 1.00 - - 0.18 1.00 - -

BA 0.12 1.00 -0.09 1.00 0.06 0.99 -0.05 0.99 0.04 0.68 -0.07 0.80

FCC 0.13 1.00 -0.09 1.00 0.15 1.00 -0.06 0.98 0.14 1.00 -0.17 0.99

Table 4. Posterior mean parameter estimations, and the probability that they are different from zero, for

the environmental and forest structure variables in each environmental zone (see Figure 4).

Page 107: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

93

4. DISCUSSION

Processes affecting tree species richness appeared to have different patterns (i.e., varying

magnitude and, in some cases, function shape of the relationships) depending on the

regional conditions of aridity and elevation range, as defined by the main environmental

zones into which the study area was divided (Mediterranean mountains, Mediterranean

North and Mediterranean South). By structuring model parameters and hyperparameters,

Hierarchical Bayes provided inferences at the site level, while accounting for factors

influencing species richness at other organizational (life history traits of the dominant tree

species and silvicultural disturbances) and spatial (environmental region) levels.

4.1. Regeneration treatments and tree species richness

In general, forests managed through clearcutting practices had fewer tree species than

unmanaged forests, while selection-cutting forests had more tree species than unmanaged

forests, under similar regional environmental conditions and for the same functional group,

with some exceptions. It is important to note that these results were obtained after

accounting for variability due to succession dynamics and differences in biomass (i.e. basal

area and FCC). Clearcutting is only applied in the study area to Pinus sylvestris and Pinus

pinaster stands, managed for timber production. This intensive forest management

promotes mono-specific forests, in some cases even through artificial regeneration.

Additionally, mechanical site preparation hampers abundance and growth of residual

sources of shade-tolerant species (by affecting the seed bank, and their advance and

vegetative reproduction) (e.g. Newmaster et al., 2007). Finally, canopy removal in

clearcutting results in drastic environmental changes that enhance disturbance-tolerant,

superior competitor species hindering the coexistence of shade-tolerant species (e.g.

Halpern and Spies, 1995). By contrast, canopy gaps created by selection cuttings are

characterized by the maintenance of the influence of surrounding trees and of the main

environmental properties of the site (Aunós, 2005; Wagner et al., 2011). Actually,

intermediate light intensity conditions have been found to benefit seedling establishment,

rather than closed or open conditions, for many Iberian tree species (e.g. Rousset and

Lepart, 2000; Arrieta and Suárez, 2005; Rodríguez-Calcerrada et al., 2010; see also

references in Wagner et al., 2011). The new microclimate conditions in the understory and

the resulting increase in resources availability (nutrients and light) (Aussenac, 2000)

probably promoted the establishment of fast-growing, pioneer species in dense hardwood

forests subjected to selection cutting, while allowing tolerant species to persist. In the case

Page 108: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

94

of coniferous forests, they have been shown to support oak species regeneration in Spain,

although saplings depend on canopy gaps to grow to adulthood (Gómez, 2003; Urbieta et

al., 2011); this phenomenon might explain in part the increased tree species richness in the

northern Mediterranean region after selection cutting. The higher tree species richness in

selection-cutting versus unmanaged forests was consistent with results found in Catalonia

(NE Spain) by Torras and Saura (2008). However, environmental changes in gaps also

favor the herbaceous cover, generating competition interactions with tree seedlings

(Aussenac, 2000). This competition, together with photodamage, might compromise

seedling viability under strong radiation and water stress conditions, as in southern-

Mediterranean summers. Actually, numerous studies suggest that canopy facilitation, rather

than competition, prevails in biotic interactions in semi-arid regions (Castro et al., 2004;

Pausas et al., 2004; Rey-Benayas et al., 2005; Puerta-Piñeiro et al., 2007; Gómez-Aparicio

et al., 2008; Valladares et al., 2008). These are probably the mechanisms that explain the

neutral or negative effect of canopy removal on tree species richness in the southern

Mediterranean region, even at intermediate gap sizes. A longer period of shade during plant

life history is normally required for most species growing under water deficit conditions

(Aussenac, 2000); however, as shown here, as canopy closure occurs competitive exclusion

is enhanced in the southern region (Figure 7). The highest canopy densities in this region

occur in the latest successional stage in forests of shade-tolerant tree species, where most

Mediterranean, light-demanding tree species difficultly survive.

In Mediterranean mountains, non-managed coniferous forests had more species of trees

than coniferous forests managed through selection cutting. Apart from Pinus pinaster

species, coniferous mountain forests in the study area receiving selection cuttings are

mainly composed by Pinus nigra and Pinus sylvestris species (Table 1). The negative

impact of selection cuttings might be explained by the fact that P. sylvestris and P. nigra are

the conifers with best timber quality in Spain. Although the main differences among

regeneration treatments here considered come from canopy-gap size, each of them

represents a wider set of forestry operations, such as elimination of competitor species.

These operations may be more severe in forests with higher commercial value, and in some

cases can counteract the potential positive effect of canopy opening on tree species richness.

Finally, conifer stands currently managed through selection cutting, which are mainly

located in eastern mountains in the study area, have historically been harvested through

shelterwood treatments until late XX century (del Campo Sanchís and Solís Camba, 1993),

which can influence composition of current plant communities (Hermy and Verheyen,

2007).

Page 109: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

95

4.2. Management in the neighbouring landscape

Our results showed that management disturbances not only affected the forest stand where

they were applied, but also they influenced the number of tree species of forests in the

neighboring landscape. Forest stands embedded in landscapes with increasing proportions

of even-aged management (mainly clearcutting) tended to have fewer tree species.

Similarly, species richness in selection cutting stands diminished when this treatment was

extensively applied in their landscape. The elimination of some shade-tolerant species from

the landscape, may hinder their recolonization of small canopy gaps, given distance

limitation in seed dispersal (Bengtsson et al., 2000). This fits the findings in chapter 2,

where gamma richness of tree species decreased when more than 60% of the landscape was

managed through selection cuttings. Finally, the increased species richness in stands

surrounded by forest-dominated landscapes supports the mentioned metacommunity

processes, but is probably also explained by a common ecological history (Legendre and

Legendre, 1998; Hermy and Verheyen, 2007).

4.3. Dehesas, plantations and riparian forests

Results seem to support the hypothesis that large numbers of species present in the

landscape or broader region (gamma diversity) are important to allow colonization and

enrichment of a particular habitat type. The proportion in the landscape of the woodland

covers here considered influenced gamma species richness and ultimately local species

richness. The remarkably high richness of tree species in riparian forests of the study area

reflects the role of river floodplains in the persistence of many temperate deciduous species

non-adapted to dryer Mediterranean habitats. Despite their slight representation in the

landscapes (maximum 5%), riparian forests had a positive impact (p= 0.90) on tree species

richness of surrounding forests (Figure 6). Notwithstanding the singular herbaceous and

animal diversity in dehesas (Díaz et al., 2003; Pérez-Soba et al., 2007; Ramírez and Díaz,

2008) contrasted with poor tree species richness (one or two: mainly Quercus ilex or

Quercus suber) probably due to management practices that select for only these two woody

species. Metapopulation processes could be also the origin of the poorer forests observed in

landscapes where dehesas are abundant. However, it is very probable that ecological history

largely explains the observed patterns. Many forest plots in landscapes dominated by

dehesas were also exploited until some decades ago through practices typical of these

traditional grazing systems (tree clearing and pruning, periodical ploughing, livestock

grazing). Recent changes in the farming industry since the 1950s are pushing these types of

Page 110: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

96

agricultural systems towards intensification in the more productive and populated areas and

abandonment in marginal zones, even at local scales (Peco et al., 2006; Ramírez and Díaz,

2008). And, although abandoned dehesas have great importance in the sustainability of oak

populations in the region (Ramírez and Díaz, 2008), the presence of saplings or young trees

of species other than the main one (e.g. Q. ilex) in dehesas abandoned at least 20-30 years

ago is negligible (Peco et al., 2006; Ramírez and Díaz, 2008; Tárrega et al., 2009).

Finally, pine woodlands reforested within the second part of the 20th century in Spain had

also lower tree species richness than the rest of forests, as defined in this study. These

plantations have scarcely been colonized by late-successional broadleaved species, because

of their high density and dense shade (Gómez-Aparicio et al., 2009) and seldom

management through silvicultural post-plantation operations (Valbuena-Carabaña et al.,

2010). Despite the negative effect on natural forests diversity suggested by our results,

appropriate silvicultural disturbances, in the context of forests dominated landscapes, could

maintain and even enhance forest biodiversity in plantations, while their persistence would

not be compromised (Pausas et al., 2004).

4.4. Environmental and forest structure variables

Climatic factors are expected to be the main drivers of species richness at extents larger

than 1000 km (Field et al., 2009). Here, energy and water variables had also a prominent

role explaining tree species richness at the intermediate extent defined by the environmental

regions (ca. 350 km). Differences among regions in the range of values encompassed might

explain, in part, variability in the relevance of each climate variable (e.g. July temperature)

as a control of tree species richness. But the results also probably reflect the primacy of

water availability under marked hydric stress conditions, very extreme in the southern

Mediterranean region, whereas mountains (MDM) distributed in warmer regions still

benefit from lower temperatures and therefore lower evapotranspiration rates. Such

conditions in mountain areas have allowed the refuge of some temperate deciduous tree

species since the Mid-Holocene (Benito et al., 2007). Also in the case of Mediterranean

mountains and northern Mediterranean region, higher species richness was associated to

highest summer and winter temperatures, and intermediate altitudes. Given the high

altitudinal correlation of winter temperature (Spearman r= -0.7 in each region) and the

latitudinal decreasing gradient of annual and summer temperature (see patterns in chapter

2), each of these associations needs to be interpreted in the context of the other

environmental variables. Therefore, the highest species richness was predicted in upper

Page 111: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 3

97

areas, with colder winters, preferentially distributed in the southern-most mountain ranges,

with warmer summers. Nonetheless, similar winter temperatures can be found at varying

elevations depending on the latitude in the study area. For a homogeneous range of

temperature conditions in winter, the upper peaks of southern mountains probably host

fewer species, due to harsh conditions and grazing land use, than forests found at

intermediate elevations at northern latitudes. Therefore, the patterns of elevation-species

richness relationships observed in Figure 6 should not be interpreted as the typically

observed hump-shaped patterns at local scales (Rahbek, 2005; Nogués-Bravo et al., 2008),

since they accounted for climate variability among different mountain ranges within a same

region (Austin et al., 1984).

Page 112: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 113: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

Page 114: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 115: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

101

CHAPTER 4

Landscape species pools and connectivity patterns

influence tree species richness in both managed and

unmanaged stands*

Based on results in chapter 3, we hypothesized that both species richness of trees in a stand

and its response to a silvicultural depend on the pool of species in the surroundings able to

colonize that site. In this chapter we specifically tested this potential metacommunity

dynamics using a graph theoretical approach. We estimated the potential diversity of seed

fluxes that could be received at a given stand through long distance dispersal from other

source tree populations, and evaluated the relevance of such flux diversity in a top-down

hierarchical modelling structure.

* This chapter has not yet been published. It will be submitted in the near future as an article for a

scientific journal with the following authors: Martín-Queller, E. and Saura, S.

Page 116: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

102

ABSTRACT

Intermediate disturbances and metapopulation dynamics are among the theoretical

mechanisms explaining species coexistence at the local and landscape scale. Both

mechanisms might interact, such that the consequences of local disturbances might depend

on long distance dispersal events. In this study we examined whether the diversity of tree

species potentially able to colonize a locality from the surroundings is associated with the

tree species richness observed in that locality, and/or with the response of that richness to

selection cutting treatments. The study was located in a Mediterranean region in central

Spain were forests selectively cut had been found to have more tree species than

unmanaged forests. We used a top-down hierarchical modelling structure to account for the

effect of other factors such as climate, lithology and amount of forest cover at the landscape

scale. Species richness of trees was strongly associated to annual precipitation, and was

maximized at intermediate rainfall levels. Under homogeneous climate and lithological

conditions, the composition and connectivity of seed sources in the landscape seemed to

play a more relevant role explaining tree species richness than the amount of forest habitat

in the surroundings. Particularly, higher species richness was observed in forest stands

potentially receiving a higher diversity of seed fluxes. Patterns in the response of species

richness to selection cutting were less clearly explained by differences in diversity of

potential seed fluxes, but time lags in the responses, or differences in the proportion of

shade-tolerant species in the landscape could mediate this interaction. Stronger importance

of the amount of forest habitat and diversity of potential seed flux may be masked by their

correlations with precipitation gradients in the study area.

Page 117: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

103

1. INTRODUCTION

One of the fundamental objectives of ecology is the search for the general mechanisms that

control the spatio-temporal pattern of species coexistence. Physical environmental gradients

have been found to highly correlate with patterns of species richness of plant communities

at large spatial (extent and/or grain) scales (e.g. Field et al., 2009). Climate and lithology

determine resource availability and influence plant physiology, thus directly driving plant

community composition according to the niche theory (Hutchinson, 1957). But apart from

the equilibrium mechanisms based on the partitioning of spatio-temporal niche

heterogeneity (MacArthur, 1972; Chesson, 1983; Tilman, 1986), many other dynamic-

equilibrium mechanisms for species coexistence have been offered (reviewed in Palmer,

1994; Huston, 1999; and Wright, 2002). For instance, periodical disturbances -when

intermediate in intensity, frequency and/or extent- are hypothesised to maximize plant

species richness in a local community by preventing the competitive dominance of few

species (Connell, 1978; Shea et al., 2004). At the same time, it is increasingly accepted the

necessity to incorporate processes occurring at landscape or regional scales to understand

patterns of local community composition (Gardner and Engelhardt, 2008). Actually,

coexistence of some plant species may occur at landscape rather than local scales due to

metapopulation dynamics (see Leibold et al., 2004). It has been suggested that inter-specific

differences in competitive versus dispersal ability allow inferior competitors to persist by

colonizing more rapidly or efficiently other localities (Horn and Mac Arthur, 1972);

similarly, differences in the abilities to perform under different habitat conditions can also

explain landscape coexistence (see Leibold et al., 2004). Both mechanisms, local

intermediate disturbances and trade-offs in ecological traits of species may be linked. Thus,

disturbances might allow persistence of inferior competitor plant species both from the local

community and from surrounding communities. In this context, the response of a forest

community composition to a particular disturbance will depend both on local attributes and

on the composition of surrounding communities and the degree of connection with them

(Niemelä ,1999; Bengtsson et al., 2000; Ricotta et al., 2008).

Metacommunity dynamics, that is, connections among local communities by dispersal of

multiple potentially interacting species (Wilson, 1992), seems to have therefore a crucial

role in the maintenance of both landscape and local species richness. It has even been

hypothesized that the available pool of species in a region is the main determinant of local

species richness (Eriksson, 1993; Cornell and Lawton, 1992; Pärtel and Zobel, 1999). The

species pool hypothesis suggests that patterns of local richness are mainly explained by

evolutionary and historic extrinsic processes, which determine landscape or regional species

Page 118: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

104

richness, rather than by intrinsic local processes such as disturbances or biotic interactions

(but see Akatov et al., 2005).

Large-scale connectivity of populations depends on long-distance dispersal events (LDD).

These can be defined for each plant species in terms of a threshold distance, or as the tail

from an empirically estimated dispersal density function or kernel (See Cain et al., 2000).

However, LDD events are rare compared to local dispersal and frequently rely on

nonstandard dispersal vectors (or combinations of several of them) or result from rare

behavior of the (standard) dispersal vector (Higgins et al., 2003). These particularities make

LDD events difficult to sample, and thus empirical data is limited almost exclusively to

short-distance events (Nathan and Muller-Landau, 2000). Even so, there is compelling

evidence from many plant species that effective LDD do occur far beyond the otherwise

observed dispersal distances (Nathan, 2006). For instance, studies have reported plausible

dispersal distances for many plants up to 10-20 km (e.g. Clark et al., 1999; Cain et al., 2000;

Jordano et al., 2007).

Disentangling the role of metacommunity dynamics in the generation of the observed

patterns of local species richness is a challenge. In fact, species richness patterns result from

complex interactions of factors operating at different spatial scales (Whittaker et al., 2001;

Field et al., 2009). These interactions are commonly accepted to occur mainly in a top-down

hierarchical structure (Levin, 1992; McMahon and Diez, 2007). Particularly, coarse-scale

patterns of climate and lithology are assumed to be hardly affected by ecological processes

that take place at smaller spatial scales (e.g. Sarr et al., 2005; Kallimanis et al., 2007),

whereas these ecological processes or even human induced changes in the landscape or

ecosystem structure are strongly determined by the environmental context.

In previous studies we found that intermediate silvicultural disturbances, such as selection

cutting or improvement treatments, were associated to an increased richness of woody plant

species, when compared to unmanaged stands (chapters 2, 3), at least in some

Mediterranean areas in central Spain. However, results suggested that this benefit varied

depending on the land uses in the surrounding landscape. For instance, species richness in

stands treated with selection cuttings diminished along with increasing clearcutting in the

neighborhood. Accordingly, we hypothesized that the increase in species richness after

these disturbances is more important where more species are potentially able to recolonize

the opened gaps. This hypothesis was not however explicitly addressed or tested in previous

studies. For this reason, here we focused on a Mediterranean region (c.a. 88,000 km2) where

species richness had been previously found to be higher in selectively-cut than in

Page 119: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

105

unmanaged stands (chapter 3), and addressed the following questions: (1) to what extent the

richness and relative abundance of species potentially able to colonize a locality from the

surroundings is associated with the species richness observed in that locality? and (2) to

what extent the species richness in stands disturbed by management is affected by

(metacommunity) processes from outside the community? In order to explore these

questions, we analyzed the relationship between the diversity (or richness) of propagules

potentially colonizing a forest stand through LDD (here ≥ 1 km) from the surrounding

landscape and the number of tree species of the stand, differentiating between unmanaged

and selectively-cut forest stands. Potential seed flux was estimated through a graph theory

approach for connectivity analysis. We based our analysis in a top-down hierarchical

framework, assessing the pure effect of seed flux diversity on local richness, after

accounting for factors that may shape local species richness directly (climate and lithology),

or indirectly by being related to the observed patterns of seed flux diversity (landscape

amount of forest cover) (e.g. Fahrig, 2003; Montoya et al., 2010). Standard multiple

regression models cannot incorporate any causal knowledge we might have about the

reasons for the correlations among predictors (Shipley, 2000). In order to incorporate the

theoretical top-down, hierarchical structure of the factors influencing species richness, we

applied regression analyses hierarchically separated into three different steps, using

residuals from the previous step as the response variable in the next one.

To our knowledge, this is the first study adopting a graph-theoretical approach to evaluate

the association between plant species richness in a local community and the diversity of

potential seed sources from the surrounding landscape. Additionally, the evaluation of the

effect on tree species richness of the interaction between silvicultural disturbances and

potential seed flux diversity is also novel; previous studies have focused on the analysis of

the effects of either the landscape context or the management treatments but have not

jointly considered both.

2. METHODS

2.1. Study area

The study area encompasses the forests located within the Mediterranean North Region

(according to the European Environmental Stratification by Metzger et al. (2005)) of the

Spanish regions of Castilla y León, Castilla-La Mancha and Madrid (Figure 1). We selected

this climatic region (Mediterranean North Region) because under these conditions

selection-cutting stands had been found to have more tree species than unmanaged stands

Page 120: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

106

(see chapter 3). This makes it a particularly interesting study area for further exploring and

understanding the effects of management on Mediterranean tree species richness. Further

details about the study area are provided in chapter 2 and in appendices A and B.

Figure 1. Analysed sink plots in the Mediterranean North Region (MDN) in the study area (A) and detail

of the neighbour source forest plots within a radius of 17 km around the analysed sink forest plots (B).

The study is based on data from the third Spanish National Forest Inventory (3SNFI)

(Ministerio de Medio Ambiente, 1997-2007). Plots in the 3SNFI were located

systematically in the intersections of the 1 km x 1 km UTM grid that fall inside forests and

other woodlands. Plots were circular and the inventory of tree stems depended on their

diameter at breast height (DBH) and distance to the plot centre, which ranged from 5 m for

trees with DBH from 7.5 cm to 12.5 cm up to a maximum radius of 25 m for trees with

DBH of at least 42.5 cm. Silvicultural treatments are surveyed in the 3SFNI according to

direct or indirect evidences assessed in the field, such as stumps or slash, normally not older

than ten years. Forests were defined here as woodlands with a minimum Forest Canopy

Cover (FCC) of 5%, excluding plantations, dehesas, riparian or burnt forests. The selection

of the study plots was made in order to ensure that a set of managed and unmanaged stands

with similar environmental conditions, metacommunity dynamics, and ecological and

evolutionary histories were represented. Firstly, all forest plots dominated by coniferous

species and where selection cutting treatments had been applied were selected (696 plots).

We focused on coniferous forests, for which sample size of managed plots was much bigger

than for other functional groups. Secondly, all unmanaged, coniferous forest plots located

within a 2.5 km radius around selection cutting plots were also selected for the analysis

Page 121: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

107

(849 plots). We call this set of 1545 managed and unmanaged plots ‘sink plots’, since

diversity and richness of potential seed flux to each of them was subsequently evaluated.

All 3SNFI plots surrounding each sink plot within a certain influence radius, independently

of their functional group or type of management, were also included in the analysis to

estimate the potential seed flux that could originate from them (source plots), allowing

some propagules to reach the sink plots (see the following section).

2.2. Estimation of potential seed flux

We used an approach based on graph theory to estimate the potential seed flux received by

each sink plot (either selectively-cut or unmanaged). Graph theory is a intuitive and very

powerful approach to evaluate the functional connectivity between spatially structured

ecological elements (individuals, populations, communities), as testifies the increasing

number of ecological studies based on it (Neel, 2008; Pascual-Hortal and Saura, 2008;

Urban et al., 2009; Erös et al., 2011; García-Feced et al., 2011; Pereira et al., 2011). A

graph represents the landscape as a set of nodes functionally connected to some degree by

links that join pairs of nodes (Urban and Keitt, 2001). A total of 49 graphs were built, one

for each native tree species present in at least one of the sink plots according to the 3SNFI

surveys (see Table 1). These graphs were composed of two different types of nodes: (1) the

nodes represented by ‘sink plots’, that is, forest stands where the effects of potential seed

flux on species richness were subsequently evaluated; and (2) the nodes represented by

‘source plots’, that is forest stands susceptible to disperse seeds to the sink populations.

Each graph, therefore, represented a set of populations (source plots) of a particular species

distributed across the study area and susceptible to disperse seeds to the sink populations.

All 3SNFI plots (including conifer, deciduous and sclerophyllous forests, dehesas,

plantations or burnt forests) were potential source populations as long as at least one adult

potentially generating a significant seed output of the species represented in the graph was

present, according to the 3SNFI field plots. This implies that sink plots could also act as

seed sources for other sink plots. In order to exclude as a potential seed source those

individuals that had not reached sexual maturity, a threshold of age, dominant height or

DBH should be used. However, this type of data was hard to obtain for each of the 49 tree

species in this study. Furthermore, the sampling design in the 3SNFI further complicated

the issue; although all species within 25 m radius were surveyed in the 3SNFI, tree-level

measurements were only done for individuals with DBH ≥ 7.5 cm; and in the case of the

smallest individuals, those with 7.5 cm ≤ DBH ≤ 12.5 cm, only for those occurring within a

Page 122: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

108

radius of 5 m (see details of 3SNFI sampling design above). Indeed, less frequent tree

species occurring in the dominated overstorey normally fall within the smallest diameter

class; individuals from these species are less probably characterized in the 3SNFI than

individuals from dominant species reaching larger DBH and therefore inventoried in a

wider radius (up to 25 m). As a consequence, for many tree species, excluding individuals

with DBH < 7.5 cm would imply removing 80%, and in most cases 90%, of the plots where

they were present in the study area, according to the 3SNFI. For these species, using a DBH

threshold would be too restrictive. We applied a threshold of 7.5 cm for those dominant

species with a large number of tally trees in the study area (Table 1): these were defined as

those species in which at least 30% of the plots (in most cases >50%) where they were

present (according to the 3SNFI) had one or more individuals with this minimum DBH

value. Sexual reproduction of many of the main, dominant tree species in Spain begins at

the age of 15-25 years (Madrigal et al., 1999; Montero et al., 200; Serrada et al., 2008),

which corresponds to very variable DBH values, from 5 cm to 20 cm, depending on the

species and on site quality. Ribbens et al. (1994), when modelling local spatial recruitment

of temperate forests, considered that individuals with a DBH lower than 10 cm were

presumed to produce few or no recruits. In this study, we considered that, given the scale of

grain used, 1x1 km, the presence of an individual with 7.5 cm DBH indicates a very

probable presence of a population in the neighborhood susceptible to disperse seeds at long

distances. Also, high densities of young trees with low seed production may render a large

crop size on a population basis. For the non-dominant species (Table 1) we did not apply

any DBH threshold, and they were considered as potential seed sources in every plot in

which they were present within the 25 m sampling radius.

Internode seed dispersal probabilities (pij) were estimated as a negative exponential function

of distance (Bunn et al. 2000):

· ijk dijp e−=

Where k is a constant that was specified by setting a probability p=0.01 at the maximum

dispersal distance of 17 km (see Jordano et al., 2007). Median distance (p=0.5) according to

this function corresponds to 2.6 km. Probabilities below 0.01 were not considered in the

analysis. The potential seed flux of a given species to each sink plot j was equal to the sum

of the probabilities of dispersal from the i nodes (source plots) within a 17 km-radius (i.e.,

p≥0.01). Finally, for each sink plot we estimated seed flux richness, as the number of

species with a potential seed flux to the plot above zero; and seed flux diversity, as the

Page 123: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

109

Shannon diversity of potential seed fluxes of each tree species reaching the sink plot. We

used the Conefor Sensinode software (CS) (Saura and Torné, 2009,

http://www.conefor.org) to estimate this flux, by using the Flux (F) metric (Bunn et al.,

2000; Urban and Keitt, 2001).

Dominant species Non-dominant species

Alnus glutinosa Acer monspessulanum

Castanea sativa Acer pseudoplatanus

Fraxinus angustifolia Amelanchier ovalis

Fraxinus excelsior Arbutus unedo

Juglans regia Buxus sempervirens

Juniperus thurifera Celtis australis

Pinus halepensis Cornus sanguinea

Pinus nigra Crataegus monogyna

Pinus pinaster Crataegus spp.

Pinus pinea Frangula alnus

Pinus sylvestris Juniperus communis

Populus alba Juniperus oxycedrus

Populus nigra Juniperus phoenicia

Quercus faginea Malus sylvestris

Quercus ilex Myrtus communis

Quercus pyrenaica Olea europaea

Quercus suber Pistacia terebinthus

Salix alba Prunus avium

Prunus spinosa

Prunus spp.

Pyrus spp.

Rhamnus alaternus

Salix atrocinerea

Salix purpurea

Salix spp.

Sambucus nigra

Sorbus aucuparia

Sorbus spp.

Sorbus torminalis

Taxus baccata

Ulmus minor

Table 1. List of tree species present in at least one of the 1545 3SFNI plots included in the analysis,

differentiating between dominant and non-dominant species for which a DBH threshold of 7.5 and 0 cm

has been considered, respectively (see methods).

Page 124: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

110

2.3. Estimation of habitat amount and connectivity

Despite the term habitat is by definition species-specific, we used forest cover as a proxy of

the amount of habitat for all tree species present in the study area, as in Montoya et al.

(2010). We estimated an indicator that accounts for the patterns of forest amount and

connectivity in the landscape around each sink plot. This indicator therefore should

represent processes derived from habitat loss and fragmentation (Fahrig, 2003), such that

diversity of potential seed sources unambiguously reflects community composition in the

surrounding landscape. In order to achieve this, estimation of this indicator was also based

on graph theory. We used the same Flux (F) index and negative exponential function, with

the same decay rate (i.e. assigning a probability 0.001 at 17 km). In this case, a unique

graph included all the 3SNFI plots, except plantations, regardless of species composition.

The variable was calculated for each of the 1545 target (‘sink’) plots as the sum of

probabilities of direct connection with each plot in the neighbourhood (radius of 17 km).

Hereafter we call this variable ‘forest cover’. The forest cover variable was therefore

calculated in an analogous way to the seed fluxes but considering all the forest cover

irrespectively of the individual tree species present in the different plots.

2.4. Statistical analysis

We used Bayesian models to explain the patterns of tree species richness in the sink plots.

Multivariate models evaluate the relative effect of explanatory variables, once considered

their covariation with the rest of predictors in the model. However, from an ecological point

of view, the influence between explanatory variables may only occur in one sense.

Particularly, in a top-down hierarchy paradigm, large scale climate and lithology patterns

not only are expected to influence directly local species richness, but also they may be the

template that defines other processes shaping landscape configuration and composition,

which themselves affect local species richness. That is, environmental conditions might

indirectly influence local community composition by shaping habitat amount and

connectivity and landscape pool of species. In order to integrate this theoretical structure in

our analyses, and evaluate the role of forest cover and diversity of potential seed flux not

explained by processes operating at larger scales, explanatory variables were evaluated

hierarchically by introducing them in three different steps according to specific hypotheses.

Spearman correlation values between all pair of variables, and potential interactions with

categorical variables, were calculated to evaluate possible interactions between predictors of

species richness and to discuss causality of the estimated relationships.

Page 125: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

111

First step. In this step, the number of tree species was modeled with a Poisson distribution

and log-link function against a set of climatic, topographic and lithological factors.

In order to reduce collinearity, and to represent environmental variability in a minimum

number of variables, we performed a Principal Component Analysis (PCA) with the

explanatory variables. The 13 environmental variables used in the PCA were: elevation,

slope, annual and seasonal (spring, summer, autumn and winter) precipitation, maximum

July temperature, minimum January temperature, mean annual radiation, Thornthwaite’s

annual potential evapotranspiration (PET), drought length and Dry Season Water Deficit

(DSWD). Climatic data were obtained from the Climatic Atlas of the Iberian Peninsula at a

resolution of 200 m (Ninyerola et al., 2005), and topographic data from the official Spanish

Digital Elevation Model at a resolution of 25 m (Ministerio de Fomento, 1999). Drought

length was calculated as the number of dry months, that is, when the value of the monthly

amount of precipitation (mm) was less than, or equal to, twice the value of the average

temperature (ºC) of the month considered (Allué-Andrade, 1990). Monthly water deficit

was calculated as the difference between monthly precipitation and PET, and DSWD was

the accumulated water deficit for all dry months (Dufour-Dror and Ertas, 2004). PCA was

conducted using the prcomp function in R software.

Scores of the two first components of the PCA and an additional categorical variable,

lithology, were used as explanatory variables in this step. Lithology had two categories:

calcareous versus siliceous soils; this variable was obtained from the 3SNFI for each plot,

where it was determined based on hydrochloric acid test. The shape of the relationship

between each PC and tree species richness was explored and accordingly, linear or

quadratic functions were included in the model.

Second step. Once eliminated the patterns of tree species richness that could be explained

by covariation with environmental patterns, the residuals from the previous step were

modeled with a Normal distribution using forest cover as explanatory variable, and

including an intercept parameter in the model.

Third step. Residuals from the second step, now used as response variable, represented the

distribution patterns of tree species richness that were not explained by environmental

conditions of local stands, or by differences in the forest cover distribution. We included in

Page 126: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

112

the model intercept parameters, rα , for each category of regeneration treatment (selection

cutting or unmanaged) in order to evaluate the difference in species richness between these

two types of stands. Different coefficient parameters for the diversity of potential seed

flux, rβ , were also included for each category of management.. Both intercept and

coefficient parameters were estimated in a hierarchical fashion, and were assigned a

common non-informative prior distribution:

~ (0,10000)

~ (0,10000)r

r

Normal

Normal

αβ

In order to evaluate the difference in species richness between selection-cutting stands and

unmanaged stands, we estimated within the model the difference, D, between the two

intercepts. Thus we obtained the posterior distribution of this difference, and the probability

that selection cutting resulted in higher species richness than no management, or p(D>0),

was then calculated as one minus the cumulative density of D up to zero (1-p(D<0). The

same process was done for the two coefficients for the diversity of potential seed flux. This

is a Bayesian form of a ‘test statistic’ from which the significance of differences due to

silviculture was calculated.

If enrichment after selection cutting with regard to unmanaged stands depends on the

availability of large and diverse seed pools, then the slope coefficient of diversity of

potential seed flux would be higher in selection-cutting than in unmanaged stands (see

Figure 2A). Conversely, if changes observed in species richness after selection-cutting are

independent of the diversity of seed flux, then this coefficient will be similar for both

managed and unmanaged stands (see Figure 2B).

We included in this step another covariable, FCC, to account for differences in successional

stage and biotic interactions, although this variable might reflect also differences in forest

types across the study area.

Given that spatial autocorrelation in residuals violates the assumption of independence and

invalidates inferences from a model (Legendre and Legendre, 1998), and that other factors

not considered here could underlie the observed patterns of species richness, we included

spatially explicit random effects in the model. We did not include them in previous steps

because the processes object of study, metacommunity dynamics, are inherently spatially

structured, and spatial effects could have accounted for this structure. For reasons of

computational limitations, plots were grouped in 8x8 km cells. Spatial effects for each of

these cells were assigned to a multivariate Gaussian distribution with covariance matrix

Page 127: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

113

expressed as a negative exponential function of the distance between cell centroides (e.g.

Diggle et al., 1998).

Figure 2. Assuming a positive effect of selection-cutting on species richness with respect to unmanaged

stands: A1. Increase in species richness due to higher diversity of potential seed flux only occurs when

new niche opportunities are created after disturbance; A2. Positive effects of selection-cutting are more

evident with enough diversity of potential seed flux, although the benefits of surrounding pool of species

is also significant in unmanaged stands; B1.The diversity of potential seed flux does not influence local

species richness; B2. Local species richness increases with increasing diversity of seed flux, but this

increase occurs in the same amount no matter whether the stands are managed or not.

We used Bayesian measures of explained variance (Gelman and Pardoe, 2006) to quantify

the explanatory power of the model at each step. In the case of the model in the third step,

we also estimated the explained variance when excluding the spatial term, in order to have

an approximate idea of the explained variance attributable exclusively to the analyzed

explanatory variables. We used the software OpenBUGs 3.1.2 for the statistical analyses

(Thomas et al., 2006). All variables (including residuals) except tree species richness were

standardized to zero means and unit variances. All parameters were given non-informative

priors, allowing the data to drive their estimation. Models were run for 150,000 iterations,

and convergence of the posterior means of the estimated parameters was assessed from two

chains. Pre-convergence “burn-in” iterations (10,000 in steps 1 and 2; 40,000 in step 3)

were discarded in the calculation of the parameters. Finally, we assessed the degree of

shade-tolerance of the main analyzed tree species as reported by Montero et al. (2003) and

Serrada et al. (2008), in order to evaluate if such species traits could have some role in the

Page 128: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

114

interaction between the gaps opened in the managed and unmanaged stands and the seed

fluxes potentially received in the sink plots.

3. RESULTS

The first two principal components extracted by the PCA accounted for 74% of the original

variability in environmental (climate and topography) conditions (Figure 3). PC1

represented a gradient of decreasing annual (spring, autumn and winter) water precipitation,

independently of temperature conditions (Figure 3). The second component, PC2,

represented a decreasing summer water deficit, associated to diminishing temperatures and

increasing precipitation levels in summer.

Figure 3. Eigenvectors for the first two principal components extracted by the PCA. Standard deviation

(i.e. squared root of the eigenvalues) and explained variance is shown in the parenthesis in the axis labels.

Ann_P, Spr_P, Sum_P, Aut_P and Wint_P: Annual, Spring, Summer, Autumn and Winter precipitation;

MinJanT: Minimum January Temperature; MxJulT: Maximum July Temperature; DSWD: Dry Season

Water Deficit; Elev: Elevation; Rad: Radiation; PET: Potential Evapotranspiration.

Spearman correlations between factors included in this study as potential drivers of species

richness reflected a relatively important covariation between them (Figure 4). The

environmental conditions, summarized by the first two components of the PCA, not only

were originally correlated with tree species richness, but also, in the case of the PC1,

showed correlations of r= -0.34 with forest cover (habitat amount and connectivity) and r= -

Page 129: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

115

0.83, with the diversity of seed flux. Forest cover was at the same time positively correlated

with diversity of seed flux (r=0.23), and had a slight negative correlation with tree species

richness. Forest cover in the study area was significantly higher in siliceous than in

calcareous soils (t-student=-10.8; p<0.001); when this interaction with lithology was

considered, Spearman correlation between forest cover and tree species richness was

positive (Figure 4).

Figure 4. Spearman correlation between variables considered in this study. The structure represents a top-

down hierarchy in which coarse patterns of physical environmental conditions control other processes at

smaller scales, which themselves influence local species richness. The arrows reflect the hypothesized

sense of the causality.

Model results in each step are described below:

- Step 1. The posterior mean estimate for the intercept in the first model was

significantly higher (p=1.0) in calcareous soils than in siliceous soils (Table 2). Tree

species richness had a quadratic relationship with the first axis of the PCA.

Increased scores in PC1 were associated to higher number of species, but after a

certain threshold (0.31 in the normal standard PC1 variable) the number of species

tended to decrease. The posterior mean for the PC2 slope coefficient was not

Page 130: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

116

significantly different from zero (95% CI: -0.04 – 0.06; Table 2). This model

explained 38.5% of the variance in tree species richness.

- Step 2. Forest cover was negatively associated (Table 2) with the residuals of step 1.

This model explained 1.8% of the variance of residuals from the previous step,

which corresponds to 1.1% of the total variance of tree species richness.

mean sd

Step1

Calcareous 1.61 0.02

Siliceous 1.38 0.02

PC1 -0.77 0.04

PC1^2 -0.82 0.05

PC2 0.01 0.03

Step2

Forest cover -0.14 0.03

Step3

FCC 0.14 0.02

Selection cutting -0.04 0.05

No management -0.06 0.05

Diversity of flux (SEL) 0.12 0.07

Diversity of flux (UN) 0.19 0.07

Table 2. Posterior mean estimates and their standard deviation (sd) for the parameters used in the

estimation of tree species richness, according to the models in each step. SEL: Selection-cutting stands;

UN: unmanaged stands.

- Step 3. The correlation of diversity of potential seed flux with tree species richness

was twice that of richness of potential seed flux (Spearman r =0.53 and 0.25,

respectively). For this reason, we selected diversity of potential seed flux as the

indicator of metacommunity processes explaining species richness patterns.

According to the posterior means for the intercepts, residuals from the model in step

2 were higher in selectively-cut than in unmanaged forest plots (p=0.86; Table 2).

Given that all variables were standardized to zero mean, this difference corresponds

to intermediate conditions of FCC and potential flux diversity. Conversely, the

slope coefficient for diversity of seed flux was stronger in unmanaged forests than

in those treated with selection cuttings (p=0.94; Table 2). Figure 5 shows the

posterior mean and credible interval for the predicted species richness when FCC is

57% (mean FCC in the study area) according to model in step 3. Results suggest

Page 131: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

117

that the difference in species richness in managed versus unmanaged plots is higher

at low levels of seed flux diversity, and becomes blurred at the highest levels of flux

diversity. This model explained 28.1% of the variance in the response variable, that

is, 16.9% of the total variance in species richness. However, when estimating the

same model without the spatial effects the explained variance was 7.3% (4.4% of

the variance in step 1). When the diversity of potential seed fluxes was low (first

quartile), a 89% of the total amount of seed fluxes corresponded to shade-intolerant

species, while this percentage decreased down to 45% when the seed fluxes were

more diverse (forth quartile).

Figure 5. Predicted tree species richness (standardized residuals from step 2) for the 95% central range of

the diversity of potential seed flux in managed (selection cuttings) and unmanaged forests, while

maintaining constant FCC (mean value of 57%). Parameters for the regression coefficients were

estimated in step 3 with both dependent and explanatory variables standardized, although in order to

facilitate interpretation, the original potential seed flux values are shown. Uncertainty on the parameter

estimates is represented by the 95% central interval of their posterior distributions.

4. DISCUSSION

4.1. Environmental gradients and tree species richness

Patterns of tree species richness in conifer forests of the Mediterranean North Region in

Spain showed a strong association with a gradient of annual precipitation. The maximum

levels of richness observed at intermediate precipitation levels concur with other analyses in

a similar study area (chapters 2, 3), and agree with the frequently observed hump-shaped

Page 132: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

118

relationship between productivity and plant species richness (Mittelbach et al., 2001), given

that in the Mediterranean region water availability is the main constraining factor of

productivity. But precipitation patterns in the study area also covaried with other considered

factors. Indeed, areas with lower precipitation levels, independently of temperature

conditions, tended to have a sparser forest cover, and less diverse and connected

surrounding tree communities, with non-linear relationships (Figure 4). This might be

explained by a higher ecosystem vulnerability to human disturbances under low

productivity conditions (Kondoh, 2001), or by the tendency of the cultivated lands to be

located in lowland flat areas within the Mediterranean North Region, which in our study

area have lower annual precipitation (Figure 3). Direct causality of the estimated unimodal

association between annual precipitation and species richness is therefore not

straightforward. It is most likely that the PC1 regression coefficient in the first step reflexes

a complex combination of all these factors, including a direct climatic effect on tree species

richness and an indirect effect through habitat loss, fragmentation or landscape diversity

patterns. In any case, assuming there is a top-down hierarchical structure, that is, that

climate context shapes other small-scale processes influencing species richness patterns,

direct and indirect effects of climate on tree species richness patterns were controlled in the

first step of the analysis.

4.2. Does a higher amount of landscape forest cover explain by itself an increased

diversity in local stands?

Contrarily to the commonly reported decrease in species richness with habitat loss (e.g.

Fahrig, 2003; Montoya et al., 2008; Montoya et al., 2010), an increase in surrounding forest

cover was negatively related to those patterns of local species richness not explained by

environmental patterns. In other words, under relatively homogenous climatic and

lithological conditions, and once accounted for the effects of forest cover indirectly driven

by climate, regions with more abundant and connected forest patches tended to have lower

local species richness. The expected positive effect on species richness might somehow be

captured in the estimated coefficients for PC1 in the first step, as supported by the initial

Spearman correlations (Figure 4), although this was not specifically tested in this study.

Some stands located in productive regions with abundant forest have relatively high

economic value. These stands could have lower species richness than predicted by

environmental conditions due to a higher intensity of human homogenization, favouring

monospecific stands. By contrast, forests in more arid locations with a poor productivity,

Page 133: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

119

when not replaced by other land uses, might remain closer to their natural composition as

expected in those environmental conditions. In any case, it should be noted that this

negative relationship accounted for a minor proportion of variance in tree species richness,

and that the other analysed factors were comparatively much more determinant of the stand-

scale diversity in the study area.

4.3. Landscape diversity of well-connected seed sources as a determinant of local

species richness

The present study adopted a conservative approach, as only the pure effect of diversity of

potential seed flux was assessed, excluding its possible covariation with environmental or

forest cover. Even so, we found that those localities where the potential flux of dispersed

seeds was more diverse tended to have more tree species. This result suggests that the

number of species in a local community is influenced by the degree of complexity of the

community network to which it is connected through long distance dispersal, independently

of its climatic and lithological context and the amount of surrounding forest.

The degree of connectivity through seed dispersal is not a general characteristic of plant

communities but instead should be differentiated for individual species, given differences in

dispersal ability among species (Ozinga et al., 2009). Nonetheless, intrinsic differences in

ecological traits for each of the forty nine species studied here were to some extent

incorporated in the estimation of potential seed flux by using Shannon diversity instead of

richness. Because of natural spatial autocorrelation of environmental conditions, the relative

abundance of a particular species in a region can be, at least in part, an indicator of its

competitive ability, that is, its degree of adaptation to the local environmental conditions in

the ‘sink’ forest stand. Following this reasoning, low potential flux for a particular species

is not necessarily associated to low probabilities of recruitment, but might indicate the

potential colonization by less frequent, inferior competitor species, that are however more

efficiently dispersed. For instance, in the study area, some species such as Pinus pinaster or

Pinus pinea reached values of potential seed flux about 35 to 40 times higher than the

maximum reached by many other species such as Acer monspessulanum. The higher the

number of species potentially colonizing a forest stand, the higher the probability that more

species have suitable functional traits to exploit the ecological niches in that stand, and this

was accounted by both richness and Shannon diversity metrics. However, Shannon diversity

also penalized the excessive seed flux of superior competitors, whereas recognized the

Page 134: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

120

importance of the contribution of species with lower potential seed flux in the absence of an

evident dominance of other species.

As in previous studies (chapters 2, 3), we found that tree species richness was higher in

conifer forests treated with selection cutting that in unmanaged conifer forests. This

difference occurred under intermediate conditions of FCC and diversity of potential seed

flux. We had hypothesized that the new niche opportunities created by the disturbance

would result in a greater enrichment of selectively-cut stands with regard to unmanaged

forests when the diversity of potentially colonizing species increased. Although we found

that the selectively-cut stands presented more species when the diversity of potential seed

fluxes from the surrounding communities was higher (Figure 5), the increase in local

species richness compared to unmanaged stands was not maintained along the regional

gradient of landscape diversity of seed sources; that is, we did not observe any of the

patterns illustrated in Figure 2. Contrarily to our hypothesis, the difference in species

richness between selectively-cut and unmanaged stands was lower in more diverse and

well-connected regions, and even became non-significant in those stands surrounded by the

most diverse communities (Figure 5). These results on the potential seed diversity and

species enrichment after disturbances are difficultly explained by colonization dynamics.

The establishment of additional tree species after disturbances is expected to be more

probable when the sampling pool is enlarged (Grime, 1998). We suggest that this

phenomenon could be in part attributable to site-specific characteristics of the treatments

applied, independent of the diversity of seed flux of the region, given that we are analysing

anthropogenic disturbances. Actually, selection-cutting treatments can include a set of

silvicultural operations additional to partial harvesting that may limit the number of

coexisting species. Differences in the response of species richness to selection cutting might

also be explained by a more evident statistical significance of recruitment of only one or

two new species in selection-cutting stands in those regions where many unmanaged stands

are mono-specific. Finally, and more importantly, we observed that under low flux diversity

conditions the potential seed fluxes were dominated by shade-intolerant species, while

when the surrounding communities were more diverse, this entailed a comparatively much

stronger representation of shade-tolerant species. These shade-tolerant species are present

only in a sufficient amount as seed sources in the more diverse landscape metacommunities.

Thus, they could find more appropriate conditions for its establishment in the stands not

affected by silvicultural treatments. In these stands only natural factors and tree mortality

not directly induced by humans (wind, pests, etc.) might create smaller and fewer gaps and

therefore sufficiently closed canopies. Although the selection cuttings are commonly

Page 135: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

121

regarded as less appropriate for the establishment of helophytes, at least when compared to

other treatments such as clearcutting or shelterwood cuttings, it should be noted that the

initial stand conditions before the application of the silvicultural treatments may already

correspond to relatively open stand with significant amounts of light penetration (average

FCC in the analysed stands is 57%), with selection cuttings further promoting light

exposure for potential seedlings.

The processes addressed in this study, metacommunity dynamics and local disturbances,

probably operate at different temporal scales. Metacommunity processes, that we showed

are responsible, at least in part, for the number of species found in a local community, have

probably operated at a historical or even evolutionary scale. At the spatial resolution used

here (1 km2), and the long distance dispersal analysed, local disturbance events are probably

linked to metacommunity dynamics in a much longer temporal scale than the approximate

period of ten years considered here. In such time scale, historical disturbance regime rather

than individual disturbances might have interacted with long distance dispersal events to

shape current metacommunities structures or landscape pool of species. This fits the

statement by the species pool hypothesis (Eriksson, 1993; Cornell and Lawton, 1992).

Moreover, given the initial correlations presented in Figure 4, diversity of seed flux could

be the main driver of local species richness, and the observed environmental effects in step

1 could be driven by common historical and evolutionary conditions of localities under

similar current climate conditions (Ricklefs, 2004). Actually, the species pool hypothesis

argues that past opportunity for origination of adapted species and hence, contemporary

resident species richness, should be generally highest in habitats with somewhat less than

intermediate productivity, the most common habitat productivity type over evolutionary

time (Schamp et al., 2002). After the main part of the species pool has developed, local

processes in ecological time such as disturbances can also modify local richness, as shown

here (Pärtel and Zobel, 1999). Colonization of the opened gaps might depend on the more

frequent, local dispersal events. In addition, seedlings of many common Mediterranean

species, such as Quercus spp., are frequently dispersed to conifer forests by jays (Garrulus

glandarius) but have a limited survival and growth under closed canopies, and therefore

rely on canopy gaps to establish successfully (e.g. Gómez, 2003).

4.4. Managing forests for enhancing plant diversity

Our findings have important implications for foresters, since direct changes in species

diversity derived from silvicultural practices might influence in the long term the diversity

Page 136: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CHAPTER 4

122

of other communities located some kilometres away. Lindborg and Eriksson (2004)

observed that historical rather than current connectivity explained current patterns of local

plant diversity in grasslands in Sweden. Actually, a time lag in the response of sessile, long-

lived tree communities surrounding the managed forest can be expected, and therefore these

consequences are difficult to capture within the temporal scope of most management plans.

Additionally, the targets of enhancing local biodiversity by management plans should also

recognize limitations emerging from the surrounding landscape. Indeed, our results show

that the diversity on the managed stands will not just depend on the treatments that may (or

not) be applied locally but in the diversity of the forest communities in a wider spatial

context, which may very well extend far beyond the limits of the public or private lands

being considered in a particular management plan. A significant issue to consider also when

predicting the response of a community to a particular silvicultural treatment is the degree

of shade tolerance of the species assemblage present in the neighbouring landscape. The

response of species richness to canopy opening will depend on the amount of species

adapted to the generated light conditions that are present in the landscape. This might result

in landscape-dependent responses of species richness to the same harvest intensity. The

interactive effect on tree species richness of local disturbances and diversity of potential

seed flux should be evaluated at shorter spatial scales, or in larger time lags. This would

provide a deeper understanding of the consequences of silviculture on plant species

richness, and the underlying mechanisms. In any case, we contend that precautionary forest

management for conserving and enhancing plant diversity requires a perspective much

beyond the spatio-temporal scope of most of the current forest management plans. This may

be better addressed through a wider landscape-scale forest management perspective (Perera

et al., 2006; Lafortezza et al., 2008; Saura, 2010) and, in the case of Spain, through the

district-level forest plans PORFs (Planes de Ordenación de los Recursos Forestales) that

have been established by the Spanish Forest law (passed in 2003) and are being increasingly

recognized and applied as a basic planning instrument to broaden the traditional

management scale focused on individual forest lands.

Page 137: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

Page 138: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 139: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

125

DISCUSSION

1. FACTORS EXPLAINING FOREST BIODIVERSITY

Selection of the most relevant explanatory variables in each model constructed

throughout this PhD thesis has been constrained in some cases by their high correlation

and by an advisable restriction in the number of variables considered as an input for the

models. For instance, elevation, temperature (annual or monthly), precipitation (annual

or seasonal), or the standard deviation of elevations were moderately or even highly

correlated in our study area, or in part of it (as in reduced spatial extents in

Geographically Weighted Regression models). In the first chapters we avoided potential

bias due to collinearity among explanatory variables (Graham, 2003) by limiting the

maximum Spearman correlation among them (chapters 1, 2). In this context, it should

be noted that the differences between chapters 1 and 2 in spatial and environmental

extents, and in spatial versus habitat definition of the units of analysis (grain)

determined some changes in the set of variables selected to represent energy and water

availability, topographic heterogeneity and altitudinal gradients. These categories of

variables have been commonly found to correlate with landscape patterns of plant

species richness (see the general introduction). However, it has been suggested that

within a reasonably well specified model and when adequate thought is given to the

ecological mechanisms behind modelled predictors, partial coefficients from multiple

regression models are useful measures of effect strength even for correlated predictors

(Smith et al., 2009). Therefore, in chapter 3 we included a wider set of more

mechanistic drivers of species richness (Austin, 2007); i.e., variables reflecting

particular ecophysiological processes such as summer stress (maximum July temperature)

or frosts (minimum January temperature). Finally, in the last chapter we improved this

somehow mechanistic exploration of environment-species richness relationships by

including variables that explicitly measure the interaction between energy and water (a

better representation of summer water deficit), seasonal variation in precipitation levels,

potential evapotranspiration as a measure of available environmental energy, and lithology.

Despite the heterogeneity in the assessed variables throughout this PhD, considerable

consistency was found in the relationships between species richness of woody plants and

Page 140: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

126

climate/productivity variables. Potential underlying mechanisms are discussed in next

section.

In general, our results support previous research on the drivers of species richness (see the

general introduction). At intermediate grain scales, we found that topographical complexity

was a large determinant of diversity patterns of woody plants in Spain (chapters 1, 2),

confirming the important role of environmental heterogeneity in the regional coexistence of

species (Gardner and Engelhardt, 2008). Also, landscape configuration indices were shown

to be useful indicators of biodiversity, since they correlate with the distribution of some

structural and compositional components of biodiversity (chapter 1). Nonetheless, many

problems arise when trying to generalize the relationships found in a given study area

between a particular landscape index and an ecological processes. According to Tischendorf

(2001), some of these problems derive from the thresholds, nonlinearity, ambiguity and

sensitivity to scale associated to the use of landscape indices. A similar situation occurs for

the association between elevation and species richness, which cannot easily extrapolated to

other areas. Altitudinal variation of plant species richness may be detected through the

associated temperature decline along altitudinal gradients at relatively small spatial extents

(chapter 2), but patterns are more difficult to interpret when working at extents covering

wider latitudinal, and therefore climatic, ranges (chapters 2, 3). As argued in chapter 3, the

inclusion of elevation was intended to account for as much environmentally or human

driven variation as possible, thus minimizing the effect of confounding factors not included

in the models. It is also noteworthy the important role that had in the models the variables

representing biotic interactions occurring at the stand scale, which have been less frequently

examined in the literature at the spatial scales here analyzed. Forest structure features such

as basal area or canopy cover are key to understand the distribution of woody plant species

richness in central Spain at the landscape scale (chapters 1, 2), and of stand-scale tree

species richness even when analyzing large spatial extents (chapters 3, 4).

2. WATER, ENERGY AND PRODUCTIVITY

The relative importance of climate/productivity as an explanatory variable of spatial

patterns in species richness increases in large-grain and large-extent studies (Field et al.,

2009). Actually, the role of climate/productivity can be remarkable even at smaller grain

sizes (e.g. below 10 km2) when larger spatial extents are covered (e.g. > 1000 km)

(Field et al., 2009). Our results showed a noteworthy association of climate/productivity

and woody plant species richness at intermediate grain sizes (25 km2 or forest habitat

Page 141: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

127

level; chapters 1, 2), or even local grain scales (2000 m2 plot size with resolution 1x1

km; chapters 3, 4), but relatively large spatial extents (500-1000 km). This is true even

when controlling for the variability due to forest types (chapter 1). The reason why climatic

factors related with water and energy availability are strongly associated with patterns of

species richness has been explained in terms of productivity (Currie and Paquin, 1987). In

the case of the Mediterranean region, where energy availability is not a constraining factor,

the increase in annual precipitation may be used as a surrogate of increased productivity

rates (Begon et al., 2006). Higher productivity is expected to result in coexistence of a

higher number of species by different theoretical mechanisms (Currie et al., 2004). But

depending on the spatial scale, the highest number of species is more frequently found at

intermediate rather than at maximum productivity rates (chapters 2, 3, 4). At intermediate

and local grain scales, as those here studied, biotic competition in highly productive sites

could explain richness reduction beyond a certain productivity threshold. At larger, regional

scales, variability in species richness due to species interactions is generally overridden, and

the theoretically expected positive linear association with species productivity is more

frequently observed (Sarr et al., 2005). But the rate of increase in productivity along a water

availability gradient depends on energy conditions. Because water deficit limits

regeneration and growth of many plant species, the same precipitation levels might result in

lower productivity rates under most extreme temperature conditions with respect to

moderate temperatures within the Mediterranean region. In the warmer, southern

Mediterranean (according to the regionalization by Metzger et al. (2005)), where maximum

annual precipitation, and therefore productivity, is lower than in other Mediterranean

regions, processes of biotic competition will occur above a precipitation threshold higher

than in more mesic regions (chapters 2, 3). Furthermore, facilitation rather than competition

interactions may prevail in such conditions (Callaway et al., 2002), resulting in a positive

linear association between productivity and species richness (chapter 3). By contrast, energy

availability might limit productivity in most cool, humid regions, such as Mediterranean

mountains. In these regions, productivity, and therefore, species richness, will increase

along with increasing temperature conditions (chapter 3). It follows that the variation in

coarse patterns of climate conditions will result in the non-stationarity of the relationships

between water-energy variables and species richness (chapters 2, 3). Careful interpretation

of climate-species richness relationships in large study areas, encompassing wide climatic

gradients should be done, since averaged trends may override potential mechanisms derived

from productivity gradients and biotic interactions. In the context of the predicted

environmental changes under future climate warming (Metzger et al., 2008), the

Page 142: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

128

dependence of tree species richness on precipitation patterns will probably be

accentuated, and forest ecosystems in the most xeric areas in the Mediterranean may be

more vulnerable to potential local extinctions of tree species. Although numerous

hypotheses theoretically explain the relationship between species richness and productivity

(Currie et al., 2004), other processes may underlie the relevant role of climate shaping

species richness patterns. From a top-down hierarchical framework, coarse climate patterns

affect many other processes that take place at intermediate or local scales (chapters 3, 4).

For instance, productivity conditions determine community responses to local disturbances

in terms of tree species richness (chapter 3). Additionally, landscape patterns of forest

amount and connectivity across a region are associated to the coarse variation in annual

precipitation, at least in central Spain (chapter 4). Finally, current climate might reflect

historical or even evolutionary conditions and the hump-shaped relationship between

productivity and species richness may be explained in terms of processes that have operated

at these temporal scales rather than at ecological time scales (Schamp et al., 2002; chapter

4).

3. PROCESS OF PLANT REGENERATION IN THE MEDITERRANEAN

We have focused the discussion of results obtained in this PhD on two important stages of

the process of plant regeneration: propagule dispersal and seedling recruitment. However,

successful plant recruitment depends on the synergistic action of several demographic

stages that are sequentially connected. The collapse of any of them can blur the spatial

patterns derived from previous stages (Herrera et al., 1994; Jordano et al., 2004). In the

Mediterranean, the main factors constraining establishment of plant species are summer

drought (e.g. Herrera et al., 1994; Rey and Alcántara, 2000; Castro et al., 2004; Gómez-

Aparicio et al., 2008; Mendoza et al., 2009; Rodríguez-García et al., 2011a) and a high

herbivore pressure on seeds and seedlings (e.g. Gómez et al., 2003; Herrera et al., 2004;

Matías et al., 2009). Because they alleviate drought stress for seedling establishment, shade

conditions may be crucial for the regeneration of many Mediterranean species (e.g. Castro

et al., 2004; Pausas et al., 2004; Puerta-Piñeiro et al., 2007; Gómez-Aparicio et al., 2008).

Nonetheless shade potentially leads to light limitation in later recruitment stages of seedling

growth (Marañón et al., 2004; Gómez-Aparicio et al., 2008). The temporal need of shade

(i.e. facilitation interaction among species) during the establishment process actually

depends on water availability of the site (Rodríguez-Calcerrada et al., 2010). In general, the

tolerance to different degrees of light or water deficit defines broadly the regeneration niche

Page 143: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

129

of many species in the Mediterranean (Marañón et al., 2004). All these considerations

highlight the determinant influence of canopy opening by silviculture treatments on

Mediterranean plant community dynamics (see next section).

The most commonly reported seed predators in Spain are the wild boar (Sus scrofa), for oak

seeds, and the wood mouse (Apodemus sylvaticus) (e.g. Gómez et al., 2003; Matías et al.,

2009), although these animals may also function as seed dispersers (e.g. Perea et al., 2010;

Perea et al., 2011a). Strong predation rates have been observed in many studies in Spain

(e.g. Herrera et al., 1994; Castro et al., 1999; Alcántara et al., 2000; Matías et al., 2009).

However, these rates are species and/or site specific (see García et al., 2005a and references

therein), which indeed might favour species richness when strong pressure exists on the

most dominant species (e.g. Orrock et al., 2003). Additionally, seed loss by predation may

have been overestimated since partially eaten seeds have been shown to contribute

significantly to natural regeneration (Perea et al., 2011b). Finally, postdispersal seed

predation may override patterns of seed dispersal (i.e. uncouple seed rain and recruitment)

only under seed limitation conditions (García et al., 2005a), and when both processes

(predation and dispersal) occur at the same spatial scale (García et al. 2005b). At least in the

case of the wood mouse, which has a relatively small home range, it is reasonable to expect

that the spatial patterns of predation become noise at the scale of study used here to analyze

dispersal processes (chapter 4). It should be mentioned as well the potentially significant

influence of wild and domestic ungulates on plant regeneration through trampling and

browsing of seedlings (Gómez et al., 2003; Castro et al., 2004; Zamora et al., 2004). The

inclusion of all these factors at the spatial extent used in this PhD thesis, is a hard task.

However, considering seed and seedling predation patterns will undoubtedly improve our

understanding of the mechanisms explaining post-disturbance plant community dynamics.

Recent studies have reported important effects of seed dispersal in forest plant communities

(e.g. Svenning and Skov, 2002; McEuen and Curran, 2004; Gonzalez et al., 2009). For

instance, it has been suggested that when a superior competitor species fails to establish in a

site due to dispersal limitation, the opportunity created for inferior competitors results in a

increased diversity. This mechanism is known as “winning by forfeit” (Hurtt and Pacala

1995). However, a net increase in species richness with higher connectivity levels has also

been found (e.g. Malanson and Armstrong, 1996; Gilbert et al., 1998; González et al., 1998;

Lindborg and Eriksson, 2004; Damschen et al., 2006). A balanced proportion of potential

seed flux of superior and inferior competitors colonizing a site might increase the

probability that the latter win some sites by forfeit. In such case, rather than the amount of

seed sources in the landscape, it seems that it is the diversity of well connected seed sources

Page 144: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

130

which will be most associated to increased local species richness (chapter 4). However, seed

dispersal is not only influenced by the quantity and quality (landscape structure) of the

surrounding habitat, as considered in this work, but depends also on the abundance and

behavior of dispersal vectors (Damschen et al., 2008). The role of animals in seed dispersal,

particularly in the connection of plant populations through long distance dispersal, is

especially relevant in the Mediterranean (Jordano et al., 2004). For instance, the abundance,

dynamics and spatial distribution of mammals such as the red fox (Vulpes vulpes), stone

marten (Martes foina), and wild boar (Sus scrofa) (e.g., Jordano et al., 2007; Matías et al.,

2010); migratory birds wintering in Spain (e.g., Santos et al., 1999; García and Ortiz-Pulido,

2004; Tellería et al., 2005); or frugivorous birds such as the European Jay (Garrulus

landarius) (e.g. Gómez, 2003; Purves et al., 2007) are critical for the successful long

distance dispersal of many plant species in Spain. Effects of forest loss and fragmentation

on the populations of these animals might have an indirect, negative impact on animal-

dispersed plant species (Bascompte and Jordano, 2007; Montoya et al., 2008), although this

was not observed for the total tree species richness in our study area (chapter 4). As in the

case of post-dispersal predation, animal-plant mutualistic interactions as dispersal are

crucial processes influencing plant community dynamics and they should be considered in

management plans for forest biodiversity.

4. SILVICULTURE, BIODIVERSITY AND IMPLICATIONS FOR MANAGEMENT

The intermediate disturbance hypothesis (Connell, 1978; Shea et al., 2004) assumes that

competitive exclusion in late-successional stages results in an impoverishment in the

number of species coexisting in a system (Figure 1). For example, in forest ecosystems as

canopy density increases, water, space, light and other resources are intercepted by

dominant plant species in the overstory (Aussenac, 2000; Serrada et al., 2008). Actually,

some studies have observed that, along successional stages, the maximum richness of plant

species in the understory occurs immediately before canopy closure, in both previously

clearcut stands (e.g. Aubert et al., 2003; Onaindia et al., 2004) and natural stands (Halpern

and Spies, 1995). At this stage, plants typical of early-successional stages can survive in

canopy gaps and coexist with some shade-tolerant species (Shea et al., 2004).

Mediterranean forests selectively cut in the last decade seem to fit this scheme, at least

under mesic climate conditions (chapters 3, 4).

Page 145: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

131

Figure 1. Intermediate Disturbance Hypothesis. In this hypothesis, the number of species in an ecosystem

is maximized if the intensity, frequency, or extent of ecological disturbances is moderate. At moderate

levels, competitive exclusion is controlled, heterogeneity is favoured, and a balance with the consequent

environmental stress is achieved. Although the frequency of the disturbance is higher in selection cutting

than in clearcutting treatments, the intensity and extent of the disturbance is lower. From a landscape

perspective, the extent of the disturbance also depends on the proportion of managed forest in the

landscape.

Because the intensity of a disturbance is generally assessed in terms of the response of a

species or a community to that disturbance, intermediacy of a disturbance is often defined in

terms of the conditions under which diversity is maximized (Shea et al., 2004).

Intermediacy in the extent, frequency or intensity of a disturbance is therefore defined by

the characteristics of the species in a community and of the system in question (see Shea et

al., 2004). Although intermediacy, within this PhD thesis, seems not to depend on life-

history traits of the dominant tree species in a forest, it does depend on coarse-scale climatic

conditions (chapter 3). Thus, while “intermediate-intensity” disturbances such as selection

cutting or improvement treatments benefit tree species coexistence under moderate

Mediterranean climate (chapters 3, 4), the frequency or intensity of these silvicultural

disturbances seems to be excessive under more xeric conditions in the South, at least for

tree species (chapter 3). Indeed, in ecosystems subjected to stressful conditions facilitative

interactions between plants may be more common than competitive interactions (Callaway

et al., 2002). Although under water deficit conditions competitive exclusion may still occur

in closed-canopy forests (chapter 3), the maximum number of coexisting tree species does

Page 146: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

132

not occur in selectively harvested stands in the Southern Mediterranean part of the study

area (chapter 3). Many species may require the protection of the overstorey for successful

recruitment and growth (e.g. Rousset and Lepart, 2000; Rodríguez-Calcerrada et al., 2010),

while being eventually displaced by tolerant, late-successional stages. It should be noted as

well that facilitation is frequently attributed to nurse shrubs (e.g. Gómez et al., 2003; Castro

et al., 2004; Plieninger et al., 2010; Rodríguez-García et al., 2011b), and the results

observed in chapter 3 might be explained in some cases by shrub removal operations, or by

the synergic combination of reducing both overstorey and understory facilitation of tree

seedlings. In any case, these findings call for caution in the management of Mediterranean

forests subjected to most extreme summer conditions.

Tree species are essential components of forest ecosystems, greatly influencing many of

their structural and functional features. Enhancing diversity in the overstory will promote

heterogeneity of light and litter conditions, vertical structure, etc., and the increased number

of microhabitats will positively influence diversity of many other organisms inhabiting the

forest (e.g. Vesseby et al., 2002; Gil-Tena et al., 2007; Kissling et al., 2008; Sobek et al.,

2009). Furthermore, complementarity in the types of functional response to fire

disturbances will also be promoted with increasing number of tree species in a forest. For

instance, faster growth and earlier seed production of some Mediterranean pines versus

efficient resprouting capacity of oaks are complementary mechanisms that will foster

compensation among species after fire disturbances (Pausas et al., 2004; Hooper et al.,

2005). Therefore, apart from helping to prevent propagation of forest fires through their

reduced biomass, stands selectively cut or subjected to improvement treatments may

promote ecosystem resilience to fire disturbances by maximizing local richness of tree

species, at least in the northern Mediterranean or in the Mediterranean mountains in central

Spain. This is especially relevant in the context of current rural abandonment and increased

fire hazard and propagation.

But, as discussed in the introduction, species richness and diversity are not the only

components of biodiversity. Other ecosystem processes that may or not be influenced by

species richness, such as carbon sequestration or biogeochemical cycles, and that take place

at different levels of biological organization, from genes to communities, can be designated

as indicators of forest ecosystem sustainability (Zavala et al., 2004). For instance, Aubert et

al. (2003) found that in managed temperate forests in France richness of plant species was

higher before canopy closure and declined in most mature stages. They argued that when

priority is given to compositional indicators of biodiversity (i.e., plant species richness),

management efforts should focus on the maintenance of system instability, or heterogeneity.

Page 147: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

133

Conversely, when priority is given to functional indicators of biodiversity (i.e. system

organization) emphasis should be made on maintaining stable mature communities.

Actually, different components of biodiversity might be driven by different factors, or may

respond differently to the same factor (chapter 1). Because management strategies

prioritizing different aspects of forest sustainability might not be easily reconciled, a trade-

off between all or part of them needs to be chosen (Bodin and Wiman, 2007). In fact,

despite a generalized benefit of management for vascular plant species richness (see Paillet

et al., 2010a), mature, unmanaged stands are essential for the maintenance of species

diversity of many other organisms. Particularly, structural features of old-growth forests are

required by species with slow growth, limited dispersal, and dependence on substrates as

dead wood at different decay stages, or on moist microhabitats, among others (Bengtsson et

al., 2000). This explains the decline of species richness of bryophytes, lichens and

saproxylic fungi in managed compared to unmanaged forests; see Paillet et al. (2010a, b)

and Halme et al. (2010) for a review on the effects of forest management on species

richness of different taxonomic groups.

In the disturbance-succession cycle of forest communities, pioneer tree species are

gradually replaced by late-successional tree species. This successional gradient is

continuously fed by the exchange of colonists among neighbouring communities at different

stages of the cycle or in different habitats (metacommunity dynamics; Leibold et al., 2004).

The colonization of tree species adapted to the successive new conditions in the community

will depend on the presence of neighbouring communities in similar maturation stages, or

inhabiting similar habitat conditions. An extensive application of selection cutting in a

landscape may result in the homogenization of the successional stages present in a

landscape, hindering the persistence of more demanding shade-tolerant tree species. High

percentages of selection cutting (or clearcutting) forest in the landscape can thereby reduce

the total number of tree species present in the landscape (chapter 2; Figure 1). The reduction

in the described exchange of species among communities at different stages will eventually

reduce also species richness of trees in each local community present in the landscape

(chapter 4). This might explain the decline in tree species richness of forest stands

surrounded by forest habitats poor in tree species, such as dehesas or plantations, in central

Spain; or the increase with the presence of riparian or unmanaged forests in the landscape

(chapter 3). These considerations lead us to acknowledge the necessity and importance of

widening the spatio-temporal scope when evaluating the consequences of any silvicultural

operation on local species richness. If we are to protect and enhance species richness in a

forest stand we must have an integral view of the whole span of successional stages in that

Page 148: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

134

forest and of the neighbouring communities upon which persistence of some species may

depend in the long term. Even when forests subjected to partial harvest seem to foster

richness of woody plant species in some Mediterranean areas, the preservation of

unmanaged forests intermingled with them in the landscape is of vital importance.

Unmanaged forests will assure persistence of species dwelling in old-growth forests, while

providing sources for recolonization of disturbed sites along the succession cycle.

Additionally, an appropriate infrastructure of dispersal vectors needs to be assured, for

instance, by avoiding excessive forest cleaning, since dispersing birds will be attracted by

the abundance of fruiting shrubs (Tellería et al., 2005; García et al., 2010); or by the

promotion of forest landscape functional connectivity for mammals’ populations (e.g.

Minor and Lookingbill, 2010).

The time-scale at which the local extinction/colonization trade-off takes place depends in

part on the distance between the implicated communities in a forest stand, and the dispersal

capacity of species composing them. Although we observed a decline of the number of tree

species in a managed stand when increasing the proportion of managed forests around it

(radius of 5.5 km; chapter 3), we can not clearly confirm that this is explained by a lower

diversity in the neighbourhood of species potentially colonizing the disturbed sites (median

and maximum dispersal distance of 2.6 and 17 km, respectively) (chapter 4). Because long-

distance dispersal events are rare compared to local dispersal events, dispersal rates at the

studied spatial scale may not be large enough to influence significantly the dynamics of a

tree community within ten years after disturbance. Now, an important question is whether

we are addressing an isolated silvicultural disturbance in a particular site or a historical

sequence of periodical silvicultural disturbances. Certainly, many of the sites that are

currently harvested have a historical forestry planning, and have therefore been harvested

for decades, or even for more than a century in some cases (e.g. Bravo et al., 2010).

However, the use of regeneration systems such as selection cuttings has become more

widespread in Spain relatively recently. This is explained by the growing demand for less

intense silvicultural treatments that better conciliate production with other environmental

services such as biodiversity conservation, physical protection, recreation or landscape

aesthetics. Therefore, many of these forests which have been historically managed and are

currently selectively cut had been previously subjected to more intensive silviculture

systems such as clearcutting or shelterwood cutting. The potential long-lasting effects of

past management (e.g. Vandekerkhove et al., 2009) together with the possible time lags in

community responses to landscape conditions (Lindborg and Eriksson, 2004) hinders the

identification of a potential interaction between current disturbances and colonization

Page 149: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

DISCUSSION

135

events from surrounding communities. In fact, stands historically clearcut may present a

relatively important proportion of light-demanding tree species, while not enough time has

passed for the colonization and fully effective establishment of more tolerant species

potentially colonizing the smaller gaps and conditions created after selection cutting

(chapter 4). Results of this PhD evidence an association between the number of tree species

in a forest stand and the diversity and connectivity of species potentially colonizing it

through long distance dispersal. The time scale at which processes underlying this

association occur is not specifically evaluated within this PhD thesis. However,

management plans should keep in mind the potential future impact of silviculture treatments

beyond the current spatial scope of these operations, and vice versa.

Page 150: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 151: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CONCLUSIONS

Page 152: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 153: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CONCLUSIONS

139

CONCLUSIONS 1. The relative effect of different (environmental, landscape structure, biotic)

explanatory factors varied considerably for different forest biodiversity indicators in Spanish habitats. In general, compositional indicators (richness and diversity of woody plants) were better explained by environmental factors than structural indicators (percentage of uneven-aged stands and snag abundance).

2. Climate factors were strongly associated with patterns of woody plant species

richness at intermediate grain sizes (25 km2 or forest habitat level), or even at local grain scales (ca. 2000 m2 plot size) when relatively large spatial extents (500-1000 km) were considered. The highest species richness of trees occurred at intermediate productivity levels. In the Mediterranean, productivity was assumed to depend on water availability in most xeric areas, and on energy availability in mountains.

3. In central Spain, a higher woody plant species richness in the landscape (25 km2)

was observed with an increased proportion of forests managed through selection cuttings or improvement treatments, up to a certain limit of about 40-60% beyond which the positive effect of such management operations was reversed.

4. There was a decline in the richness of woody plant species in the landscape (25

km2) with increasing proportions of clearcut forests. Across environmental gradients and forest functional types (coniferous, broadleaved deciduous or sclerophyllous evergreen) in central Spain, clearcut forest stands were poorer in tree species than unmanaged stands.

5. The relative importance of many environmental, biotic or management factors and

the form of their relationships with local and landscape species richness of trees or shrubs varied spatially across an area of approximately 150,000 km2 in central Spain.

6. Our results support a top-down hierarchical structure in which broad, regional

climatic patterns condition ecological processes influencing local tree species richness and operating at intermediate or local spatial scales (e.g. the response to silvicultural disturbances).

7. Forest stands selectively cut had, in general, more tree species than unmanaged

stands under similar conditions in mesic Mediterranean environments. However, this benefit was not evident in most xeric Mediterranean regions, suggesting that canopy facilitation, rather than competition, was more important for allowing the regeneration of a diversity of species in those areas.

8. Tree species richness in a forest stand was associated with the proportion of forests

with different management intensities in the surrounding landscape. For instance, tree species richness declined with an increasing proportion of dehesas, plantations or clearcut forests in the neighbouring lands. On the contrary, stands in landscapes with larger amounts of unmanaged forests or riparian forests tended to have more tree species.

Page 154: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CONCLUSIONS

140

9. Tree species richness in a locality increased along a gradient of increasing diversity and connectivity of potential seed fluxes from the surrounding landscape.

10. Our results show that woody plant species richness in a managed forest will not just depend on the treatments that may (or not) be applied locally, as accounted for in a typical management plan for an individual forest land. Species richness also depends on relevant spatiotemporal interactions with other forest communities located some kilometres away. This advocates for a broader landscape-scale forest management perspective to sustain and enhance biodiversity and, therefore, the resilience of forests in an increasingly changing environmental and social context.

Page 155: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CONCLUSIONES

141

CONCLUSIONES

1. El efecto relativo de diferentes factores explicativos (ambientales, de estructura de paisaje, bióticos) varió considerablemente entre los distintos indicadores de biodiversidad forestal analizados en los hábitats forestales españoles. En general, los indicadores de composición (riqueza y diversidad de plantas leñosas) fueron mejor explicados por factores ambientales que los indicadores estructurales (porcentaje de rodales con estructura irregular y abundancia de árboles muertos en pie).

2. Los factores climáticos estuvieron fuertemente asociados a los patrones de riqueza

de especies de leñosas a tamaños de grano intermedios (25 km2 o a nivel de hábitat forestal), o incluso locales (tamaño de parcela de aproximadamente 2000 m2) cuando se consideraron extensiones espaciales relativamente grandes (500-1000 km). La mayor riqueza en especies arbóreas se dio en niveles intermedios de productividad; se asumió que la región mediterránea la productividad depende de la disponibilidad de agua en las zonas más xéricas, y de la disponibilidad energética en las montañas.

3. Se observó una mayor riqueza de especies de plantas leñosas en los paisajes (25

km2) del centro de España al aumentar la proporción de paisajes gestionados por entresaca o tratamientos de mejora hasta un umbral de aproximadamente 40-60%.

Porcentajes mayores de dichas prácticas de gestión en el paisaje tuvieron un efecto contrario.

4. La riqueza de especies de plantas leñosas en el paisaje (25 km2) disminuyó al

aumentar la proporción de bosques gestionados con cortas a hecho. Los rodales cortados a hecho fueron más pobres en especies arbóreas que los no gestionados para los distintos gradientes ambientales y grupos funcionales (bosques de coníferas, frondosas caducifolias o perennes esclerófilas) del centro de España.

5. La importancia relativa de muchos factores ambientales, bióticos o de gestión y el

tipo de asociación con la riqueza de especies arbóreas o arbustivas a escala local o e paisaje varió espacialmente a lo largo de unos aproximadamente 150,000 km2 en el centro de España.

6. Nuestros resultados apoyan una estructura jerárquica según la cual los patrones

climáticos regionales condicionan los procesos ecológicos que afectan a la riqueza local de especies arbóreas y que operan a escalas espaciales intermedias o locales (ej. la respuesta a perturbaciones selvícolas).

7. Los rodales entresacados tuvieron en general más especies arbóreas que los rodales

sin gestionar en condiciones similares en ambientes mésicos mediterráneos. Sin embargo, dicho beneficio no se evidenció en las regiones mediterráneas más xéricas. Este hecho sugiere que la facilitación por parte del dosel arbóreo, y no la competencia, es más determinante para permitir la regeneración de una diversidad de especies en ambientes xéricos.

8. La riqueza de especies arbóreas en un rodal estuvo asociada con la proporción de

bosques sometidos a gestión de distinta intensidad en el paisaje del entorno. Por ejemplo, la riqueza de especies arbóreas disminuyó al aumentar la proporción de dehesas, plantaciones o bosques cortados a hecho en las zonas cercanas. Por el

Page 156: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

CONCLUSIONES

142

contrario, en paisajes con mayor proporción de bosques sin gestionar o de ribera la riqueza de árboles en los rodales tendió a aumentar.

9. La riqueza de especies arbóreas en el rodal fue más elevada cuanto mayor fue la

diversidad y conectividad del flujo potencial de semillas procedente del paisaje circundante.

10. Nuestros resultados muestran que la riqueza de especies leñosas en un bosque

gestionado no depende únicamente del tipo de tratamientos que se puedan (o no) aplicar localmente, tal y como se contempla en los proyectos de ordenación o planes de gestión tradicionales para un monte individual. Dicha riqueza también depende de las interacciones espaciotemporales que se establecen con otras comunidades boscosas que pueden estar situadas a varios kilómetros de distancia. Ello sugiere la necesidad de ampliar el alcance de la gestión forestal hasta la escala de paisaje para propiciar una mejor conservación y fomento de la biodiversidad forestal, y por tanto, de la resiliencia de los bosques en un contexto ambiental y social cambiante.

Page 157: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

ACKNOWLEDGEMENTS/AGRADECIMIENTOS

143

ACKNOWLEDGEMENTS

Agradecimientos

Este trabajo ha sido financiado por el Ministerio de Ciencia e Innovación a través de los

proyectos IBEPFOR (CGL2006-00312/BOS), MONTES CONSOLIDER (CSD2008-

00040), Restauración y Gestión Forestal (PSE-310000-2009), y DECOFOR (AGL2009-

07140/FOR). Los datos del Tercer Inventario Forestal Nacional han sido proporcionados

por el Ministerio de Medio Ambiente y Medio Rural y Marino. Mi estancia en la

Universidad de Michigan fue en parte financiada por una bolsa de viaje de la Universidad

de Lleida.

Para agradecer a la gente que considero ha formado parte de esta tesis de alguna manera

voy a robarle la idea a un buen amigo de seguir un orden cronológico.

En primer lugar a mi padre, por transmitirme desde niña sus ganas de conocer y por su

tenacidad en asegurar una buena educación de sus hijos. Y a mi madre, por ser una Madre

con mayúsculas y apoyarme en todos y cada uno de los momentos y decisiones desde el

inicio de esta etapa.

A la gente del Instituto Pirenaico de Ecología (IPE). Si bien ésta es una etapa anterior al

inicio de esta tesis, la última no se puede entender sin la primera. Con ellos descubrí la

Ciencia, en el despacho, en el laboratorio, en las reuniones, en los cafés de la Cadiera, en las

excursiones por el monte (recuerdo aquella visita de un amigo que me dijo: “me siento

como en una excursión de la carrera”), en Monegros, en las cenas entre amigos.

A César Pedrocchi, por darme la oportunidad de empezar en la investigación e introducirme

en el mundo Monegrino (Mordor), que tanto cariño llegué a coger. A Juan (Juanillo) y

Guillermo Sanz, por adoptarme en mi candidez y enseñarme, no solo ciencia sino muchas

cosas de la vida. Al Brown (David Moreno), por orientarme en aquel mar de incertidumbre,

e incluso desesperación en algunos momentos, que supuso para mi el enfrentarme a mi

primer artículo; por su perseverancia en revisar una y otra vez cada versión. A los

investigadores, técnicos y administrativos del IPE, aunque no os nombre a todos, formasteis

una parte importante de aquella etapa.

A Guille, por su conversación siempre interesante, y por ser un ejemplo de compañerismo y

de honestidad; a él y a la Isabelita (mis “Curie”) por transmitirme su pasión por la ciencia, y

Page 158: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

ACKNOWLEDGEMENTS/AGRADECIMIENTOS

144

a Isa: por su alegría; a Aranchilla, por enseñarme su filosofía de vida; de Maite aprendí su

gran espíritu crítico, a las dos: por instruirme en el mundo de la montaña; a Gonzalo, por su

naturalidad, “¡aupa con la chavalería!”; y a Felipe, contigo aprendí muchas cosas, por

mencionar solo una, las lecciones de botánica en nuestras numerosas e inolvidables

excursiones. También a Marcos, Edel, Olatz, Alex, Esther, Jose, Nachete, Luismi, Sara.

Más allá de la ciencia, a todos los “jaqueses” por su extraordinaria calidad humana y su

amistad.

A Santiago Saura, por respetar mi trabajo y mis aportaciones; esto que parece obvio, pero

que una descubre no es tan frecuente al escuchar a otros precarios. Por su capacidad de

entender rápidamente lo que a mi me ha llevado mucho tiempo, desenredar la complejidad

con soltura y proporcionarme en cada intercambio de ideas nuevos caminos.

A Assu, por abrirme sin reparos la caja de conocimientos laboriosamente acumulados a lo

largo de su tesis y explicármelos con paciencia. Y por su ayuda en el cálculo de algunas de

las variables utilizadas en esta tesis.

A Pilar Martín de Agar, por iniciarme en la disciplina de la ecología del paisaje, y por sus

valiosas aportaciones en mi DEA.

A Cristina Vega, por acogerme incondicionalmente en “el lab”, por su calidad humana y

por preocuparse sinceramente por nuestro bienestar.

A todos los que han pasado de manera temporal por el lab, pero que sin duda proporcionado

una grata compañía y han dejado huella: Paco, Evelina, Cheikh, Haifa y Sara, Paula, Esra y

Marc. A Marconni (Padilla), se te ha echado mucho de menos en la última etapa de la tesis,

no es fácil encontrar compañeros de trabajo y amigos como tú. A Sergi, por tus payasadas,

por las bromas y los buenos momentos que hemos compartido. A Lidón, porque si pudiera

volver a atrás y elegir la persona con la que compartiría más horas a lo largo de esta tesis sin

duda volverías a ser tú. Por las risas que nos hemos echado.

A los otros doctorandos, químicos y cerdólogos (me lo acabo de inventar, sé que no os

gustará :p): Sandrine, Rebeca, Roger, Nora, Harry, Sol, Carmen, Edinson, Olalla, Toni, por

esas comidas en la cafetería, que muchas veces fueron el mejor momento del día; por

vuestra complicidad, amistad y los buenos momentos vividos fuera del campus.

A Alex Escolà, por dar un breve toque de alegría a nuestra rutina diaria con solo venir a

prepararse un té.

Gracias también a Sandra, una pena no coincidir más. Recuerdo con mucho cariño el

congreso de Bragança, ¡a pesar de mi jetlag!

Page 159: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

ACKNOWLEDGEMENTS/AGRADECIMIENTOS

145

Quiero agradecer también a Inés Ibáñez por Acogerme, con mayúsculas, en su laboratorio y

hacerme sentir una más en el grupo de investigación. Por esos brillantes y espontáneos

despliegues de flechas y letras griegas en la pizarra que a pesar de desconcertarme en un

principio, me aportaron mucha claridad.

I would like also to thank Jeff Diez. I certainly learned a lot with you. You patiently helped

me to disentangle the ideas and questions I had in my head, and taught me to seriously

question whether my model was exactly answering the particular ecological hypothesis I

wanted to test. This may seem obvious, but when dealing with varying slopes and varying

intercepts that come from prior distributions that… Thank you also for your nice and

clarifying drawings. And finally, thank you for making lunch time such a pleasant moment.

I am very grateful too to all people in Dana building, Ann Arbor; I miss those happy hours!

And especially to Kyle and Sarah, for their daily happiness and friendship.

Page 160: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 161: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

Page 162: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 163: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

149

REFERENCES

Akatov, V.; Chefranov, S.; Akatova, T. 2005. The relationship between local species richness

and species pool: A case study from the high mountains of the Greater Caucasus. Plant

Ecology 181: 9-22.

Alados, C.L.; Navarro, T.; Komac, B.; Pascual, V.; Rietkerk, M. 2010. Dispersal abilities and

spatial patterns in fragmented landscapes. Biological Journal of the Linnean Society

100: 935-947.

Alberdi, I.; Saura, S.; Martínez-Millán, F.J. 2005. El estudio de la biodiversidad en el tercer

inventario forestal nacional. Actas de la I reunión de inventario y teledetección forestal.

Cuadernos de la Sociedad Española de Ciencias Forestales 19: 11-19.

Albert, D.A. and Barnes, B.V. 1987. Effects of clearcutting on the vegetation and soil of a

sugar maple-dominated ecosystem, western upper Michigan. Forest Ecology and

Management 18: 283-298.

Alcántara, J.M.; Rey, P.J.; Sánchez-Lafuente, A.M.; Valera, F. 2000. Early effects of rodent

post-dispersal seed predation on the outcome of the plant-seed disperser interaction.

Oikos 88: 362-370.

Allué-Andrade, J.L. 1990. Atlas fitoclimático de España. Taxonomías. Ministerio de

Agricultura, Pesca y Alimentación, Instituto Nacional de Investigaciones Agrarias,

Madrid, Spain.

Andrews, P. and O’Brien, E.M. 2000. Climate, vegetation, and predictable gradients in

mammal species richness in southern Africa. Journal of Zoology 251: 205-231.

Aparicio, A.; Albaladejo, R.G.; Olalla-Tárraga, M.Á.; Carrillo, L.F.; Rodríguez, M.Á. 2008.

Dispersal potentials determine responses of woody plant species richness to

environmental factors in fragmented Mediterranean landscapes. Forest Ecology and

Management, 255: 2894-2906.

Arrieta, S. and Suárez, F. 2005. Spatial patterns of seedling emergence and survival as a

critical phase in holly (Ilex aquifolium L.) woodland recruitment in Central Spain.

Forest Ecology and Management 205: 267–282.

Arroyo, J.; Carrión, J.; Hampe, A.; Jordano, P. 2004. La distribución de las especies a

diferentes escalas espacio-temporales. In: Valladares, F. (Ed.), Ecología del bosque

mediterráneo en un mundo cambiante. Ministerio de Medio Ambiente, EGRAF SA,

Madrid, pp. 27-67.

Page 164: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

150

Atauri, J.A.; de Pablo, C.L.; de Agar, P.M.; Schmitz, M.F.; Pineda, F.D. 2004. Effects of

management on understory diversity in the forest ecosystems of Northern Spain.

Environmental Management 34: 819-828.

Atlegrim, O. and Sjöberg, K. 2004. Selective felling as a potential tool for maintaining

biodiversity in managed forests. Biodiversity and Conservation 13: 1123-1133.

Aubert, M.; Alard, D.; Bureau, F. 2003. Diversity of plant assemblages in managed temperate

forests: A case study in Normandy (France). Forest Ecology and Management 175:

321-337.

Auerbach, M. and Shmida, A. 1987. Spatial scale and the determinants of plant species

richness. Trends in Ecology and Evolution 2: 238-242.

Aunós, A. 2005. Configuración y potencialidades de los bosquetes en las estructuras

irregulares. Actas del IV Congreso Forestal Español, Zaragoza.

Aussenac, G. 2000. Interactions between forest stands and microclimate: Ecophysiological

aspects and consequences for silviculture. Annals of Forest Science 57: 287-301.

Austin, M.P. 2007. Species distribution models and ecology theory: A critical assessment and

some possible new approaches. Ecological Modelling 200: 1-19.

Austin, M.P.; Cunningham, R.B.; Fleming, P.M. 1984. New approaches to direct gradient

analysis using environmental scalars and statistical curve-fitting procedures. Vegetatió

55: 11-27.

Baessler, C. and Klotz, S. 2006. Effects of changes in agricultural land-use on landscape

structure and arable weed vegetation over the last 50 years. Agriculture, Ecosystems

and Environment 115: 43-50.

Bascompte, J. and Rodríguez, M.Á. 2001. Habitat patchiness and plant species richness.

Ecology Letters 4: 417-420.

Bascompte, J. and Jordano, P. 2007. Plant-animal mutualistic networks: The architecture of

biodiversity. Annual Reviews of Ecology and Systematic 38: 567-593.

Bauer-Manderscheid, E. 1991. Los montes de España en la Historia. Fundación Conde del

Valle de Salazar y Ministerio de Agricultura, Pesca y Alimentación, Madrid.

Begon, M.; Harper, J.L.; Townsend, C.R. 2006. Ecología. Individuos, poblaciones y

comunidades. Omega, Barcelona.

Bengtsson, J.; Nilsson, S.G.; Franc, A.; Menozzi, P. 2000. Biodiversity, disturbances,

ecosystem function and management of European forests. Forest Ecology and

Management 132: 39-50.

Page 165: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

151

Benito Garzón, M.; Sánchez de Dios, R.; Sáinz-Ollero, H. 2007. Predictive modelling of tree

species distributions on the Iberian Peninsula during the Last Glacial Maximum and

Mid Holocene. Ecography 30: 120-134.

Benjamini, Y. and Yekutieli, D. 2001. The control of the False Discovery Rate in Multiple

Testing under dependency. The Annals of Statistics 29: 1165-1188.

Blondel, J. 2006. The 'design' of Mediterranean landscapes: A millennial story of humans and

ecological systems during the historic period. Human Ecology 34: 713-729.

Blondel, J. and Aronson, J. 1999. Biology and wildlife of the Mediterranean region. Oxford

University Press, Oxford.

Bock, M.D. and Van Rees, C.J. 2002. Mechanical site preparation impacts on soil properties

and vegetation communities in the Northwest Territories. Canadian Journal of Forest

Research 32: 1381-1392.

Bodin, P. and Wiman, B.L.B. 2007. The usefulness of stability concepts in forest management

when coping with increasing climate uncertainties. Forest Ecology and Management

242: 541-552.

Brashears, M.B.; Fajvan, M.A.; Schuler, T.M. 2004. An assessment of canopy stratification

and tree species diversity following clearcutting in Central Appalachian hardwoods.

Forest Science 50(1): 54-64.

Bravo, F.; Del Río, M.; Del Peso, C. (Eds.). 2002. El Inventario Forestal Nacional. Elemento

clave para la Gestión Forestal Sostenible. Fundación general de la Universidad de

Valladolid, Valladolid.

Bravo, F.; Osorio, L.F.; Pando, V.; Del Peso, C. 2010. Long-term implications of traditional

forest regulation methods applied to Maritime pine (Pinus pinaster Ait.) forests in

central Spain: a century of management plans. iforest-Biogeosciences and Forestry 3:

33-38.

Briggs, J.M.; Spielmann, K.A.; Schaafsma, H.; Kintigh, K.W.; Kruse, M.; Morehouse, K.;

Schollmeyer, K. 2006. Why ecology needs archaeologists and archaeology needs

ecologists. Frontiers in Ecology and the Environment 4: 180-188.

Brokaw, N. and Lent, R. 1999. Vertical structure. In: Hunter, M.L. (Ed.), Maintaining

biodiversity in forest ecosystems. Cambridge University Press, Cambridge, pp. 373-

399.

Bunn, A.G.; Urban, D.L.; Keitt, T.H. 2000. Landscape connectivity: A conservation

application of graph theory. Journal of Environmental Management 59: 265-278.

Page 166: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

152

Burke, D.M.; Elliott, K.A.; Holmes, S.B.; Bradley, D. 2008. The effects of partial harvest on

the understory vegetation of southern Ontario woodlands. Forest Ecology and

Management 255: 2204-2212.

Burnham, K.P. and Anderson, D.R. 2002. Model Selection and Multimodel Inference: A

Practical Information-Theoretical Approach. Springer-Verlag, New York.

Cain, M.L.; Milligan, B.G.; Strand, A.E. 2000. Long-distance seed dispersal in plant

populations. American Journal of Botany 87: 1217-1227.

Callaway, R.M.; Brooker, R.W.; Choler, P.; Kikvidze, Z.; Lortie, C.J.; Michalet, R.; Paolini,

L.; Pugnaire, F.I.; Newingham, B.; Aschehoug, E.T.; Armas, C.; Kikodze, D.; Cook,

B.J. 2002. Positive interactions among alpine plants increase with stress. Nature 417:

844-848.

Castro, J.; Gómez, J.M.; García, D.; Zamora, R.; Hódar, J.A. 1999. Seed predation and

dispersal in relict Scots pine forests in southern Spain. Plant Ecology 145: 115-123.

Castro, J.; Zamora, R.; Hodar, J.A.; Gómez, J.M. 2004. Seedling establishment of a boreal

tree species (Pinus sylvestris) at its southernmost distribution limit: Consequences of

being in a marginal Mediterranean habitat. Journal of Ecology 92: 266-277.

Castro-Parga, I.; Saiz, J.C.M.; Humphies, C.J.; Williams, P.H. 1996. Strengthening the natural

and national park system of Iberia to conserve vascular plants. Botanical Journal of the

Linnean Society 121(3): 189-205.

Charlton, M.; Fotheringham, S.; Brunsdon, C. 2003. GWR 3. Software for geographically

weighted regression. Spatial analysis research group, Department of Geography,

University of Newcastle upon Tyne, UK.

Chase, J.M. and Leibold, M.A. 2003. Ecological niches: linking classical and contemporary

approaches. University of Chicago Press, Chicago.

Chesson, P.L. 1983. Coexistence of competitors in a stochastic environment: The storage

effect. In: Freedman, H.I. and Strobeck, C. (Eds.), Population biology. Lecture notes in

Biomathematics 52. Springer, New York, pp. 188-1988.

Christensen, N.L.; Bartuska, A.M.; Brown, J.H.; Carpenter, S.; D'Antonio, C.; Francis, R.;

Franklin, J.F.; MacMahon, J.A.; Noss, R.F.; Parsons, D.J.; Peterson, C.H.; Turner,

M.G.; Woodmansee, R.G. 1996. The report of the Ecological Society of America

committee on the scientific basis for ecosystem management. Ecological Applications

6: 665-691.

Page 167: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

153

Clark, J.S. 2005. Why environmental scientists are becoming Bayesians. Ecology Letters 8: 2-

14.

Clark, J.S.; Silman, M.; Kern, R.; Macklin, E.; HilleRisLambers, J. 1999. Seed dispersal near

and far: Patterns across temperate and tropical forests. Ecology 80: 1475-1494.

Connell, J.H. 1978. Diversity in tropical forests and coral reefs. Science 199: 1302-1309.

Cornell, H.V. and Lawton, J.H. 1992. Species interactions, local and regional processes, and

limits to the richness of ecological communities: A theoretical perspective. Journal of

Animal Ecology 61: 1-12.

COST E43. 2004-2008. Harmonization of national forest inventories in Europe: techniques

for common reporting [on line]. Available in http://www.metla.fi/eu/cost/e43/ [last

access 12th Sep. 2011].

Crow, T.R. and Perera, A.H. 2004. Emulating natural landscape disturbance in forest

management - An introduction. Landscape Ecology 19: 231-233.

Currie, D.J. and Paquin, V. 1987. Large-scale biogeographical patterns of species richness of

trees. Nature 329: 326-327.

Currie, D.J.; Mittelbach, G.G.; Cornell, H.V.; Field, R.; Guégan, J.F.; Hawkins, B.A.;

Kaufman, D.M.; Kerr, J.T.; Oberdorff, T.; O'Brien, E.M.; Turner, J.R.G. 2004.

Predictions and tests of climate-based hypotheses of broad-scale variation in taxonomic

richness. Ecology Letters 7: 1121-1134.

Damschen, E.I.; Haddad, N.M.; Orrock, J.L.; Tewksbury, J.J.; Levey, D.J. 2006. Corridors

increase plant species richness at large scales. Science 313: 1284-1286.

Damschen, E.I.; Brudvig, L.A.; Haddad, N.M.; Levey, D.J.; Orrock, J.L.; Tewksbury, J.J.

2008. The movement ecology and dynamics of plant communities in fragmented

landscapes. Proceedings of the Natural Academy of Sciences 105: 19078-19083.

Dauber, J.; Hirsch, M.; Simmering, D.; Waldhardt, R.; Otte, A.; Wolters, V. 2003. Landscape

structure as an indicator of biodiversity: Matrix effects on species richness. Agriculture,

Ecosystems and Environment 98: 321-329.

De Angelis, D.L. and Waterhouse, J.C. 1987. Equilibrium and Nonequilibrium concepts in

ecological models. Ecological Monographs 57: 1-21.

Decocq, G.; Aubert, M.; Dupont, F.; Alard, D.; Saguez, R.; Wattez-Franger, A.; De Foucault,

B.; Delelis-Dusollier, A.; Bardat, J. 2004. Plant diversity in a managed temperate

deciduous forest: Understorey response to two silvicultural systems. Journal of Applied

Ecology 41: 1065-1079.

Page 168: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

154

Degen, T.; Devillez, F.; Jaquemart, A. L. 2005. Gaps promote plant diversity in beech forests

(Luzulo-Fagetum), North Vosges, France. Annals of Forest Science 62(5): 429-440.

Del Campo Sanchos, A and Solís Camba, A. 1993. Influencia de la ordenación en las masas

de Pinus nigra Arn. y Pinus sylvestris L. en el desarrollo rural de la comarca del Alto

Tajo. Actas del I Congreso Forestal Español. Lourizán, Pontevedra.

di Castri, F. 1981. Mediterranean-type shrublands of the world. In: di Castri, F.; Goodall,

D.W.; Specht, R.C. (Eds.), Mediterranean-type shrublands. Elsevier, Amsterdam, pp. 1-

52.

Díaz, L. 2006. Influences of forest type and forest structure on bird communities in oak and

pine woodlands in Spain. Forest Ecology and Management 223: 54-65.

Díaz, M.; Pulido, F.J.; Marañón, T. 2003. Diversidad biológica y sostenibilidad ecológica y

económica de los sistemas adehesados. Ecosistemas 3. Available in:

http://www.revistaecosistemas.net/pdfs/193.pdf [last access 12th Sep. 2011].

Díaz-Varela, E.R.; Marey-Pérez, M.F.; Rigueiro-Rodríguez, A.; Álvarez-Álvarez, P. 2009.

Landscape metrics for characterization of forest landscapes in a sustainable

management framework: Potential application and prevention of misuse. Annals of

Forest Science 66: 301.

Diez, J.M. and Pulliam, H.R. 2007. Hierarchical analysis of species distributions and

abundance across environmental gradients. Ecology 88: 3144-3152.

Diggle, P.J.; Tawn, J.A.; Moyeed, R.A. 1998. Model-Based Geostatistics. Journal of the

Royal Statistical Society. Series C: Applied Statistics 47: 299-350.

Drever, C.R.; Peterson, G.; Messier, C.; Bergeron, Y.; Flannigan, M.; 2006. Can forest

management based on natural disturbances maintain ecological resilience? Canadian

Journal of Forest Research 36: 2285-2299.

Duelli, P. and Obrist, M.K. 2003. Biodiversity indicators: The choice of values and measures.

Agriculture, Ecosystems and Environment 98: 87-98.

Duffy, D.C. and Meier, A.J. 1992. Do Appalachian herbaceous understories ever recover

from clearcutting? Conservation Biology 6: 196-201.

Dufour-Dror, J.M. and Ertas, A. 2004. Bioclimatic perspectives in the distribution of Quercus

ithaburensis Decne. subspecies in Turkey and in the Levant. Journal of Biogeography

31: 461-474.

Dupouey, J.L.; Dambrine, E.; Laffite, J.D.; Moares, C. 2002. Irreversible impact of past land

use on forest soils and biodiversity. Ecology 83: 2978-2987.

Page 169: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

155

Ellison, A.M. 2004. Bayesian inference in ecology. Ecology Letters 7: 509-520.

Eriksson, O. 1993. The species-pool hypothesis and plant community diversity. Oikos 68:

371-374.

Erös, T.; Schmera, D.; Schick, R.S. 2011. Network thinking in riverscape conservation: a

graph-based approach. Biological Conservation 144(1): 184-192.

Fabbio, G.; Merlo, M.; Tosi, V. 2003. Silvicultural management in maintaining biodiversity

and resistance of forest in Europe- the Mediterranean region. Journal of Environmental

Management 67: 67-76.

Fahrig, L. 2001. How much habitat is enough? Biological Conservation 100(1): 65-74.

Fahrig, L. 2002. Effect of habitat fragmentation on the extinction threshold: A synthesis.

Ecological Applications 12: 346-353.

Fahrig, L. 2003. Effects of habitat fragmentation on biodiversity. Annual Review of Ecology,

Evolution and Systematics 34: 487-515.

Fernández-Juricic, E. 2000. Forest fragmentation affects winter flock formation of an

insectivorous guild. Ardea 88: 235-241.

Field, R.; O'Brien, E.M.; Whittaker, R.J. 2005. Global models for predicting woody plant

richness from climate: Development and evaluation. Ecology 86: 2263-2277.

Field, R.; Hawkins, B.A.; Cornell, H.V.; Currie, D.J.; Diniz-Filho, J.A.F.; Guégan, J.F.;

Kaufman, D.M.; Kerr, J.T.; Mittelbach, G.G.; Oberdorff, T.; O'Brien, E.M.; Turner,

J.R.G. 2009. Spatial species-richness gradients across scales: A meta-analysis. Journal

of Biogeography 36: 132-147.

Finkeldey, R. and Ziehe, M. 2004. Genetic implications of silvicultural regimes. Forest

Ecology and Management 197: 231-244.

Forman, R.T.T. 1995. Land mosaics: The ecology of landscapes and regions. Cambridge

University Press, Cambridge.

Forman, R.T.T. and Godron, M. 1986. Landscape ecology. John Wiley & Sons, New York.

Fortin, M.J. and Dale, M. 2005. Spatial analysis. A guide for ecologists. Cambridge

University Press, Cambridge.

Fotheringham, A.S.; Brunsdon, C.; Charlton, M. 2002. Geographically Weighted Regression.

The analysis of spatially varying relationships. Wiley, Chichester.

Page 170: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

156

García, D. and Ortiz-Pulido, R. 2004. Patterns of resource tracking by avian frugivores at

multiple spatial scales:Two case studies on discordance among scales. Ecography 27:

187-196.

García, D.; Obeso, J.R.; Martínez, I. 2005a. Rodent seed predation promotes differential

recruitment among bird-dispersed trees in temperate secondary forests. Oecologia 144:

435-446.

García, D.; Obeso, J.R.; Martínez, I. 2005b. Spatial concordance between seed rain and

seedling establishment in bird-dispersed trees: Does scale matter? Journal of Ecology

93: 693-704.

García, D.; Zamora, R.; Amico, G.C. 2010. Birds as suppliers of seed dispersal in temperate

ecosystems: Conservation guidelines from real-world landscapes. Conservation Biology

24: 1070-1079.

García-Feced, C.; Saura, S.; Elena-Rosselló, R. 2011. Improving landscape connectivity in

forest districts: A two-stage process for prioritizing agricultural patches for

reforestation. Forest Ecology and Management 261: 154-161.

García-López, J.M. and Allué-Camacho, C. 2008. Phytoclimatic versatility and potential

diversity of natural arboreal forest cover in peninsular Spain. Forest Systems 17: 297-

307.

Gardner, R.H. and Engelhardt, K.A.M. 2008. Spatial processes that maintain biodiversity in

plant communities. Perspectives in Plant Ecology, Evolution and Systematics 9: 211-

228.

Gaston, K.J. 1996. Biodiversity. A biology of numbers and difference. Blackwell, Oxford.

Gaston, K.J. 2000. Global patterns in biodiversity. Nature 405: 220-227.

Gelman, A. and Pardoe, I. 2006. Bayesian measures of explained variance and pooling in

multilevel (hierarchical) models. Technometrics 48: 241-251.

Gelman, A. and Hill, J. 2007. Data analysis using regression and multilevel/hierarchical

models. Cambridge University Press, Cambridge.

Gil-Tena, A.; Saura, S.; Brotons, L. 2007. Effects of forest composition and structure on bird

species richness in a Mediterranean context: Implications for forest ecosystem

management. Forest Ecology and Management 242: 470-476.

Gil-Tena, A.; Brotons, L.; Saura, S. 2009. Mediterranean forest dynamics and forest bird

distribution changes in the late 20th century. Global Change Biology 15: 474-485.

Page 171: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

157

Gil-Tena, A.; Brotons, L.; Saura, S. 2010. Effects of forest landscape change and management

on the range expansion of forest bird species in the Mediterranean region. Forest

Ecology and Management 259: 1338-1346.

Gilbert, F.; Gonzalez, A.; Evans-Freke, I. 1998. Corridors maintain species richness in the

fragmented landscapes of a microecosystem. Proceedings of the Royal Society B:

Biological Sciences 265: 577-582.

Gilliam, F.S.; Turrill, N.L.; Adams, M.B. 1995. Herbaceous-layer and overstory species in

clear-cut and mature central Appalachian hardwood forests. Ecological Applications 5:

947-955.

Gómez, J.M. 2003. Spatial pattern in long-distance dispersal of Quercus ilex acorns by jays in

a heterogeneous landscape. Ecography 26: 573-584.

Gómez, J.M.; García, D.; Zamora, R. 2003. Impact of vertebrate acorn- and seedling-

predators on a Mediterranean Quercus pyrenaica forest. Forest Ecology and

Management 180: 125-134.

Gómez-Aparicio, L.; Pérez-Ramos, I.M.; Mendoza, I.; Matías, L.; Quero, J.L.; Castro, J.;

Zamora, R.; Marañón, T. 2008. Oak seedling survival and growth along resource

gradients in Mediterranean forests: Implications for regeneration in current and future

environmental scenarios. Oikos 117: 1683-1699.

Gómez-Aparicio, L.; Zavala, M.A.; Bonet, F.J.; Zamora, R. 2009. Are pine plantations valid

tools for restoring Mediterranean forests? An assessment along abiotic and biotic

gradients. Ecological Applications 19: 2124–2141.

Gonzalez, A.; Lawton, J.H.; Gilbert, F.S.: Blackburn, T.M.; Evans-Freke, I. 1998.

Metapopulation dynamics, abundance, and distribution in a microecosystem. Science

281: 2045-2047.

Gonzalez, M.; Deconchat, M.; Balent, G. 2009. Woody plant composition of forest layers:

The importance of environmental conditions and spatial configuration. Plant Ecology

201: 305-318.

González-Alday, J.; Martínez-Ruiz, C.; Bravo, F. 2009. Evaluating different harvest

intensities over understory plant diversity and pine seedlings, in a Pinus pinaster Ait.

natural stand of Spain. Plant Ecology 201: 211-220.

González-Alday, J.; Martínez-Ruiz, C.; Marrs, R.H.; Bravo, F. 2010. Influence of harvesting

intensity on the floristic composition of natural Mediterranean maritime pine forest.

Acta Oecologica 36: 349-356.

Page 172: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

158

Götmark, F.; Paltto, H.; Nordén, B.; Götmark, E. 2005. Evaluating partial cutting in

broadleaved temperate forest under strong experimental control: Short-term effects on

herbaceous plants. Forest Ecology and Management 214: 124-141.

Graham, M.H. 2003. Confronting multicollinearity in ecological multiple regression. Ecology

84(11): 2809-2815.

Grime, J.P. 1998. Benefits of plant diversity to ecosystems: Immediate, filter and founder

effects. Journal of Ecology 86: 902-910.

Guo, L.; Ma, Z.; Zhang, L. 2008. Comparison of bandwidth selection in application of

geographically weighted regression: A case study. Canadian Journal of Forest Research

38: 2526-2534.

Hall, M. 2000. Comparing damages: Italian and American concepts of restoration. In:

Agnoletti, M. and Anderson, S. (Eds.), Methods and approaches in forest history.

CABI, Wallingford, pp. 165-172.

Halme, P.; Toivanen, T.; Honkanen, M.; Kotiaho, J.S.; Mönkkönen, M.; Timonen, J. 2010.

Flawed meta-analysis of biodiversity effects of forest management. Conservation

Biology 24: 1154-1156.

Halpern, C.B. and Spies, T.A. 1995. Plant species diversity in natural and managed forests of

the Pacific Northwest. Ecological Applications 5: 913-934.

Hannerz, M. and Hinell, B. 1997. Effects on the flora in Norway spruce forests following

clearcutting and shelterwood cutting. Forest Ecology and Management 90:29-49.

Hanski, I. 1999. Metapopulation Ecology. Oxford University Press, Oxford.

Hawkins, B.A.; Albuquerque, F.S.; Araújo, M.B.; Beck, J.; Bini, L.M.F.; Cabrero-Sañudo, J.;

Castro-Parga, I.; Diniz-Filho, J.A.F.; Ferrer-Castán, D.: Field, R.; Gómez, J.F.; Hortal,

J.; Kerr, J.T.; Kitching, I.J.; León-Cortés, J.L.; Lobo, J.M.; Montoya, D.; Moreno, J.C.;

Olalla-Tárraga, M.Á.; Pausas, J.G.; Qian, H.; Rahbek, C.; Rodríguez, M.Á.; Sanders,

N.J.; Williams, P. 2007. A global evaluation of metabolic theory as an explanation for

terrestrial species richness gradients. Ecology 88: 1877-1888.

Heikkinen, R.K. and Neuvoenen, S. 1997. Species richness of vascular plants in the subarctic

landscape of northern Finland: Modelling relationships to the environment. Biodiversity

and Conservation 6: 1181-1201.

Hermy, M. and Verheyen, K. 2007. Legacies of the past in the present-day forest biodiversity:

A review of past land-use effects on forest plant species composition and diversity.

Ecological Research 22: 361-371.

Page 173: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

159

Herrera, C.M.; Jordano, P.; López-Soria, L.; Amat, J.A. 1994. Recruitment of a mast-fruiting,

bird-dispersed tree: Bridging frugivore activity and seedling establishment. Ecological

Monographs 64: 315-344.

Higgins, S.I.; Nathan, R.; Cain, M.L. 2003. Are long-distance dispersal events in plants

usually caused by nonstandard means of dispersal? Ecology 84: 1945-1956.

Hill, J.L. and Curran, P.J. 2005. Fragment shape and tree species composition in tropical

forests: A landscape level investigation. African Journal of Ecology 43: 35-43.

Honnay, O.; Hermy, M.; Coppin, P. 1999. Effects of area, age and diversity of forest patches

in Belgium on plant species richness, and implications for conservation and

reforestation. Biological Conservation 87: 73-84.

Honnay, O.; Piessens, K.; Van Landuyt, W.; Hermy, M.; Gulinck, H. 2003. Satellite based

land use and landscape complexity indices as predictors for regional plant species

diversity. Landscape and Urban Planning 63: 241-250.

Hooper, D.U.; Chapin III, F.S.; Ewel, J.J.; Hector, A.; Inchausti, P.; Lavorel, S.; Lawton, J.H.;

Lodge, D.M.; Loreau, M.; Naeem, S.; Schmid, B.; Setälä, H.; Symstad, A.J.;

Vandermeer, J.; Wardle, D.A. 2005. Effects of biodiversity on ecosystem functioning:

A consensus of current knowledge. Ecological Monographs 75: 3-35.

Horn, H.S. and MacArthur, R.H. 1972. Competition among fugitive species in a harlequin

environment. Ecology 53: 749-752.

Hunter, M.L. 1990. Wildlife, forests, and forestry: Principles of managing forests for

biological diversity. Prentice-Hall, Englewood Cliffs, New Jersey.

Hurtt, G.C. and Pacala, S.W. 1995. The consequences of recruitment limitation: Reconciling

chance, history and competitive differences between plants. Journal of Theoretical

Biology 176: 1-12.

Huston, M.A. 1979 A general hypothesis of species diversity. American Naturalist 113: 81-

101.

Huston, M.A. 1994. Biological diversity: The coexistence of species on changing landscapes.

Cambridge University Press, Cambridge.

Huston, M.A. 1999. Local processes and regional patterns: Appropriate scales for

understanding variation in the diversity of plants and animals. Oikos 86: 393–401.

Hutchinson, G. E. 1957. The ecological theatre and the evolutionary play. Yale University

Press, New Haven.

Page 174: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

160

Jordano, P.; Pulido, F; Arroyo, J.; García-Castaño, J.L.; García-Fayos, P. 2004. Procesos de

limitación demográfica. In: Valladares, F. (Ed.), Ecología del bosque mediterráneo en

un mundo cambiante. Ministerio de Medio Ambiente, EGRAF SA, Madrid, pp. 229-

248.

Jordano, P.; García, C.; Godoy, J.A.; García-Castaño, J.L. 2007. Differential contribution of

frugivores to complex seed dispersal patterns. Proceedings of the Natural Academy of

Sciences 104: 3278-3282.

Jump, A.S.; Marchant, R.; Peñuelas, J. 2009. Environmental change and the option value of

genetic diversity. Trends in Plant Science 14: 51-58.

Kallimanis, A.; Ragia, V.; Sgardelis, S.; Pantis, J. 2007. Using regression trees to predict

alpha diversity based upon geographical and habitat characteristics. Biodiversity and

Conservation 16: 3863-3876.

Kallimanis, A.; Halley, J.; Vokou, D.; Sgardelis, S. 2008. The scale of analysis determines the

spatial pattern of woody species diversity in the Mediterranean environment. Plant

Ecology 196: 143-151.

Keane, R.E.; Hessburg, P.F.; Landres, P.B.; Swanson, F.J. 2009. The use of historical range

and variability (HRV) in landscape management. Forest Ecology and Management 258:

1025-1037.

Kern, C.C.; Palik, B.J.; Strong, T.F. 2006. Ground-layer plant community responses to even-

age and uneven-age silvicultural treatments in Wisconsin northern hardwood forests.

Forest Ecology and Management 230: 162-170.

Kissling, W.D.; Field, R.; Böhning-Gaese, K. 2008. Spatial patterns of woody plant and bird

diversity: Functional relationships or environmental effects? Global Ecology and

Biogeography 17: 327-339.

Kondoh, M. 2001. Unifying the relationships of species richness to productivity and

disturbance. Proceedings of the Royal Society B: Biological Sciences 268: 269-271.

Kozlowski, T.T. 2002. Physiological ecology of natural regeneration of harvested and

disturbed forest stands: Implications for forest management. Forest Ecology and

Management 158: 195-221.

Kraft, L.S.; Crow, T.R.; Buckley, D.S.; Nauertz, E.A.; Zasada, J.C. 2004. Effects of

harvesting and deer browsing on attributes of understory plants in northern hardwood

forests, upper Michigan, USA. Forest Ecology and Management 199: 219-230.

Page 175: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

161

Krzic, M.; Newman, R.F.; Broersma, K. 2003. Plant species diversity and soil quality in

harvested and grazed boreal aspen stands of northeastern British Columbia. Forest

Ecology and Management 182: 315-325.

Kuuluvainen, T. 2009. Forest management and biodiversity conservation based on natural

ecosystem dynamics in Northern Europe: The complexity challenge. Ambio 38: 309-

315.

Lafortezza, R., Chen, J., Sanesi, G., Crow, T.R. (Eds.). 2008. Patterns and processes in forest

landscapes: Multiple use and sustainable management. Springer, New York.

Lafortezza, R.; Coomes, D.A.; Kapos, V.; Ewers, R.M. 2010. Assessing the impacts of

fragmentation on plant communities in New Zealand: Scaling from survey plots to

landscapes. Global Ecology and Biogeography 19(5): 741-754.

Laikre, L.; Allendorf, F.W.; Aroner, L.C.; Baker, C.S.; Gregovich, D.P.; Hansen, M.M.;

Jackson, J.A.; Kendall, K.C.; McKelvey, K.; Neel, M.C.; Olivieri, I.; Ryman, N.;

Schwartz, M.K.; Bull, R.S.; Stetz, J.B.; Tallmon, D.A.; Taylor, B.L.; Vojta, C.D.;

Waller, D.M.; Waples, R.S. 2010. Neglect of genetic diversity in implementation of the

convention on biological diversity. Conservation Biology 24: 86-88.

Lavorel, S. 1999. Ecological diversity and resilience of Mediterranean vegetation to

disturbance. Diversity and Distributions 5: 3-13.

Leathwick, J.R.; Burns, B.R.; Clarkson, B.D. 1998. Environmental correlates of tree alpha-

diversity in New Zealand primary forests. Ecography 21: 235-246.

Legendre, P. and Legendre, L. 1998. Numerical ecology. Elsevier, Amsterdam.

Leibold, M.A.; Holyoak, M.; Mouquet, N.; Amarasekare, P.; Chase, J.M.; Hoopes, M.F.;

Holt, R.D.; Shurin, J.B.; Law, R.; Tilman, D.; Loreau, M.; Gonzalez, A. 2004. The

metacommunity concept: A framework for multi-scale community ecology. Ecology

Letters 7: 601-613.

Levin, S.A. 1992. The problem of pattern and scale in ecology. Ecology 73: 1943-1967.

Lindborg, R. and Eriksson, O. 2004. Historical landscape connectivity affects present plant

species diversity. Ecology 85: 1840-1845.

Lindenmayer, D.B.; Margules, C.R.; Botkin, D.B. 2000. Indicators of biodiversity for

ecologically sustainable forest management. Conservation Biology 14: 941-950.

Lloret, F. 2004. Régimen de incendios y regeneración. In: Valladares, F. (Ed.), Ecología del

bosque mediterráneo en un mundo cambiante. Ministerio de Medio Ambiente, EGRAF

SA, Madrid, pp. 101-126.

Page 176: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

162

Lloret, F.; Lobo, A.; Estevan, H.; Maisongrande, P.; Vayreda, J.; Terradas, J. 2007. Woody

plant richness and NDVI response to drought events in Catalonian (Northeastern Spain)

forests. Ecology 88: 2270-2279.

Lloret, F.; Solé, A.; Pino, J.; Vayreda, J.; Font, X.; Terradas, J. 2009. Patterns of species

impoverishment in managed forests of Catalonia (NE Spain). Journal of Vegetation

Science 20: 675-685.

Lobo, J.M.; Castro, I.; Moreno, J.C. 2001. Spatial and environmental determinants of vascular

plant species richness distribution in the Iberian Peninsula and Balearic Islands.

Biological Journal of the Linnean Society 73: 233-253.

Loepfe, L.; Martinez-Vilalta, J.; Oliveres, J.; Piñol, J.; Lloret, F. 2010. Feedbacks between

fuel reduction and landscape homogenisation determine fire regimes in three

Mediterranean areas. Forest Ecology and Management 259: 2366-2374.

Long, J.N. 2009. Emulating natural disturbance regimes as a basis for forest management: A

North American view. Forest Ecology and Management 257: 1868-1873.

Loreau, M.; Naeem, S.; Inchausti, P.; Bergtsson, J.; Grime, J.P.; Hector, A.; Hooper, D.U.;

Huston, M.A.; Raffaelli, D.; Shmid, B.; Tilman, D.; Wardle, D.A. 2001. Biodiversity

and ecosystem functioning: Current knowledge and future challenges. Science 294:

804-808.

Loyn, R.H. and Kennedy, S.J. 2009. Designing old forest for the future: Old trees as habitat

for birds in forests of Mountain Ash Eucaliptus regnans. Forest Ecology and

Management 258: 504-515.

MacArthur, R.H. 1972. Geographical Ecology. Harper and Row, New York.

MacArthur, R.H. and MacArthur, J.W. 1961. On bird species diversity. Ecology 42: 594-598.

Madrigal, A. 2008. Gestión forestal: Ayer, hoy, mañana. Lección de Primavera del Curso

académico 2007-2008. ETSI Montes, Universidad Politécnica de Madrid.

Madrigal, A.; Alvárez González, J.G.; Rodríguez Soalleiro, R.; Rojo Alboreca, A. 1999.

Tablas de producción para los montes españoles. Fundación Conde del Valle de

Salazar, Madrid.

Magurran, A.E. 1989. Diversidad ecológica y su medición. Vedrà, Barcelona.

Malanson, G.P. and Armstrong, M.P. 1996. Dispersal probability and forest diversity in a

fragmented landscape. Ecological Modelling 87: 91-102.

Marañón, T.; Camarero, J.J.; Castro, J.; Díaz, M.; Espelta, J.M.; Hampe, A.; Jordano, P.;

Valladares, F.; Verdú, M.; Zamora, R. 2004. Heterogeneidad ambiental y nicho de

Page 177: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

163

regeneración. In: Valladares, F. (Ed.), Ecología del bosque mediterráneo en un mundo

cambiante. Ministerio de Medio Ambiente, EGRAF SA, Madrid, pp. 69-99.

Matías, L.; Mendoza, I.; Zamora, R. 2009. Consistent pattern of habitat and species selection

by post-dispersal seed predators in a Mediterranean mosaic landscape. Plant Ecology

203: 137-174.

Matías, L.; Zamora, R.; Mendoza, I.; Hodar, J.A. 2010. Seed dispersal patterns by large

frugivorous mammals in a degraded mosaic landscape. Restoration Ecology 18: 619-

627.

McEuen, A.B. and Curran, L.M. 2004. Seed dispersal and recruitment limitation across

spatial scales in temperate forest fragments. Ecology 85: 507-518.

McDonald, R.I.; Motzkin, G.; Foster, D.R. 2008. The effect of logging on vegetation

composition in Western Massachussetts. Forest Ecology and Management 255: 4021-

4031.

McMahon, S.M. and Diez, J.M. 2007. Scales of association: Hierarchical linear models and

the measurement of ecological systems. Ecology Letters 10: 437-452.

MCPFE. 2003. State of Europe’s Forests: The MCPFE Report on Sustainable Forest

Management in Europe. Ministerial Conference on Protection of Forests in Europe.

Liaison Unit, Viena.

Meier, A.J.; Bratton, S.P.; Duffy, D.C. 1995. Possible ecological mechanisms for loss of

vernal-herb diversity in logged eastern deciduous forests. Ecological Applications 5:

935-946.

Mendoza, I.; Gómez-Aparicio, L.; Zamora, R.; Matías, L. 2009. Recruitment limitation of

forest communities in a degraded Mediterranean landscape. Journal of Vegetation

Science 20: 367-376.

Metzger, F. and Schultz, J. 1984. Understory response to 50 years of management of a

northern hardwood forest in upper Michigan. American Midland Naturalist 112: 209-

223.

Metzger, M.J.; Bunce, R.G.H.; Jongman, R.H.G.; Mücher, C.A.; Watkins, J.W. 2005. A

climatic stratification of the environment of Europe. Global Ecology and Biogeography

14: 549-563.

Metzger, M.J.; Bunce, R.G.H.; Leemans, R.; Viner, D. 2008. Projected environmental

shifts under climate change: European trends and regional impacts.

Environmental Conservation 35: 64-75.

Page 178: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

164

Millennium Ecosystem Assessment. 2005. Ecosystems and human well-being: Current state

and trends. Condition and Trends Working Group, Island, Washington, DC.

Miller, J.N.; Brooks, R.P.; Croonquist, M.J. 1997. Effects of landscape patterns on biotic

communities. Landscape Ecology 12: 137-153.

Ministerio de Fomento. 1999. Modelo Digital del Terreno 1: 25000. Dirección General del

Instituto Geográfico Nacional. Centro Nacional de Información Geográfica, Madrid.

Ministerio de Medio Ambiente. 1997-2007. Tercer Inventario Forestal Nacional. Dirección

General de Conservación de la Naturaleza, Madrid.

Ministerio de Medio Ambiente, 2006. Incendios Forestales en España. Año 2006. Technical

Report, Ministerio de Medio Ambiente, Madrid. Available in:

http://www.mma.es/secciones/biodiversidad/defensa_incendios/estadisticas_incendios/p

df/publicacion_2006.pdf [last access 12th Sep. 2011]

Minor, E.S. and Lookingbill, T.R. 2010. A multiscale network analysis of protected-area

connectivity for mammals in the United States. Conservation Biology 24(6): 1549-

1558.

Mittelbach, G.G.; Steiner, C.F.; Scheiner, S.M.; Gross, K.L.; Reynolds, H.L.; Waide, R.B.;

Willig, M.R.; Dodson, S.I.; Gough, L. 2001. What is the observed relationship between

species richness and productivity? Ecology 82: 2381-2396.

Montero, G. and Madrigal, A. 1999. La Selvicultura y la Ordenación de Montes. In: Madrigal,

A. (Ed.), Ciencias y Técnicas Forestales: 150 años de aportaciones de los ingenieros de

montes. Fundación Conde del Valle de Salazar, Madrid, pp. 23-32.

Montero, G.; Cisneros, O.; Cañellas, I. 2003. Manual de selvicultura para plantaciones de

especies productoras de madera de calidad. Mundi-Prensa, Madrid.

Montoya, D.; Zavala, M.A.; Rodríguez, M.A.; Purves, D.W. 2008. Animal versus wind

dispersal and the robustness of tree species deforestation. Science 320: 1502-1504.

Montoya, D.; Alburquerque, F.S.; Rueda, M.; Rodríguez, M.A. 2010. Species' response

patterns to habitat fragmentation: Do trees support the extinction threshold hypothesis?

Oikos 119: 1335-1343.

Moreira, F. and Russo, D. 2007. Modelling the impact of agricultural abandonment and

wildfires on vertebrate diversity in Mediterranean Europe. Landscape Ecology 22:

1461-1476.

Moreno-Saiz, J. and Lobo, J. 2008. Iberian-Balearic fern regions and their explanatory

variables. Plant Ecology 198: 149-167.

Page 179: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

165

Moser, D.; Zechmeister, H.G.; Plutzar, C.; Sauberer, N.; Wrbka, T.; Grabherr, G. 2002.

Landscape patch shape complexity as an effective measure for plant species richness in

rural landscapes. Landscape Ecology 17: 657-669.

Myers, N.; Mittermeier, R.A.; Mittermeier, C.G.; da Fonseca, G.A.B.; Kent, J. 2000.

Biodiversity hotspots for conservation priorities. Nature 403: 853-858.

Nadrowski, K.; Wirth, C.; Scherer-Lorenzen, M. 2010. Is forest diversity driving ecosystem

function and service? Current Opinion in Environmental Sustainability 2: 75-79.

Nathan, R. 2006. Long-distance dispersal of plants. Science 313: 786-788.

Nathan, R. and Muller-Landau, R. 2000. Spatial patterns of seed dispersal, their determinants

and consequences for recruitment. Trends in Ecology and Evolution 15: 278-285.

Navarro, F.B.; Jiménez, M.N.; Gallego, E.; Ripoll, M.A. 2010. Short-term effects of overstory

reduction and slash mulching on ground vegetation in a Mediterranean Aleppo pine

woodland. European Journal of Forest Research 129: 689-696.

Neel, M.C. 2008. Patch connectivity and genetic diversity conservation in the federally

endangered and narrowly endemic plant species Astragalus albens (Fabaceae).

Biological Conservation 141: 938-955.

Newmaster, S.G.; Parker, W.C.; Bell, F.W.; Paterson, J.M. 2007. Effects of forest floor

disturbances by mechanical site preparation on floristic diversity in a central Ontario

clearcut. Forest Ecology and Management 246: 196-207.

Niemelä, J. 1999. Management in relation to disturbance in the boreal forest. Forest Ecology

and Management 115: 127-134.

Ninyerola, M.; Pons, X. ; Roure, J.M. 2005. Atlas Climático Digital de la Península Ibérica.

Metodología y aplicaciones en bioclimatología y geobotánica. Universidad Autónoma

de Barcelona, Bellaterra.

Nogués-Bravo, D.; Araújo, M.B.; Romdal, T.; Rahbek, C. 2008. Scale effects and human

impact on the elevational species richness gradients. Nature 453: 216-220.

Noss, R.F. 1990. Indicators for monitoring biodiversity: A hierarchical approach.

Conservation Biology 4: 355-364.

Noss, R. F. 1991. Landscape connectivity: Different functions at different scales. In: Hudson,

W.E. (Ed.), Landscape linkages and biodiversity. Island Press, Covelo, California, pp.

27-39.

Page 180: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

166

O’Brien, E.M. 1993. Climatic gradients in woody plant species richness: Towards an

explanation based on an analysis of southern Africa’s woody flora. Journal of

Biogeography 20: 181-198.

O’Brien, E.M. 1998. Water-energy dynamics, climate, and prediction of woody plant species

richness: An interim general model. Journal of Biogeography 25: 379-398.

O’Brien, E.M.; Field, R.; Whittaker, R.J. 2000. Climatic gradients in woody plant (tree and

shrub) diversity: Water-energy dynamics, residual variation, and topography. Oikos 89:

588-600.

Odion, D.C. and Sarr, D.A. 2007. Managing disturbance regimes to maintain biological

diversity in forested ecosystems of the Pacific Northwest. Forest Ecology and

Management 246: 57-65.

Okland, T.; Rydgren, K.; Okland, R.H.; Storaunet, K.O.; Rolstad, J. 2003. Variation in

environmental conditions, understorey species number, abundance and composition

among natural and managed Picea abies forest stands. Forest Ecology and Management

177: 17-37.

Onaindia, M.; Dominguez, I.; Albizu, I.; Garbisu, C.; Amezaga, I. 2004. Vegetation diversity

and vertical structure as indicators of forest disturbance. Forest Ecology and

Management 195: 341-354.

Orrock, J.L.; Danielson, B.J.; Burns, M.J.; Levey, D.J. 2003. Spatial ecology of predator-prey

interactions: Corridors and patch shape influence seed predation. Ecology 84: 2589-

2599.

Ortega, M.; Bunce, R.G.H.; García del Barrio, J.M.; Elena Roselló, R. 2008. The relative

dependence of Spanish landscape pattern on environmental and geographical variables

over time. Forest systems 17: 114-129.

Osborne, P.E.; Foody, G.M.; Suárez-Seoane, S. 2007. Non-stationarity and local approaches

to modelling the distributions of wildlife. Diversity and Distributions 13: 313-323.

Ozinga, W.A.; Römermann, C.; Bekker, R.M.; Prinzing, A.; Tamis, W.L.M.; Schaminée,

J.H.J.; Hennekens, S.M.; Thompson, K.; Bakker, J.P.; van Groenendael, J.M. 2009.

Dispersal failure contributes to plant losses in NW Europe. Ecology Letters 12: 66-74.

Paillet, Y.; Bergès, L.; Hjältén, J.; Ódor, P.; Avon, C.; Bernhardt-Römermann, M.; Bijlsma,

R.; De Bruyn, L.; Fuhr, M.; Grandin, U.; Kanka, R.; Lundin, L.; Luque, S.; Magura, T.;

Matesanz, S.; Mészáros, I.; Sebastià, M.T.; Schmidt, W.; Standovár, T.; Tóthmérész,

B.; Uotila, A.; Valladares, F.; Vellak, K.; Virtanen, R. 2010a. Biodiversity differences

Page 181: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

167

between managed and unmanaged forests: Meta-analysis of species richness in Europe.

Conservation Biology 24: 101-112.

Paillet, Y.; Bergès, L.; Hjältén, J.; Ódor, P.; Avon, C.; Bernhardt-Römermann, M.; Bijlsma,

R.; De Bruyn, L.; Fuhr, M.; Grandin, U.; Kanka, R.; Lundin, L.; Luque, S.; Magura, T.;

Matesanz, S.; Mészáros, I.; Sebastià, M.T; Schmidt, W.; Standovár, T.; Tóthmérész, B.;

Uotila, A.; Valladares, F.; Vellak, K.; Virtanen, R. 2010b. Compromises in data

selection in a meta-analysis of biodiversity in managed and unmanaged forests:

Response to Halme et al. Conservation Biology 24: 1157-1160.

Palmer, M.W. 1994. Variation in species richness: Towards a unification of hypotheses. Folia

Geobotanica et Phytotaxonomica 29: 511-530.

Pärtel, M. and Zobel, M. 1999. Small-scale plant species richness in calcareous grasslands

determined by the species pool, community age and shoot density. Ecography 22: 153-

159.

Pascual-Hortal, L.and Saura, S. 2008. Integrating landscape connectivity in broad-scale forest

planning through a new graph-based habitat availability methodology: Application to

capercaillie (Tetrao urogallus) in Catalonia (NE Spain). European Journal of Forest

Research 127: 23-31.

Pausas, J.G. 1994. Species richness patterns in the understorey of Pyrenean Pinus sylvestris

forests. Journal of Vegetation Science 5: 517-524.

Pausas, J.C.; Carreras, J.; Ferré, A.; Font, X. 2003. Coarse-scale plant species richness in

relation to environmental heterogeneity. Journal of Vegetation Science 14: 661-668.

Pausas, J.C.; Bladé, C.; Valdecantos, A.; Seva, J.P.; Fuentes, D.; Alloza, J.A.; Vilagrosa, A.;

Bautista, S.; Cortina, J.; Vallejo, R. 2004. Pines and oaks in the restoration of

Mediterranean landscapes of Spain: New perspectives for an old practice - A review.

Plant Ecology 171: 209-220.

Peco, B.; Sánchez, A.M.; Azcárate, F.M. 2006. Abandonment in grazing systems:

Consequences for vegetation and soil. Agriculture, Ecosystems and Environment 113:

284-294.

Peltzer, A.D.; Bast, M.L.; Wilson, S.D.; Gerry, A.K. 2000. Plant diversity and tree responses

following contrasting disturbances in boreal forest. Forest Ecology and Management

127: 191-203.

Page 182: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

168

Perea, R. San Miguel-Ayanz, A.; Gil-Sánchez, L. 2009. Dispersión secundaria de semillas de

robles (Quercus petraea y Quercus pyrenaica) en el Hayedo de Montejo (Madrid).

Actas del V Congreso Forestal Español, Ávila.

Perea, R.; San Miguel-Ayanz, A.; Gil, L. 2011a. Acorn dispersal by rodents: The importance

of re-dispersal and distance to shelter. Basic and Applied ecology 12(5): 432-439.

Perea, R.; San Miguel, A.; Gil, L. 2011b. Leftovers in seed dispersal: ecological implications

of partial seed consumption for oak regeneration. Journal of Ecology 99: 194-201.

Pereira, M.; Segurado, P.; Neves, N. 2011. Using spatial network structure in landscape

management and planning: A case study with pond turtles. Landscape and Urban

Planning 100: 67-76.

Perera, A.H., Buse, L.J., Crow, T.R. (Eds.). 2006. Forest Landscape Ecology. Transferring

knowledge to practice. Springer, New York.

Pérez-Soba, M.; San Miguel, A.; Elena-Rosselló, R. 2007. Complexity in the simplicity: The

Spanish dehesas. The secret of an ancient cultural landscape with high nature value still

functioning in the 21st century. In: Pedroli, B.; van Doorn, A.; de Blust,

G..; Paracchinni, M.L.;Wascher, D.; Bunce, F. (Eds.), Europe´s Living Landscapes

. KNNV Publishing, The Netherlands, pp. 369-384.

Plieninger, T. 2006. Habitat loss, fragmentation, and alteration - Quantifying the impact of

land-use changes on a Spanish dehesa landscape by use of aerial photography and GIS.

Landscape Ecology 21: 91-105.

Plieninger, T.; Rolo, V.; Moreno, G. 2010. Large-scale patterns of Quercus ilex, Quercus

suber, and Quercus pyrenaica regeneration in Central Western Spain. Ecosystems 13:

644-660.

Poyatos, R.; Latron, J.; Llorens, P. 2003. Land use and land cover change after agricultural

abandonment: The case of a Mediterranean Mountain area (Catalan Pre-Pyrenees).

Mountain Research and Development 23: 362-368.

Puerta-Piñero, C., Gómez, J.M., Valladares, F. 2007. Irradiance and oak seedling survival and

growth in a heterogeneous environment. Forest Ecology and Management 242: 462-

469.

Purves, D.W.; Zavala, M.A.; Ogle, K.; Prieto, F.; Rey Benayas, J.M. 2007. Environmental

heterogeneity, bird-mediated directed dispersal, and oak woodland dynamics in

Mediterranean Spain. Ecological Monographs 77: 77-97.

Page 183: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

169

Rahbek, C. 2005. The role of spatial scale and the perception of large-scale species-richness

patterns. Ecology Letters 8: 224-239.

Rahbek, C. and Graves, G.R. 2001. Multiscale assessment of patterns of avian species

richness. Proceedings of the National Academy of Sciences of the United States of

America 98: 4534-4539.

Ramírez, J.A. and Díaz, M. 2008. The role of temporal shrub encroachment for the

maintenance of Spanish holm oak Quercus ilex dehesas. Forest Ecology and

Management 255: 1976-1983.

Ramovs, B.V. and Roberts, M.R. 2003. Understory vegetation and environment responses to

tillage, forest harvesting, and conifer plantation development. Ecological Applications

13: 1682-1700.

Rands, M.R.W.; Adams, W.M.; Bennun, L.; Stuart, H.M.B.; Clements, A.; Coomes, D.;

Entwistle, A.; Hodge, I.; Kapos, V.; Sharlemann, J.P.W.; Sutherland, W.J.; Vira., B.

2010. Biodiversity conservation: Challenges beyond 2010. Science 329: 1298-1303.

Rex, K.D. and Malanson, G.P. 1990. The fractal shape of riparian forest patches. Landscape

Ecology 4(4): 249-258.

Rey, P.J. and Alcántara, J. 2000. Recruitment dynamics of a fleshy-fruited plant (Olea

europaea): connecting patterns of seed dispersal to seedling establishment. Journal of

Ecology 88: 622-633.

Rey-Benayas, J.M. 1995. Patterns of diversity in the strata of boreal forest in British

Columbia. Journal of Vegetation Science 6: 95-98.

Rey-Benayas, J.M. and Scheiner, S.M. 2002. Plant diversity, biogeography and environment

in Iberia: Patterns and possible causal factors. Journal of Vegetation Science 13: 245-

258.

Rey Benayas, J.M.; Navarro, J.; Espigares, T.; Nicolau, J.M.; Zavala, M.A. 2005. Effects of

artificial shading and weed mowing in reforestation of Mediterranean abandoned

cropland with contrasting Quercus species. Forest Ecology and Management 212: 302–

314.

Ribbens, E.; Silander, J.A.; Pacala, S.W. 1994. Seedling recruitment in forests: Calibrating

models to predict patterns of tree seedling dispersion. Ecology 75(6): 1794-1806

Ricklefs, R.E. 1987. Community diversity: Relative roles of local and regional processes.

Science 235: 167-171.

Page 184: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

170

Ricklefs, R.E. 2004. A comprehensive framework for global patterns in biodiversity. Ecology

Letters 7: 1-15.

Ricotta, C.; Godefroid, S.; Celesti-Grapow, L. 2008. Common species have lower taxonomic

diversity. Diversity and Distributions 14: 530-537.

Rivas-Martínez, S. 1987. Memoria del mapa de series de vegetación de España 1:400.000.

ICONA, Ministerio de Agricultura, Pesca y Alimentación, Madrid.

Roberts, M.R. and Gilliam, F.S. 1995. Patterns and mechanisms of plant diversity in forested

ecosystems: Implications for forest management. Ecological Applications 5: 969-977.

Roberts, M.R. and Zhu, L. 2002. Early response of the herbaceous layer to harvesting in a

mixed coniferous-deciduous forest in New Brunswick, Canada. Forest Ecology and

Management 155: 17-31.

Rodríguez-Calcerrada, J.; Cano, F.J.; Valbuena-Carabaña, M.; Gil, L.; Aranda, I. 2010.

Functional performance of oak seedlings naturally regenerated across microhabitats of

distinct overstorey canopy closure. New Forests 39: 245-259.

Rodríguez-García, E., Gratzer, G., Bravo F. 2011a. Post-disturbance natural

establishment of Pinus pinaster and environmental factors in five Mediterranean

populations of central Spain. Annals of Forest Science 68(4): 811-823.

Rodríguez-García, E., Ordóñez, C., Bravo F. 2011b. Effects of shrub and canopy cover on the

relative growth rate of Pinus pinaster Ait. seedlings of different sizes. Annals of Forest

Science (published online march 3rd 2011) DOI: 10.1007/s13595-011-0039-5.

Rodríguez-Sánchez, F.; Hampe, A.; Jordano, P.; Arroyo, J. 2010. Past tree range dynamics in

the Iberian Peninsula inferred through phylogeography and palaeodistribution

modelling: A review. Review of Palaeobotany and Palynology 162: 507-521.

Rosenzweig, M.L. 1995. Species diversity in space and time. Cambridge University Press,

Cambridge.

Rousset, O. and Lepart, J. 2000. Positive and negative interactions at different life stages of a

colonizing species (Quercus humilis). Journal of Ecology 88: 401-412.

Rowland, E.L.; White, A.S.; William, H.L. 2005. A literature review of the effects of

intensive forestry on forest structure and plant community composition at the stand and

landscape levels. Maine Agricultural and Forest Experiment Station, University of

Maine 754.

Page 185: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

171

Rozas, V.; Zas, R.; Solla, A. 2009. Spatial structure of deciduous forest stands with

contrasting human influence in northwest Spain. European Journal of Forest Research

128: 273-285.

San Miguel-Ayanz, A. 1994. La dehesa española. Origen, tipología, características y gestión.

Fundación Conde del Valle de Salazar, Madrid.

Santos, T.; Tellería, J.L.; Virgós, E. 1999. Dispersal of Spanish juniper Juniperus thurifera by

birds and mammals in a fragmented landscape. Ecography 22: 193-204.

Sarr, D.A.; Hibbs, D.E.; Huston, M.A. 2005. A hierarchical perspective of plant diversity. The

Quarterly Review of Biology 80: 187-212.

Saura, S. 2010. Del rodal al paisaje: un cambio de escala, nuevas perspectivas para la

planificación y ordenación forestal. Conferencias y Ponencias del V Congreso Forestal

Español. Ponencia Invitada de la Mesa Temática 8 (Monte, paisaje y territorio:

planificación y ordenación forestal). Cuadernos de la Sociedad Española de Ciencias

Forestales 31: 213-239.

Saura, S. 2011. Las instrucciones de ordenación de monte. In: Prieto, A.; Saura, S.; Díaz-

Balteiro, L., Martínez, S. Manual para la redacción de Instrumentos de Gestión Forestal

Sostenible (Ordenación de Montes) en Castilla-La Mancha. Junta de Comunidades de

Castilla-La Mancha, Consejería de Agricultura y Medio Ambiente, Dirección General

de Política Forestal. Serie Forestal nº 8, pp. 33-53.

Saura, S. and Carballal, P. 2004. Discrimination of native and exotic forest patterns through

shape irregularity indices: An analysis in the landscapes of Galicia, Spain. Landscape

Ecology 19: 647-662.

Saura, S.; Torras, O.; Gil-Tena, A.; Pascual, L. 2008. Shape irregularity as an indicator of

forest biodiversity and guidelines for metric selection. In: Lafortezza, R., Chen, J.,

Sanesi, G.; Crow, T.R. (Eds.), Patterns and processes in forest landscapes: Multiple use

and sustainable management. Springer, New York, pp. 167-189.

Saura, S. and Torné, J. 2009. Conefor Sensinode 2.2: A software package for quantifying the

importance of habitat patches for landscape connectivity. Environmental Modelling and

Software 24: 135-139.

Sawada, M. 1999. ROOKCASE: An Excel 97/2000 Visual Basic (VB) add-in for exploring

global and local spatial autocorrelation. Bulletin of the Ecological Society of America

80: 231-235.

Page 186: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

172

Scarascia-Mugnozza, G.; Oswald, H.; Piussi, P.; Radoglou, K. 2000. Forests of the

Mediterranean region: Gaps in knowledge and research needs. Forest Ecology and

Management 132: 97-109.

Schamp, R.; Laird, R.A.; Aarssen, L.W. 2002. Fewer species because of uncommon habitat?

Testing the species pool hypothesis for low plant species richness in highly productive

habitats. Oikos 97: 145-152.

Scheller, R.M. and Mladenoff, D.J. 2002. Understory species patterns and diversity in old-

growth and managed northern hardwood forests. Ecological Applications 12(5): 1329-

1343.

Schmid, B.; Joshi, J.; Schlipfer, E. 2002. Empirical evidence for biodiversity-ecosystem

functioning relationships. In: Kinzig, A.; Tilman, D.; Pacala, S. (Eds.), Functional

consequences of biodiversity: Experimental progress and theoretical extensions.

Princeton University Press, Princeton, New Jersey, pp. 120-150.

Schumann, M.A.; White, A.S.; Witham, J.W. 2003. The effects of harvest-created gaps on

plant species diversity, composition, and abundance in a Maine oak-pine forest. Forest

Ecology and Management 176: 543-561.

Serrada, R.; Montero, G.; Reque, J.A. 2008. Compendio de selvicultura aplicada en España.

Fundación Conde del Valle Salazar, Madrid.

Shachak, M.; Boeken, B.; Groner, E.; Kadmon, R.; Lubin, Y.; Meron, E.; Neeman, G.;

Perevolotsky, A.; Shkedy, Y.; Ungar, E.D. 2008. Woody species as landscape

modulators and their effect on biodiversity patterns. Bioscience 58: 209-221.

Shea, K.; Roxburgh, S.H.; Rauschert, E.S.J. 2004. Moving from pattern to process:

Coexistence mechanisms under intermediate disturbance regimes. Ecology Letters 7:

491-508.

Shipley, B. 2000. Cause and correlation in biology. Cambridge, UK.

Sluiter, R. and de Jong, S.M. 2007. Spatial patterns of Mediterranean land abandonment and

related land cover transitions. Landscape Ecology 22: 559-579.

Smith, A.C.; Koper, N.; Francis, C.M.; Fahrig, L. 2009. Confronting collinearity: Comparing

methods for disentangling the effects of habitat loss and fragmentation. Landscape

Ecology 24: 1271-1285.

Sobek, S.; Steffan-Dewenter, I.; Scherber, C.; Tscharntke, T. 2009. Spatiotemporal changes of

beetle communities across a tree diversity gradient. Diversity and Distributions 15: 660-

670.

Page 187: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

173

Sousa, W.P. 1984. The role of disturbance in natural communities. Annual Review of

Ecology and Systematics 15: 353-391.

Spanos, K.A.; Feest, A.; Petrakis, P.V. 2009. Improving the assessment and monitoring of

forest biodiversity. Management of Environmental Quality 20: 25-63.

Spiegelhalter, D.T.A. and Best, N. 2000. WinBUGS User Manual. MRC Biostatistics Unit,

Cambridge.

Spies, T.A. 1998. Forest structure: A key to the ecosystem. Northwest Science 72(2): 34-39.

Stephens, P.A.; Buskirk, S.W.; Hayward, G.D.; Del Rio, C.M. 2005. Information theory and

hypothesis testing: A call for pluralism. Journal of Applied Ecology 42: 4-12.

Suárez-Seoane, S.; Osborne, P.E.; Baudry, J. 2002. Responses of birds of different

biogeographic origins and habitat requirements to agricultural land abandonment in

northern Spain. Biological Conservation 105: 333-344.

Svenning, J.C. and Skov, F. 2002. Mesoscale distribution of understorey plants in temperate

forest (Kalo, Denmark): The importance of environment and dispersal. Plant Ecology

160: 169-185.

Tapias, R.; Climent, J.; Pardos, J.A.; Gil, L. 2004. Life histories of Mediterranean pines. Plant

Ecology 171: 53-68.

Tárrega, R.; Calvo, L.; Taboada, A.; García-Tejero, S.; Marcos, E. 2009. Abandonment and

management in Spanish dehesa systems: Effects on soil features and plant species

richness and composition. Forest Ecology and Management 257: 731-738.

Taylor, P.D.; Fahrig, L.; Henein, K.; Merriam, G. 1993. Connectivity is a vital element of

landscape structure. Oikos 68: 571-573.

TEEB. 2009. The economics of ecosystems and biodiversity for national and international

policy makers—Summary: Responding to the value of nature. Welzel and Hardt,

Wesseling.

Tellería, J.L.; Ramírez, A.; Pérez-Tris, J. 2005. Conservation of seed-dispersing migrant birds

in Mediterranean habitats: Shedding light on patterns to preserve processes. Biological

Conservation 124: 493-502.

Thomas, A.; O’Hara, R.; Ligges, U.; Sturts, S. 2006. Making BUGS open. R News 6: 12-17.

Tilman, D. 1986. Evolution and differentiation in terrestrial plant communities: The

importance of the soil resource: light gradient. In: Diamond, J. and Case, T.J. (Eds.),

Community Ecology. Harper and Row, New York, pp.359-380.

Page 188: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

174

Tischendorf, L. 2001. Can landscape indices predict ecological processes consistently?

Landscape Ecology 16: 235-254.

Torras, O. and Saura, S. 2008. Effects of silvicultural treatments on forest biodiversity

indicators in the Mediterranean. Forest Ecology and Management 255: 3322-3330.

Torras, O.; Martín-Queller, E.; Saura S. 2009. Relating landscape structure, environment and

management to biodiversity indicators estimated from forest inventory data in

Catalonia. Forest systems 18(3): 322-337.

Turner, M.G. 1989. Landscape ecology: The effect of pattern on process. Annual Review of

Ecology, Evolution and Systematics 20: 171-197.

Urban, D.L. and Keitt, T. 2001. Landscape connectivity: A graph-theoretic perspective.

Ecology 82(5): 1205-1218.

Urban, D.L.; Minor, E.S.; Treml, E.A.; Schick, R.S. 2009. Graph models of habitat mosaics.

Ecology Letters 12: 260-273.

Urbieta, I.R.; Zavala, M.A.; Marañón, T. 2008. Human and non-human determinants of forest

composition in southern Spain: Evidence of shifts towards cork oak dominance as a

result of management over the past century. Journal of Biogeography 35: 1688-1700.

Urbieta, I.R.; García, L.V.; Zavala, M.A.; Marañón, T. 2011. Mediterranean pine and oak

distribution in southern Spain: Is there a mismatch between regeneration and adult

distribution? Journal of Vegetation Science 22: 18-31.

Valbuena-Carabaña, M.; López de Heredia, U.; Fuentes-Utrilla, P.; Gónzalez-Doncel, I.; Gil,

L. 2010. Historical and recent changes in the Spanish forests: A socio-economic

process. Review of Palaeobotany and Palynology 162: 492-506.

Valladares, F.; Camarero, J.J.; Pulido, F.J.; Gil-Pelegrín, E. 2004a. El bosque mediterráneo,

un sistema humanizado y dinámico. In: Valladares, F. (Ed.), Ecología del bosque

mediterráneo en un mundo cambiante. Ministerio de Medio Ambiente, EGRAF SA,

Madrid, pp. 13-25.

Valladares, F.; Aranda, I.; Sánchez-Gómez, D. 2004b. La luz como factor ecológico y

evolutivo para las plantas y su interacción con el agua. In: Valladares, F. (Ed.),

Ecología del bosque mediterráneo en un mundo cambiante. Ministerio de Medio

Ambiente, EGRAF SA, Madrid, pp. 335-369.

Valladares, F.; Zaragoza-Castells, J.; Sánchez-Gómez, D.; Matesanz, S.; Alonso, B.;

Portsmuth, A.; Delgado, A.; Atkin, O.K. 2008. Is shade beneficial for Mediterranean

Page 189: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

175

shrubs experiencing periods of extreme drought and late-winter frosts? Annals of

Botany 102(6): 923-933.

Valverde, T. and Silvertown, J. 1997. Canopy closure rate and forest structure. Ecology 78:

1555-1562.

Vandekerkhove, K.; De Keersmaeker, L.; Menke, N.; Meyer, P.; Vershchelde, P. 2009. When

nature takes over from man: Dead wood accumulation in previously managed oak and

beech woodlands in North-western and Central Europe. Forest Ecology and

Management 258: 425-435.

Vega-García, C. and Chuvieco, E. 2006. Applying local measures of spatial heterogeneity to

Landsat-TM images for predicting wildfire occurrence in Mediterranean landscapes.

Landscape Ecology 21: 595-605.

Vesseby, K.; Söderström, B.O.; Glimskar, A.; Svensson, B. 2002. Species-richness

correlations of six different taxa in Swedish seminatural Grasslands. Conservation

Biology 16(2): 430-439.

Vetaas, O.R. and Ferrer-Castán, D. 2008. Patterns of woody plant species richness in the

Iberian Peninsula: Environmental range and spatial scale. Journal of Biogeography 35:

1863-1878.

Vilà, M.; Vayreda, J.; Comas, L.; Ibáñez, J.J.; Mata, T.; Obón, B. 2007. Species richness and

wood production: A positive association in Mediterranean forests. Ecology Letters 10:

241-250.

Vogiatzakis, I.N.; Mannion, A.M.; Griffiths, G.H. 2006. Mediterranean ecosystems: Problems

and tools for conservation. Progress in Physical Geography 30: 175-200.

Wagner, H.H. and Fortin, M. 2005. Spatial analysis of landscapes: Concepts and statistics.

Ecology 86: 1975-1987.

Wagner, S.; Fischer, H.; Franka, H. 2011. Canopy effects on vegetation caused by harvesting

and regeneration treatments. European Journal of Forest Research 130: 17-40.

Wheeler, D. and Tiefelsdorf, M. 2005. Multicollinearity and correlation among local

regression coefficients in geographically weighted regression. Journal of Geographical

Systems 7: 161-187.

Whittaker, R.J.; Willis, K.J.; Field, R. 2001. Scale and species richness: Towards a general,

hierarchical theory of species diversity. Journal of Biogeography 28: 453-470.

Whittaker, R.J. and Heegaard, E. 2003. What is the observed relationship between species

richness and productivity? Comment. Ecology 84: 3384-3390.

Page 190: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

176

Whittaker, R.J.; Nogués-Bravo, D.; Araújo M.B. 2007. Geographical gradients of species

richness: A test of the water-energy conjecture of Hawkins et al. (2003) using European

data for five taxa. Global Ecology and Biogeography 16: 76-89.

Willig, M.R.; Kaufman, D.M.; Stevens, R.D. 2003. Latitudinal gradients of biodiversity:

Pattern, process, scale and synthesis. Annual Review of Ecology, Evolution and

Systematics 34: 273–309.

Wilson, D.S. 1992. Complex interactions in metacommunities, with implications for

biodiversity and higher levels of selection. Ecology 73: 1984-2000.

Wilson, E.O. and Willis, E.O. 1975. Applied biogeography. In: Cody, M.L. and Diamond,

J.M. (Eds.), Ecology and Evolution of Communities. Belknap Press, Cambridge.

Wohlgemuth, T. 1998. Modelling floristic species richness on a regional scale: A case study

in Switzerland. Biodiversity and Conservation 7: 159-177.

Wright, J.P. and Jones, C.G., 2006. The concept of organisms as ecosystem engineers ten

years on: Progress, limitations, and challenges. BioScience 56: 203-209.

Wright, S.J. 2002. Plant diversity in tropical forests: A review of mechanisms of species

coexistence. Oecologia 130: 1–14.

Wu, J.; Jones, K.B.; Li, H.; Loucks, O. 2006. Scaling and uncertainty analysis in ecology.

Springer, Netherlands.

Zamora, R.; García-Fayos, P.; Gómez-Aparicio, L. 2004. Las interacciones planta-planta y

planta animal en el contexto de la sucesión ecológica. In: Valladares, F. (Ed.), Ecología

del bosque mediterráneo en un mundo cambiante. Ministerio de Medio Ambiente,

EGRAF SA, Madrid, pp. 371-393.

Zavala, M.A. 2004. Estructura, dinámica y modelos de ensamblaje del bosque mediterráneo:

entre la necesidad y la contingencia. In: Valladares, F. (Ed.), Ecología del bosque

mediterráneo en un mundo cambiante. Ministerio de Medio Ambiente, EGRAF SA,

Madrid, pp. 249-277.

Zavala, M.A.; Zamora, R.; Pulido, F.J.; Blanco, J.A.; Imbert, J.B.; Marañón, T.; Castillo, F.J.;

Valladares, F. 2004. Nuevas perspectivas en la conservación, restauración y gestión

sostenible del bosque mediterráneo. In: Valladares, F. (Ed.), Ecología del bosque

mediterráneo en un mundo cambiante. Ministerio de Medio Ambiente, EGRAF SA,

Madrid, pp. 509-529.

Page 191: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

REFERENCES

177

Zenner, E.K.; Kabrick, J.M.; Jensen, R.G.; Peck, J.E.; Grabner, J.K. 2006. Responses of

ground flora to a gradient of harvest intensity in the Missouri Ozarks. Forest Ecology

and Management 222: 326-334.

Page 192: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 193: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

APPENDIX A

179

Appendix A. Distribution patterns of the main tree species in the study area, as analyzed

in chapters 2, 3 and 4. The points represent those 3SNFI plots where each tree species

was dominant.

Page 194: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes
Page 195: PATRONES DE DIVERSIDAD DE ESPECIES LEÑOSAS EN LOS …oa.upm.es/9873/1/Emilia_Martin_Queller.pdf · patrones de diversidad de especies leÑosas en los bosques espaÑoles a diferentes

APPENDIX B

181

Appendix B. Distribution patterns of the main regeneration treatments applied in the

study area, as analyzed in chapters 2, 3 and 4.