© 2016 judit ungvari-martin - university of...

168
ENVIRONMENTAL HETEROGENEITY AND THE ORIGIN AND MAINTENANCE OF REGIONAL AVIAN DIVERSITY IN WESTERN AMAZONIA By JUDIT UNGVARI-MARTIN A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2016

Upload: others

Post on 22-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

ENVIRONMENTAL HETEROGENEITY AND THE ORIGIN AND MAINTENANCE OF REGIONAL AVIAN DIVERSITY IN WESTERN AMAZONIA

By

JUDIT UNGVARI-MARTIN

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2016

Page 2: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

© 2016 Judit Ungvari-Martin

Page 3: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

To my trusty external hard drive

Page 4: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

4

ACKNOWLEDGMENTS

I want to thank my parents in my mother tongue: Köszönöm szépen a rengeteg

türelmet! Additionally, I also want to thank my committee for their patience, especially J.

Gordon Burleigh who pushed me through when needed, Scott K. Robinson who

supported me throughout this arduous but fun journey, Bette Loiselle, David Reed and

Per Lundberg who made themselves available during the short periods of time when

everyone could be around and provided valuable knowledge, feedback and support. Kitt

Heckscher and Keith Hobson were essential for the successful publication of the Grey-

cheeked Thrush data. The funding sources were the following: the Wilson Ornithological

Society, the Cooper Ornithological Society, Sigma Xi, the American Ornithologists

Union, the Florida Museum of Natural History Ordway Endowment, and the National

Science Foundation Graduate Research Fellowship (DGE-0802270). SERNANP

granted permits to perform the fieldwork and data collections. Information in Chapter 2

was provided by NatureServe (www.natureserve.org) and its network of natural heritage

member programs, a leading source of information about rare and endangered species,

and threatened ecosystems. I must thank immensely my family near and far, my lab

mates, my friends, loved ones, and the counseling center who kept me borderline sane,

and my pets who provided the much needed affection. Families in the communities of El

Dorado, Nueva Esperanza, San Martin, Mishana, Yuto and Porvenir hosted our field

crews for weeks on end, and fed us nourishing meals made of simple ingredient.

Throughout the four years of fieldwork, 44 field assistants helped out in Peru with data

collections, without their sweat and tears this work would not have been possible.

Page 5: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

5

TABLE OF CONTENTS page

ACKNOWLEDGMENTS .................................................................................................. 4

LIST OF TABLES ............................................................................................................ 7

LIST OF FIGURES .......................................................................................................... 9

LIST OF ABBREVIATIONS ........................................................................................... 12

ABSTRACT ................................................................................................................... 13

CHAPTER

1 MOTIVATION ......................................................................................................... 15

Overview ................................................................................................................. 15

Outline .................................................................................................................... 18 Purpose .................................................................................................................. 20

2 PATTERNS OF SPECIES CO-OCCURRENCE FROM A PHYLOGENETIC PERSPECTIVE ....................................................................................................... 22

Introduction ............................................................................................................. 22

Methods .................................................................................................................. 27

Species Lists .................................................................................................... 27

Species Composition Comparisons .................................................................. 28 Phylogenetic Composition Comparisons .......................................................... 29

Results .................................................................................................................... 31 Species richness .............................................................................................. 31 Betadiversity ..................................................................................................... 32

Phylobetadiversity ............................................................................................ 32 Discussion .............................................................................................................. 35

3 STRUCTURE OF UNDERSTORY BIRDS COMMUNITIES IN FORESTS ON TWO CONTRASTING SOIL TYPES IN AMAZONIAN PERU ................................. 50

Introduction ............................................................................................................. 50

Methods .................................................................................................................. 54 Study Region .................................................................................................... 54 Bird Surveys ..................................................................................................... 55 Avian Guild Classification ................................................................................. 55 Data Summary and Analyses ........................................................................... 57

Results .................................................................................................................... 59 Community-wide Patterns ................................................................................ 59 Vegetation Structure ......................................................................................... 62

Page 6: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

6

Avian Guilds ..................................................................................................... 63

Discussion .............................................................................................................. 64

4 MOVEMENT, ISOLATION AND GENE FLOW AMONG PATCHES OF WHITE SAND FORESTS .................................................................................................... 86

Introduction ............................................................................................................. 86 Methods .................................................................................................................. 90

Study Landscape .............................................................................................. 90

Species-area Relationships .............................................................................. 90 Molecular Analysis ........................................................................................... 90 Population structure analysis ............................................................................ 92

Results .................................................................................................................... 94 Movements detected by recaptures ................................................................. 94

Species – area relationships ............................................................................ 95 Genetic diversity and population assignment tests ........................................... 95

Discussion .............................................................................................................. 96

5 INTER-ANNUAL SITE FIDELITY AND BREEDING ORIGINS OF GRAY-CHEEKED THRUSHES IN WHITE SAND FORESTS OF THE PERUVIAN AMAZON .............................................................................................................. 107

Introduction ........................................................................................................... 107

Methods ................................................................................................................ 109 Study Site ....................................................................................................... 109

Field Sampling ................................................................................................ 109 Stable Isotope Analysis .................................................................................. 111

Isotopic Assignment ....................................................................................... 112 Results .................................................................................................................. 112

Discussion ............................................................................................................ 114

6 CONCLUDING REMARKS ................................................................................... 129

APPENDIX

A LIST OF SPECIES CAPTURED IN AMNR FORESTS ......................................... 133

B SUMMARY OF RICHNESS ESTIMATORS .......................................................... 138

C BETWEEN-YEAR RETURN RATES IN WSF AND CLAY FOREST .................... 139

D SUPPLEMENTARY INFORMATION OF DDRADSEQ ANALYSIS ...................... 141

LIST OF REFERENCES ............................................................................................. 148

BIOGRAPHICAL SKETCH .......................................................................................... 168

Page 7: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

7

LIST OF TABLES

Table page 2-1 Ecoregions and habitat types used in this study ................................................. 42

2-2 Summary of predictions of the four hypotheses tested. ...................................... 43

3-1 Vegetation structure environmental variables used for comparisons of the four habitat types. All values represent mean values for habitat types ± standard deviations ............................................................................................ 71

3-2 Dietary guild comparisons of the four habitat types. We assessed community composition using broad dietary categories and no differences were found in the trophic organization of the understory bird community. ................................ 72

3-3 Capture rate comparisons for all birds and for select guilds or functional groups in the four habitat types using individuals per 500 net hour units. All values represent mean number of individual birds ............................................. 73

3-4 Diversity indices comparisons of bird communities of the four habitat types for all species and for white sand forest specialists alone from the AMNR capture data........................................................................................................ 74

4-1 The twenty most captured bird species, and the recaptures and movement proportions of those species. Movements are defined as recaptures on a different netline. .................................................................................................. 99

4-2 Localities, area of WSF patches, distances to nearest patch, and area of nearest patch to the sampled patch. Lat/lon coordinates are in UTM geographic coordinate system, UTM 18S. ........................................................ 100

4-3 Model selection results of the explanatory values of area of WSF patches, distances to nearest patch, and area of nearest patch to the sampled patch. .. 101

4-4 Summary of sampling locality, samples in the locality, and summary of raw sequence information for those samples. ......................................................... 102

4-5 Summary of population differentiation results based on FST calculations. ........ 103

5-1 Captured and recaptured Gray-cheeked Thrushes (Catharus minimus) in Amazonian white sand forests (wsf) and clay terra firme forests (clay) in Allpahuayo-Mishana National Reserve ............................................................. 120

A-1 List of species captured in forest in Allpahuayo-Mishana during 2009-2012 with raw capture numbers for each species ..................................................... 133

Page 8: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

8

B-1 Incidence-based richness estimators with standard deviations of the four different habitat types in AMNR. ....................................................................... 138

C-1 Comparison of return rates for all species with a minimum of 5 captures. All values represent mean number of individual birds in the habitat types ± standard deviations. Habitat specialists are indicated in the last column. ........ 139

D-1 Run metrics summary from the High throughput NextSeq 500 run of single end 150 bp reads. ............................................................................................. 142

D-2 Raw sequencing information for samples after quality filtering steps from the catalog of RAD loci. .......................................................................................... 143

D-3 Summary statistics for the 129 individuals included in the final filtered dataset, 2240 biallelic loci included in the dataset. ........................................... 147

Page 9: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

9

LIST OF FIGURES

Figure page 1-1 The edge of the Nueva Esperanza (AMNR, Loreto, Peru) field site in 2011,

where forest stood in previous years. Photo courtesy of Cristian Gallego Carmona. ............................................................................................................ 21

2-1 Map of the three ecoregions used for assembling the regional communities. The labeled and outlined ecoregions all belong to the moist tropical forest biome. ................................................................................................................. 44

2-2 Network representation of the cluster of communities that are taxonomically more similar to each other than expected by chance. The color of the spheres represents the ecoregion (red=NAPO, blue=SW, black=GUI). ............. 45

2-3 Cluster dendrograms using taxonomic betadiversity and phylobetadiversity for all local bird communities in different habitats across the three ecoregions .. 46

2-4 Clusters of communities in habitats that are phylogenetically more similar to each other than expected by chance. This network representation of all the phylobetadiversity comparison of the 33 habitats in 3 ecoregions. ..................... 47

2-5 Within-ecoregion betadiversity phylogenetic betadiversity. A) Betadiversity results using taxonomic turnover. B) Phylogenetic betadiversity results within each ecoregion. .................................................................................................. 48

2-6 Relationship between phylogenetic betadiversity (PBD, black points) betadiversity (BC, blue points), and species richness. The lines represent best fit lines corresponding to linear regressions. ............................................... 49

3-1 Comparison of number of species between two general habitat types. Slightly more species were captured in clay soil forests than in white sands forests per sampling event (t-test, p= 0.01043) .............................................................. 75

3-2 Rank abundance curves for white sand forests and clay forest captures. Species are arranged according to their rank order abundance in clay soil forest, since terra firme is the dominant forest type. ........................................... 76

3-3 Forest canopy height in the four different habitat types, letter codes refer to statistical differences (ANOVA, p<0.05), red points indicate outlier values, letter codes a-b-c-d indicate statistically different values. ................................... 77

3-4 Cumulative distribution of species occurrences over all netlines (n=93 sampling events). Each point represent a species, and the number of sampling events at which that species was recorded. ........................................ 78

Page 10: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

10

3-5 Species accumulation curves for four different habitat types. Rarefaction analysis was performed to standardize comparisons with effort. ........................ 79

3-6 Turnover, dissimilarity and nestedness between white sand and clay soil forest bird communities. ..................................................................................... 80

3-7 Ordination diagram based on Bray-Curtis Dissimilarity using netlines as sites for all WSF and clay habitats. The grey lines connecting the points result from the hierarchical clustering analysis using the average linkage method. ..... 81

3-8 Comparison of capture rates and number of observed species of flocking or non-flocking shrub gleaner insectivores. A) Capture rates B) Observed number of species per guild in each habitat type ................................................ 82

3-9 Comparisons of captures and species according to foraging methods. A) Capture rates B) The observed number of species per guild ............................. 83

3-10 Comparisons of captures and species number based on dietary categories. A) Capture rates of nectarivores and frugivores B) The observed number of species per guild. ................................................................................................ 84

3-11 Rarefaction based species accumulation curves show the differences for white sand forest habitats and clay habitat. Overlapping 95% confidence intervals are interpreted as no difference. .......................................................... 85

4-1 Map of white sand forest patches including the 150 m buffer from the edge for each forest island. The names of the putatively isolated regions are included on the map. ........................................................................................ 104

4-2 The relationship between species richness and the area of the sampled white sand patch, the area of the nearest WSF patch and the distance to the nearest WSF patch. Blue points represent WSF specialist species. ................ 105

4-3 Population genetics results based on k-means clustering as well as principal component analysis based on allele frequencies indicate little differentiation between regional populations, and a single genetic cluster. ............................. 106

5-1 Location of Allpahuayo-Mishana National Reserve (outline), Department of Loreto, southeast of Iquitos, Peru. White sand forests are the small polygons outlined in black (image source: Google Earth). ............................................... 124

5-2 Weekly numbers of Gray-cheeked Thrushes by age caught using constant effort mistnet transects at Allpahuayo-Mishana National Reserve, Department of Loreto, Peru, 13 October 2010 to 8 December 2012. ............... 125

5-3 Cumulative number of Gray-cheeked Thrushes (Catharus minimus) and Swainson’s Thrushes (C. ustulatus) captured over a nine week period in 2010, 2011, and 2012 at Allpahuayo-Mishana National Reserve ..................... 126

Page 11: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

11

5-4 Cumulative recaptures by week for Gray-cheeked Thrushes using constant effort mistnet transects at Allpahuayo-Mishana National Reserve, Department of Loreto, Peru. ............................................................................. 127

5-5 Depictions of likely breeding or natal origins of Gray-cheeked Thrushes at Allpahuayo-Mishana National Reserve, Department of Loreto, Peru. The spatially explicit assignment of the population (n=12)....................................... 128

A-1 Comparisons of WSF and clay and the four specific habitat types by capture rate. Neither A) Capture rates between clay and WSF nor B) captures rates among the four specific habitat types in AMNR differed. .................................. 137

B-1 Graphical representation of incidence based richness estimators with confidence intervals for all sampling localities and habitats combined. ............ 138

Page 12: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

12

LIST OF ABBREVIATIONS

AMNR Allpahuayo Mishana National Reserve

BCD Bray-Curtis dissimilarity

bp Base pairs

c.i. Confidence Interval

CC Cocha Cashu

dbh Diameter at breast height

ddRADseq Double Digest RAD Sequencing

Gbp Giga base pairs

GUI Guianan Moist Forest Ecoregion

HWE Hardy - Weinberg Equilibrium

km kilometer

LD Linkage Disequilibrium

m Meter

MCMC Markov chain Monte Carlo

ML Maximum Likelihood

PBD Phylobetadiversity

PCR Polymerase chain reaction

SD Standard Deviation

SW Southwest

WSF White sand forest

δ2H Stable hydrogen isotope

Page 13: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

13

Abstract of Dissertation Presented to the Graduate School of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

ENVIRONMENTAL HETEROGENEITY AND THE ORIGIN AND MAINTENANCE OF

REGIONAL AVIAN DIVERSITY IN WESTERN AMAZONIA

By

Judit Ungvari-Martin

May 2016

Chair: Scott K Robinson Major: Biology

The number of species that can co-occur in a place depends on environmental

complexity and on species interactions. Although the vast region of Amazonia has a

relatively uniform climate, the rich landscape heterogeneity creates a patchwork of

habitats with different bird communities. The objective of this study was to examine the

origin and maintenance of Amazonian biodiversity. First, the distribution of bird species

was examined using taxonomic betadiversity and phylogenetic betadiversity to explore

broad-scale patterns of diversity in different habitats across ecoregions. Then, on a

landscape level I examined how habitat specialization affects patterns of species

composition and trophic structure, focusing on two contrasting habitats: clay soil forests

and white sand forests (WSF). These bird communities differed in species richness,

which may be related to habitat productivity, but the community structure was similar.

White sand forests are patchy, though widely distributed, and they occur in an

island-like landscape with different-sized areas and degrees of isolation. If species

disperse little among patches, their isolation may promote speciation. We performed a

mark-recapture study in a reserve in western Amazonia to examine the movement of

understory birds among white sand forest patches and found that 3.6% of all recaptures

Page 14: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

14

involved movements among different netlines. However, we found evidence of high

gene flow among isolated populations of an obligate WSF specialist species, which

suggests that the surrounding clay soil forests do not pose a strong barrier to dispersal

for the obligate WSF specialist Neopelma chrysocephalum on the scale of 1-10km. We

also found evidence that WSF are important for long-distance migrants, and identified

that Catharus minimus uses WSF.

Together, these data suggest that although diversity of habitats is critical to the

maintenance of the high regional species diversity of western Amazonia, communities in

forested habitats are similar in structure with the addition of a few habitat specialists and

limited turnover. Also, forested habitat matrix is permeable for some habitat specialist

species, which may slow the rates of speciation, even in seemingly isolated patches.

Page 15: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

15

CHAPTER 1 MOTIVATION

Overview

The search for patterns in the spatial organization of species diversity and the

identification of the mechanisms that underlie those patterns have long been central

concerns of community ecologists. Community ecology has undergone significant

paradigms shifts, and theory has developed faster than empirical tests (Logue et al.

2011). Ricklefs proposed the concept of “regional communities” that highlights

communities made up of populations of species that are themselves distributed across

potential ecological and geographical gradients (Ricklefs 2008). The regional

community concept, which is strongly analogous to the concept of metacommunity

(Leibold et al. 2004, Holyoak et al. 2005), has expanded the scope of community

ecology both spatially and temporally by integrating global-scale analyses with multiple

time scales that approach the realm of biogeography (Leibold et al. 2004, Ricklefs 2008,

Ricklefs 2011).

Ecoregions, defined based on large-scale patterns of floristic and zoogeographic

variation, provide broad categories of all major terrestrial, freshwater and marine habitat

types spanning all continents and oceans (Olson et al. 2001). Ecoregions provide the

context for well-defined regional communities. Local communities are the traditional

focus of community ecology; these local communities consist of a subset of the species

found within the regional species pool. Local communities within a region can vary

greatly in composition, even when they are close together. We use the concept of local

communities as the subsets of species that live in a particular habitat within the tropical

moist forest biome of Amazonia.

Page 16: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

16

The Amazon basin is over seven million square kilometers, five million of which

are covered in tropical rainforest. The region displays some variation in rainfall,

temperature and geology, and some researchers have emphasized the biogeographical

and evolutionary importance of this variation in explaining high levels of regional

diversity (Tuomisto and Ruokolainen 1997, Tuomisto 2007). An extensive river system

dominates the region, and there are well known chemical, physical and biological

differences between eutrophic, nutrient rich white-water rivers and oligotrophic, nutrient

poor black-water and clear-water rivers (Janzen 1974, Tuomisto et al. 1995).

These rivers are of great importance; they alter the soil conditions by seasonally

flooding some areas, which deposits nutrient-rich sediments (in the case of white-water

rivers) or renders the soils infertile with high humic acid concentrations (in the case of

black-water rivers), thus creating a mosaic of patches in the region (Janzen 1974).

Environmental heterogeneity produces habitat diversity in the Amazonian region, which,

in turn, contributes to regional species diversity (-diversity) (Borges 2004, Cramer and

Willig 2005, Andersen et al. 2010, Tuomisto 2010a). Within heterogeneous landscapes,

many, if not most habitat types harbor endemic plant and animal species (Steege et al.

2000, Fine et al. 2005, Ruokolainen et al. 2007).

One of the most dramatic recent discoveries is the extent to which communities

that grow on certain kinds of soils differ in their species composition. White sands

forests (hereafter WSF), in particular, have been shown to have many specialist birds

and plants (Alvarez Alonso and Whitney 2001, Fine et al. 2005, Poletto and Aleixo

2005, Fine et al. 2010, Alvarez Alonso et al. 2013, Fine and Baraloto 2016) and to be

lower in productivity than soils growing on more nutrient-rich clay soils that dominate

Page 17: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

17

most sections of the Amazon Basin (Tuomisto and Poulsen 1996, Asner et al. 2000,

Fine et al. 2006, Higgins et al. 2014). White-sand forests are patchily distributed

throughout northern South America in Brazil, Peru, Guyana, Venezuela, French Guiana,

Suriname and Colombia as well. The sediments making up these soils eroded from the

Guiana Shield and were transported westward into the basin of the Nanay River of Peru

and to the Upper Rio Negro region of Brazil during the early Miocene (Janzen 1974,

Kalliola et al. 1993). WSF soil is composed almost solely of sand and has less than 5%

clay (Anderson 1981). There are a variety of vernacular terms to describe sand-based

vegetation forms have a variety of vernacular terms describing them depending on the

regions. They are called varillales in Peru, caatingas or campinas/campinaranas

sometimes also humirizales, carrascales, chavascales or charavascales in Brazil, muri

bush or wallaba forest in Guyana, and bana, cunuri or yaguácanan in Venezuela

(Anderson 1981, Adeney et al. 2016, Daly et al. 2016).

White-sand soils are extremely nutrient-poor (Janzen 1974) and also quick-

drying because sand has little water-holding capacity. The rapid passage of water

through the soil leaches organic matter and clay particles; making the soils deficient in

minerals (Anderson 1981, Tuomisto 2006, Andersen et al. 2010). Even within the

climatically almost uniform rain forests it has been observed that very few plant species

can grow on all kinds of soil (Brown and Prance 1987, Coomes and Grubb 1996,

Benavides et al. 2006, Andersen et al. 2010, Garcia-Villacorta et al. 2016). These

nutrient-poor white sand soils that cover less than three percent of Amazonia support

resource-limited plant species that have slow growth rates. This heterogeneity in

vegetation influences the evolution and interactions among species through niche-

Page 18: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

18

based processes and habitat based biotic filtering even in the absence of strong climatic

environmental gradients (Chase and Myers 2011, Kraft et al. 2014).

Outline

Studies of community assembly seek to identify the processes that determine

how large regional species pools are filtered into local communities, and conversely,

how processes that occur in the local community affect regional biodiversity. On both

regional and local scales, community assembly is influenced by ecological processes

such as environmental heterogeneity, interspecific interactions, and dispersal limitation.

The following questions were explored in this dissertation: (1) what are the patterns of

species co-occurrence in local bird communities of the different habitat types that are

found in lowland rainforests and (2) how we can interpret and explain these patterns of

distribution in terms of ecological and evolutionary processes. The objective of this

project was to improve our understanding of the incredibly high alpha, beta and gamma

diversity of western Amazonian birds in a region that is widely considered to be the

most regionally diverse lowland avian community on earth where more than 700

species occur in lowland forests that vary little in abiotic factors (elevation, rainfall and

temperature) (Stotz 1996, Schulenberg et al. 2007).

The first specific objective was to examine the taxonomic and phylogenetic

turnover of bird assemblages at the regional spatial scale in various habitats within

different ecoregions in Amazonia (Webb et al. 2002, Cramer and Willig 2005, Gotelli

and Ulrich 2012). Understanding the turnover of species composition between habitat

types (beta diversity) within a region (gamma diversity) can help elucidate the

mechanisms driving the origin and maintenance of species diversity (Kraft et al. 2011,

Meynard et al. 2011). We used an integrated framework for linking variation in the

Page 19: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

19

species and phylogenetic composition of the ecoregions and habitats within a broader

regional community. Habitat preference plays an important role in speciation and

diversification (Futuyma and Moreno 1988, Barnagaud et al. 2011). Among birds,

specialization to a specific habitat evolved multiple times, because species from various

distant lineages can exhibit similar preferences for a certain habitat type (Terborgh et al.

1990, Kratter 1997, Alvarez Alonso and Whitney 2003).

Our second objective was to compare community composition in the WSFs of the

Peruvian Amazon to the dominant forest type of terra firme to improve our

understanding of the selective forces that may lead to the evolution of habitat

specialization. Chapter 3 discusses the regional dynamics and differentiation of bird

communities in the Allpahuayo-Mishana Natural Reserve (AMNR), where fieldwork was

conducted between 2009 and 2012. This reserve harbors some of the most extensive

systems of WSF patches in Western Amazonia and is a site in which 5 newly described

species were recently discovered in a forests less than 30 km from Iquitos, a city of

more than 500,000 people (Alvarez Alonso and Whitney 2001, Alvarez Alonso 2002,

Alvarez Alonso and Whitney 2003, Salo and Pyhala 2007, Alvarez Alonso et al. 2012).

In Chapter 4 we evaluated how patch size and isolation affected species richness

of WSF specialists whether gene flow is maintained among patches that vary widely in

their degree of isolation, and whether the degree of habitat specialization affects the

amount of gene flow among populations. In Chapter 5 we report the discovery that white

sand forests are important wintering sites for migratory thrushes. These latter chapters

focus on selected white sand forest habitat specialists.

Page 20: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

20

Purpose

Despite the overall climatic stability in the lowland tropical moist forest biome,

habitat heterogeneity contributes to the biological diversity observed in the Amazon

basin. Many endemic species are associated with just one or a few habitats such as

bamboo or Mauritia palm swamps, but other species seem to be generally associated

with habitats of lower productivity. The variation of Amazonian habitats has contributed

to the processes of speciation and diversification in the region (Brumfield and

Capparella 1996, Nores 2000, Marks et al. 2002, Fine et al. 2005, Borges 2007, Naka

2011).

WSFs had received little attention until the last decade, and their animal

communities are still not well studied. Recent studies, however, have reported new bird

species (Alonso and Whitney 2001), endemic plant species (Frasier et al. 2008, Fine

and Baraloto 2016), and the presence of restricted-range bird species in WSFs (Alonso

and Whitney 2003). WSF is a unique ecosystem that is particularly vulnerable to human

disturbance; the tough woody vegetation is desirable for construction, and nearby cities

and towns threaten the forest with increased logging, and the extraction of sand and

gravel from below ground (Fig. 1-1). Amazonian WSFs are rare and naturally

fragmented. Therefore, they should be a conservation priority for their value for

harboring unique species that increase the regional biodiversity of the lowland tropical

moist forest biome.

Page 21: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

21

Figure 1-1. The edge of the Nueva Esperanza (AMNR, Loreto, Peru) field site in 2011,

where forest stood in previous years. The forest was bulldozed away, all trees and topsoil were removed, and the sand and gravel were extracted for construction. Photo courtesy of Cristian Gallego Carmona.

Page 22: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

22

CHAPTER 2 PATTERNS OF SPECIES CO-OCCURRENCE FROM A PHYLOGENETIC

PERSPECTIVE

Introduction

The ever-expanding spatial and temporal scope of community ecology makes it

essential to consider the evolutionary processes responsible for species distributions,

which play an important role in community assembly and composition (Vellend 2010).

Applying concepts from evolutionary biology and phylogenetic approaches to

community ecology can help us understand the factors governing community assembly,

and has yielded new insights into the biodiversity of many taxa (Emerson and Gillespie

2008, Cavender-Bares et al. 2009, Graham et al. 2009, Mayfield and Levine 2010, Fine

and Kembel 2011, Kooyman et al. 2011, Peixoto et al. 2014, Weinstein et al. 2014). In

this Chapter, we seek to expand these studies to include the role of habitat

specialization in originating and maintaining regional diversity. The turnover of avian

species composition between different habitat types is governed by factors such as

landscape and vegetation (Jankowski et al. 2009, Pomara et al. 2012, Stegen 2013,

Bennett and Gilbert 2016). Various mechanisms can lead to speciation in birds, and

habitat alone generally does not (Edwards 2005), although ecological segregation

across habitats between closely related species (Veen 2010, Tene Fossog 2015) can

be a significant factor maintaining high biodiversity in small geographic regions.

In this study, we ask how habitat selection mediates the establishment of local

communities from regional species pools. Habitat choice and habitat specialization have

predictable consequences for community structure (Morris 1996, Cavender-Bares et al.

2009). For example, piscivorous organisms live near water, but they could be

specialized on different bodies of water (lake, river, small streams). We examine

Page 23: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

23

taxonomic β diversity (variation of species identities between sites) and

phylobetadiversity (variation in phylogenetic relatedness of communities between sites)

in three terrestrial ecoregions within the lowland tropical moist forest biome of Northern

South America, which harbors a large proportion of Earth’s terrestrial biodiversity (Olson

et al. 2001). Biodiversity is usually plotted as taxonomic richness of a geographic area

described by a certain index in reference to species numbers. Phylogenetic diversity

measures include additional information on phylogenetic relationships among species.

Using phylogenies can serve as a tool to generate hypotheses about processes that

lead to the assemblages that we observe.

Our study system is well known for high α diversity in single localities (Terborgh

et al. 1990). Environmental gradients (Pianka 1966, Brown et al. 2001, Tuomisto 2006,

Jankowski et al. 2009, Chase and Myers 2011, Kraft et al. 2011, Jetz and Fine 2012)

certainly play a role in the distribution of organisms; however, it is not clear how the little

variation in abiotic factors such as temperature, precipitation, or elevation affect

phylogenetic diversity and speciation in lowland tropical moist forests biomes such as

the Amazon basin. Climatically the Amazon basin is relatively uniform with few strong

abiotic gradients and there is little variation in elevation (<400m), annual mean

temperature, and precipitation (>2000mm) based on WorldClim v.1.4 data; yet, the

region is home to the most diverse terrestrial ecoregions on earth (Hijmans et al. 2005).

Amazonia contains 7 million square kilometers of land (Malhado et al. 2013), and

the formation of the current Amazon basin spans close to 60 million years with a

complex geologic history that is responsible for the isolation and connectivity of

replicated sets of habitats and communities (Hoorn et al. 2010). During this time major

Page 24: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

24

contributors to the formation of the basin were rivers that have changed their course,

the Andean uplift, formation of new rivers and wetlands, climatic changes, and global

sea level drops during the glacial periods (Hoorn et al. 2010). The area known today as

Amazonia was once part of a much larger “pan-Amazonian” region, which included the

area of the present Amazon, Orinoco, and Magdalena drainage basins. A diverse fauna

existed in this region, elements of which are now restricted to the Amazon basin. The

three ecoregions in our study - Napo Moist Forest (NAPO), Southwestern Amazonian

Moist Forest (SW) and Guianan Moist Forest (GUI) – include large expanses of

relatively undisturbed forest, and they all share the dynamic geologic history of the

formation the Amazon basin. Based on the large regional species pool, diversity of

habitats, heterogeneity of the landscapes and the relatively low variation of abiotic

factors, we can use the ecoregions of lowland tropical forest as a replicated study

system to evaluate the effects of habitat specialization on ecological and phylogenetic

dissimilarities of local communities.

We centered our investigation on four biological processes that are known to

affect community assembly: drift, speciation, dispersal, and selection. Our null

hypothesis is that the turnover of species between or within the local communities is not

different from random, which indicates that drift is responsible for species distributions.

Under this hypothesis we would predict that betadiversity and phylobetadiversity within

and between ecoregions and habitats would not be significantly different from random

sampling of the species pool. We test for an influence of deterministic processes by

comparing observed phylogenetic turnover between assemblages to a stochastic

expectation. Significant deviations from the stochastic expectation indicate that

Page 25: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

25

deterministic processes more strongly influence community composition. We use drift

here synonymously with stochastic processes (Orrock and Watling 2010, Chase and

Myers 2011). We then test for evidence of several deterministic processes underlying

community organization. First, the speciation hypothesis is based on a strong influence

of local diversification processes. Local speciation can have profound effects on species

richness (Emerson and Kolm 2005), betadiversity, and phylobetadiversity (Emerson et

al. 2011) and is quite plausible given the extended periods during which various

sections of the Amazon have been isolated from each other, mainly by rivers (Hoorn et

al. 2010). If local diversification dominates community structure and organization, then

we would predict: (a) high betadiversity (few shared species) between habitats within an

ecoregion and among ecoregions; and (b) low phylobetadiversity within ecoregions and

(c) random patterns of phylobetadiversity among ecoregions.

The second hypothesized deterministic mechanism is dispersal, which is based

on the effects of the movement of organisms across space and their influence on

species richness and turnover patterns (Donoghue 2008, Wang et al. 2013, Peixoto et

al. 2014). This process, which is analogous to the processes of gene flow, hypothesizes

that patterns of community assembly are determined by limitations in dispersal rather

than selection imposed by the different habitats themselves. Under this scenario,

distance, connectivity, and species-specific dispersal abilities would be the main factors

underlying species turnover. Specifically, if dispersal processes dominate community

composition, then we would predict high species turnover between habitats that are not

connected within an ecoregion, but low betadiversity between habitats that are

connected within and between ecoregions. For example, riverine habitats have high

Page 26: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

26

connectivity across large distances; local communities in river edge habitat, river

habitat, or along streams, therefore, should have lower turnover than expected by

chance. In contrast, isolated habitats such as white sand forests, bamboo patches or

inselbergs should be more different than any of the habitats we use for comparisons,

and terra firme forests that are isolated by rivers should also have higher species

turnover. Thus, the dispersal hypothesis predicts: (a) low phylobetadiversity among

habitats within an ecoregion, (b) low phylobetadiversity between habitats that occur in

multiple ecoregions, and (c) high phylobetadiversity in isolated habitats in all

ecoregions.

Third, the environmental filtering effect is based on the hypothesis that selection

by abiotic and biotic processes causes deterministic fitness differences between taxa

within a habitat or ecoregion (Swenson et al. 2012). Within the biome in question in this

study, temperature, precipitation and elevation do not differ significantly, but edaphic

conditions, and habitat heterogeneity do create differences. If environmental filtering

dominates community organization, then we would predict high betadiversity between

habitats within an ecoregion and between ecoregions, high phylobetadiversity in

comparison of unique habitats within ecoregions (e. g., bamboo, white sands), and low

phylobetadiversity between habitats that occur in multiple ecoregions.

The three ecoregions in our study offer both unique and replicate habitats. These

ecoregions are large areas of relatively uniform climate that harbor characteristic sets of

species and ecological communities that were defined based on expert opinions (Olson

et al. 2001). Investigating patterns of species richness and phylobetadiversity in diverse

landscapes in the framework of evolutionary biology (Vellend 2010) can elucidate how

Page 27: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

27

habitat choice contributes to species sorting from a regional pool into local communities

and thereby the origin of the communities.

Methods

Species Lists

We assembled ecoregion species lists using InfoNatura NatureServe bird

distribution data (Natureserve 2016). We used species range maps and overlapped

them with the delimitation of the Napo Moist Forest Ecoregion, the Southwestern

Amazonian Moist Forest Ecoregion and the Guianan Moist Forest Ecoregion to create

the regional species pool lists. We ensured that the elevational ranges were

incorporated in the dataset, because range maps often do not include that information.

Marginally overlapping species were also removed, because shapefiles of ranges of

extralimital species sometimes overlapped with the boundaries of a specific ecoregion,

thereby artificially inflating the regional species pool.

To generate the habitat level classification, we selected three sites: Allpahuayo

Mishana National Reserve in the Napo Ecoregion, Cocha Cashu Biological Station in

Southwestern Amazonian Moist Forest Ecoregion, and Nouragues Biological Station, in

the Guianan Moist Forest Ecoregion. The bird communities of these three sites have

been intensively studied, which allowed us to generate appropriate habitat-level

classifications within the regional species pools. We used a combination of published

checklists (Robinson and Terborgh 1990, Thiollay 1994, Alvarez Alonso et al. 2012) and

data from detailed field observations and habitat descriptions from field guides (Stotz

1996, Schulenberg et al. 2007) to assemble the species lists for the three regional

species pools and the habitat pools. We assembled species lists for a total of 33 habitat

based local communities, and we removed vagrants (species that were only recorded

Page 28: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

28

as very infrequently observed, wandering individuals that had no known local

populations within the regions) from species lists.

Species Composition Comparisons

We first compared the species composition of communities using the Bray-Curtis

dissimilarity (BCD; (Bray and Curtis 1957)). The BCD is the sum of the absolute value of

the differences in the number of species in two communities divided by the sum of the

total number of species at the two sites. A BCD of 0 indicates that two communities

have the same species composition, whereas a BCD of 1 indicates that the sites share

no species. We calculated the BCD for all pairs of communities based solely on species

composition, not abundance. This is equivalent to calculating BCD when the abundance

of each species in each community is either 1 (present) or 0 (absent).

Differences in richness among localities can bias comparisons using raw BCD

values. For example, when comparing a community with many species to a community

with few species, the BCD will be large simply due to the difference in the number of

species, even if the smaller community represents a subset of the larger community. To

account for this potential bias, we estimated the BCD that we would expect with the

same sampling, if all of the localities represented a single, homogenous community.

This was done by randomly permuting the species observation data to generate 10,000

data sets representing the distribution of species under the null model of a single,

homogeneous community. We permuted the list of species observed at each of the

localities using a random reshuffling technique that preserves the total number of

species observed in each locality. A single reshuffling move consists of randomly

picking two localities, and then randomly choosing a species that occurs at each

locality. If the species from the first locality does not occur in the second and the

Page 29: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

29

species in the second locality does not occur in the first, we swap the locations of the

two species. To generate each randomly permutated data set, we performed this

reshuffling 10,000,000 times. The calculation of the BCD distances and the creation of

the randomly permuted data sets were done using Perl scripts written by J Gordon

Burleigh and followed the same methods as described in (Oswald et al. 2016)

For each pairwise comparison of localities, we divided the observed BCD

(calculated from the species lists) by the expected BCD (the average BCD calculated

from the 10,000 randomly permuted replicates). This ratio of observed to expected BCD

accounts for the differences in size and sampling effort among localities. If the ratio is

<1 or >1, the composition of the two localities is either more similar or more different,

respectively, than would be expected if all localities represented a single, homogenous

community. The BCDs calculated from the 10,000 random replicates also provide a null

distribution to test whether any two localities are more similar or different than expected

by chance.

Phylogenetic Composition Comparisons

Two localities may share few or no species and yet have a very similar

phylogenetic composition (e.g.(Graham and Fine 2008). To evaluate the phylogenetic

distance between all pairs of localities (i.e., the phylobetadiversity), we used the UniFrac

measure (Lozupone et al. 2011), which estimates the unique fraction of the phylogeny

represented by each locality. Given a phylogenetic tree that contains only species from

two localities, the UniFrac distance is the percentage of branch lengths in a

phylogenetic tree that are unique to a single locality. For example, if Locality A was

composed of species that formed a clade that was sister to a different clade composed

of the species in Locality B, the two localities would have a UniFrac distance of 1

Page 30: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

30

because there is no shared phylogenetic history between the species in the two

localities. In contrast, if all species in Community A had closely related sister species in

Community B and all the species in Community B had closely related sister species in

Community A, the UniFrac distance would be small (close to 0), even if the localities

share no species.

Estimates of phylobetadiversity depend on a phylogenetic tree, usually with

ultrametric branch lengths. We pruned the avian phylogenetic tree of (Burleigh et al.

2015) to include only the 764 species from the detailed species lists. The tree contained

641 (83.9%) of these species. This tree and 100 bootstrap trees were built from a 29-

locus supermatrix using maximum likelihood (ML), implemented in RAxML (Stamatakis

2006),and the branch lengths were made ultrametric using penalized likelihood (with a

smoothing parameter of 100) implemented in r8s (Sanderson 2003). We first estimated

the Unifrac distances between all localities based only on the species ML tree. To

examine the possible effects of topology and branch length uncertainty, we also

estimated the Unifrac distances using all 100 ML bootstrap trees. To account for the

123 missing species, we generated 100 new trees by randomly inserting the missing

taxa into the ML tree in a position from the stem branch preceding the most recent

common ancestor of the species from the same genus, or if missing, the same family, in

the tree to the tips descending from the stem.

Like the BCD, the size of a locality or the sampling effort can greatly affect the

raw Unifrac distances. Therefore, for each comparison between two localities, we also

calculated the Unifrac distance using 1,000 randomly permuted data sets, created using

the same methods described for the BCD analyses. For each comparison between

Page 31: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

31

localities, we divided the observed Unifrac distance by the average Unifrac distance

from the 1,000 random replicates, and this ratio was used as the pairwise distance

measure between localities.

Statistical analysis were performed using the R version 3.1.0 (R Development

Core Team 2016). Mantel tests were performed to measure the association between

the elements of two matrices, by taking into account the autocorrelation that exists

between the elements of each matrix (Mantel 1967) were performed using the ade4

package (Dray and Dufour 2007). This test is often used to test for a significant

association between different distance matrices. For network analyses the package

hclust and library igraph (Csardi and Nepusz 2006) we used, applying various linkage

methods to find the appropriate number of groupings.

Results

Species richness

The analyses revealed high regional species richness and variable local species

richness in different habitats. Overall, 1486 species were included in the three

ecoregions associated with the habitat types in the 33 assemblages compiled in this

study, 1194 species in Southwest Amazonia, 1072 species in the Napo Ecoregion, and

606 species in the Guiana Moist Forest. These high totals include some Andean foothill

species as well that have partially overlapping distributional maps, these were not used

in the final analyses. The AMNR dataset included 496 species, 473 at Cocha Cashu

and 420 species in Nouragues. The average number of bird species per habitat was 76

(range: 13 to 235). Primary forested habitats had the highest species richness.

Page 32: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

32

Betadiversity

Of the 528 pairwise comparisons of Bray-Curtis dissimilarities of species

composition, 100 pairs (18.9%) of local communities showed results that were more

similar to each other than expected by chance (Fig. 2-2). Of these 100 pairs, 96

remained in the results when repeating the analysis with the 641 species that were in

the phylogenetic tree. All of these habitats were also more similar than expected by

chance to at least one other habitat, forming one large cluster with various

interconnecting links. There were 134 (25.4%) pairwise combinations that were

significantly different based on taxonomic betadiversity.

The hierarchical cluster analysis of the set of Bray-Curtis Dissimilarities identified

two large clusters with all agglomeration algorithms and grouping methods (Fig.2-3).

The forest habitats and the non-forest habitats form the two distinct clusters within

which the different habitats are arranged in similar dendrograms produced by ‘average’,

‘complete’, ‘single’, and ‘ward.D’ groupings (Fig. 2-3). The cluster of forested habitats

included local communities from all three ecoregions.

Phylobetadiversity

Of the 528 pairwise comparisons of phylobetadiversity of the 33 different habitats

(Table 2-2), 30 pairs (5.7%) of local communities (as defined by habitat preferences)

were statistically more similar to each other than expected by chance (Fig.2-4). Of these

30, 23 were also significantly more similar in the bootstrap replicates (adding a new

pair: h7-h18: Napo lakes and SW rivers), and in the maximum likelihood tree that also

included all species. Running the analysis through the bootstrap replicates and the all-

species tree also showed that secondary forest habitat from Guiana (h27) was similar to

Page 33: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

33

(forested habitats from the Napo ecoregion (h3, h4, h5). High phylobetadiversity was

found in 227 pairwise comparisons (43% of all comparisons).

The local community comprised of habitat generalists (h1, any species that is

equally likely to occur in four or more habitat types) was phylogenetically more similar to

the community of birds mostly observed above canopy (overhead) (h26) in the SW, and

the open habitat of large openings (h32) in Guiana, represented relatively open habitats

from all three ecoregions. When comparing forested habitats, Napo secondary forests,

forest gaps, and primary forest edges (h3) were more similar to periodically inundated

forest in the same ecoregion (h5) and the periodically inundated forest (h14) and the

zabolo, canebrakes, and vines (h23) in the SW Amazonian ecoregion. The Napo

primary, terra firme forest (h4) was found to be phylogenetically more similar to the

periodically inundated forests in the same ecoregion (h5), and both of these mature

forest types were more phylogenetically similar to 4 other mature forest habitats: the

white sand forests in Napo (h12), the high ground forest (h13) and the periodically

inundated forest (h14) in the SW, as well as the Guianan mature primary forest (h28).

White sand forests (h12) were more similar to SW Amazonian high ground forest (h13),

periodically inundated forests (h14) and the latter two were also more similar to each

other, and to the Guianan mature forest (h28) as well.

Lakes and rivers of the Napo ecoregion were phylogenetically more similar to

each other, and the Napo river habitat was also linked to SW Amazonian river and lake

habitats (h7, h11, h18, h20 clustered together). Similarly, Napo lake edge and river

edge were both similar to SW Amazonian lake edge habitat, which was similar to the

river edge of the same ecoregion (h8, h10, h19, h21). The two aguajales communities

Page 34: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

34

(birds living in Mauritia palm stands in broad, marshy clearings) also grouped together

with low phylogenetic turnover (h2 and h25). SW Amazonian swamp forest (h15)

clustered with forest stream margins (h16) of the same ecoregion. The SW forest

openings from treefall gaps (h17) clustered with Guianan disturbed or secondary forest

(h27), and the SW bamboo habitat (h24) was more similar to the Guianan treefall gaps

edges (h29).

Within ecoregions there were 18 pairs of communities with low betadiversity in

the NAPO, 19 within the SW and 3 within GUI ecoregions (27%, 21% and 14% of

comparisons, respectively). There was a higher number of pairwise comparisons with

high taxonomic turnover among these habitats: 26 in NAPO, 18 in SW and 8 within GUI;

39%, 20%, and 36% respectively (Fig.2-5a). Similarly, there were very few communities

that were phylogenetically more similar to each other than expected by chance: 5 within

NAPO (8%), 5 within SW (5.5%), none within GUI, and more communities had higher

phylobetadiversity than expected by chance: 30 within NAPO (45%), 49 within SW

(54%), and 4 within GUI (18%), respectively (Fig.2-5b).

The distance matrices based on Unifrac distances between local communities of

all species, only species in the phylogenetic tree, and the bootstrapped tree were all

significantly correlated with the BCD matrix based on species composition (Mantel test

results: all species, r=0.60, P<0.001; only in tree species r=0.68, P<0.001, bootstraps

r=0.57, P<0.001). All three sets of Unifrac distances were also significantly correlated

(Mantel test results: all species to in tree only, r=0.98, P<0.001; only in tree species to

bootstraps r=0.84, P<0.001, bootstraps to all species r=0.85, P<0.001). The relationship

between species richness, betadiversity and phylobetadiversity (Fig. 2-6) showed a

Page 35: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

35

stronger correlation between species richness and phylobetadiversity results (lm,

p<0.01, R2 = 0.14).

Discussion

Our results demonstrate the role of beta diversity in maintaining the high regional

diversity of Amazonia. Using an approach that combined different spatial scales in three

different regions with phylogenetic comparison, we documented high betadiversity and

high phylobetadiversity in our pairwise comparison within ecoregions, and also among

regions. Yet, we also documented that a great many habitats that showed little

betadiversity or phylobetadiversity. Taken together, these results suggest that none of

our hypothesized processes underlying avian biodiversity in the Amazon can explain all

or even the majority of the patterns we documented. Instead, we argue that context-

dependent deterministic processes shape community assemblages in different habitats

with different levels of connectivity. Overall, we found strong evidence that species were

not simply randomly distributed; a large proportion of communities were either more

similar or more different than expected by chance, and the results did not support the

drift hypothesis that species were simply randomly distributed across habitats.

The results indicate that although many habitats shared species, only ~20% of

habitat comparisons had low betadiversity (were significantly more similar than

expected by chance) across ecoregions. Within the two western Amazonian ecoregions

there were more pairs of local communities that were significantly similar to each other

than expected by chance, but not in the Guianan ecoregion. These differences may be

explained by the broader distribution of species in the western Amazonian region, which

appears to contain species that are more likely to share habitats than they are in the

Guianan region. The Napo and SW Amazonian ecoregions harbor diverse local

Page 36: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

36

communities that share more species with each other than with the Guianan moist

forest ecoregion, which could be explained by their proximity and shared

biogeographical histories.

Some of the patterns uncovered in our study are driven by widespread species

that occur across the region (Graham and Fine 2008), making species composition

between sites similar (low betadiversity and low phylobetadiversity). For example, the

similar patterns discovered in riverine communities and lakes can be explained by the

presence of species with wide distributions. In addition, riverine habitats are intrinsically

connected (Remsen and Parker 1983, Puhakka et al. 1992, dos Anjos et al. 2007,

Gillies and Clair 2008, Adams and Burg 2015); therefore, dispersal limitation is less

likely to lead to differences in communities. It is interesting to note that species living in

forested habitats, where dispersal limitation has been hypothesized as an important

mechanism, also showed less turnover than expected by chance. This can be explained

by habitat filtering or by strong biotic interactions between species, which we did not

test.

Taxonomic and phylogenetic betadiversity and bird species composition in

forested habitats were all more similar to each other than expected by chance,

regardless of whether they were in the same or different ecoregions. These

assemblages were in similar environments, but many were separated by large

distances. These results suggest that forest communities throughout the lowland

tropical moist forest biome have similar bird communities as shown by a comparison of

the 100-ha census results of Terborgh et al. (1990) and Thiollay (1988, 1994), which

generally only differ in details of species composition. Even with all of the complex

Page 37: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

37

changes in connectivity through time, the communities of forests have diverged mainly

in species composition rather than in deeper phylogenetic composition, and even

forests with different levels of disturbance (floodplain versus terra firme) still are very

similar overall in phylo- and betadiversity. These results suggest that niche processes

structure these communities, which strongly supports the deterministic hypotheses and

suggest that drift and dispersal limitation are not strong factors affecting community

structure in forested habitats. It is also possible that the reduced phylogenetic turnover

among habitats is a direct result of the high species diversity of these habitats.

We found three larger clusters of habitat types ( forested habitats, open water,

and water edge) that form distinct clusters, and a few other identical habitat types that

form small clusters across ecoregions. Bird communities that share water edge habitats

such as lake edges and river edges in Western Amazonia also showed low

phylogenetic turnover, and lake and river communities also formed a cluster based on

low phylogenetic betadiversity. The bird communities living in the aguajales, dominated

by Mauritia flexuosa palms, also clustered together, as they share a large number of

habitat specialist species, and these habitats are also connected because they occur

along major river floodplains. Aguajales occur more in Western Amazonia, and were

missing from our GUI dataset. These patterns were only revealed using the

phylogenetic betadiversity metrics. These clusters could also be explained by strong

effects of biotic, niche-based processes determined by elements of vegetation for the

bird assemblages of the communities in these particular habitats (Emerson and

Gillespie 2008, Graham and Fine 2008, Kraft et al. 2014, Weinstein et al. 2014).

Page 38: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

38

Habitats that form along linear habitat features such as rivers are connected

across large distances, and in Amazonia in general terra firme forest are the dominant

forest type that comprises the matrix in which other habitats are embedded. However,

there are certain Amazonian habitats that form on essentially permanent soil formations

such as WSFs and consist of small, isolated patches in a matrix of terra firme, which

can create a potential barrier to gene flow. Despite their island-like configuration

(Prance 1996) and the number of habitat specialist species that do not occur in other

habitats (Alvarez Alonso 2002, Alvarez Alonso and Whitney 2003, Alvarez Alonso et al.

2013), the bird communities of white sand forests were also found to be similar to other

forested habitat types. This result suggests that despite the differences in floristic

composition (Tuomisto et al. 2003, Fine et al. 2010, Pomara et al. 2012), there is little

evidence for a unique bird community in WSFs other than the relatively small proportion

(less than 10%) of specialists. Local communities varied in species richness, and as

species richness increased, phylogenetic turnover decreased among communities;

richer communities were more similar to each other, although there was no strong

congruence between PBD and species richness. This could also explain why all the

forest habitats seem to be more similar to each other, because these communities have

the highest number of species.

When combining all three ecoregions in the analysis, over a quarter of the

communities were more different from each other across ecoregions than expected by

chance, and the differences among local communities within ecoregions in terms of

taxonomic diversity were even higher. The results of sets of communities that showed

both high phylogenetic betadiversity and high betadiversity can be explained by the

Page 39: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

39

presence of a high proportion of small-range species, at least some of which come from

more distantly related lineages. Close to two thirds of these comparisons, however,

were based on pairs of habitats that contrasted a forested habitat type with an “aquatic”

habitat type. Bird communities in those habitats rarely share species; therefore, the

strong patterns of turnover can be explained by the intrinsic differences of the habitats

themselves.

Our results indicate that local communities were both taxonomically and

phylogenetically more distinct than expected by chance both within and among

ecoregions, not an unexpected result in such a highly diverse region. We did not find

that habitats within ecoregions formed clusters. In fact, habitats from distant ecoregions

were more similar than expected by chance, and only very few (0-8%) pairwise

comparisons were more similar to each other within ecoregions. Therefore, we found no

support for the speciation hypothesis, which predicted that the greatest differences

would occur among regions in which geographic isolation had led to speciation, and

within ecoregions it predicted low phylogenetic turnover.

Our plot data can be considered complete. There are few complete datasets from

the Neotropics due to the high diversity and the expertise needed to census these

areas; however, the number of well-studied sites is growing. Our study includes species

lists assembled from sites distributed over a vast geographic area, and from sites

separated by great distances (more than 900 km between Napo and SW, more than

2300 km between Napo and French Guiana, and more than 2500 km between SW and

French Guiana). Distribution ranges of most bird species in the analyses do not span

Page 40: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

40

these distances, and dispersal distances for individual forest species is known to be

limited (Blake 2007, Moore et al. 2008).

In conclusion, we found strong evidence that communities have deterministic

assembly rules within the Amazon Basin, and results of comparisons are context-

dependent. The relative roles of dispersal and niche processes differ in specific kinds of

habitats that vary strongly in resources and habitat structure (e. g., aquatic and forest

habitats), in connectivity (riverine versus more isolated patches of permanent soil

formations), and in habitats with generally similar structure and resource availability (e.

g., the overall similarity in forests of all ecoregions). Our phylogenetic analyses allowed

us to disentangle turnover due to replacements of closely related species in regions

isolated by rivers from environmental filtering in habitats in which fundamentally different

processes shape bird communities (e. g., aquatic versus terrestrial habitats).

Phylogenies therefore served as a tool to generate further hypotheses about processes,

even if they do not necessarily allow us to test processes directly (Mayfield & Levine

2010).

To elucidate the proximate mechanisms of habitat filtering, we would like to

investigate what traits are shared among species in habitats that show clustering and

low phylogenetic turnover across distant regions. We also need to investigate if other

taxa show similar patterns to birds, and have already started collecting species list

information for plant communities. Strong shared patterns across multiple groups or in

the same geographic region could help us identify centers of endemism (Gonzalez-

Orozco et al. 2015).With the fast-growing publication of detailed phylogenetic trees and

advanced computational approaches, it will be possible to continue investigating this

Page 41: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

41

question, explore whether there is turnover across entire lineages, and how the size of

species ranges affect patterns of turnover..

Page 42: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

42

Table 2-1. Ecoregions and habitat types used in this study

Habitat NAPO SW GUIANA

Generalist (more than 4 habitats) h1 h26

primary terra firme forest h4 h13 h28

secondary forest h3 h27

inundated forest h5 h14 h33

white sand forest h12

agricultural/disturbed area h6 h32

aguajal h2 h25

forest opening h17 h29

bamboo h24

zabolo h23

marsh h22

swamp h15

bare rock h30

lake h7 h20

lake edge h8 h21

river h11 h18 h31

river edge h10 h19

forested creek edge h9 h16

Page 43: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

43

Table 2-2. Summary of predictions of the four hypotheses tested.

Hypothesis Betadiversity Prediction Phylobetadiversity Prediction

Null (drift) Random Random Speciation/Biogeography (historical isolation)

High within ecoregions High between ecoregions

Low/random within ecoregions High between ecoregions

Dispersal

Low within ecoregion Low between ecoregions Low between habitats

Low within ecoregion Low between ecoregions Low between habitats

Environmental filtering

High within ecoregions High between ecoregions High/Low between habitats context dependent

Low within ecoregions High/Low between ecoregions context dependent

Page 44: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

44

Figure 2-1. Map of the three ecoregions used for assembling the regional communities.

The labeled and outlined ecoregions all belong to the moist tropical forest biome.

Page 45: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

45

Figure 2-2. Network representation of the cluster of communities that are taxonomically

more similar to each other than expected by chance. The color of the spheres represents the ecoregion (red=NAPO, blue=SW, black=GUI).

Page 46: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

46

Figure 2-3. Cluster dendrograms using taxonomic betadiversity and phylobetadiversity

for all local bird communities in different habitats across the three ecoregions, using average, complete, single and ward.D linkage methods. Labels correspond to habitat codes from Table 2-1.

Page 47: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

47

Figure 2-4. Clusters of communities in habitats that are phylogenetically more similar to

each other than expected by chance. This network representation of all the phylobetadiversity comparison of the 33 habitats in 3 ecoregions. The color of the spheres represents the ecoregion (red=NAPO, blue=SW, black=GUI). Links represent two habitats that were phylogenetically more similar to each other than expected by chance.

Page 48: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

48

A)

B)

Figure 2-5. Within-ecoregion betadiversity phylogenetic betadiversity. Only statistically

significant pairwise comparisons within each ecoregion are shown. A) Betadiversity results using taxonomic turnover. B) Phylogenetic betadiversity results within each ecoregion. Black linkages represent pairs of communities more similar to each other than expected by chance. Red linkages represent pairs of communities more different from each other than expected by chance.

Page 49: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

49

Figure 2-6. Relationship between phylogenetic betadiversity (PBD, black points)

betadiversity (BC, blue points), and species richness. The lines represent best fit lines corresponding to linear regressions.

Page 50: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

50

CHAPTER 3 STRUCTURE OF UNDERSTORY BIRDS COMMUNITIES IN FORESTS ON TWO

CONTRASTING SOIL TYPES IN AMAZONIAN PERU

Introduction

Amazonia hosts the highest terrestrial forest diversity in the world in spite of its

low climatic and topographic variability. One hypothesized mechanism for the

maintenance of this diversity is the extraordinary range of different habitat types present

in the region ((Rosenzweig 1987, 1995, Borges 2013, Arellano et al. 2014, Castano-

Villa et al. 2014), and Chapter 2). These habitat types may be sufficiently different in

resources and structure that they support different bird communities, including many

habitat specialists. Habitat differences may further enhance diversity by allowing closely

related species to coexist through habitat segregation, which can be either passive or

active (Robinson and Terborgh 1995, Nores 1999).

Extensive geological and botanical studies show that the soils in Western

Amazonia are composed of a mosaic of geologically different substrates with different

chemical properties that influence the forest flora (Ruokolainen et al. 1997, Tuomisto et

al. 2002, Ruokolainen et al. 2005, Ruokolainen et al. 2007, Asner et al. 2014, Higgins et

al. 2014). The avifauna also co-varies with both floristic composition and the soil

nutrients in upland forests (Pomara et al. 2012). One of the most distinctive soil types is

on white sand soils, which are associated with plant communities that differ from those

on the more dominant clay soil. Five types of white sand forests are generally

recognized in Peru, each of which differs in structure and physiognomy,: high-dry-open

(varillal alto seco, tall dry varillal), high-wet-open (varillal alto húmedo, tall humid

varillal), low-dry-dense (varillal bajo seco, low dry varillal), low-wet-open (varillal bajo

húmedo, , low humid varillal) and low, dense forest (chamizal) (García Villacorta et al.

Page 51: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

51

2003, Garcia-Villacorta and Hammel 2004, Coronado et al. 2015). For the purpose of

this study we reduced the five WSF categories to three, humid and dry varillales and

chamizal. Each of these white sand forest types differs from the others but all share

characteristics that distinguish them from forests on clay soils. Vegetation changes

among the habitat types both in composition (Fine 2010) and structure (Jirka et al.

2007, Ferreira et al. 2010). White sand soil forests contain many habitat specialist

plants that tend to be slow-growing and short, traits associated with the lower

productivity of sandy soils, which do not hold water well and are often leached of their

nutrients (Fine et al. 2005, Fine et al. 2006). The trees in most white sand forests are

also more slender, shorter, and they often have very hard, sclerophyllous leaves. The

understory is relatively sparse in WSF, with fewer herbs except for terrestrial ferns,

which are very common - especially Trichomanes spp., Elaphoglossum spp. and

Lindsaea divaricata (Gentry 1988). The humid varillales have relatively poor drainage in

comparison with other WSF, not much organic matter, exposed roots covering the

ground and low tree species diversity. Caraipa utilis (Clusiaceae) is often dominant in

these forests. Dry varillales have better drainage, little organic matter and many more

tree species. Chamizal vegetation is shorter, wet, usually 3-5 m high, with dispersed

emergent trees up to 8-12 m tall. It is dominated by shrubs such as Graffenrieda limbata

(Melastomataceae), and interspersed throughout with small palms, e.g. Mauritia and

Mauritiella, and only a few species of ferns (Gentry 1981, García Villacorta et al. 2003).

Given these differences between white sand and clay soil forests and among the

different kinds of white sand forests, it is likely that the bird communities associated with

these different habitats will also differ in composition, trophic and guild structure and

Page 52: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

52

abundance. We documented in Chapter 2 that forested habitats harbor unique,

taxonomically distinct communities, some with specialists, and comparisons of the

community structure may shed additional light on the selective environments underlying

the role of these white sands plant communities in maintaining the high beta diversity of

Amazonia. In this study we used mist nets to sample adjacent communities on different

kinds of white sands forest and compared them with the clay soil matrix terra firme in

which the white sands patches are embedded in the AMNR. We compared bird species

richness, capture rates, turnover (species replacement) among communities, and

nestedness (differences due to species loss or addition among different sites) among

white sands forests and with forests growing on clay soils.

Our first hypothesis is that habitat productivity is the primary determinant of the

overall abundance and diversity of birds in communities. A simple prediction of this

hypothesis is that clay soil forests, which are generally more productive (Higgins et al.

2014) than white sands forests would have the highest abundance and richness of

species, followed by humid varillal, dry varillal, and chamizal in order of decreasing soil

nutrient content and vegetation height. According to this hypothesis, the richness of

most of the trophic guilds should vary in parallel with fewer species and captures in all

guilds, reflecting the overall reduced availability of resources in all trophic levels due to

lower productivity. Alternatively, the lower stature of the forest in white sands may lead

to higher capture rates because the mistnets used to capture birds are more likely to

sample the entire community rather than just the birds living in the understory (Remsen

Jr and Good 1996). In this case, the capture rate of canopy-living species should

increase with decreasing vertical habitat structure.

Page 53: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

53

An alternative to the productivity hypothesis is that community composition will

more closely reflect idiosyncratic differences in vegetation structure and the resources

available in the different habitats in relation to the details of plant community

composition and structure (Holmes et al. 1979, Robinson and Holmes 1982, Bush and

Holmes 1983, Holmes and Recher 1986, Kornan et al. 2013) or the associated fruit and

insect resources that are available to birds (Kissling et al. 2007). We predicted that

guilds will vary independently with local, habitat-specific vegetation structure and

resources available. We first hypothesized that the composition of frugivorous birds,

which are captured readily in understory mistnets (Karr et al. 1991, Loiselle and Blake

1999, Wunderle et al. 2005, Kissling et al. 2007, Blake and Loiselle 2009), will change

with the composition of the forest (Hasui et al. 2007, Vidal et al. 2014). Specifically, we

predicted that: (a) there will be relatively higher capture rates and a higher proportion of

frugivores in the understory of WSF reflecting the dominance of fleshy fruited

Melastomes, a bird-dispersed fruit; and (b) based on the dominance of Melastomes, we

further predicted that frugivorous species known to prefer Melastomes (e.g., Mionectes

flycatcher and manakins) would be proportionately more abundant and caught at higher

rates in WSF.

Second, we predicted that white sands forests will also contain a higher overall

proportion and capture rate of ant followers reflecting the general dominance of

hymenopterans (Saaksjarvi et al. 2004, Fine et al. 2006, Saaksjarvi et al. 2006, Gomez

et al. 2015) in the drier white sands soils relative to other groups of insects, which tend

to be less abundant on white sands soils (Vormisto et al. 2000, Saaksjarvi et al. 2006,

Fine 2010, Lamarre et al. 2012, Lamarre et al. 2016b). And third, we predicted that

Page 54: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

54

multi-species flocking species and non-flocking foliage gleaning insectivores will be

caught at lower rates in white sands forests because of the low density of understory

foliage. Conversely, we predicted that sally gleaning insectivores, which tend to forage

in more open habitats, would be captured more often in the open understory of white

sand forests.

Methods

Study Region

Our study site was the Allpahuayo-Mishana National Reserve (AMNR) in the

Department of Loreto, northeastern Peru (latitude -3.922904, longitude -73.516645).

The study area is accessible from the Iquitos-Nauta highway, a paved road on the

southern border of the reserve, or the Nanay River along the northern border. Sampling

sites were established in closed-canopy white sand forest patches that were accessible

on foot (elevation: mean = 126.5m, range = 91.9 - 165.9m), and some sampling was

simultaneously completed in adjacent closed-canopy forests growing on alluvial terra

firme clay soils (elevation: mean = 116.7 m, range = 106.1 – 130.6 m).

Vegetation structure was measured at each patch and around each netline within

square 10 m by 10 m plots using a protocol adapted from the breeding biology research

and monitoring database field protocol (Martin et al. 1997)). Variables summarized from

vegetation structure data generally describe aspects of forest vertical structure or

understory vegetation density. Forest vertical structure variables included average

canopy height, number of trees greater than 10 cm dbh (tallied within a 20 by 20 m plot

surrounding each 10 by 10 m plot); and total tree basal area. Understory vegetation

variables included density of small stems (smaller than 2.5 cm dbh; less than 1 m in

Page 55: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

55

height), density of large stems (greater than 2.5 cm dbh), and leaf litter depth

(summarized in Table 3-1).

Bird Surveys

We evaluated community composition of understory birds in an intensive mist-

netting and mark-recapture program from June 2009 through December 2012. We used

constant-effort mist-netting (Dunn and Ralph 2004, Nur et al. 2004) to survey understory

birds in the reserve. All nets were 38-mm mesh, 12 m x 3 m, tethered with five

trammels. We created transects through the forest interior using a line of 20 nets set on

existing foot trails in primary forest or older secondary growth forest contiguous with

primary forest. Nets were separated from one another by <1 m when possible. The

minimum distance between transect sites was 250 m. The greatest distance between

any two transects was 18 km. Each net transect was opened for ~500 net hours during

three consecutive days. We opened nets by 06:00 and checked nets approximately

every 30 min until 16:00 on the first two days and 11:00 on the third day. To reduce net

avoidance, we made sure that a site was not sampled for longer than 3 consecutive

days. We have limited data from open areas and second-growth as well. Through the

studies of Alvarez Alonzo, Juan Diaz and Noam Shany, a complete description of the

avifauna of this reserve is available (Alvarez Alonso et al. 2012).

Avian Guild Classification

All captured birds were identified, aged, sexed, and, if appropriate, marked with a

metal band, or by clipping the tip of a tail feather to recognize recaptures. Birds were

classified into feeding guilds using references on diets of tropical birds, and first-hand

observations of foraging in the field. Our classification also follows Terborgh et al (1990)

for species occurring in lowland tropical forest and a number of studies with diet

Page 56: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

56

information compiled by Wilman et al. (Terborgh et al. 1990, Wilman et al. 2014).

Species were assigned to specific guilds based on foraging stratum, diet and substrate

according to the same references (Robinson and Terborgh 1990, Wilman et al. 2014).

We designated species as frugivores, nectarivores, granivores, omnivores, and

insectivores. Carnivorous species such as raptors, and piscivorous species such as

kingfishers were combined in our foraging guild analysis as they do not typically get

caught in mist nets and are undersampled (Remsen Jr and Good 1996). This guild

classification is based on the dominant trophic level for each species; we recognize that

many species simultaneously occupy more than one trophic guild. Even nectarivorous

hummingbirds supplement their diet with insects. Assignment of species that did not

have such ecological data available in the literature was characterized based on

personal observations.

Obligate, near obligate or facultative habitat specialization was assigned based

on the description of the bird community by Alvarez et al. (Alvarez Alonso 2002, Alvarez

Alonso and Whitney 2003, Alvarez Alonso et al. 2013). A guild of 17 species that

participate in mixed-species foraging flocks was also constructed; canopy species were

treated in a separate analysis. Sally-gleaning species that search more in the open and

fly relatively long distances to attack prey (Remsen and Robinson 1990) were also

distinguished from those that glean from nearby foliage (shrub gleaners that are not in

mixed species flocks). Ant followers and terrestrial insectivores were identified using

field observations, and ecological information from literature (Robinson and Terborgh

1990, Stotz et al. 1996, Schulenberg et al. 2007)

Page 57: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

57

Data Summary and Analyses

Sampling effort differed between WSF (37000 net hours: 20500 in humid, 13500

in dry and 3000 in chamizal) and terra firme clay (9500 net hours). Therefore, we

compared captures and species richness across sampling events standardized to 500

net hours per event. We used raw species richness and rarefied species richness, and

they produced similar rankings. Our analyses are based on capture rate, excluding

same-day recaptures, which we calculated for each sampling period. Capture rate is

interpreted as a measure of bird activity in the understory (Karr 1979, 1981, Martin and

Karr 1986, Karr et al. 1991), which is potentially useful for comparisons over time and

among patches. We recognize biases with mistnet sampling (Remsen Jr and Good

1996) and will discuss their potential influence on some of our comparisons, especially

those dealing with habitats with very different canopy heights. All study sites were in

forest and had substantial undergrowth; therefore we assumed that habitat-specific

capture rates among guilds did not vary in such a way to alter the validity of the results.

Analyses were performed at the level of the sampling event (N=93, at 500 net hours

each), at the level of netlines (N= 42; hereafter “netline level”) and for particular

questions at the level of white sand forest patches (N=8 WSF patches; hereafter “patch

level”). We calculated percentages for each category from the number of captures, and

number of individuals. We also calculated species accumulation curves for the two

general habitat types (clay vs sand) and for the four specific habitat types (terra firme,

humid varillal, dry varillal and chamizal) to check for adequate sampling and rarefaction

analyses based on the minimum number of captures in the habitat type.

We examined patterns of species diversity (overall and within each guild) across

netlines and patches by the following metrics: total number of species, number of white

Page 58: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

58

sand specialists, and number of rare species. We examined dissimilarity values

between white sand netlines and adjacent terra firme forest netlines to assess species

turnover and nestedness between contrasting habitat types. Measurements of change

in species composition across sites (e.g., beta diversity) that are expressed as

dissimilarities can reflect two separate processes: change due to species loss

(nestedness) and change due to species replacement (turnover) (Ruokolainen et al.

2005, Tuomisto 2010a, b, Baselga and Orme 2012, Baselga and Leprieur 2015). We

performed rarefaction analyses with standard samples for all species, and 95%

confidence intervals were calculated from rarefaction curves. The lack of overlap among

them was interpreted as significant differences in diversity among samples (Gotelli and

Colwell 2001, Gotelli and Entsminger 2001). We summarized pairwise bird dissimilarity

values of turnover and nestedness for each netline and patch. In case of replicates

across multiple years we also calculated averages of pairwise comparison across the

years. We also performed cluster analyses using the beta diversity index and

hierarchical clustering algorithm. Several grouping methods were used to visualize the

cluster dendrograms. Dissimilarity matrices for beta diversity were calculated in R v.

3.1.0 (The R Foundation for Statistical Computing 2015) using the vegan package.

We used a non-metric multidimensional scaling (NMDS) ordination to

characterize the compositional diversity of each habitat using capture data. The capture

data were standardized by the total number of captures in each site. Similarity matrices

were built using Bray–Curtis index. We applied a one-way analysis of similarity

(ANOSIM) to different similarity matrices to test the null hypothesis that the WSF and

clay habitats have no significant differences in species composition.

Page 59: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

59

Results

Community-wide Patterns

During 22 months of banding, we recorded 8328 captures in forested habitats of

5057 marked individual birds. We captured 142 species with mist-nets (approximately

29% of the list of the area) in 46,500 net hours (Appendix A Table A-1.). The 142

species were from 114 genera and 32 families captured in the forest at least once

(including both clay and white sands). In white sands, 126 species were caught in 6421

capture events, and in clay-based forests 105 species were caught in 1907 captures.

The average capture rate for 500 net hours in WSF was 89.6 ± 29.2 SD individuals

compared with 100.4 ± 31.6 SD in clay soils. We recorded 1907 captures in clay during

19 sampling events, 2400 in humid varillal during 27 sampling events, 3621 in dry

varillal during 41 sampling events, and 400 captures in chamizal during 5 sampling

events. On average, overall capture rates were not significantly different between white

sand forest and clay forest netlines (Shapiro-Wilk normality test, Welch’s Two sample t-

test, p=0.099, Appendix A Fig. A-1A) or among the four habitat types overall (Shapiro-

Wilk normality test, ANOVA, p=0.092, Appendix A Fig. A-1B), although there was a

trend towards decreasing capture rates with decreasing productivity. The small sample

size of netlines in chamizal (N=6) limits the power of this analysis.

The average number of species caught per sampling event in WSF was 25.3 ±

5.8 SD. In the clay soil forests, the average number of species was 31.5 ± 8.8 per

sampling event. The number of species captured in the two general habitats differed

significantly (Welch’s Two sample t-test using log transformed species number data,

p=0.01043, Fig. 3-1). The number of species captured also differed among the four

Page 60: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

60

different habitat types (log transformed species data passed the Shapiro-Wilk normality

test, ANOVA, p=0.0177).

The species captured in WSF and in terra firme clay also differed in their rank-

ordered distributions using proportion of captures in the particular habitat type (Fig. 3-2).

The most abundant bird in the sample of both habitats was the bark-foraging

Glyphorynchus spirurus, representing greater than 20% of captures in WSF. Whereas

Dixiphia pipra was commonly captured in both habitats (second most common in WSF

and fifth most common in clay), the second most abundant species in clay is a different

manakin species: Lepidothrix coronata. This particular manakin species was previously

reported as absent from WSF (Alvarez Alonso 2002), but we captured 58 individuals in

WSF suggesting that they regularly use this forest type.

Fifteen species that were captured in clay-based terra firme were never captured

in white sand forest, but each of these species represented less than 0.5% of clay forest

captures, and cumulatively made up only about 2% of the overall clay captures:

Automolus infuscatus, Bucco tamatia, Crypturellus soui, Electron platyrynchus,

Pteroglossus azara, Pygiptila stellaris, Ramphocaenus melanurus, Rhynchocyclus

olivaceus, Xiphorhynchus guttatus, Saltator grossus, Monasa morphoeus, Euphonia

xanthogaster, Myiobius barbatus, Thamnophilus schistaceus, Hypocnemis peruviana.

Given the much greater sampling effort in WSF, it is likely that these species are very

rare or absent in WSF.

In contrast, 36 species were captured in white sand forests that were never

captured in clay (Accipiter superciliosus, Ancistrops strigilatus, Campephilus rubricollis,

Cyanerpes caeruleus, Cyanerpes cyaneus, Hylocharis sapphirina, Micromonacha

Page 61: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

61

lanceolata, Microrhopias quixensis, Oporornis agilis, Poecilotriccus latirostris,

Thamnophilus doliatus, Tityra semifasciata, Trogon rufus, Turdus lawrencii, Cacicus

cela, Caprimulgus nigrescens, Cercomacra nigrescens, Chrysuronia oenone,

Deconycura longicauda, Lophotriccus vitiosus, Neoctantes niger, Notarchus ordii,

Pachyramphus marginatus, Phoenicircus nigricollis, Celeus grammicus, Chlorostilbon

mellisugus, Claravis pretiosa, Philydor erythrocercum, Sclerurus caudatus, Phaeothlypis

fulvicauda, Rhytipterna simplex, Attila citriniventris, Hylocharis cyanea, Heterocercus

aurantiivertex, Cnemotriccus duidae, Neopelma chrysocephalum). Most noteworthy of

the above list is the species Neopelma chrysocephalum, which makes up ~3% of all

captures in white sands but was never present in clay forest. N. chrysocephalum is an

obligate white sand forest specialist, and is locally common in the habitat. These

apparent habitat specialists constituted ~4.7% of all captures in white sand forest, but

17 of these species typically forage in the canopy. Forest canopy is shorter in WSF (Fig.

3-3, ANOVA, p<0.05 for all comparisons), which may increase the odds of their being

captured in mist nets. Canopy birds that occasionally forage in the understory were also

more likely to be caught more with increasing netting effort. There were marginally more

canopy species captured in the shortest chamizal forest (17% in chamizal compared to

~13% in all other habitats), but there were no significant differences (Pearson’s chi-

squared test statistic=0.5031, p=0.9182).

Most species occurred only at a few sampling sites, and only relatively few

species were widely distributed enough to be caught in the majority of sampling events

(Fig. 3-4). The 20 most abundant species represented ~80% of captures in WSF and

~60% of captures in clay forest, a fairly typical distribution in avian communities in which

Page 62: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

62

most species are rare or rarely encountered. These data further suggest an oligarchy of

a few very abundant species with many rare species (Robinson et al. 2000); 63 species

were captured on fewer than 5 sampling occasions and 29 species (20% of all species

sampled) were only recorded on a single netline.

Species accumulation curves based on rarefaction using all species varied

among specific habitat types (Fig.3-5), with clear separation of the three white sand

forest habitat types and the clay forest. The combined species accumulation curves

based on the 42 sampling sites do not show a clear asymptote (Appendix B Fig. B-1)

with any of the estimators (observed species richness, Chao, jackknife, or bootstrap)

comparing the specific habitats of terra firme clay and humid varillal, dry varillal and

chamizal (Appendix B Table B-1).

Compared to turnover, nestedness contributed much less to dissimilarity

between white sand and clay soil forest bird communities, and the two habitat types

show overlapping results in all dissimilarity measures (Fig. 3-6). Using Bray-Curtis

dissimilarity between WSF and terra firm clay for an ordination plot using netlines as

sites, we did not find clear separation between the communities, which overlapped

broadly (Fig. 3-7).

Vegetation Structure

Vegetation structure was similar in the white sand habitats, but different from the

terra firme forest (summarized in Table 3-1). Leaf litter depth was shallowest in clay soil

forest, and all white sand forest habitats had significantly deeper leaf litter than the terra

firme forest. Canopy was tallest in clay, followed by humid varillal, dry varillal, and

significantly shorter in chamizal than in any other habitat, although the only statistically

significant differences were in canopy height of chamizal compared to all others. The

Page 63: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

63

average DBH of trees was significantly smaller is dry varillal and chamizal than in humid

varillal or terra firme clay, and the number of small stems (less than 2.5 cm DBH)

followed the same pattern, while the number of large stems was different in chamizal

only.

Avian Guilds

The dietary trophic guild structure of the samples was very similar among

habitats. Insectivores dominated most habitats proportionally (Table 3-2) and the

proportions of samples in different broad dietary categories did not differ significantly

when comparing clay to WSF (Pearson’s Chi-squared test, p=0.62) or among all four

habitat types (p=0.998).

We also compared the differences in capture rates and number of species per

habitat per guild in birds that are frequent members of mixed-species foraging flocks,

non-flocking shrub gleaners, army ant followers, terrestrial insectivores, dead-leaf

probers, sally-gleaners, non-canopy frugivores, and nectarivores using MANOVA

among the four habitat types (Table 3-3). We found significant differences in capture

rates among habitats for shrub-gleaning species, members of mixed species flocks (Fig.

3-8), and terrestrial insectivores (Fig. 3-9), and also marginal differences in nectarivores

(Fig. 3-10), but no differences in ant followers and sally-gleaners. The number of

species in these guilds was lower in chamizal for most categories, although the small

number of netlines in chamizal limited the power of these comparisons. We excluded

canopy birds from the comparisons, and rarefaction analysis (Fig. 3-11) for just the non-

canopy species showed similar asymptotic species accumulation curves as well as

differences between clay soil and the WSF habitats.

Page 64: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

64

Discussion

We predicted reduced species richness and abundance with lower resource

availability in forests growing on less productive white sand soils, especially in chamizal

habitat, although with specific differences for certain groups such as an increase in

frugivores, and more army ant followers. As predicted, the bird communities on sandy,

poorer soils showed lower numbers of species and individuals than on richer clay soils,

confirming that productivity predicts species richness in the understory bird community

(Appendix A, Figure A-1A). The three white sand forest types did not vary significantly in

overall capture rates, although the trend was in the predicted direction; the low number

of samples in chamizal may have limited the power of this analysis (Appendix A, Figure

A-1B). We found partial support for our hypothesis that if resource availability and

habitat structure vary in contrasting habitats, then guild structure should also differ: the

proportions of dietary trophic guilds were the same in all four habitats, but some more

specific guilds based on foraging behavior differed significantly. Habitat heterogeneity is

one of the classical mechanisms promoting regional diversity, and the results of

analyses of species turnover in our system confirm that white sand forests add to the

regional diversity of the area by providing a different mix of niches available to birds,

some of which were habitat specialists.

In our surveys, we recorded 26% of the species classified as rare or absent from

white sands in our white sand sites. These results show the extent to which mistnet

samples potentially differ from song censuses. Even if some species do not actively

defend white sands forests vocally, they may use them for feeding and disperse through

them (see also Chapter 4).The examination of the between-year return rates in our data

(Appendix B Table B-2), however, do not support the hypothesis that WSFs act as

Page 65: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

65

holding areas for transient or dispersing young birds. The return rates of birds from all

categories, including species previously characterized as species that avoid WSFs,

were the same as recapture rates for species considered to be residents of WSFs

(Alvarez Alonso and Whitney 2003, Álvarez Alonso et al. 2013).

We did not detect different trophic guild composition in bird communities living in

forests on different soil types, which suggests that there were few differences in the

overall availability of the major resource groups (fruit, nectar, insects) in the forest types.

Therefore, the declines in richness with soil productivity were probably not related to

major changes in the availability of food at different trophic levels (Kissling et al. 2007).

Rather, it appears that species loss was restricted to certain guilds defined by a

combination of foraging behavior and diet.

Both flocking and non-flocking species had higher capture rates in terra firme

clay, and had fewer species in chamizal habitat (14 species in clay, 11 in humid, 10 in

dry and 4 in chamizal), a possible reflection of the sparser foliage of the forest

understory in WSFs. Understory mixed-species flock species and non-flocking shrub

gleaners were significantly more common in terra firme clay forest types (Fig.3-8). The

loss of multi-species flocking species followed a nested distribution with the loss of the

sentinels and some of the foliage-gleaning insectivores as the soils became

progressively poorer. Chamizal forest, for example only retained two flocking species

compared with 7 in terra firme. The sentinel species, Thamnomanes, which depends

upon insects flushed by other species (Munn and Terborgh ) dropped out of flocks

entirely in chamizal, perhaps reflecting the lack of enough individuals of other species

available to flush insects. Epinecrophila haematonota,a dead-leaf gleaner, also dropped

Page 66: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

66

out, but other frequent flock members such as Myrmotherula axillaris and

Xiphorhynchus elegans did not drop out of even the driest, most stunted forest. These

species were observed foraging apart from mixed-species flocks in these drier habitats.

The non-flocking shrub gleaners were caught less often, but there were also a number

of species replacements, including some of the WSF specialist species (Myrmeciza

castanea, Percnostola arenarum), suggesting the possibility of competitive exclusion.

Insectivores dominated all white sand and clay habitats.

Insectivores dominated all white sand and clay habitats. Within this dietary guild,

we expected that army ant followers would be overrepresented in WSF, especially

because we observed more frequent, larger ant swarms in this habitat, but there were

no differences in capture rates of army ant followers, although the number of species

dropped off in the chamizal habitat (Fig. 3-8A for capture rates and Fig. 3-8B for number

of species per guild per habitat). Ant-following insectivores are sensitive to the isolation

of forest fragments (Stouffer and Bierregaard 1995), but they did not differ in capture

rates. In general, the ant followers showed a tendency towards nestedness with the loss

of the larger obligate ant followers (fewer Phlegopsis and Myrmeciza fortis) in WSF with

declining soil productivity.

Given the greater thickness of the leaf litter in WSFs (Table 3-1), we might have

predicted higher capture rates of terrestrial insectivores in this habitat, but the pattern

actually showed the opposite trend. The thicker, sclerophyllous leaves of WSF and the

dry soil may slow the decay of the leaf litter, which may reduce the productivity of this

litter layer such that it is thicker, but less productive. Dead leaf probing gleaners had

approximately the same number of species in clay, humid and dry varillal, but fewer in

Page 67: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

67

chamizal. Thryothorus coraya and Hyloctistes subulatus were not detected in chamizal,

but were captured in other habitats. Other dead leaf probers such as Ancistrops

strigilatus and Automolus infuscatus were only captured once, and were very rare in our

samples. In chamizal, fewer species were captured in all three of these foraging

substrate categories. Sally-glean foraging species that should prefer more open habitats

did not show any differences in capture rates among the habitats, and the number of

species did not vary either, which suggests that both forest types offer rich and varied

opportunities for species that search for insects far from their perches.

While capture rates did not differ for nectarivores and frugivores, both of these

foraging guild showed higher observed species richness in dry and humid varillal forests

(Fig. 3-9A & Fig. 3-9B). Nectarivore capture rates also showed a marginal trend towards

higher capture rates in terra firme clay forests. However, these idiosyncratic patterns

can be explained by the frequent movement of these birds through the forest, because

most species were traplining hermit species (Stiles 1975). Contrary to our predictions,

primarily frugivorous species were not captured more often in either habitat, but this

guild was more speciose in white sands. We predicted higher abundance of frugivores

in WSF because the understory vegetation is known to be richer in melastomes in WSF

than in clay soil forests (Ruokolainen et al. 1997, Thessler et al. 2005, Ruokolainen et

al. 2007, Pomara et al. 2012). Nevertheless, the diversity of frugivores was high in WSF

and one migratory frugivore that had a previously unknown winter distribution and

habitat preference, was almost entirely restricted to WSF (Ungvari-Martin et al. 2016).

Many specialist plants in white sand forests exhibited mast fruiting (Janzen 1974,

Macedo and Prance 1978, Arbelaez and Parrado-Rosselli 2005), producing cyclical

Page 68: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

68

resources and periods of fruit scarcity. The commonest species detected by Alonso et

al, Claravis pretiosa, which was essentially absent during our study years, belongs to a

genus that exploits masting species such as bamboo. A canopy frugivorous cotinga

(Xipholena purpurata) that was common in 2006 (Robinson, pers.obs) was only rarely

observed.

Some of the biases associated with mistnet samples were not as strong as we

had predicted. There was a slight increase in capture rates of canopy birds in the

shortest WSF, but this pattern was not statistically different. We acknowledge that mist-

netting provides a biased sample of the bird community (Remsen Jr and Good 1996).

For these reasons, we excluded canopy birds from the comparisons and rarefaction

analyses (Fig. 3-11) for non-canopy species only show similar asymptotic species

accumulation curves and similar differences between clay soil and the WSF habitats.

Although our results from mistnet captures were generally in accord with the

results from song censuses from this particular Peruvian white sand forest site (Alonso

and Whitney 2003, Alvarez Alonso et al. 2012), there were some notable differences.

The results described above are not completely in concordance with the initial published

findings. We confirmed that Electron platyrhynchus, Bucco tamatia, Xiphorhynchus

guttatus, Pigyptila stellaris, Myiobius barbatus, and Saltator grossus were species rare

or absent from white sand forest patches, but present in the AMNR. However, in WSF

we did consistently capture Thamnomaes ardesiacus and T. caesius, Myrmeciza fortis,

Machaeropterus regulus, Manacus manacus, Lepidothrix coronata and one individual of

Turdus lawrencii. We also recorded recaptures for these species, suggesting that WSFs

are not simply serving as holding areas for young, non-breeding floaters that are just

Page 69: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

69

moving through this habitat or residing in it for short periods while waiting for territories

in clay forest to open up. Thus, the WSFs were not as species-poor as originally

thought, at least in the taller humid WSF. More recent publications from this area also

discuss a growing list of apparent habitat specialists from the site (Álvarez Alonso et al.

2013).

Floristic composition, forest structure and arthropod assemblages vary across

different Amazonian forest types (Vormisto et al. 2000, Saaksjarvi et al. 2006, Fine

2010, Lamarre et al. 2012, Lamarre et al. 2016b). Most arthropod functional groups

showed higher relative abundances in terra firme clay forests than in WSF on a large

regional scale (Lamarre et al. 2016b), and the same pattern holds for Lepidoptera

assessed on a local scale as well (Lamarre et al. 2016a). However, the abundance and

species richness of Hymenoptera differ from the above pattern, and the species

composition of this group is highly heterogeneous and does not correlate to floristic

composition or productivity patterns (Saaksjarvi et al. 2004, Fine et al. 2006, Saaksjarvi

et al. 2006, Gomez et al. 2015). These differences in forest structure, composition, and

arthropod community composition can scale up to explain the slight differences we

found in the bird communities inhabiting WSF habitat and terra firme clay forests.

Further analyses in fine scale community patters, and more chamizal sampling could

elucidate the differences and the mechanisms behind them.

The most important finding from our dataset is how similar WSF is to the

surrounding terra firme matrix. Despite the differences in vegetation, the low productivity

forest growing on sandy soil is able to maintain a diverse community of birds.

Understanding the factors that regulate patterns of species diversity, turnover and

Page 70: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

70

richness of communities is a fundamental ecological problem, and productivity has been

a focus of many studies (Mittelbach et al. 2001). Without direct measure of productivity,

vegetation structure can serve as a proxy for comparing forested habitats, but the

differences in productivity change so many different aspects of community structure that

it is not surprising that we found a variety of guild-specific responses to plant

communities growing on different soil types. Apart from floristic differences, it may be

structural complexity alone driving bird communities and bird species diversity, just as

described by MacArthur’s classic and highly cited hypothesis that the vertical profile of

foliage drives bird species diversity (MacArthur 1961, Macarthur 1964, Karr and Roth

1971, Pearman 2002).

Page 71: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

71

Table 3-1. Vegetation structure environmental variables used for comparisons of the four habitat types. All values represent mean values for habitat types ± standard deviations, and these variables had significant differences among habitat types.

HABITAT Canopy Height (m)

Leaf litter Depth (cm)

DBH trees (cm)

Number of Large Stems

Number of Small Stems

Clay 17.38 ± 4.6a 36.12 ± 20.1a 21.06 ± 5.3a 35.25 ± 9.8a 45.21±31.5a

Humid 15.71 ± 3.9b 46.8 ± 24.1bc 19.43 ± 4.9a 14.71 ± 10.1a 46.48±25.7a

Dry 14.34 ± 3.5c 41.72 ± 20.5bc 18.87 ± 4.5b 13.66 ± 7.4a 59.58±37.4b

Chamizal 7.5 ± 1.15d 50.5 ± 12.7ab 16.48 ± 4.8b 12.19 ± 7.9b 69.56±18.3b

Different superscripts abcd denote statistically different values.

Page 72: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

72

Table 3-2. Dietary guild comparisons of the four habitat types. We assessed community composition using broad dietary categories and no differences were found in the trophic organization of the understory bird community.

Dietary Guild clay humid dry chamizal

Insectivores 0.613208 0.59596 0.622642 0.54717

Omnivores 0.188679 0.181818 0.141509 0.169811

Frugivores 0.084906 0.10101 0.103774 0.169811

Insectivores-Nectarivores 0.066038 0.090909 0.103774 0.09434

Carnivores and piscivores 0.04717 0.030303 0.028302 0.018868

Page 73: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

73

Table 3-3. Capture rate comparisons for all birds and for select guilds or functional groups in the four habitat types using individuals per 500 net hour units. All values represent mean number of individual birds for the guild in the habitat types ± standard deviations.

Guilds Clay Humid Dry Chamizal MANOVA

Total 100.37±31.61 88.89±23.32 88.32±31.74 66.67±9

Mixed Species Flocks 10.22±6.15 6.52±3.65 6.35±3.57 5±2.28 **

Shrub Gleaners 42.68±13.86 29.3±11.12 28.9±10.09 20.33±4.5 ***

Ant Followers 12.26±5.67 11.81±5.31 14.98±7.13 9.67±5.35

Terrestrial Insectivores 2.33±1.78 2±1.23 1.76±0.97 1.5±0.71 ***

Dead Leaf Probers 2.94±1.55 3.13±1.46 2.11±1.1 2.2±1.3 *

Sally-Gleaners 6.23±3.13 5.54±2.59 6.4±3.22 5.5±1.38

Nectarivores 6.21±2.59 3.89±2.5 5.32±3.42 2.6±1.52 *

Non-canopy Frugivores 7±3.53 6.27±2.65 5.46±2.69 5±2.53

MANOVA Significance codes: ‘***’p< 0.001, ‘**’ p<0.01, ‘*’ p<0.05

Page 74: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

74

Table 3-4. Diversity indices comparisons of bird communities of the four habitat types for all species and for white sand forest specialists alone from the AMNR capture data. The last column indicates significance levels of ANOVA statistical test, numbers represent mean values with standard deviations.

Comparison Subset Clay (n=19)

Humid (n=27)

Dry (n=41)

Chamizal (n=6) P

Mean Capture all species 100.37±31.61 88.89±23.32 88.32±31.74 66.67±9

Recaptures all species 0.18±0.16 0.22±0.22 0.2±0.29 0.17±0.33

No. Species all species 39.55±15.44a 37.36±9.91a 36.65±10.1a 30.33±5.51b *

Species Recaptured all species 0.21±0.47 0.32±0.4 0.31±0.43 0.27±0.61

Shannon Index all species 3.11±0.3a 2.7±0.3b 2.8±0.21b 2.7±0.21b ***

Simpson Index all species 0.94±0.03a 0.89±0.05b 0.91±0.02b 0.90±0.03ab *

Fisher’s Alpha all species 19.69±6.1a 12.39±3.7b 12.7±3.1b 13.1±5.1b ***

Mean Capture specialists 4±2.67 6.16±4.13 6.88±3.51 7.17±3.19

Mean No. Species specialists 1.5±0.85 2.16±1.18 2.76±1.24 4.17±0.41

Recaptures specialists 0.18 ± 0.33 0.09 ± 0.15 0.14 ± 0.26 0.2 ± 0.18

Shannon Index specialists 0.29 ± 0.48a 0.63 ± 0.55b 0.91 ± 0.49b 1.42 ± 0.09c ***

Simpson Index specialists 0.18 ± 0.3a 0.39 ± 0.31b 0.54 ± 0.25b 0.76 ± 0.02c ***

Fisher’s Alpha specialists 21.04±061a 21.4±0.85a 21.15±0.99a 22.33±0.73b ***

Significance codes: ‘***’p< 0.001, ‘**’ p<0.01, ‘*’ p<0.05 Different superscripts abc denote statistically different values.

Page 75: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

75

Figure 3-1. Comparison of number of species between two general habitat types.

Slightly more species were captured in clay soil forests than in white sands forests per sampling event (t-test, p= 0.01043), raw data are represented in the figures, log transformed data were used for statistical test.

Page 76: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

76

Figure 3-2. Rank abundance curves for white sand forests and clay forest captures.

Species are arranged according to their rank order abundance in clay soil forest, since terra firme is the dominant forest type. WSF ranks show differences in a few species.

Page 77: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

77

Figure 3-3. Forest canopy height in the four different habitat types, letter codes refer to

statistical differences (ANOVA, p<0.05), red points indicate outlier values, letter codes a-b-c-d indicate statistically different values.

Page 78: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

78

Figure 3-4. Cumulative distribution of species occurrences over all netlines (n=93

sampling events). Each point represent a species, and the number of sampling events at which that species was recorded. The number of species with very few occurrences is high, and the distribution is skewed, where most species were only recorded at a few sampling events.

Page 79: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

79

Figure 3-5. Species accumulation curves for four different habitat types. Rarefaction

analysis was performed to standardize comparisons with effort.

Page 80: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

80

Figure 3-6. Turnover, dissimilarity and nestedness between white sand and clay soil

forest bird communities.

Page 81: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

81

Figure 3-7. Ordination diagram based on Bray-Curtis Dissimilarity using netlines as

sites for all WSF and clay habitats. The grey lines connecting the points result from the hierarchical clustering analysis using the average linkage method.

Page 82: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

82

A)

B)

Figure 3-8. Comparison of capture rates and number of observed species of flocking or

non-flocking shrub gleaner insectivores. A) Species that are members of mixes-species foraging flocks, and non-flocking shrub gleaners were captured more in terra firme clay than in the WSF habitats (p= 0.0057 **, and p= 0.000014 ***,mean captures per sampling event with standard deviations). B) The observed number of species per guild in each habitat type varied little except for the notable loss of species in chamizal habitat.

Page 83: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

83

A)

B)

Figure 3-9. Comparisons of captures and species according to foraging methods. A)

Capture rates of dead-leaf probing species and terrestrial insectivores varied somewhat between terra firme clay and in the WSF habitats, army ant follower and sally-gleaners were captured at similar rates. B) The observed number of species per guild is each group dropped considerably in the chamizal habitat. The highest number of sally gleaner species was captures in dry varillal.

Page 84: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

84

A)

B)

Figure 3-10. Comparisons of captures and species according to diet A) Both

nectarivore species and frugivores were captured at approximately equal rates on average in terra firme clay and in the WSF habitats, though somewhat fewer captures in humid varillal B) The observed number of species per guild increased between clay, humid and dry varillales, and dropped considerably in the chamizal habitat.

Page 85: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

85

Figure 3-11. Rarefaction based species accumulation curves show the differences for

white sand forest habitats and clay habitat. Overlapping 95% confidence intervals are interpreted as no difference.

Page 86: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

86

CHAPTER 4 MOVEMENT, ISOLATION AND GENE FLOW AMONG PATCHES OF WHITE SAND

FORESTS

Introduction

One of the most critical questions in the origin and maintenance of species

diversity is the extent to which isolation of habitat patches affects species composition,

dispersal, gene flow and, thereby, incipient speciation, local evolution and

differentiation. In the Amazon, some habitat patches such as riverine habitats (igapo,

varzea, Mauritia palm swamps) are intrinsically connected as they form along linear

habitat features such as rivers, but other habitats such as white sands forests (see

Chapter 2) occur in isolated patches embedded in a matrix of terra firme clay soils

(Pitman et al. 2008, Stropp et al. 2011). Bamboo and river islands are ephemeral, can

vary greatly in size, and require life-history adaptations that maximize dispersal ability to

find patches (e. g., when the bamboo dies) (Kratter 1997, Santana and dos Anjos

2010).

White sands forests form on essentially permanent soil formations and often

consist of small, isolated patches or archipelagos (Anderson 1981, Brown and Prance

1987, Prance 1996, Borges et al. 2016). While the white sand ecosystem is widely

distributed throughout Northern Amazonia, the area of patches ranges from as small as

less than one square kilometer to thousands of square kilometers (Adeney et al. 2016).

Terra firme clay soils surround these patches (Anderson 1981, Stropp et al. 2011),

potentially creating barriers to gene flow among white sand patches (Capurucho et al.

2013). Such habitat patches therefore have the potential to be crucibles of evolution in

which local differentiation can take place if gene flow is indeed limited (Frasier et al.

2008). The small size and isolation of many WSF patches probably contributes to the

Page 87: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

87

large number of endemic species and their low species richness in general (Steege et

al. 2000, Adeney et al. 2016). Yet, the community analyses in Chapter 2 and Chapter 3

suggested that WSF patches have very similar community composition to other forest

types, which suggests that there may be fewer barriers to gene flow. In this chapter, we

take this analysis one step further by examining the extent to which patches of one

habitat type—white sands forests—are interconnected within a region.

AMNR forms the largest archipelago of white sand forest islands in Western

Amazonia, and this area is somewhat isolated from the larger expanses of sandy

Amazonian habitats within the same biogeographical region (Fine et al. 2010, Álvarez

Alonso et al. 2013, Adeney et al. 2016). The reserve is located west of the Napo River,

and north of the Amazon River, within the Napo ecoregion. A subset of white sand

forest specialist bird species inhabit the reserve; not all species can be found here that

are considered WSF specialists (Alvarez Alonso 2002, Alonso et al. 2013). Comparative

studies have been carried out in the past decade (Pomara et al. 2012, Álvarez Alonso et

al. 2013) evaluating bird species composition in the region We used detailed maps of

the white sand islands to evaluate the species-area relationship in the white sand

habitats and the effects of the degree of isolation on the patches.

The questions we were interested in were the following: First, does patch size

and distance between patches affect species richness in white sand forests? Second,

do white sand forest specialists ever occur in the matrix outside of WSF patches? And

third, can we find evidence of dispersal or gene flow among these patches? We

hypothesized that species richness of habitat specialists in a habitat patch could be

predicted by its area, and therefore we predicted that larger white sand forest patches

Page 88: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

88

would harbor more species of white sand specialists (MacArthur and Wilson 1967,

Connor 1979, Lomolino 2000, Lomolino 2001). If species were randomly distributed

among patches, the resulting curve of species composition and patch area would rise

gradually. If only small subsets of species survive on small patches, then we would

expect an S shaped curve, indicating a threshold patch size where species richness

would level out.

Patch isolation should also affect community composition if WSF specialists are

dispersal-limited. If specialists are less likely to move longer distances across the matrix

habitats, we predicted that isolated patches would have fewer WSF specialists.

Alternatively, there may be no difference in community composition, which would

suggest that the clay soil matrix does not act as a significant barrier to dispersal. In this

case, the patchy nature of their habitat may select for a bimodal distribution of dispersal

distances in WSF specialists with a high incidence of very short movements that keep

the birds within their patches, and a smaller number of very long-distance movements

between different patches compared with habitat generalists, which should have more

intermediate length movements.

In addition to field observations of potential movement, we inferred dispersal

among WSF patches by estimating gene flow among the populations. The genome

provides a record of the demographic processes acting on a given lineage. The diversity

in a population and the divergence between individuals allows us to elucidate the

processes that shape the diversity. Isolation-by-distance describes the tendency of

individuals to find mates from nearby populations rather than distant populations (Wright

1943). As a result of this tendency, populations that live near each other are genetically

Page 89: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

89

more similar than populations that live further apart. Alternatively, patchy habitats can

produce signatures of small island population models, which predict abrupt changes

among the populations on different islands (Wright 1943, Slatkin 1993). Populations of

obligate white-sand specialists living on different white sand forest patches are

connected by dispersal on a local scale but may become increasingly isolated on the

regional scale.

We examined genetic variation and migration rates within and among

populations and we examined these variables for a white-sand specialist focal species

(Neopelma chrysocephalum) living in a landscape in which patches of white sands

forests are embedded in a matrix of clay sand forest. The genus Neopelma exhibits

largely allopatric, almost circum-Amazonian distribution with five recognized species;

together with the genus Tyranneutes, which is sister to all other taxa of manakins, they

form one of the basal lineages of the family Pipridae (Tello et al. 2009, McKay et al.

2010). N. chrysocephalum is widely distributed across the Guiana Shield, but it was only

recently documented in Northeastern Peru in white sand forests, where it is fairly

common (Alonso and Whitney 2003), its behavior and distribution throughout the range

remain poorly known (Kirwan and Green 2011). If the clay forest matrix in which these

patches are embedded act as dispersal barriers, we predict that genetic differentiation

among populations would increase with distance. Alternatively, if the matrix is not an

effective barrier, then we predicted that there would be little difference in the genetic

structure of populations among patches.

Page 90: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

90

Methods

Study Landscape

We used the AMNR study areas, regions and netlines represented on the map

(Fig. 4-1). We evaluated species composition as a function of patch size and insularity

or isolation from other patches of similar habitat. We plotted each sampling event for

each area cumulatively. We plotted all species as well just the white sands habitat

specialists as defined by Alvarez et al. (2002). The cumulative number of species using

the Chao estimate of richness (Gotelli and Chao 2013) was plotted against area of each

patch and region. Area calculations were carried out with the raster calculator in

ArcMAP version 10.0 (ESRI 2009). We then used several approaches based on species

richness estimations to examine variation in bird species composition and to evaluate

correspondence of bird communities with patch size.

Species-area Relationships

We fitted different models for species richness and patch area, patch isolation,

and the area of the nearest neighboring patch, and applied AIC model selection to test

for area effects within white sand habitats. A model selection approach based on AIC

(Burnham & Anderson 2002) was applied to select among possible hypotheses

regarding the source of variation in diversity or composition indices (i.e., effects of area,

distance, or both). Species totals for each “island” were averaged and linear models of

species richness as a function area were constructed with log-transformed variables,

models were compared using AIC.

Molecular Analysis

Blood samples taken from 141 individuals during mist-netting sessions were

used to quantify between-patch gene flow on different spatial scales in the habitat

Page 91: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

91

specialist species Neopelma chrysocephalum. Genomic DNA was extracted from blood

samples using Qiagen DNeasy kits (Quiagen, Inc., Valencia, California, USA) following

the manufacturers protocol. All samples’ DNA concentrations were quantified using

nanodropping, and the mean fragment size of each sample was further assessed using

gel electrophoresis.

We used a double digest RAD sequencing (ddRADseq) approach to obtain

variable genetic markers for the population level sampling (Barber et al. 2004). The

ddRADseq libraries for each individual were prepared using a slightly modified version

of the protocol from Peterson et al (2012). Each sample was assigned a unique 8-9-10-

or 14 basepair (bp) long inline barcode, equimolar concentration of uniquely barcoded

individuals were pooled and double indexed by 11 cycles of high fidelity PCR. The PCR

products were pooled and sequences in one lane of a high throughput run of on a

NextSeq 500 sequencer (Illumina, San Diego, CA, USA) at the ICBR of University of

Florida, producing 150-bp single-end reads. A description of detailed methods on

sample preparation is available in Appendix C.

All samples were sequenced at the University of Florida Cancer and Genetics

Research Center in a single lane of the NextSeq 500 following a sequencing protocol

for 150bp single-end reads. We prepared the raw Illumina FASTQ files for analysis by

processing the sequence data using the program process_radtags in STACKS version

1.35 (Catchen et al. 2013). First, we sorted the read pairs by barcode, and removed any

reads that did not contain both a correct barcode and the remaining bases of the

restriction site sequence. When demultiplexing the samples, we did not allow for

sequencing error in the barcodes and discarded reads that included low quality sites

Page 92: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

92

using the Phred quality scoring system, only keeping reads with greater than 33 quality

score reads. We retained 83.52% of the raw reads, and 91.2% of the reads had higher

than 30 quality scores (Appendix D, Table D-1). We retained 316,455,808 reads 130 bp

or greater in length, and trimmed all reads to 130 bp for the STACKS pipeline that

requires uniform read length. The reads were merged into a single file per individuals,

followed by catalog building using denovo_map.pl Perl script. We used a minimum

depth of coverage (-m) of 2 to create stacks, a maximum nucleotide distance between

stacks (-M) of 3 when identifying the loci, and a maximum number of mismatches

between sample tags (-n) of 2 when generating the catalog (Catchen et al. 2013).

These parameters were chosen following Catchen et al 2013. We enabled the

Deleveraging and Removal algorithms to filter out highly repetitive likely paralogous loci.

These analyses were carried out on the University of Florida HPC server (HyperGator).

Population structure analysis

We eliminated some individuals due to low sequence coverage, then clustered

the reads into RAD loci, identified the RAD loci containing SNPs and genotyped all

samples by calling SNPs in each individual (Appendix D, Table D-2) using a maximum

likelihood statistical model implemented in STACKS (Davey et al. 2011, Hohenlohe et

al. 2011, Hohenlohe et al. 2012). Then, we used load_radtags.pl and index_radtags.pl

to locally populate and index a MYSQL database of loci. Using the scripts export_sql.pl

and populations.pl, we selected the polymorphic loci that were present in 90% of the

individuals. We only selected the first biallelic or triallelic SNP at loci that contained

multiple SNPs to reduce the possible effects of linkage. In addition, using the script

populations.pl we calculated several genetic diversity parameters (expected and

observed heterozygosity, nucleotide diversity and inbreeding coefficient) based on one

Page 93: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

93

SNP for each RAD locus for each of the putative geographic populations and the whole

landscape level metapopulation.

Information on the RAD loci identified will be submitted to Dryad, including

information on each locus present in at least 120 individuals and consensus genotypes

for each individual. Appendix D Table D-1 shows the number of reads obtained in the

run and after the quality filtering steps. We only used loci present in all putative

populations, and in at least 75% of the samples. We also calculated several genetic

diversity parameters for separate groups and overall, basic indices of genetic diversity

within populations, such as degree of heterozygosity, number of alleles or number of

polymorphic loci (Appendix D Table D-3). STACKS assembled a catalog containing a

raw total of 349,166 loci. 144,003 of the total were polymorphic loci, an after filtering

procedure in populations with the specific resulted in levels of completeness ranging

from 80-100%. After filtering for a minimum stack depth of 5 and presence in at least

75% of the individuals of all populations 87125 loci were retained. From this set of loci,

at least one SNP variable sites were found among 49155 of the loci, resulting in a data

matrix was 75 % complete across 132 individuals in the 8 putative populations defined

by WSF patches. We further filtered loci for low minor allele frequency across the

dataset (>10%) to avoid not being able to tell the difference between rare alleles and

sequencing error.

We conducted three different tests to identify population structure. First we

calculated population differentiation FST statistics with STACKS using the population

program. Then we used the program STRUCTURE, version 2.3.4 (Pritchard et al. 2000)

to establish the number of distinct populations through cluster analyses based on the

Page 94: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

94

genotype data alone, as well as GENODIVE (Meirmans and Van Tienderen 2004),

which has unique functions that are not implemented in other programs including

Analysis of Molecular Variance. We evaluated genetic diversity and allelic richness in

populations that occur in different patches. In STRUCTURE we used possible K values

(number of populations) ranging from 1-8, one population per sampling region. We used

20 runs for each K value. The Markov chain was run for 5.0 × 105 steps per analysis,

with a burn-in of 1.0 × 105. Log likelihood values were averaged across runs for each K

value, and these values were evaluated along with patterns of individual assignment to

clusters to draw inference regarding the optimal K value for each dataset. Results were

summarized with STRUCTURE HARVESTER (Earl and vonHoldt 2012), figures were

prepared using the R package adegenet (Jombart 2008, Jombart and Ahmed 2011).

Results

Movements Detected by Recaptures

Most species in the sample were recaptured on the same netline and within the

same patch across years. We intentionally scheduled sampling dates to correspond

with the previous year’s sampling dates as close as possible. Only Dendrocincla

merula, a large ant-follower woodcreeper, was consistently recaptured on different

netlines; over 25% of recaptures of this species between years occurred on a different

netline (Table 4-1). Across all species, most movements were relatively short distance,

between 300-700m, but some individuals were captured as far as 6km from the initial

capture site. The second widest-ranging species, Pithys albifrons, was also an obligate

ant-follower species that was not restricted to white sands forests (Table 4-1). Of the

WSF specialists, only Neopelma chrysocephalum was recaptured on a different netline

(see below).

Page 95: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

95

Species – Area Relationships

We estimated species richness for all captured species as well as just the habitat

specialists on each WSF habitat patch. Habitat patch sizes varied from as small as 6

hectares to over 800 hectares (Table 4-2). White sand forest habitat patches were

separated from each other by distances ranging from 435m to 3.3km, and each patch in

our analysis had another WSF patch at least 5 hectares in size within those distances

(all localities and spatial data are summarized in Table 4-2). According to the AIC model

testing, the best model to explain the patterns of species richness for both sets was the

full model that included all three factors (patch area, the area of the nearest patch, and

the distance to the nearest patch) with a Poisson distribution (Table 4-3). The results do

not show strong support for patch dynamics in this system, and none of the spatial

variables explained the variation in species richness. White sand forest specialist

species can be found in all of the patch sizes, and estimated species richness in this

subset of species does not increase with patch area, as we predicted (Fig. 4-2).

Genetic Diversity and Population Assignment Tests

We obtained estimates of genetic diversity of Neopelma chrysocephalum in each

population, examining polymorphic loci only, for the most complete dataset of 2240

biallelic loci and 129 individuals (Table 4-4). Pairwise FST values (Table 4-5) calculated

for the data set were consistent across all methods, and showed little to no genetic

differentiation among the populations. We found evidence for panmixia, and identified 9

individuals that were migrants in the population using GenoDive, as well as using

STRUCTURE’s admixture analysis. Genetic distances between inferred populations

and the observed populations were consistent between STRUCTURE and GenoDive.

We tested population differentiation, and consistently, across all datasets including all

Page 96: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

96

genotyped individuals and various subsets of markers, the best clustering was a single

cluster of individuals k=1 (Fig. 4-3) according to Bayesian Information Criterion (BIC)

using among clusters sums of squares from an AMOVA, and also using Euclidean

distances calculated from within-individuals allele frequencies implemented in GenoDive

(Meirmans and Van Tienderen 2004). Taken together the analyses all suggest high

levels of gene flow throughout the populations.

Discussion

In this study, we found evidence that habitat specialization does not necessarily

prevent birds from moving across the matrix of a different habitat. One individual of the

obligate habitat specialist Neopelma chrysocephalum exhibited an inter-patch

movement (~6.5 km) that can be considered a dispersal event. In addition, white sand

specialist species were found in all patch sizes, and the distance between patches also

did not explain their presence and the estimated species richness of specialists,

indicating that the birds can move across other forested habitats as well. In combination

with the strong similarity in communities of forested habitats documented in Chapter 2,

we can conclude that dispersal does not strongly constrain bird communities of WSF.

Two species of ant-following birds exhibited the largest proportion of birds

recaptured at distant sites: Dendrocincla merula and Pithys albifrons. These data imply

either very large home ranges or a large area over which young birds disperse. Large

home ranges are especially likely because these species are known to move through

the forest understory searching for ant swarms (Willis and Oniki 1992, Schulenberg et

al. 2007). We observed only a single Neopelma chrysocephalum moving across

patches: this bird moved 6.5 km between 2010 and 2011, and then another 670 m

between 2011 and 2012. Based on the brown blotches in the eyes, we identified this

Page 97: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

97

bird as a juvenile in 2010. In the subsequent years, the bird’s iris color changed to

creamy yellow. The same bird was captured twice in 2012 on the same netline. Other

individuals of Neopelma that moved only traveled between 350m-450m distances from

initial to subsequent capture sites (n=3).

The circum-Amazon distribution of the genus Neopelma implies that this species

either disperses readily now or did so in the past. There are currently no studies of the

phylogenetics or the population genetics of any members of this genus. Recently, a

study examined the population genetics of another white sand habitat specialist

manakin species, Xenopipo atronitens (Capurucho et al. 2013) on a larger geographic

scale. Consistent with our findings: they found no strong isolation-by-distance among

sites within the landscape, and geographic distance explained very little of the variation

in genetic distance.

Interestingly, landscape genetic analyses indicate that terra firme forests are less

permeable to dispersal, acting as a barrier to gene flow for X. atronitens. According to

our results of mark-recapture as well as genetics, Neopelma chrysocephalum can move

through the terra firme forest, at least on the spatial scale of white sand archipelagos.

Inventories in campina habitats reveal a widespread bird community with little local

endemism, meanwhile the Western Amazonian white sands harbor seemingly disjunct

populations of some species. Therefore, further genetic studies of white sand specialist

birds can complement the current knowledge about Amazonian landscape history,

revealing patterns that have not been detected so far.

Studies of migration rate often rely on two approaches: a coalescent approach,

which uses the genealogical information contained in DNA sequences, and a multilocus

Page 98: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

98

genotype approach. Generally, the coalescent approaches (e.g., those implemented in

the program MIGRATE; Beerli and Felsenstein 2001) estimate long term evolutionary

parameters in populations geographically isolated with little migration over longer time-

scales, whereas the genotype approach estimates short-term parameters in populations

in the same geographical area with potential migrants every generation. We can

implement our data in the future using coalescent approaches to estimate effective

population size and assess the long term stability of the population.

The persistence of a population depends on dispersal across fragmented

habitats (Hanski 1994, Moilanen 1998, Thomas and Kunin 1999, Van Houtan et al.

2007). Biogeography and ecology in Amazonia reflect consequences of regionally

varying environmental factors, both past and present. Better understanding of these

factors is critical to reconstructing the evolution of Amazonian biota, and to be able to

plan and manage Amazonian ecosystems in a sustainable manner. WSF bird

communities appear to be well adapted to persist as long as some dispersal corridors

remain to connect the patches, and even small patches can harbor the habitat specialist

species.

Page 99: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

99

Table 4-1. The twenty most captured bird species, and the recaptures and movement proportions of those species. Movements are defined as recaptures on a different netline.

Species Number of Banded birds

Recaptures between years (%)

Movements (%)

Year-to-year movements (%)

Dendrocincla merula 138 52.90 44.20 25.36 Pithys albifrons 230 29.13 20.43 8.26 Dendrocincla fuliginosa 78 15.38 10.26 7.69 Willisornis poecilonotus 270 27.04 5.56 4.44 Xenops minutus 46 26.09 8.70 4.35 Xiphorhynchus elegans 138 21.01 5.80 4.35 Myrmoborus myotherinus 98 12.24 4.08 4.08 Dixiphia pipra 469 16.20 5.54 4.05 Schiffornis turdina 58 20.69 6.90 3.45 Mionectes oleagineus 332 8.73 3.31 3.31 Glyphorynchus spirurus 875 26.51 5.14 3.09 Gymnopithys leucaspis 209 21.05 9.57 2.87 Neopelma chrysocephalum 141 25.53 3.55 2.84 Lepidothrix coronata 145 8.28 2.76 2.76 Myrmotherula axillaris 159 9.43 3.14 1.89 Myrmotherula hauxwelli 127 10.24 2.36 1.57 Megastictus margaritatus 135 11.85 5.19 1.48 Automolus ochrolaemus 74 18.92 2.70 1.35 Corythopis torquatus 45 24.44 2.22 0.00 Percnostola arenarum 34 20.59 2.94 0.00

Page 100: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

100

Table 4-2. Localities, area of WSF patches, distances to nearest patch, and area of nearest patch to the sampled patch. Lat/lon coordinates are in UTM geographic coordinate system, UTM 18S.

Lon Lat Area (ha)

Distance to near patch (m)

Area of near patch(ha) REGION Habitat

674665.6 9560322 6.77 1643.98 15.19 km 28 Dry

672846.1 9558681 51.75 981.49 68.86 km 31 Dry

660513.5 9566505 87.16 909.15 51.91 Porvenir Humid

667166.6 9568501 102.72 954.15 7.63 Mishana Dry

659549.4 9566840 116.18 1020.57 87.16 Porvenir Humid

673338 9564324 124.30 1623.89 14.00 km 28 Chamizal

673243.4 9570139 148.43 1206.41 37.69 San Martin Humid

675195.2 9570028 186.09 781.53 19.70 San Martin Humid

662287.1 9567414 208.84 435.16 208.84 Yuto Dry

674455.4 9568191 219.07 1378.30 15.29 Nueva Esperanza

Humid

676660.3 9563815 255.21 3361.10 124.30 km 25 Humid

670579.1 9568112 847.83 2091.88 37.69 Mishana Dry

Page 101: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

101

Table 4-3. Model selection results of the explanatory values of area of WSF patches, distances to nearest patch, and area of nearest patch to the sampled patch.

Models Distribution Intercept Area Near Area Near Dist AIC

Area Poisson 3.716 0.0004534 NA NA 49.5279

Area Negative Binomial

3.722 0.0004314 NA NA 50.1910

Near Area Poisson 3.960 NA -0.0013457 NA 51.7558

Near Area Negative Binomial

3.958 NA -0.0013171 NA 51.0828

Near Dist Poisson 4.008 NA NA -0.0001009 51.4814

Near Dist Negative Binomial

4.022 NA NA -0.0001101 50.8433

Full Poisson 4.079 0.0006489 -0.0012541 -0.0002117 45.1498

Full Negative Binomial

4.079 0.0006489 -0.0012541 -0.0002117 47.1499

Page 102: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

102

Table 4-4. Summary of sampling locality, samples in the locality, and summary of raw sequence information for those samples.

Region Samples Mean raw reads

Mean unique stacks

Mean polymorphic loci

km 25 9 1.77x105 4.97x104 2.19x104

km 28 14 2.19x105 6.81x104 2.50x104

km 31 6 1.13x105 9.76x104 1.29x104

Mishana 34 1.66x105 5.13x104 1.95x104

Nueva_Esperanza 31 1.48x105 6.12x104 1.69x104

Porvenir 27 1.62x105 4.92x104 1.79x104

San_Martin 20 1.68x105 5.35x104 1.94x104

Yuto 37 1.79x105 4.84x104 2.40x104

Page 103: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

103

Table 4-5. Summary of population differentiation results based on FST calculations.

KM_28 San_Martin Mishana Yuto Nueva_ Esperanza km31 Porvenir km_25

KM_28 0 0.0082 0.0067 0.007 0.0061 0.0133 0.0067 0.0147

San_Martin 0 0.0043 0.0049 0.0046 0.0113 0.0046 0.0103

Mishana 0 0.0041 0.004 0.009 0.0045 0.0085

Yuto 0 0.0038 0.0093 0.0046 0.0073

Nueva_Esperanza 0 0.0095 0.0043 0.0103

km31 0 0.0097 0.0204

Porvenir 0 0.0102

Page 104: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

104

Figure 4-1. Map of white sand forest patches including the 150 m buffer from the edge

for each forest island. The names of the putatively isolated regions are included on the map.

Page 105: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

105

Figure 4-2. The relationship between species richness and the area of the sampled

white sand patch, the area of the nearest WSF patch and the distance to the nearest WSF patch. Blue points represent WSF specialist species.

Page 106: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

106

Figure 4-3. Population genetics results based on k-means clustering as well as

principal component analysis based on allele frequencies indicate little differentiation between regional populations, and a single genetic cluster.

Page 107: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

107

CHAPTER 5 INTER-ANNUAL SITE FIDELITY AND BREEDING ORIGINS OF GRAY-CHEEKED

THRUSHES IN WHITE SAND FORESTS OF THE PERUVIAN AMAZON 1

Introduction

The annual migration of Gray-cheeked Thrushes (Catharus minimus) from boreal

forests in northern North America and eastern Siberia to Amazonian rainforest is among

the longest by a Nearctic-Neotropical migratory songbird with some individuals making

a 20,000 km round trip. Most of the annual cycle of Gray-cheeked Thrushes is spent in

South America yet we are lacking knowledge about their non-breeding ecology in the

tropics, including movements, ecological roles, and seasonal habitat use. Recent

advances in tracking technology have revealed that Catharus thrushes may undertake

significant intratropical movement than can span much of their non-breeding season

(Heckscher et al. 2011, Cormier et al. 2013, Heckscher et al. 2015, Hobson et al.

2015). Although we lack details regarding movements of individual Gray-cheeked

Thrushes in South America, the possibility of seasonal movements hinders our ability to

draw inference regarding the habitat associations of sedentary birds because banding,

specimen, or sight records may represent individuals at stopover or temporary sites

rather than settled birds. This gap in our knowledge limits our ability to delimit accurate

winter ranges, assess threats, and convey conservation needs. The breeding origin of

Nearctic-Neotropical migratory birds occupying specific regions in South America and

1 Reprinted with permission from Ungvari-Martin, J., C. M. Heckscher, and K. A. Hobson. 2016. Inter-

annual site fidelity and breeding origins of Gray-cheeked Thrushes in white sand forests of the Peruvian Amazon. Journal of Field Ornithology 87:55-64.

Page 108: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

108

the connectivity between breeding regions and non-breeding sites is of interest to

conservationists (Hobson et al. 2014).

Only recently have advances in tracking technology and stable isotope analyses

begun to reveal that connectivity (Fraser et al. 2012, Cormier et al. 2013, Gallo et al.

2013, Hobson et al. 2014). Although the understanding of connectivity between

wintering and breeding regions is widely considered of paramount importance (Faaborg

et al. 2010)(Webster and Marra 2005, Faaborg et al. 2010), from a conservation

perspective, understanding habitat associations and the movement and settlement

patterns of Nearctic-Neotropical migrants in tropical regions is equally important

especially if populations depend on rare or uncommon tropical ecosystems.

White sand terra firme forests are scattered throughout lowlands of the western

Amazon basin often amid weathered clay terra firme forests. White sand systems differ

from clay forests in their unusual floristic composition, sandy substrate, scleromorphic

physiognomy, and low nutrient content (Anderson 1981). These forests cover just 3% of

the Amazon basin (Pennington 2009) and effectively function as ecological islands in

the vast forests of Amazonia (Prance 1996). They maintain distinct floral assemblages,

including many endemic plant species, and have been examined for their importance in

maintaining obligate and unusual assemblages of birds (Alvarez Alonso and Whitney

2003). Importantly, the use of white sand forests in western Amazonia by Nearctic-

Neotropical migratory songbirds has not been evaluated.

We report the capture of Gray-cheeked Thrushes over three field seasons in

white sand forests in the Amazon basin near Iquitos, Peru. The thrushes were captured

during a broader study directed by JUM that focused on white sand resident understory

Page 109: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

109

bird species. We report the thrushes’ association with white sand forests at our site,

provide the approximate date of first arrival, provide the first confirmation of inter-annual

site fidelity of this species to a non-breeding non-migratory site in South America, and

report the approximate breeding region of 12 individuals (including site-faithful

individuals) inferred via stable isotope (δ2H) analysis. We argue that Gray-cheeked

Thrushes are seasonal residents at our study area. To support our contention, we

compare our capture of Gray-cheeked Thrushes to those of a transient congener –

Swainson’s Thrushes (C. ustulatus) – also captured during this time period. To the best

of our knowledge, our results are the first to report on the occurrence, habitat affiliations,

and migratory connectivity of Gray-cheeked Thrush from a known non-migratory (i.e.,

“wintering”) site in South America. Specifics regarding the range limits and habitat

associations of Gray-cheeked Thrush in South America are lacking (Lowther et al.

2001); therefore, our results are an important contribution toward an understanding of

this species’ life history and conservation needs in the tropics.

Methods

Study Site

Our study site was the Allpahuayo-Mishana National Reserve in the Department

of Loreto, northeastern Peru (-3.922904, -73.516645; Fig. 5-1). The study area is

accessible from the Iquitos-Nauta highway, a paved road on the southern border of the

reserve, or the Nanay River along the northern border.

Field Sampling

The primary objective of the fieldwork was to assess the understory bird

community composition in white sand forests and adjacent clay soil terra firme sites. We

used constant-effort mist-netting to survey understory birds in the reserve from June to

Page 110: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

110

8 December 2010-2012. Net transects were established in closed-canopy white sand

forest patches that were accessible on foot (elevation: mean = 126.5m, range = 91.9 -

165.9m). Some nets were simultaneously placed in adjacent closed-canopy alluvial

terra firme clay based forest (elevation: mean = 116.7 m, range = 106.1 – 130.6 m). All

nets were 38-mm mesh, 12 m x 3 m, tethered with five trammels. We created transects

through the forest interior using a line of 20 nets set on existing foot trails in primary

forest or older secondary growth forest contiguous with primary forest. Nets were

separated from one another by <1 m when possible. The minimum distance between

transect sites was 250 m. The greatest distance between any two transects was 18 km

(Fig. 5-1). Each net transect was opened for ~500 net hrs during three consecutive

days. We opened nets by 06:00 and checked nets approximately every 30 min until

16:00 on the first two days and 11:00 on the third day. Thrushes were weighed, aged

using criteria in Pyle (2008), and subcutaneous fat scores (0 = none to 7 = very

excessive) were recorded using the MAPS protocol (DeSante et al. 2015). A single

rectrix was collected from Gray-cheeked Thrushes. After three days, we relocated each

transect; no site was sampled for more than three days in a given year.

We used the earliest first arrival date of Gray-cheeked Thrushes (13 October) to

calculate sampling effort for thrushes in the two forest types over the three seasons.

Nineteen transects were sampled each year during the period 13 October to 8

December with a total of ~20,000 net hours in white sand forest and ~3000 net hours in

clay forest. To test for a difference in capture rate between clay and white sand forests,

we used a Chi-square Goodness-of-fit test adjusted for sample effort. Student’s t-test

was used to compare the mean fat scores of Gray-cheeked Thrushes to the congener

Page 111: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

111

Swainson’s Thrush (C. ustulatus). All statistical tests were accomplished using the R

programming language version 3.0.1.

Stable Isotope Analysis

Feathers of Gray-cheeked Thrushes were sampled (N = 12) in 2012 after we

realized the opportunity to conduct stable isotope analyses posed by the capture of

numerous thrushes at our study site. A single rectrix was collected from individuals at

the time of capture and preserved in separate paper envelopes for future stable isotope

analysis. A 2:1 chloroform:methanol solvent rinse was used to remove surface oils.

Samples were prepared for stable-hydrogen isotope analysis and analyzed at the

Stable Isotope Hydrology and Ecology Laboratory of Environment Canada in

Saskatoon, Canada. Stable-hydrogen isotope analyses of feathers were conducted

using the comparative equilibration method (Wassenaar and Hobson 2001) through

use of calibrated keratin δ2H reference materials. Stable-hydrogen isotope

measurements were performed on H2 derived from high-temperature (1350˚C) flash

pyrolysis of 350 ± 10 ug feather subsamples using continuous-flow isotope-ratio mass

spectrometry. Measurement of three keratin laboratory reference materials (CBS, SPK,

and KHS; corrected for linear instrumental drift) were both accurate and precise, with

typical mean δ2H values of –197, -121, and -54.1‰, respectively. A control keratin

reference yielded a 6-month SD of ± 3.3‰ (N = 76) and within-run SD of standards

typically < 2‰. All results are for non-exchangeable δ2H expressed in the typical delta

notation, units per mil (‰), and normalized on the Vienna Standard Mean Ocean Water

– Standard Light Antarctic Precipitation (VSMOW-SLAP) standard scale.

Page 112: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

112

Isotopic Assignment

Based on feather δ2H values (δ2Hf), our assignment to origin followed the

procedure described in Hobson et al. (Hobson et al. 2014). Briefly, our δ2Hf isoscape

was calibrated from the growing-season precipitation (δ2Hp) surface created by Bowen

et al. (Bowen et al. 2005) using the ground foraging Neotropical migrant algorithm δ2Hf

= −17.57 +0.95*δ2Hp. The probability of occurrence of a feather with a measured δ2Hf

value at a given pixel was assessed using a likelihood-based assignment method that

was confined to the breeding range in North America of Gray-cheeked Thrush.

Following the derivation of probability of origin at each pixel for each individual, a 2:1

odds ratio was applied as a criteria for including each pixel or not when summing up

surfaces for all individuals in the population. The resulting surface represents the

number of individuals so assigned for each pixel within the delimited range. All

assignments were conducted in the R statistical computing environment (R Core Team

2016).

Results

From 2010 through 2012, we captured 62 Gray-cheeked Thrushes in both clay

and white sand forests (2010: N = 24; 2011: N = 22; 2012: N = 16), with 57 (92%)

captured in white sand forests and five birds (8%) in adjacent clay forests. Excluding all

recaptures, 43% (N = 27) were aged after-hatch year, 27% (N = 17) were hatch year

birds and 29% (N = 18) were unknown (Fig. 5-2). During the same time period, we

caught 22 Swainson’s Thrushes in both clay and white sand forests (2010: N=6; 2011:

N=7; 2012: N=6), with 20 (91%) captured in white sands and two in clay forests (9%;

Fig. 5-3). Controlling for effort, we found no difference in capture rate in white sand

forests and adjacent clay terra firme forests for Gray-cheeked Thrush (χ21 =1.3, P =

Page 113: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

113

0.24) or Swainson’s Thrush (χ21 =0.29, P = 0.58; both species combined: χ2

1 =1.6, P =

0.20). Over the three-year period, the earliest capture of Gray-cheeked Thrushes in

white sand forest was 13 October and the earliest capture in clay forest was 1

December; the earliest capture of Swainson’s Thrushes in white sands was 11 October

with the earliest capture in clay forests 28 October. The latest capture of Gray-cheeked

Thrushes in either habitat was 7 December at the end of our sampling period (8

December). The latest capture of Swainson’s Thrushes was 16 November (Fig. 5-3).

Capture data, including latitude and longitude for individual Gray-cheeked Thrushes, are

reported in Supplemental Table S1.

We recaptured 12 Gray-cheeked Thrushes (19.3%) on at least one subsequent

day from that in which they were banded including three recaptures in a subsequent

year (4.8%). Dates of recapture occurred from 19 October 2011 to 6 December 2012

(Fig. 5-4). Inter-annual recaptures occurred on 19 October and 21 November 2011, and

17 November 2012. We recaptured two individual Swainson’s Thrushes on subsequent

days (9%) and none in a subsequent year. Because net transects were relocated every

three days, all intra-annual recaptures of Gray-cheeked Thrushes (N = 9) occurred

within 2 days of initial capture from 20 October to 6 December. The two Swainson’s

Thrushes were recaptured on 19 - 20 October and 14 - 15 - 16 November (one

individual). All inter-annual recaptures of Gray-cheeked Thrushes occurred at the same

site (net transect) where the individual was first captured and within 5 calendar days in

subsequent years. One bird initially captured in 2010 was recaptured in 2012 on the

same calendar date and at the same sampling location. All recaptures for both thrush

species were in white sand forests. Stable hydrogen isotope values (δ2H) ranged from -

Page 114: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

114

174 to -129 ‰ (mean = -161 ‰). Based on our limited δ2H analyses, all birds tested

were from breeding locations in northwest Canada or Alaska rather than north-central or

eastern North America (Fig. 5-5). Gray-cheeked Thrush subcutaneous fat scores were

significantly lower than Swainson’s Thrush subcutaneous fat scores (Gray-cheeked

Thrush: mean=0.54, range = 0 – 4; Swainson’s Thrush: mean = 1.4, range = 0 – 5; P =

0.02).

Discussion

We confirmed inter-annual fidelity of Gray-cheeked Thrushes to sites in white

sand forests in the western Amazon basin. Although Veeries (C. fuscescens) and

Swainson’s Thrushes (C. ustulatus) have been tracked for full annual cycles (Heckscher

et al. 2011, 2015, Hobson and Kardynal 2015, Cormier et al. 2013), geolocator data do

not have the resolution necessary to determine if individuals returning to regions in a

subsequent year are faithful to a former site. Therefore, we believe our study is also the

first to confirm inter-annual site fidelity of migratory Catharus thrushes to a wintering site

in South America. However, non-breeding season site philopatry has been reported for

Bicknell’s Thrushes (C. bicknelli) in the Dominican Republic (Rimmer and McFarland

2001) and Hermit Thrushes (C. guttatus) in the southeast United States (Brown et al.

2000). Of 271 settled Hermit Thrushes banded at their Louisiana USA wintering site,

Brown et al. (2000) reported that 25 (18%) were known to have returned in a

subsequent year of which 23 (98%) were captured within 30 m of the same net in which

they were originally banded (Brown et al. 2000). Similarly, all three of our inter-annual

recaptures occurred at the same transect and within 90 m of where the bird was banded

in a prior year.

Page 115: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

115

Although some of the 62 Gray-cheeked Thrushes captured in our study (Table 5-

1) may represent transient birds, we assume many of our captures represent birds that

are seasonally settled because: (1) a 4.8% inter-annual recapture rate, all occurring at

the same transect and ≤90m from initial capture, would be extraordinary at a migratory

stopover site (we believe our inter-annual recapture rate would have been higher had

transects remained stationary for more than three days), (2) dates of first capture and

recaptures (13 October – 6 December) were consistent with the timing of arrival and

settlement of Veeries at their winter sites in South America, a Nearctic-Neotropical

migratory congener that also migrates to the Amazon basin from North America, (15

October – 8 December; (Heckscher et al. 2015)), (3) our study site was at or near the

southern edge of the species’ known South American range (Lowther et al. 2001), (4)

the pattern of capture of Gray-cheeked Thrushes is inconsistent with birds captured at a

migratory stopover site. The temporal pattern of transients captured on stopover

typically show a unimodal distribution similar to our Swainson’s Thrush captures (Fig. 5-

2) and shown by Gómez et al. (2014) for three Catharus species at a Colombia

stopover site (Gomez et al. 2014), (5) Gray-cheeked Thrush had significantly lower fat

scores than transient Swainson’s Thrush. Birds recently arrived or settled on winter

sites are expected to have largely exhausted their migratory fat stores while birds en

route at stopover should retain more fat, and (6) a 19% intra-annual recapture rate from

sites re-sampled no more than two days in a given year would be high in comparison to

other work at migratory stopover sites and suggests the presence of settled and

perhaps territorial birds. For example, using 10 – 15 12m nets over a three year period,

Gómez et al. 2014 reported recapture rates between 1.5% and 20% for three species of

Page 116: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

116

Catharus migrants at two different stopover regions in Colombia (Gomez et al. 2014).

However, their nets remained stationary from mid-September to mid-November and

many recaptures occurred >2 days after initial capture. In contrast, we assume

Swainson’s Thrush used our study area as a migratory stopover site considering: (1)

we have no records of inter-annual recaptures, (2) dates of first capture and recaptures

are not consistent with sedentary birds (11 October – 16 of November), (3) the unimodal

pattern of capture for Swainson’s Thrushes at our sites is typical for a transient

songbird, (4) intra-annual capture rates (9%) were lower than that for Gray-cheeked

Thrush and more typical for birds on stopover (Gomez et al. 2014).

Nearctic-Neotropical migratory Catharus thrushes can move substantial

distances through tropical regions throughout much of the non-breeding season,

particularly in the vast rainforests of South America (Heckscher et al. 2015). To date,

individual Gray-cheeked Thrushes have not been tracked to determine if they exhibit

mid-season movements. The possibility that Gray-cheeked Thrushes undertake mid-

season intra-tropical movements means they may not remain in white sand forests of

the western Amazon for the entire non-breeding season. Regardless, our findings show

that despite two annual trans-hemispheric migrations, Gray-cheeked Thrushes can

show inter-annual site fidelity to specific wintering sites as they probably do for breeding

sites in boreal forest. Site fidelity presumably has selective advantages if birds returning

to familiar sites can use that familiarity to their advantage, e.g., increased foraging

efficiency due to prior knowledge of productive areas, effective evasion of predators, or

intraspecific dominance in high-quality habitat (Wunderle and Latta 2000). Site fidelity

also has implications for conservation. Species in which individuals make annual returns

Page 117: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

117

to the same site may be more vulnerable to site degradation such as habitat loss

(Warkentin 1996).

We only documented site fidelity in white sand forests, but Gray-cheeked

Thrushes were also captured in adjacent alluvial clay terra firme forests. Controlling for

effort, there was no difference in capture rates between white sand and clay forests,

but, over the three-year period, three of our five captures in clay forest occurred on one

day suggesting possible synchronous movement. Outside this day (4 December 2011),

our capture rates were higher in white sand forests (χ2=4.8, df = 1, P = 0.02). This

difference suggests that Gray-cheeked Thrushes may favor white sand forests in this

region (e.g., late arriving individuals may settle in less desirable clay forests). Thus,

although we assume that Gray-cheeked Thrushes are not solely dependent on white

sand forests for non-breeding sites given their broad distribution in South America, we

conclude that they are associated with white sand forests in this region and may favor

them over adjacent clay terra firme forest. Sexual habitat segregation may be a

contributing factor to differences in habitat use. For example, Bicknell’s Thrushes may

segregate habitat whereby males are more common than females in sites with denser

understory vegetation (Townsend et al. 2011). Unfortunately, we were not able to sex

our birds but later arrivals in clay forests may differ in age or sex; for example, late

arrivals might be young females following a pattern seen in Veery (Heckscher, unpub.

data). In our study area white sand forests do differ structurally and floristically from clay

forest: they have a higher number of vertical stems (<10 cm DBH) in the understory (<3

m) and therefore maintain a denser shrub-layer as opposed to clay forests that have

more herbaceous vegetation and an open understory (see Table 3-1). Thus, Gray-

Page 118: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

118

cheeked Thrushes may favor the comparatively dense understory of white sand forest

on their wintering grounds similar to their affiliation with dense vegetation on their

breeding grounds (Lowther et al. 2001, Marshall 2001).

Based on our δ2H analyses, all Gray-cheeked Thrushes were from breeding

locations in northwest Canada or Alaska. These results provide the first data regarding

the connectivity between boreal and tropical forest populations of Gray-cheeked Thrush.

Although our sample size is small, it seems that the connectivity of Gray-cheeked

Thrush between breeding and wintering grounds may be strong, as it is for Swainson’s

Thrushes (Cormier et al. 2013), rather than weak as it is for Veery (Heckscher et al.

2011, Hobson and Kardynal 2015).

Considering that this species winters across a broad east-west expanse of

northern South America (Lowther et al. 2001), one might assume that birds that breed in

western North America might winter in western South American regions as that pattern

represents the shortest distance between breeding and non-breeding regions. However,

songbirds do not always take the simplest possible migratory routes (Ruegg 2002).

Further, it has become apparent that Veeries breeding in eastern and western North

America occur in sympatry in Amazonia (Heckscher et al. 2011, Hobson and Kardynal

2015). Further, birds at our study site may relocate substantial distances long before

they initiate Nearctic-Neotropical migration in April. For example, Veeries in the Amazon

river basin may move >2000 km between first and second winter sites (Heckscher et al.

2015). Annual tracking of individual Gray-cheeked Thrushes is needed to more fully

assess regional patterns of settlement and movement and the connectivity between

breeding and wintering regions.

Page 119: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

119

Our results suggest that Gray-cheeked Thrushes exhibit site fidelity and may

concentrate in white sand forests – an uncommon and scattered ecosystem type in

western Amazonia. We have also established the connectivity of Gray-cheeked

Thrushes at a specific non-migratory site in South America with breeding regions in

northwestern North America. Further study is needed to better establish migratory

connectivity, temporal patterns of movement, the dependency of this species on white

sand forests, and the use of weathered clay terra firme and lowland forests in the

western Amazon basin. White sand forests are rare in Amazonia and the affiliation of

Gray-cheeked Thrush with this ecosystem type provides new and valuable information

about this ecosystem and the distribution and ecology of Gray-cheeked Thrushes in

South America.

Page 120: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

120

Table 5-1. Captured and recaptured Gray-cheeked Thrushes (Catharus minimus) in Amazonian white sand forests (wsf) and clay terra firme forests (clay) in Allpahuayo-Mishana National Reserve, Department of Loreto, Peru (2010 – 2012).

Band

Number Date Habitat Net Line and Year Longitude Latitude

Elevation

(m)

E00232 13-Oct-10 wsf Var Fun Piu_2010 -73.4443 -3.9879 154.7

E00241 20-Oct-10 wsf San Mar 1_2010 -73.4410 -3.8868 128.9

E00244 20-Oct-10 wsf San Mar 1_2010 -73.4411 -3.8881 121.3

E00262 23-Oct-10 wsf San Mar 2_2010 -73.4434 -3.8905 117.6

E00261 23-Oct-10 wsf San Mar 2_2010 -73.4433 -3.8905 110.9

E00261* 23-Oct-10 wsf San Mar 2_2010 -73.4438 -3.8903 117.2

E00266 24-Oct-10 wsf San Mar 2_2010 -73.4433 -3.8905 110.9

E00266* 24-Oct-10 wsf San Mar 2_2010 -73.4424 -3.8908 120.5

E00293 4-Nov-10 wsf Var Chico_2010 -73.4755 -3.8955 114.8

E00302 5-Nov-10 wsf Var Chico_2010 -73.4755 -3.8955 114.8

E00324 7-Nov-10 wsf Var Madre_2010 -73.4775 -3.9131 129.9

E00313 7-Nov-10 wsf Var Madre_2010 -73.4776 -3.9132 129.7

E00317 7-Nov-10 wsf Var Madre_2010 -73.4777 -3.9134 129.7

E00324* 8-Nov-10 wsf Var Madre_2010 -73.4774 -3.9128 131.8

E00327 8-Nov-10 wsf Var Madre_2010 -73.4776 -3.9133 130.1

E00313* 8-Nov-10 wsf Var Madre_2010 -73.4777 -3.9134 129.7

E00324* 9-Nov-10 wsf Var Madre_2010 -73.4774 -3.9127 127.7

E00347 17-Nov-10 wsf Var Yuto 1_2010 -73.5342 -3.9113 166.0

E00360 18-Nov-10 wsf Var Yuto 1_2010 -73.5350 -3.9122 136.9

E00365 19-Nov-10 wsf Var Yuto 1_2010 -73.5342 -3.9113 166.0

* re-captured individuals, ** Inter-annual recapture.

Page 121: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

121

Table 5-1. Continued Band

Number Date Habitat Net Line and Year Longitude Latitude

Elevation

(m)

E00374* 20-Nov-10 wsf Var Yuto 2_2010 -73.5297 -3.9105 142.2

E00366 20-Nov-10 wsf Var Yuto 2_2010 -73.5315 -3.9115 144.3

E00374* 20-Nov-10 wsf Var Yuto 2_2010 -73.5310 -3.9109 117.2

E00370* 20-Nov-10 wsf Var Yuto 2_2010 -73.5308 -3.9108 139.3

E00370* 21-Nov-10 wsf Var Yuto 2_2010 -73.5299 -3.9106 127.5

E00406 25-Nov-10 wsf Var Yuto 3_2010 -73.5364 -3.9180 79.4

E00420 26-Nov-10 wsf Var Yuto 3_2010 -73.5366 -3.9163 133.0

E00437 3-Dec-10 wsf Var Exp Porv_2010 -73.5534 -3.9192 115.2

E00447 5-Dec-10 wsf Var Blanco_2010 -73.5586 -3.9159 116.2

E00453 6-Dec-10 wsf Var Blanco_2010 -73.5591 -3.9154 122.2

E00447* 6-Dec-10 wsf Var Blanco_2010 -73.5590 -3.9156 127.0

E00455 7-Dec-10 wsf Var Blanco_2010 -73.5583 -3.9160 127.5

E00838 19-Oct-11 wsf San Mar 1_2011 -73.4411 -3.8873 121.3

E00244** 19-Oct-11 wsf San Mar 1_2011 -73.4411 -3.8871 119.8

E00915 19-Oct-11 wsf San Mar 2_2011 -73.4424 -3.8908 120.5

E00838* 20-Oct-11 wsf San Mar 1_2011 -73.4411 -3.8870 121.5

E00920 20-Oct-11 wsf San Mar 2_2011 -73.4426 -3.8908 122.5

E00922 20-Oct-11 wsf San Mar 2_2011 -73.4439 -3.8903 112.4

E00863 25-Oct-11 wsf Var Bol 1_2011 -73.4566 -3.8908 116.2

E00960 2-Nov-11 wsf Var Bol 2_2011 -73.4539 -3.8944 110.9

E00720 20-Nov-11 wsf Var Yuto 3_2011 -73.5366 -3.9175 131.1

E00716 20-Nov-11 wsf Var Yuto 3_2011 -73.5363 -3.9180 131.6

* re-captured individuals, ** Inter-annual recapture.

Page 122: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

122

Table 5-1. Continued Band

Number Date Habitat Net Line and Year Longitude Latitude

Elevation

(m)

E00741 23-Nov-11 wsf Var Yuto 1_2011 -73.5348 -3.9119 153.2

E00757 27-Nov-11 wsf Var Yuto 2_2011 -73.5310 -3.9110 137.4

E00761 30-Nov-11 wsf Var Exp Porv_2011 -73.5530 -3.9189 135.9

E00773 30-Nov-11 wsf Var Exp Porv_2011 -73.5526 -3.9183 132.1

E00777 1-Dec-11 wsf Var Exp Porv_2011 -73.5524 -3.9182 132.5

E00798 4-Dec-11 wsf Var Blanco_2011 -73.5587 -3.9159 131.3

(no band) 4-Dec-11 wsf Var Blanco_2011 -73.5582 -3.9165 128.5

E00792 4-Dec-11 wsf Var Blanco_2011 -73.5581 -3.9166 129.2

E00989 4-Dec-11 clay Porv TF_2011 -73.5524 -3.9103 113.8

E00793 4-Dec-11 wsf Var Blanco_2011 -73.5591 -3.9154 131.3

E00975 4-Dec-11 clay Porv TF_2011 -73.5522 -3.9115 114.5

E00972 4-Dec-11 clay Porv TF_2011 -73.5522 -3.9114 115.5

E00800 5-Dec-11 wsf Var Blanco_2011 -73.5591 -3.9154 131.3

E00800* 6-Dec-11 wsf Var Blanco_2011 -73.5586 -3.9159 131.3

E1009 6-Dec-11 wsf Var Blanco_2011 -73.5591 -3.9154 131.3

E1238 25-Oct-12 wsf Var Madre_2012 -73.4780 -3.9136 130.4

E1238* 25-Oct-12 wsf Var Madre_2012 -73.4768 -3.9121 121.0

E1242 25-Oct-12 wsf Var Madre_2012 -73.4769 -3.9123 90.5

E1237 25-Oct-12 wsf Var Madre_2012 -73.4771 -3.9125 91.9

J00812 16-Nov-12 wsf Var Yuto 1_2012 -73.5346 -3.9117 139.3

C0497 16-Nov-12 wsf Var Yuto 1_2012 -73.5342 -3.9112 138.1

J00811 16-Nov-12 wsf Var Yuto 1_2012 -73.5340 -3.9112 138.6

* re-captured individuals, ** Inter-annual recapture.

Page 123: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

123

Table 5-1. Continued Band

Number Date Habitat Net Line and Year Longitude Latitude

Elevation

(m)

E1534 17-Nov-12 wsf Var Yuto 2_2012 -73.5304 -3.9104 135.2

J00819 19-Nov-12 wsf Var Yuto 3_2012 -73.5365 -3.9161 135.7

J00816 19-Nov-12 wsf Var Yuto 3_2012 -73.5366 -3.9174 129.9

J00817 19-Nov-12 wsf Var Yuto 3_2012 -73.5366 -3.9166 132.5

J00818 19-Nov-12 wsf Var Yuto 3_2012 -73.5366 -3.9168 132.8

J00817* 19-Nov-12 wsf Var Yuto 3_2012 -73.5366 -3.9169 132.8

J00816* 20-Nov-12 wsf Var Yuto 3_2012 -73.5366 -3.9164 136.4

E1545 20-Nov-12 wsf Var Yuto 3_2012 -73.5366 -3.9168 132.8

E1567 30-Nov-12 wsf Var Exp Porv_2012 -73.5526 -3.9183 122.2

E1576 1-Dec-12 clay Porv TF_2012 -73.5522 -3.9112 115.0

J00838 4-Dec-12 wsf Var Blanco_2012 -73.5594 -3.9150 126.1

J00838* 6-Dec-12 wsf Var Blanco_2012 -73.5587 -3.9158 129.2

* re-captured individuals, ** Inter-annual recapture.

Page 124: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

124

Figure 5-1. Location of Allpahuayo-Mishana National Reserve (outline), Department of

Loreto, southeast of Iquitos, Peru. White sand forests are the small polygons outlined in black. Locations of mistnet transects in white sand are depicted as white triangles and those in clay terra firme forests as black circles. Inset in lower left corner is the outline of Peru with the Department of Loreto in black with the approximate location of the study sites depicted by a star (image source: Google Earth).

Page 125: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

125

Figure 5-2. Weekly numbers of Gray-cheeked Thrushes by age caught using constant

effort mistnet transects at Allpahuayo-Mishana National Reserve, Department of Loreto, Peru, 13 October 2010 to 8 December 2012.

0

1

2

3

4

5

6

7

8

9HY

AHY

Unknown

Nu

mb

er

of

Ind

ivid

ual

s

Week

Nu

mb

er

of

Ind

ivid

ual

s

Week

Page 126: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

126

Figure 5-3. Cumulative number of Gray-cheeked Thrushes (Catharus minimus) and Swainson’s Thrushes (C. ustulatus) captured over a nine week period in 2010, 2011, and 2012 at Allpahuayo-Mishana National Reserve, Department of Loreto, Peru. Numbers over columns indicate the number of individuals captured in weathered clay terra firme forests, all others were captured in white sand terra firme forests.

0

2

4

6

8

10

12

14

16

18

Gray-cheeked Thrush

Swainson's Thrush

1

Nu

mb

er

of

ind

ivid

ual

s

Week

Page 127: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

127

Figure 5-4. Cumulative recaptures by week for Gray-cheeked Thrushes using constant effort mistnet transects at Allpahuayo-Mishana National Reserve, Department of Loreto, Peru. Transects were relocated every three days, such that each transect was sampled for only three days each year. Numbers above columns indicate inter-annual recaptures for the respective weekly totals.

0

1

2

3

4

5

6

7N

um

be

r o

f in

div

idu

als

Week

1

1

Page 128: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

128

Figure 5-5. Depictions of likely breeding or natal origins of Gray-cheeked Thrushes at Allpahuayo-Mishana National Reserve, Department of Loreto, Peru. The spatially explicit assignment of the population (n=12) follows from the likelihood assignment approach described in Hobson et al. (2014). Numbers in the legend refer to number of individuals in the sample whose probability of assignment, based on a 2:1 odds ratio, at each pixel was consistent with the expected feather value for that location. Possible breeding or natal origin was delimited by the known breeding range of Gray-cheeked Thrush in North America.

Page 129: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

129

CHAPTER 6 CONCLUDING REMARKS

Tropical rainforests harbor the most species-rich communities on Earth. Many

groups of organisms reach the highest levels of species richness and largest biomass in

Neotropical rainforests, and ecological research is often focused on elucidating the

underlying mechanisms and the consequences of diversity (Gaston 2000, Storch et al.

2006). Beta diversity, or turnover between localities, in our analyses between habitats,

contributes to the high levels of species richness in the tropics (Condit et al. 2002,

Jankowski et al. 2009, Kraft et al. 2011), and the high turnover in taxonomic and

phylogenetic betadiversity is governed by various mechanisms. In this dissertation we

combined a phylogenetic approach with the observed ecological patterns on a large

geographic scale, and then focused on field data collection on a smaller spatial scale to

observe community patterns in contrasting habitat types. Floristic studies have been

more commonly used to examine this ecosystem (Fine et al. 2010, ter Steege et al.

2013, Coronado et al. 2015), but the first-ever Amazon-basin wide description of

betadiversity and phylobetadiversity study, focusing on white sand systems’ plants was

published just this year (Guevara et al. 2016). The study found high taxonomic

dissimilarity but low phylogenetic dissimilarity in pairwise community comparison.

Many questions remain regarding the diversity that white sand forests harbor, the

processes that govern community composition, structure and the basic biogeographical

patterns of this ecosystem. Amazonian white sand systems function as ecological

islands in the sea of the tropical moist forest biome; endemic plants are abundant and

various habitat specialists show disjoint ranges that are specific to this particular habitat

type (Fine 2016, Fine and Baraloto 2016). Plant community assemblages strongly

Page 130: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

130

correlate with differences in soil (Poulsen et al. 2006). In Western Amazonia, on the

gradient of soil types from nutrient poor sandy soils to relatively fertile clay soils, most

plants, especially terrestrial pteridophytes, some palms, and some species of ferns, are

restricted to specific soil conditions, and only a few can grow on all soil types (Tuomisto

and Poulsen 1996, Svenning 2001, Fine et al. 2005, Svenning et al. 2006, Fine et al.

2010).

Despite differences in plant communities in the different forest types, we did not

find strong evidence of fundamentally different avian community composition due to

habitat filtering, nor did we find population structure in an obligate habitat specialist

species. There is much to explore in the system that still requires further research.

Habitat specialization does not seem to be an isolating mechanism in most groups of

taxa. For example, within plants, only ~23% of WSF forest plants from Western

Amazonia are actually habitat specialists (Garcia-Villacorta et al. 2016). This study did

not explore the contribution of interspecific contest competition of habitat segregation

and how that plays a role in beta diversity, even though we initially predicted that the

proliferation of birds on different habitats would be governed by competition between

closely related species.

Natural communities vary on a scale from local dynamics to regional species

distribution patterns. Studies of white sands have explored the vegetation community,

but no large-scale ecological studies have been conducted on their bird communities.

There are no concise analyses of the resource and habitat use of white sand forest

specialist birds, and no data on their potential dispersal ability. There have been no

population genetics studies on isolated populations of the specialist species to

Page 131: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

131

determine whether gene flow exists between the patches, but there is a study in

progress (using samples collected from our fieldwork) at the Field Museum focusing on

the phylogeography of poor soil specialist bird species all across their distributional

ranges (Capurucho, Joao Marcos, pers. comm.).

If we compare species lists from studies in Loreto (North bank of the Amazon in

Peru) with studies from Jau National Park (North bank of the Amazon in Brazil) and in

the upper Jurua River region in Acre (South bank of the Amazon in Brazil), there is

surprisingly high overlap between the species inhabiting these stunted forests (Alonso

and Whitney 2003, Borges 2004, Guilherme and Borges 2011), indicating that perhaps

at some point they may have been more widespread or connected by dispersal

corridors. An integrative, well-designed approach emphasizing the number and spatial

distribution of samples combined possibly with long-term manipulative experiments

would allow us to evaluate macroevolutionary and biogeographic processes (Urban et

al. 2008), and to make predictions of changes and consequences relevant for

conservation biology.

Long-term studies also allow for the study of cryptic or poorly known species.

During the constant effort mist netting and mark-recapture study, we sampled a WSF

habitat specialist in high enough numbers to be able to conduct a population genetics

study. We collected blood samples, some of which are serving as the first tissue

samples of species from this region, and are already being used by collaborators for

future studies on the biogeography of Amazonian habitat specialists. The blood samples

are also being used in an avian malaria study (Pulgarin-R, P. pers. comm). In addition,

we captured 62 Catharus minimus and 17 Catharus ustulatus individuals throughout our

Page 132: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

132

fieldwork, and apparently, this is the greatest number of C. minimus reported from any

single South American site (Ungvari-Martin et al. 2016) and may exceed the number of

published accounts from all of South America. White sand forests are proving to

become an important site migratory bird species as well as residents.

Long periods spent in the field can yield data for various future publications. Our

dataset, not included in this dissertation, will allow us to compare age structure, survival

estimates, morphological differences, and even further detailed vegetation dynamics in

the future. In addition, we also collected a large number of ectoparasites and fecal

samples that can serve for co-evolutionary studies on host-parasite interactions using

lice and their avian hosts, and the fecal samples can be used for diet studies. Overall,

the four content chapters of this dissertation barely start the exploration of all the

information collected from AMNR, and we hope to contribute with these data for further

publications in years to come.

There is no one correct way to do ecology. Mathematical models, model ecosystems, field manipulation experiments and the search for large-scale patterns are all valid approaches, and all have their strengths and weaknesses. They are simply tools to help us understand nature; like all tools, each approach does some things well, some things, badly, and other things not at all (Lawton 1996).

Page 133: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

133

APPENDIX A LIST OF SPECIES CAPTURED IN AMNR FORESTS

Table A-1. List of species captured in forest in Allpahuayo-Mishana during 2009-2012 with raw capture numbers for each species

Species terra firme

humid varillal dry varillal chamizal

Glyphorynchus spirurus 239 515 725 87

Dixiphia pipra 104 228 290 42

Willisornis poecilonotus 105 158 260 26

Pithys albifrons 90 111 230 25

Gymnopithys leucaspis 81 137 216 18

Dendrocincla merula 19 65 157 15

Mionectes oleagineus 106 155 118 19

Neopelma chrysocephalum 0 63 116 12

Xiphorhynchus elegans 28 38 107 12

Myrmotherula axillaris 40 75 89 14

Myrmotherula hauxwelli 62 46 88 2

Phaetornis bourcieri 43 44 82 3

Megastictus margaritatus 31 54 81 8

Phaetornis superciliosus 62 33 66 9

Automolus ochrolaemus 7 50 51 12

Schiffornis turdina 4 21 49 8

Dendrocincla fuliginosa 11 30 48 7

Myrmoborus myotherinus 86 14 46 0

Catharus minimus 4 33 42 0

Lepidothrix coronata 118 28 41 2

Pipra erythrocephala 45 12 38 3

Thalurania furcata 8 21 35 0

Corythopis torquatus 4 36 33 1

Xenops minutus 25 15 32 4

Malacoptila fusca 7 2 31 0

Sclerurus rufigularis 7 13 31 0

Formicarius colma 3 14 30 6

Percnostola arenarum 1 17 25 5

Geotrygon montana 13 10 22 1

Threnetes leucurus 9 4 22 1

Epinecrophylla haematonota 28 37 21 1

Hylophylax naevius 10 19 21 0

Microcerculus marginatus 24 19 21 0

Myrmeciza castanea 3 10 20 5

Page 134: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

134

Table A-1. Continued

Species terra firme

humid varillal dry varillal chamizal

Thamnomanes caesius 35 18 19 0

Catharus ustulatus 2 9 18 0

Terenotriccus erythrurus 29 11 17 0

Cnemotriccus spp. 0 4 14 8

Sclerurus mexicanus 12 14 14 1

Heliodoxa aurescens 4 2 13 0

Phlegopsis erythroptera 20 5 12 0

Machaeropterus regulus 24 5 11 0

Synallaxis rutilans 23 23 11 0

Attila spadiceus 6 6 10 1

Dendrocolaptes certhia 2 2 10 0

Thamnophilus murinus 8 7 10 1

Chloroceryle aenea 5 7 9 0

Platyrinchus saturatus 4 0 9 0

Turdus albicollis 4 7 9 0

Deconychura stictolaema 3 2 8 0

Ramphotrigon ruficauda 3 3 8 0

Myrmeciza fortis 26 4 7 0

Neopipo cinnamomea 1 2 7 0

Nonnula brunnea 27 6 7 5

Heterocercus aurantiivertex 0 1 6 4

Manacus manacus 31 15 6 3

Rhegmatorhina melanosticta 7 7 6 0

Florisuga mellivora 2 3 5 1

Frederickena unduligera 4 3 5 0

Glaucis hirsutus 1 2 4 1

Hylocharis cyanea 0 4 4 0

Onychorhynchus coronatus 2 3 4 1

Attila citriniventris 0 3 3 0

Rupornis magnirostris 1 0 3 2

Celeus elegans 2 2 3 4

Laniocera hypopyrra 1 1 3 0

Rhytipterna simplex 0 2 3 0

Veniliornis affinis 1 0 3 1

Capito auratus 3 1 2 1

Caprimulgus nigrescens 0 0 2 0

Cercomacra serva 3 0 2 0

Page 135: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

135

Table A-1. Continued

Species terra firme

humid varillal dry varillal chamizal

Chrysuronia oenone 0 0 2 0

Conopophaga peruviana 11 3 2 0

Cyanocompsa cyanoides 9 0 2 0

Galbula albirostris 9 1 2 0

Gymnopithys lunulatus 1 0 2 0

Hyloctistes subulatus 19 4 2 0

Hypocnemis hypoxantha 6 12 2 3

Lophotriccus vitiosus 0 0 2 0

Momotus momota 2 3 2 1

Sporophila angolensis 1 1 2 0

Pachyramphus marginatus 0 0 2 0

Phaeothlypis fulvicauda 0 1 2 1

Sclateria naevia 1 1 2 0

Tangara chilensis 1 1 2 0

Accipiter superciliosus 0 0 1 0

Ancistrops strigilatus 0 0 1 0

Celeus grammicus 0 0 1 2

Chlorostilbon mellisugus 0 2 1 0

Claravis pretiosa 0 1 1 1

Cyanerpes caeruleus 0 0 1 0

Deconycura longicauda 0 1 1 0

Hylocharis sapphirina 0 0 1 0

Micromonacha lanceolata 0 0 1 0

Microrhopias quixensis 0 0 1 0

Myrmotherula menestriesii 2 1 1 0

Neoctantes niger 0 1 1 0

Notarchus ordii 0 0 1 1

Oporornis agilis 0 0 1 0

Percnostola leucostigma 17 10 1 3

Philydor erythrocercum 0 2 1 0

Phoenicircus nigricollis 0 0 1 1

Thamnophilus doliatus 0 0 1 0

Thryothorus coraya 7 1 1 0

Trogon rufus 0 0 1 0

Automolus infuscatus 1 0 0 0

Bucco tamatia 1 0 0 0

Page 136: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

136

Table A-1. Continued

Species terra firme

humid varillal dry varillal chamizal

Cacicus cela 0 0 0 2

Cnipodectes subbrunneus 13 4 0 0

Crypturellus soui 1 0 0 0

Cyanerpes cyaneus 0 1 0 0

Electron platyrynchus 1 0 0 0

Euphonia rufiventris 1 1 0 0

Euphonia xanthogaster 5 0 0 0

Habia rubica 2 3 0 0

Hylophilus ochraceiceps 18 4 0 0

Hypocnemis peruviana 8 0 0 0

Leucopternis melanops 1 0 0 0

Lipaugus vociferans 2 1 0 0

Micrastur gilvicollis 2 2 0 0

Micrastur mirandollei 1 1 0 0

Monasa morphoeus 3 0 0 0

Myiobius barbatus 5 0 0 0

Myrmotherula longipennis 4 1 0 0

Pachyramphus polychopterus 1 0 0 1

Poecilotriccus latirostris 0 1 0 0

Pteroglossus azara 1 0 0 0

Pygiptila stellaris 1 0 0 0

Ramphocaenus melanurus 1 0 0 0

Rhynchocyclus olivaceus 1 0 0 0

Saltator grossus 2 0 0 0

Sclerurus caudatus 0 3 0 0

Selenidera reinwardtii 1 2 0 0

Tachyphonus cristatus 2 3 0 0

Tachyphonus surinamus 3 8 0 0

Thamnomanes ardesiacus 9 7 0 0

Thamnophilus schistaceus 6 0 0 0

Tityra semifasciata 0 0 0 1

Turdus lawrencii 0 0 0 1

Xiphorhynchus guttatus 1 0 0 0

Page 137: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

137

A)

B)

Figure A-1. Comparisons of WSF and clay and the four specific habitat types by capture

rate. Neither A) Capture rates between clay and WSF nor B) captures rates among the four specific habitat types in AMNR differed.

Page 138: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

138

APPENDIX B SUMMARY OF RICHNESS ESTIMATORS

Table B-1. Incidence-based richness estimators with standard deviations of the four different habitat types in AMNR.

Habitat Species chao chao.se jack1 jack1.se boot boot.se n

clay 106 145.77 18.26 137.82 14.20 120.21 8.37 11

humid varillal 99 121.49 11.04 125.36 11.22 111.46 6.35 11

dry varillal 106 134.38 14.33 132.35 8.25 117.95 4.65 17

chamizal 53 70.36 9.14 71.00 13.11 61.52 6.79 3

Figure B-1. Graphical representation of incidence based richness estimators with

confidence intervals for all sampling localities and habitats combined.

Page 139: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

139

APPENDIX C BETWEEN-YEAR RETURN RATES IN WSF AND CLAY FOREST

Table C-1. Comparison of return rates for all species with a minimum of 5 captures. All values represent mean number of individual birds in the habitat types ± standard deviations. Habitat specialists are indicated in the last column.

species N (WSF)

Percent recaptures (WSF) N (clay)

Percent recaptures (clay) SPECIALIST

Dixiphia pipra 15.95 ± 8.88 16.14 ± 13.81 14 ± 11.17 6.92 ± 9.44 WSF Neopelma chrysocephalum 8.23 ± 2.8 28.61 ± 21.49 NA NA WSF Megastictus margaritatus 7 ± 2.69 8.63 ± 11.34 7.5 ± 2.12 16.67 ± 23.57 WSF Percnostola arenarum 6 ± 0 0 ± 0 NA NA WSF Cnemotriccus spp. 5 ± 0 0 ± 0 NA NA WSF Glyphorynchus spirurus 30.05 ± 12.84 27.39 ± 12.92 19 ± 10.86 20.55 ± 11.19 NO Mionectes oleagineus 12 ± 5.13 11.44 ± 11.82 13.75 ± 2.63 5 ± 10 NO Pipra erythrocephala 12 ± 0 0 ± 0 5 ± 0 0 NO

Pithys albifrons 9.38 ± 3.1 17.02 ± 14.92 8.5 ± 0.71 28.47 ± 22.59 NO

Willisornis poecilonotus 9 ± 3.69 28.73 ± 18.82 7.67 ± 3.79 6.67 ± 11.55 NO Myrmoborus myotherinus 9 ± 5.66 17.69 ± 3.26 7.75 ± 2.87 16.07 ± 13.66 NO Thamnomanes caesius 9 ± 0 0 ± 0 5.5 ± 0.71 8.33 ± 11.79 NO Dendrocincla merula 8.55 ± 4.2 24.73 ± 21.99 NA NA NO Gymnopithys leucaspis 8.14 ± 2.91 19.43 ± 27.63 7.33 ± 4.04 26.67 ± 30.55 NO Myrmotherula axillaris 7.55 ± 1.86 9.49 ± 12.01 6.5 ± 2.12 0 NO Xiphorhynchus elegans 6.89 ± 2.71 26.17 ± 17.41 5 ± 0 0 NO Catharus minimus 6.86 ± 1.68 3.63 ± 6.26 NA NA NO Myrmotherula hauxwelli 6.17 ± 1.94 13.89 ± 13.07 8 ± 1 8.33 ± 14.43 NO Automolus ochrolaemus 6 ± 0.63 18.25 ± 22.87 NA NA NO Microcerculus marginatus 6 ± 1.41 10 ± 14.14 NA NA NO

Schiffornis turdina 5.75 ± 1.5 8.13 ± 9.87 NA NA NO

Page 140: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

140

Table C-1. Continued

species N (WSF)

Percent recaptures (WSF) N (clay)

Percent recaptures (clay) SPECIALIST

Geotrygon montana 5.5 ± 0.71 0 ± 0 NA NA NO Catharus ustulatus 5 ± 0 0 ± 0 NA NA NO Corythopis torquatus 5 ± 0 20 ± 0 NA NA NO Dendrocincla fuliginosa 5 ± 0 10 ± 14.14 NA NA NO Epinecrophylla haematonota 5 ± 0 0 ± 0 5 ± 0 0 NO Lepidothrix coronata 5 ± 0 12 ± 17.89 10.5 ± 3 9.08 ± 7.2 NO

Malacoptila fusca 5 ± 0 0 ± 0 NA NA NO

Turdus albicollis 5 ± 0 20 ± 0 NA NA NO

Xenops minutus 5 ± 0 20 ± 0 5 ± 0 40 ± NA NO

Synallaxis rutilans NA NA 5 ± 0 20 ± NA clay Machaeropterus regulus NA NA 5 ± 0 0 clay Terenotriccus erythrurus NA NA 5 ± 0 0 clay

Page 141: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

141

APPENDIX D SUPPLEMENTARY INFORMATION OF DDRADSEQ ANALYSIS

We used 6 uL volume at 20 ng/uL concentration DNA from each sample, which

were digested using 0.9 uL Cutsmart 10x buffer, 0.28 uL EcoRI enzyme, 0.12 uL MseI

enzyme, 1.7 ul water for 8 hours at 37C in a thermocycler. A ligation step was

performed overnight adding 1 uL EcoR1 [0.5uM], 0.4 ul Cutsmart 10x buffer, 1.3 uL ATP

10um, 0.2 ul T4 DNA ligase enzyme, 0.1 ul water, and 1 ul MseI [10uM]. The 14 uL

samples were sealed and incubated at 16C for 6hours in a thermocycler. Library

construction and successful ligation were verified using a PCR reaction using 8 ul NEB

One-Taq 2x Master Mix, 0.8 ul 10 um premixed forward and reverse Illumina primers,

6.2 ul water, and 1 ul of the restriction/ligation product. This PCR was run following the

protocol: 94C for 2min then 20 cycles of of 94C for 30sec, 60C for 30sec, 68C for

45sec, and holding at 10 C. Amplification was verified on a 1.5% agarose gel using

electrophoresis, resulting in smears of visible DNA fragments in the 50-700 bp range.

We combined 7 ul of restriction/ligation product from every sample in a single

Eppendorf tube, and used the ICBR core genomics facilities to perform concentration of

the product, size selection using Pippin and Bioanalyzer evaluation (Submission ID

134735). In order to test the fractions that we wanted to use, we tested samples at

8,10,12,14 PCR cycles using 3 ul Q5 buffer 5x, 0.3ul DNTP, 1 ul of forward and reverse

primers combined at 5 um concentration, 0.15 ul Q5 Taq, 10.5 ul water, and 1 ul of size

selected sample. We ran the PCR with the following conditions on a thermocycler with a

heated lid: 98C for 30 sec, 8-12 cycles of (98C for 15sec, 60C for 30sec, 72C for

30sec), hold at 10C. While generally lower number of cycles are sufficient, we tested

the size selected samples at 12 cycles. We used fractions 5,6,7,8,9,10 for testing and

Page 142: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

142

chose fraction 9+8+7+6 for a total number of 60 tubes (15 each for each fraction for the

4 fractions). Finally, we pooled the PCR products after gel electrophoresis and

requested sequencing.

Table D-1. Run metrics summary from the High throughput NextSeq 500 run of single end 150 bp reads.

Cycles Yield (Gbp)

Aligned (%)

Error rate (%) %≥q30

Total sequences

Sequences ≥q33

151 58.51 21.26 0.54 91.49 378,893,170 316,455,808

Page 143: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

143

Table D-2. Raw sequencing information for samples after quality filtering steps from the catalog of RAD loci.

Id Region Unique Stacks

Polymorphic Loci SNPs Found

NECH_V1278 km 25 102201 5138 6543

NECH_V222 km 25 197872 26141 33281

NECH_V238 km 25 202132 28242 35871

NECH_V284 km 25 204567 28197 36231

NECH_12E2 km 28 188577 22119 28161

NECH_12E7 km 28 323328 29917 38232

NECH_3104 km 28 208297 32972 42257

NECH_3122 km 28 170905 18095 22799

NECH_3185 km 28 131250 9849 12308

NECH_3297 km 28 319495 29838 38128

NECH_7395 km 28 208023 30449 38994

NECH_V1315 km 28 201981 26835 34106

NECH_3491 km 31 19173 411 510

NECH_3492 km 31 147042 12336 15582

NECH_7356 km 31 190046 21359 27245

NECH_FP30 km 31 210112 30351 39151

NECH_1M89 Mishana 198712 28185 35959

NECH_2ME29 Mishana 141227 10050 12818

NECH_2ME6 Mishana 50408 1762 2229

NECH_2ME63 Mishana 190878 24558 31429

NECH_3845 Mishana 132826 9074 11467

NECH_3863 Mishana 141378 10971 13905

NECH_3908 Mishana 98588 4858 6136

NECH_3915 Mishana 184009 21889 27718

NECH_CH112 Mishana 185199 20909 26878

NECH_CH60 Mishana 202458 27534 35231

NECH_JM27 Mishana 175384 17742 22686

NECH_M188 Mishana 208023 31554 40292

NECH_M523 Mishana 203909 28073 35942

NECH_M530 Mishana 192394 24024 30409

NECH_ME029 Mishana 201557 28474 36321

NECH_ME030 Mishana 184392 22353 28391

NECH_ME062 Mishana 205801 30736 38978

NECH_ME063 Mishana 148887 12457 15818

NECH_ME079 Mishana 192682 24079 30801

NECH_TN01 Mishana 204922 27564 35154

NECH_TN02 Mishana 163834 15887 19986

Page 144: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

144

Table D-2. Continued

Id Region Unique Stacks

Polymorphic Loci SNPs Found

NECH_TN03 Mishana 204990 29644 37949

NECH_TN04 Mishana 187119 19668 25554

NECH_TN05 Mishana 160181 15263 19329

NECH_2NEN1 Nueva Esperanza 187165 28245 35158

NECH_2NEN2 Nueva Esperanza 209569 36127 45688

NECH_3301 Nueva Esperanza 130099 8104 10351

NECH_3317 Nueva Esperanza 75170 3233 4108

NECH_3341 Nueva Esperanza 171456 17620 22127

NECH_3372 Nueva Esperanza 153887 13945 17518

NECH_3399 Nueva Esperanza 159466 14891 18472

NECH_3418 Nueva Esperanza 184639 22403 28663

NECH_3431 Nueva Esperanza 134339 9727 12357

NECH_3434 Nueva Esperanza 196369 25642 32900

NECH_5047 Nueva Esperanza 204438 29085 36958

NECH_5078 Nueva Esperanza 151918 14916 18811

NECH_5097 Nueva Esperanza 182605 20584 26204

NECH_5121 Nueva Esperanza 188289 22880 29227

NECH_5213 Nueva Esperanza 194621 24513 31231

NECH_5297 Nueva Esperanza 157192 13673 17313

NECH_5306 Nueva Esperanza 34777 1069 1340

NECH_CU13 Nueva Esperanza 157618 13691 17383

NECH_CU17 Nueva Esperanza 177242 20764 26219

Page 145: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

145

Table D-2. Continued

Id Region Unique Stacks

Polymorphic Loci SNPs Found

NECH_CU2 Nueva Esperanza 177434 25156 32268

NECH_CU22 Nueva Esperanza 176478 19717 25007

NECH_CU48 Nueva Esperanza 198058 26569 33892

NECH_CU6 Nueva Esperanza 188522 23633 29915

NECH_NEU38 Nueva Esperanza 67507 2750 3439

NECH_4346 Porvenir 200620 26537 33821

NECH_4354 Porvenir 151186 11971 15253

NECH_4357 Porvenir 46133 1750 2151

NECH_4361 Porvenir 147811 12196 15527

NECH_4432 Porvenir 186452 22108 27866

NECH_4450 Porvenir 99218 5607 7058

NECH_4535 Porvenir 168390 16194 20492

NECH_6783 Porvenir 188707 18098 23341

NECH_6851 Porvenir 204198 28194 36215

NECH_6863 Porvenir 170691 16696 21247

NECH_6865 Porvenir 210763 31468 40371

NECH_6877 Porvenir 178144 19423 24590

NECH_6884 Porvenir 204545 29427 37591

NECH_6898 Porvenir 201463 27566 35196

NECH_6945 Porvenir 179553 19422 24557

NECH_6989 Porvenir 159510 14172 17988

NECH_BL109 Porvenir 104655 5840 7402

NECH_EX101 Porvenir 40893 1716 2164

NECH_EX31 Porvenir 137387 10925 13673

NECH_EX34 Porvenir 198545 27129 34627

NECH_EX58 Porvenir 201438 27012 34276

NECH_EX66 Porvenir 187701 20918 26645

NECH_1B34 San Martin 246815 26680 33916

NECH_2B10 San Martin 202598 28490 36175

NECH_2SM55 San Martin 180443 21892 27818

NECH_2SM73 San Martin 204547 29213 37037

NECH_3636 San Martin 193724 24912 31614

NECH_3654 San Martin 182288 20443 26013

Page 146: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

146

Table D-2. Continued

Id Region Unique Stacks

Polymorphic Loci SNPs Found

NECH_3673 San Martin 155924 12944 16476

NECH_3674 San Martin 180701 19873 25303

NECH_3675 San Martin 110075 6569 8419

NECH_3770 San Martin 112617 7363 9389

NECH_3773 San Martin 69955 3674 4553

NECH_6076 San Martin 211390 33534 42927

NECH_6092 San Martin 54400 2043 2673

NECH_VB1 San Martin 180036 19765 25201

NECH_VB2-1 San Martin 193460 24430 30865

NECH_VB2 San Martin 202649 28238 35966

NECH_1Y27 Yuto 207115 29348 37526

NECH_1Y60 Yuto 218711 39051 49685

NECH_1Y63 Yuto 193733 25073 31640

NECH_1Y64 Yuto 203044 28542 36643

NECH_1Y87 Yuto 176730 18388 23376

NECH_1Y89 Yuto 189407 23132 29445

NECH_2Y16 Yuto 193945 25179 32064

NECH_2Y22 Yuto 89269 4241 5458

NECH_2Y70 Yuto 217892 40454 51362

NECH_2Y71 Yuto 218083 40979 52197

NECH_2Y81 Yuto 182097 24177 30347

NECH_3982 Yuto 195076 24054 30633

NECH_3Y40 Yuto 215553 36340 46432

NECH_3Y69 Yuto 218965 42140 53918

NECH_4095 Yuto 205869 28582 36598

NECH_4123 Yuto 164258 15907 20093

NECH_4239 Yuto 202602 28869 36813

NECH_4291 Yuto 162226 15352 19270

NECH_4298 Yuto 165214 15439 19552

NECH_4319 Yuto 181564 21334 26998

NECH_4321 Yuto 119800 7467 9543

NECH_6562 Yuto 169024 16722 21268

NECH_6605 Yuto 205638 29902 38048

NECH_6680 Yuto 176205 20289 25722

Page 147: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

147

Table D-3. Summary statistics for the 129 individuals included in the final filtered dataset, 2240 biallelic loci included in the dataset.

Statistic Value SD* c.i.2.5%** c.i.97.5%** Description

Num 2 0 2 2 Number of alleles

Eff_num 1.412 0.005 1.402 1.422 Effective number of alleles

Hs 0.331 0.003 0.325 0.338 Heterozygosity Within Populations

Ht 0.331 0.003 0.325 0.338 Total Heterozygosity

H't 0.331 0.003 0.325 0.338 Corrected total Heterozygosity

Gis 0.84 0.005 0.829 0.85 Inbreeding coefficient

Gst -0.001 0.003 -0.006 0.005 Fixation index

G'st(Nei) -0.001 0.003 -0.008 0.006 Nei, corrected fixation index

G'st(Hed) -0.001 0.005 -0.01 0.008 Hedrick, standardized fixation index

G''st -0.001 0.005 -0.011 0.008 Corrected standardized fixation index

D_est 0 0.002 -0.004 0.003 Jost, differentiation

*Standard deviations of statistics were obtained through jackknifing over loci.

**95% confidence intervals of statistics were obtained through bootstrapping over loci.

Page 148: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

148

LIST OF REFERENCES

Adams, R. V., and T. M. Burg. 2015. Gene Flow of a Forest-Dependent Bird across a

Fragmented Landscape. Plos One 10.

Adeney, J. M., N. L. Christensen, A. Vicentini, and M. Cohn-Haft. 2016. White-sand ecosystems in Amazonia. Biotropica 48:7-23.

Alonso, J. A., and B. M. Whitney. 2001. A new Zimmerius tyrannulet (Aves : Tyrannidae) from white sand forests of northern Amazonian Peru. Wilson Bulletin 113:1-9.

Alonso, J. A., and B. M. Whitney. 2003. New distributional records of birds from white-sand forests of the northern Peruvian Amazon, with implications for biogeography of northern South America. Condor 105:552-566.

Alonso, J. Á., M. R. Metz, and P. V. A. Fine. 2013. Habitat Specialization by Birds in Western Amazonian White-sand Forests. Biotropica 45:365-372.

Alvarez Alonso, J. 2002. Characteristic Avifauna of White-Sand Forests in Northern Peruvian Amazonia. Louisiana State University.

Alvarez Alonso, J., J. Diaz Alvan, and N. Shany. 2012. Avifauna de la Reserva Nacional Allpahuayo Mishana, Loreto, Perú. Cotinga 64:61-84

Andersen, K. M., B. L. Turner, and J. W. Dalling. 2010. Soil-based habitat partitioning in understorey palms in lower montane tropical forests. Journal of Biogeography 37:278-292.

Anderson, A. B. 1981. White-sand vegetation of brazilian amazonia. Biotropica 13:199-210.

Arbelaez, M. V., and A. Parrado-Rosselli. 2005. Seed dispersal modes of the sandstone plateau vegetation of the middle Caqueta river region, Colombian Amazonia. Biotropica 37:64-72.

Arellano, G., V. Cala, and M. J. Macia. 2014. Niche breadth of oligarchic species in Amazonian and Andean rain forests. Journal of Vegetation Science 25:1355-1366.

Asner, G. P., C. B. Anderson, R. E. Martin, D. E. Knapp, R. Tupayachi, F. Sinca, and Y. Malhi. 2014. Landscape-scale changes in forest structure and functional traits along an Andes-to-Amazon elevation gradient. Biogeosciences 11:843-856.

Page 149: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

149

Asner, G. P., A. R. Townsend, and B. H. Braswell. 2000. Satellite observation of El Nino effects on Amazon forest phenology and productivity. Geophysical Research Letters 27:981-984.

Barber, B. R., J. E. Martinez-Gomez, and A. T. Peterson. 2004. Systematic position of the Socorro mockingbird Mimodes graysoniii. Journal of Avian Biology 35:195-198.

Barnagaud, J. Y., V. Devictor, F. Jiguet, and F. Archaux. 2011. When species become generalists: on-going large-scale changes in bird habitat specialization. Global Ecology and Biogeography 20:630-640.

Baselga, A., and F. Leprieur. 2015. Comparing methods to separate components of beta diversity. Methods in Ecology and Evolution 6:1069-1079.

Baselga, A., and C. D. L. Orme. 2012. betapart: an R package for the study of beta diversity. Methods in Ecology and Evolution 3:808-812.

Benavides, J. C., A. J. Duque, J. F. Duivenvoorden, and A. M. Cleef. 2006. Species richness and distribution of understorey bryophytes in different forest types in Colombian Amazonia. Journal of Bryology 28:182-189.

Bennett, J. R., and B. Gilbert. 2016. Contrasting beta diversity among regions: how do classical and multivariate approaches compare? Global Ecology and Biogeography:n/a-n/a.

Blake, J. G. 2007. Neotropical forest bird communities: A comparison of species richness and composition at local and regional scales. Condor 109:237-255.

Blake, J. G., and B. A. Loiselle. 2009. Species Composition of Neotropical Understory Bird Communities: Local Versus Regional Perspectives Based on Capture Data. Biotropica 41:85-94.

Borges, S. H. 2004. Species poor but distinct: bird assemblages in white sand vegetation in Jau National Park, Brazilian Amazon. Ibis 146:114-124.

Borges, S. H. 2007. Biogeographical analysis of the avifauna in the lower western Rio Negro region in the Brazilian Amazon. Revista Brasileira De Zoologia 24:919-940.

Borges, S. H. 2013. Bird species distribution in a complex Amazonian landscape: species diversity, compositional variability and biotic-environmental relationships. Studies on Neotropical Fauna and Environment 48:106-118.

Borges, S. H., C. Cornelius, M. Moreira, C. C. Ribas, M. Conh-Haft, J. M. Capurucho, C. Vargas, and R. Almeida. 2016. Bird communities in Amazonian white-sand vegetation patches: effects of landscape configuration and biogeographic context. Biotropica 48:121-131.

Page 150: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

150

Bowen, G. J., L. I. Wassenaar, and K. A. Hobson. 2005. Global application of stable hydrogen and oxygen isotopes to wildlife forensics. Oecologia 143:337-348.

Bray, J. R., and J. T. Curtis. 1957. An Ordination of the Upland Forest Communities of Southern Wisconsin. Ecological Monographs 27:325-349.

Brown, D. R., P. C. Stouffer, and C. M. Strong. 2000. Movement and territoriality of wintering Hermit Thrushes in southeastern Louisiana. Wilson Bulletin 112:347-353.

Brown, J. H., S. K. M. Ernest, J. M. Parody, and J. P. Haskell. 2001. Regulation of diversity: maintenance of species richness in changing environments. Oecologia 126:321-332.

Brown, K. S., Jr., and G. T. Prance. 1987. Soils and vegetation. Biogeography and Quaternary History in Tropical America:19-45.

Brumfield, R. T., and A. P. Capparella. 1996. Historical diversification of birds in northwestern South America: A molecular perspective on the role of vicariant events. Evolution 50:1607-1624.

Burleigh, J. G., R. T. Kimball, and E. L. Braun. 2015. Building the avian tree of life using a large-scale, sparse supermatrix. Molecular Phylogenetics and Evolution 84:53-63.

Bush, A. O., and J. C. Holmes. 1983. Niche separation and the broken-stick model - use with multiple assemblages. American Naturalist 122:849-855.

Capurucho, J. M. G., C. Cornelius, S. H. Borges, M. Cohn-Haft, A. Aleixo, J. P. Metzger, and C. C. Ribas. 2013. Combining phylogeography and landscape genetics of Xenopipo atronitens (Aves: Pipridae), a white sand campina specialist, to understand Pleistocene landscape evolution in Amazonia. Biological Journal of the Linnean Society 110:60-76.

Castano-Villa, G. J., S. A. Ramos-Valencia, and F. E. Fonturbel. 2014. Fine-scale habitat structure complexity determines insectivorous bird diversity in a tropical forest. Acta Oecologica-International Journal of Ecology 61:19-23.

Catchen, J., P. A. Hohenlohe, S. Bassham, A. Amores, and W. A. Cresko. 2013. Stacks: an analysis tool set for population genomics. Molecular Ecology 22:3124-3140.

Cavender-Bares, J., K. H. Kozak, P. V. A. Fine, and S. W. Kembel. 2009. The merging of community ecology and phylogenetic biology. Ecology Letters 12:693-715.

Chase, J. M., and J. A. Myers. 2011. Disentangling the importance of ecological niches from stochastic processes across scales. Philosophical Transactions of the Royal Society B: Biological Sciences 366:2351-2363.

Page 151: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

151

Condit, R., N. Pitman, E. G. Leigh, J. Chave, J. Terborgh, R. B. Foster, P. Nunez, S. Aguilar, R. Valencia, G. Villa, H. C. Muller-Landau, E. Losos, and S. P. Hubbell. 2002. Beta-diversity in tropical forest trees. Science 295:666-669.

Connor, E. F. 1979. The statistics and biology of the species-area relationship. The American naturalist 113:791.

Coomes, D. A., and P. J. Grubb. 1996. Amazonian caatinga and related communities at La Esmeralda, Venezuela: Forest structure, physiognomy and floristics, and control by soil factors. Vegetatio 122:167-191.

Cormier, R. L., D. L. Humple, T. Gardali, and N. E. Seavy. 2013. LIGHT-LEVEL GEOLOCATORS REVEAL STRONG MIGRATORY CONNECTIVITY AND WITHIN-WINTER MOVEMENTS FOR A COASTAL CALIFORNIA SWAINSON'S THRUSH (CATHARUS USTULATUS) POPULATION. Auk 130:283-290.

Coronado, E. N. H., K. G. Dexter, R. T. Pennington, J. Chave, S. L. Lewis, M. N. Alexiades, E. Alvarez, A. A. de Oliveira, I. L. Amaral, A. Araujo-Murakami, E. Arets, G. A. Aymard, C. Baraloto, D. Bonal, R. Brienen, C. Ceron, F. C. Valverde, A. Di Fiore, W. Farfan-Rios, T. R. Feldpausch, N. Higuchi, I. Huamantupa-Chuquimaco, S. G. Laurance, W. F. Laurance, G. Lopez-Gonzalez, B. S. Marimon, B. H. Marimon, A. M. Mendoza, D. Neill, W. P. Cuenca, M. C. P. Mora, N. C. A. Pitman, A. Prieto, C. A. Quesada, H. R. Angulo, A. Rudas, A. R. Ruschel, N. S. Revilla, R. P. Salomao, A. S. de Andrade, M. R. Silman, W. Spironello, H. ter Steege, J. Terborgh, M. Toledo, L. V. Gamarra, I. C. G. Vieira, E. V. Torre, V. Vos, and O. L. Phillips. 2015. Phylogenetic diversity of Amazonian tree communities. Diversity and Distributions 21:1295-1307.

Cramer, M. J., and M. R. Willig. 2005. Habitat heterogeneity, species diversity and null models. Oikos 108:209-218.

Csardi, G., and T. Nepusz. 2006. The igraph software package for complex network research. InterJournal, Complex Systems.

Daly, D. C., M. Silveira, H. Medeiros, W. Castro, and F. A. Obermuller. 2016. The white-sand vegetation of Acre, Brazil. Biotropica 48:81-89.

Davey, J. W., P. A. Hohenlohe, P. D. Etter, J. Q. Boone, J. M. Catchen, and M. L. Blaxter. 2011. Genome-wide genetic marker discovery and genotyping using next-generation sequencing. Nature Reviews Genetics 12:499-510.

Donoghue, M. J. 2008. Colloquium paper: a phylogenetic perspective on the distribution of plant diversity. Proceedings of the National Academy of Sciences - PNAS 105 Suppl 1:11549-11555.

dos Anjos, L., G. H. Volpato, E. V. Lopes, P. P. Serafini, F. Poletto, and A. Aleixo. 2007. The importance of riparian forest for the maintenance of bird species richness in

Page 152: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

152

an Atlantic Forest remnant, southern Brazil. Revista Brasileira De Zoologia 24:1078-1086.

Dray, S., and A.-B. Dufour. 2007. The ade4 package: Implementing the duality diagram for ecologists. Journal of Statistical Software 22:1-20.

Dunn, E. H., and C. J. Ralph. 2004. Use of mist nets as a tool for bird population monitoring. Studies in Avian Biology:1-6.

Earl, D. A., and B. M. vonHoldt. 2012. STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conservation Genetics Resources 4:359-361.

Edwards, S. V. 2005. Speciation in birds: genes, geography, and sexual selection. Proceedings of the National Academy of Sciences - PNAS 102 Suppl 1:6550-6557.

Emerson, B. C., F. Cicconardi, P. P. Fanciulli, and P. J. A. Shaw. 2011. Phylogeny, phylogeography, phylobetadiversity and the molecular analysis of biological communities. Philosophical Transactions of the Royal Society B: Biological Sciences 366:2391-2402.

Emerson, B. C., and R. G. Gillespie. 2008. Phylogenetic analysis of community assembly and structure over space and time. Trends in Ecology & Evolution 23:619-630.

Emerson, B. C., and N. Kolm. 2005. Species diversity can drive speciation. Nature 434:1015-1017.

Faaborg, J., R. T. Holmes, A. D. Anders, K. L. Bildstein, K. M. Dugger, S. A. Gauthreaux, P. Heglund, K. A. Hobson, A. E. Jahn, D. H. Johnson, S. C. Latta, D. J. Levey, P. P. Marra, C. L. Merkord, E. Nol, S. I. Rothstein, T. W. Sherry, T. S. Sillett, F. R. Thompson, and N. Warnock. 2010. Conserving migratory land birds in the New World: Do we know enough? Ecological Applications 20:398-418.

Ferreira, L. V., S. S. de Almeida, and P. Parolin. 2010. Amazonian white- and blackwater floodplain forests in brazil: large differences on a small scale. Ecotropica 16:31-41.

Fine, P. 2016. Special Issue: Neotropical white-sand forests. Biotropica 48:5-131.

Fine, P. V. A. 2010. A Floristic Study of the White-Sand Forests of Peru. Annals of the Missouri Botanical Garden 97:283-305.

Fine, P. V. A., and C. Baraloto. 2016. Habitat Endemism in White-sand Forests: Insights into the Mechanisms of Lineage Diversification and Community Assembly of the Neotropical Flora. Biotropica 48:24-33.

Page 153: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

153

Fine, P. V. A., D. C. Daly, G. V. Munoz, I. Mesones, and K. M. Cameron. 2005. The contribution of edaphic heterogeneity to the evolution and diversity of Burseraceae trees in the western Amazon. Evolution 59:1464-1478.

Fine, P. V. A., R. Garcia-Villacorta, N. C. A. Pitman, I. Mesones, and S. W. Kembel. 2010. A floristic study of the white-sand forests of peru. Annals of the Missouri Botanical Garden 97:283-305.

Fine, P. V. A., and S. W. Kembel. 2011. Phylogenetic community structure and phylogenetic turnover across space and edaphic gradients in western Amazonian tree communities. Ecography 34:552-565.

Fine, P. V. A., Z. J. Miller, I. Mesones, S. Irazuzta, H. M. Appel, M. H. H. Stevens, I. Saaksjarvi, J. C. Schultz, and P. D. Coley. 2006. The growth-defense trade-off and habitat specialization by plants in Amazonian forests. Ecology 87:150-162.

Fraser, K. C., B. J. M. Stutchbury, C. Silverio, P. M. Kramer, J. Barrow, D. Newstead, N. Mickle, B. F. Cousens, J. C. Lee, D. M. Morrison, T. Shaheen, P. Mammenga, K. Applegate, and J. Tautin. 2012. Continent-wide tracking to determine migratory connectivity and tropical habitat associations of a declining aerial insectivore. Proceedings of the Royal Society B-Biological Sciences 279:4901-4906.

Frasier, C. L., V. A. Albert, and L. Struwe. 2008. Amazonian lowland, white sand areas as ancestral regions for South American biodiversity: Biogeographic and phylogenetic patterns in Potalia (Angiospermae : Gentianaceae). Organisms Diversity & Evolution 8:44-57.

Futuyma, D. J., and G. Moreno. 1988. The evolution of ecological specialization. Annual Review of Ecology and Systematics 19:207-233.

Gallo, P. A., E. S. Morton, and B. J. M. Stutchbury. 2013. Prolonged spring migration in the red-eyed vireo (vireo olivaceus). Auk 130:240-246.

Garcia-Villacorta, R., K. G. Dexter, and T. Pennington. 2016. Amazonian white-sand forests show strong floristic links with surrounding oligotrophic habitats and the Guiana Shield. Biotropica 48:47-57.

Garcia-Villacorta, R., and B. E. Hammel. 2004. A noteworthy new species of Tovomita (Clusiaceae) from Amazonian white sand forests of Peru and Colombia. Brittonia 56:132-135.

García Villacorta, R., M. Ahuite Reátegui, and M. Olórtegui Zumaeta. 2003. BIODAMAZ project. Folia Amazonica 14:1-191.

Gaston, K. J. 2000. Global patterns in biodiversity. Nature (London) 405:220-227.

Page 154: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

154

Gentry, A. H. 1981. Distributional patterns and an additional species of the passiflora-vitifolia complex - amazonian species-diversity due to edaphically differentiated communities. Plant Systematics and Evolution 137:95-105.

Gentry, A. H. 1988. Tree species richness of upper amazonian forests. Proceedings of the National Academy of Sciences of the United States of America 85:156-159.

Gillies, C. S., and C. C. S. Clair. 2008. Riparian corridors enhance movement of a forest specialist bird in fragmented tropical forest. Proceedings of the National Academy of Sciences of the United States of America 105:19774-19779.

Gomez, C., N. J. Bayly, and K. V. Rosenberg. 2014. Fall stopover strategies of three species of thrush (Catharus) in northern South America. Auk 131:702-717.

Gomez, I. C., I. E. Saaksjarvi, L. Puhakka, C. Castillo, and S. Bordera. 2015. The Peruvian Amazonian species of Epirhyssa Cresson (Hymenoptera: Ichneumonidae: Rhyssinae), with notes on tropical species richness. Zootaxa 3937:311-336.

Gonzalez-Orozco, C. E., B. D. Mishler, J. T. Miller, S. W. Laffan, N. Knerr, P. Unmack, A. Georges, A. H. Thornhill, D. F. Rosauer, and B. Gruber. 2015. Assessing biodiversity and endemism using phylogenetic methods across multiple taxonomic groups. Ecology and Evolution 5:5177-5192.

Gotelli, N. J., and A. Chao. 2013. Measuring and Estimating Species Richness, Species Diversity, and Biotic Similarity from Sampling Data. Pages 195-211.

Gotelli, N. J., and R. K. Colwell. 2001. Quantifying biodiversity: procedures and pitfalls in the measurement and comparison of species richness. Ecology Letters 4:379-391.

Gotelli, N. J., and G. L. Entsminger. 2001. Swap and fill algorithms in null model analysis: rethinking the knight's tour. Oecologia 129:281-291.

Gotelli, N. J., and W. Ulrich. 2012. Statistical challenges in null model analysis. Oikos 121:171-180.

Graham, C. H., and P. V. A. Fine. 2008. Phylogenetic beta diversity: linking ecological and evolutionary processes across space in time. Ecology Letters 11:1265-1277.

Graham, C. H., J. L. Parra, C. Rahbek, and J. A. McGuire. 2009. Phylogenetic structure in tropical hummingbird communities. Proceedings of the National Academy of Sciences of the United States of America 106:19673-19678.

Guevara, J. E., G. Damasco, C. Baraloto, P. V. A. Fine, M. C. Penuela, C. Castilho, A. Vincentini, D. Cardenas, F. Wittmann, N. Targhetta, O. Phillips, J. Stropp, I. Amaral, P. Maas, A. Monteagudo, E. M. Jimenez, R. Thomas, R. Brienen, A. Duque, W. Magnusson, C. Ferreira, E. Honorio, F. d. A. Matos, F. R. Arevalo, J.

Page 155: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

155

Engel, P. Petronolli, and F. de A. Matos. 2016. Low phylogenetic beta diversity and geographic neo-endemism in Amazonian white-sand forests. Biotropica 48:34-46.

Guilherme, E., and S. H. Borges. 2011. Ornithological records from a campina/campinarana enclave on the upper jurua river, acre, brazil. Wilson journal of ornithology 123:24-32.

Hanski, I. 1994. A Practical Model of Metapopulation Dynamics. The Journal of animal ecology 63:151-162.

Hasui, E., V. S. da Mota Gomes, and W. Rodrigues Silva. 2007. Effects of vegetation traits on habitat preferences of frugivorous birds in Atlantic rain forest. Biotropica 39:502-509.

Heckscher, C. M., M. R. Halley, and P. M. Stampul. 2015. Intratropical migration of a Nearctic-Neotropical migratory songbird (Catharus fuscescens) in South America with implications for migration theory. Journal of Tropical Ecology 31:285-289.

Heckscher, C. M., S. M. Taylor, J. W. Fox, and V. Afanasyev. 2011. Veery (catharus fuscescens) wintering locations, migratory connectivity, and a revision of its winter range using geolocator technology. Auk 128:531-542.

Higgins, M. A., G. P. Asner, E. Perez, N. Elespuru, and A. Alonso. 2014. Variation in photosynthetic and nonphotosynthetic vegetation along edaphic and compositional gradients in northwestern Amazonia. Biogeosciences 11:3505-3513.

Hijmans, R. J., S. E. Cameron, J. L. Parra, P. G. Jones, and A. Jarvis. 2005. Very high resolution interpolated climate surfaces for global land areas. International Journal of Climatology 25:1965-1978.

Hobson, K. A., and K. J. Kardynal. 2015. Western Veeries use an eastern shortest-distance pathway: New insights to migration routes and phenology using light-level geolocators. Auk 132:540-550.

Hobson, K. A., S. L. Van Wilgenburg, E. H. Dunn, D. J. T. Hussell, P. D. Taylor, and D. M. Collister. 2015. Predicting origins of passerines migrating through Canadian migration monitoring stations using stable-hydrogen isotope analyses of feathers: a new tool for bird conservation. Avian Conservation and Ecology 10:50.

Hobson, K. A., S. L. Van Wilgenburg, J. Faaborg, J. D. Toms, C. Rengifo, A. L. Sosa, Y. Aubry, and R. B. Aguilar. 2014. Connecting breeding and wintering grounds of Neotropical migrant songbirds using stable hydrogen isotopes: a call for an isotopic atlas of migratory connectivity. Journal of Field Ornithology 85:237-257.

Hohenlohe, P. A., S. J. Amish, J. M. Catchen, F. W. Allendorf, and G. Luikart. 2011. Next-generation RAD sequencing identifies thousands of SNPs for assessing

Page 156: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

156

hybridization between rainbow and westslope cutthroat trout. Molecular Ecology Resources 11:117-122.

Hohenlohe, P. A., J. Catchen, and W. A. Cresko. 2012. Population genomic analysis of model and nonmodel organisms using sequenced RAD tags. Methods in molecular biology (Clifton, N.J.) 888:235-260.

Holmes, R. T., R. E. Bonney, and S. W. Pacala. 1979. Guild structure of the hubbard brook bird community - multivariate approach. Ecology 60:512-520.

Holmes, R. T., and H. F. Recher. 1986. Determinants of guild structure in forest bird communities - an intercontinental comparison. Condor 88:427-439.

Holyoak, M., M. A. Leibold, N. Mouquet, R. D. Holt, and M. F. Hoopes. 2005. Metacommunities - A framework for large-scale community ecology. Metacommunities: Spatial Dynamics and Ecological Communities:1-31.

Hoorn, C., F. P. Wesselingh, H. ter Steege, M. A. Bermudez, A. Mora, J. Sevink, I. Sanmartín, A. Sanchez-Meseguer, C. L. Anderson, J. P. Figueiredo, C. Jaramillo, D. Riff, F. R. Negri, H. Hooghiemstra, J. Lundberg, T. Stadler, T. Särkinen, and A. Antonelli. 2010. Amazonia Through Time: Andean Uplift, Climate Change, Landscape Evolution, and Biodiversity. Science 330:927-931.

Jankowski, J. E., A. L. Ciecka, N. Y. Meyer, and K. N. Rabenold. 2009. Beta diversity along environmental gradients: implications of habitat specialization in tropical montane landscapes. Journal of Animal Ecology 78:315-327.

Janzen, D. H. 1974. Tropical black water rivers animals and mast fruiting by the dipterocarpaceae. Biotropica 6:69-103.

Jetz, W., and P. V. A. Fine. 2012. Global Gradients in Vertebrate Diversity Predicted by Historical Area-Productivity Dynamics and Contemporary Environment. PLoS Biol 10:e1001292.

Jirka, S., A. J. McDonald, M. S. Johnson, T. R. Feldpausch, E. G. Couto, and S. J. Riha. 2007. Relationships between soil hydrology and forest structure and composition in the southern Brazilian Amazon. Journal of Vegetation Science 18:183-194.

Jombart, T. 2008. adegenet: a R package for the multivariate analysis of genetic markers. Bioinformatics 24:1403-1405.

Jombart, T., and I. Ahmed. 2011. adegenet 1.3-1: new tools for the analysis of genome-wide SNP data. Bioinformatics.

Kalliola, R., A. Linna, M. Puhakka, J. Salo, and M. Rasanen. 1993. Mineral nutrients in fluvial sediments from the peruvian amazon. Catena 20:333-349.

Page 157: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

157

Karr, J. R. 1979. On the use of mist nets in the study of bird communities. Inland Bird Banding 51:1-10.

Karr, J. R. 1981. Surveying birds with mist nets. Studies in Avian Biology:62-67.

Karr, J. R., S. K. Robinson, J. G. Blake, and R. O. J. Bierregaard. 1991. Birds of four neotropical forests. Gentry, a. H. (Ed.). Four Neotropical Rainforests; Meeting of the American Institute of Biological Sciences, Columbus, Ohio, USA, August 10-12, 1987. Xvi+627p. Yale University Press: New Haven, Connecticut, USA; London, England, Uk. Illus. Maps:237-270.

Karr, J. R., and R. R. Roth. 1971. Vegetation structure and avian diversity in several new world areas. American Naturalist 105:423-&.

Kirwan, G., and G. Green. 2011. Helm identification guides: contingas and manakins. Helm identification guides: contingas and manakins.:1-624.

Kissling, W. D., C. Rahbek, and K. Bohning-Gaese. 2007. Food plant diversity as broad-scale determinant of avian frugivore richness. Proceedings of the Royal Society B-Biological Sciences 274:799-808.

Kooyman, R., M. Rossetto, W. Cornwell, and M. Westoby. 2011. Phylogenetic tests of community assembly across regional to continental scales in tropical and subtropical rain forests. Global Ecology and Biogeography 20:707-716.

Kornan, M., R. T. Holmes, H. F. Recher, P. Adamik, and R. Kropil. 2013. Convergence in foraging guild structure of forest breeding bird assemblages across three continents is related to habitat structure and foraging opportunities. Community Ecology 14:89-100.

Kraft, N. J. B., P. B. Adler, O. Godoy, E. C. James, S. Fuller, and J. M. Levine. 2014. Community assembly, coexistence and the environmental filtering metaphor. Functional Ecology:n/a-n/a.

Kraft, N. J. B., L. S. Comita, J. M. Chase, N. J. Sanders, N. G. Swenson, T. O. Crist, J. C. Stegen, M. Vellend, B. Boyle, M. J. Anderson, H. V. Cornell, K. F. Davies, A. L. Freestone, B. D. Inouye, S. P. Harrison, and J. A. Myers. 2011. Disentangling the Drivers of beta Diversity Along Latitudinal and Elevational Gradients. Science 333:1755-1758.

Kratter, A. W. 1997. Bamboo Specialization by Amazonian Birds1. Biotropica 29:100-110.

Lamarre, G. P. A., D. S. Amoretti, C. Baraloto, F. Beneluz, I. Mesones, and P. V. A. Fine. 2016a. Phylogenetic overdispersion in Lepidoptera communities of Amazonian white-sand forests. Biotropica 48:101-109.

Page 158: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

158

Lamarre, G. P. A., C. Baraloto, C. Fortunel, N. Dávila, I. Mesones, J. G. Rios, M. Ríos, E. Valderrama, M. V. Pilco, and P. V. A. Fine. 2012. Herbivory, growth rates, and habitat specialization in tropical tree lineages: implications for Amazonian beta-diversity. Ecology 93:S195-S210.

Lamarre, G. P. A., B. Herault, P. V. A. Fine, V. Vedel, R. Lupoli, I. Mesones, and C. Baraloto. 2016b. Taxonomic and functional composition of arthropod assemblages across contrasting Amazonian forests. Journal of Animal Ecology 85:227-239.

Lawton, J. H. 1996. Patterns in ecology. Oikos 75:145-147.

Leibold, M. A., M. Holyoak, N. Mouquet, P. Amarasekare, J. M. Chase, M. F. Hoopes, R. D. Holt, J. B. Shurin, R. Law, D. Tilman, M. Loreau, and A. Gonzalez. 2004. The metacommunity concept: a framework for multi-scale community ecology. Ecology Letters 7:601-613.

Logue, J. B., N. Mouquet, H. Peter, and H. Hillebrand. 2011. Empirical approaches to metacommunities: a review and comparison with theory. Trends in Ecology &amp; Evolution 26:482-491.

Loiselle, B. A., and J. G. Blake. 1999. Dispersal of melastome seeds by fruit-eating birds of tropical forest understory. Ecology 80:330-336.

Lomolino, M. V. 2000. Ecology's most general, yet protean 1 pattern: the species-area relationship. Journal of biogeography 27:17-26.

Lomolino, M. V. 2001. Towards a more general species-area relationship: diversity on all islands, great and small Small island effect. Journal of biogeography 28:431-445.

Lozupone, C., M. E. Lladser, D. Knights, J. Stombaugh, and R. Knight. 2011. UniFrac: an effective distance metric for microbial community comparison. The ISME journal 5:169-172.

MacArthur, R. H. 1961. On Bird Species Diversity. Ecology (Durham) 42:594.

Macarthur, R. H. 1964. Environmental factors affecting bird species diversity. American Naturalist 98:387-397.

MacArthur, R. H., and E. O. Wilson. 1967. The Theory of Island Biogeography. Princeton University Press.

Macedo, M., and G. T. Prance. 1978. Notes on vegetation of amazonia .2. Dispersal of plants in amazonian white sand campinas - campinas as functional islands. Brittonia 30:203-215.

Page 159: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

159

Malhado, A. C. M., R. Ladle, R. Whittaker, A. Neto, Y. Malhi, and H. ter Steege. 2013. The ecological biogeography of Amazonia. Frontiers of Biogeography 5.

Mantel, N. 1967. Detection of disease clustering and a generalized regression approach. Cancer Research 27:209-&.

Marks, B. D., S. J. Hackett, and A. P. Capparella. 2002. Historical relationships among Neotropical lowland forest areas of endemism as determined by mitochondrial DNA sequence variation within the Wedge-billed Woodcreeper (Aves : Dendrocolaptidae : Glyphorynchus spirurus). Molecular Phylogenetics and Evolution 24:153-167.

Martin, T. E., and J. R. Karr. 1986. Temporal dynamics of neotropical birds with special reference to frugivores in 2nd-growth woods. Wilson Bulletin 98:38-60.

Martin, T. E., C. R. Paine, C. J. Conway, W. M. Hochachka, P. Allen, and W. Jenkins. 1997. BBIRD (Breeding biology research and monitoring database) field protocol. Available from T. E. Martin, Montana Cooperative Wildlife Research Unit, University of Montana, Missoula, MT 59812.

Mayfield, M. M., and J. M. Levine. 2010. Opposing effects of competitive exclusion on the phylogenetic structure of communities. Ecology Letters 13:1085-1093.

McKay, B. D., F. K. Barker, H. L. Mays, Jr., S. M. Doucet, and G. E. Hill. 2010. A molecular phylogenetic hypothesis for the manakins (Ayes: Pipridae). Molecular Phylogenetics and Evolution 55:733-737.

Meirmans, P. G., and P. H. Van Tienderen. 2004. GENOTYPE and GENODIVE: two programs for the analysis of genetic diversity of asexual organisms. Molecular Ecology Notes 4:792-794.

Meynard, C. N., V. Devictor, D. Mouillot, W. Thuiller, F. Jiguet, and N. Mouquet. 2011. Beyond taxonomic diversity patterns: how do alpha, beta and gamma components of bird functional and phylogenetic diversity respond to environmental gradients across France? Global Ecology and Biogeography 20:893-903.

Mittelbach, G. G., C. F. Steiner, S. M. Scheiner, K. L. Gross, H. L. Reynolds, R. B. Waide, M. R. Willig, S. I. Dodson, and L. Gough. 2001. What is the observed relationship between species richness and productivity? Ecology 82:2381-2396.

Moilanen, A. 1998. Metapopulation dynamics: effects of habitat quality and landscape structure. Ecology (Durham) 79:2503-2515.

Moore, R. P., W. D. Robinson, I. J. Lovette, and T. R. Robinson. 2008. Experimental evidence for extreme dispersal limitation in tropical forest birds. Ecology Letters 11:960-968.

Page 160: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

160

Morris, D. W. 1996. Coexistence of specialist and generalist rodents via habitat selection. Ecology 77:2352-2364.

Naka, L. N. 2011. Avian distribution patterns in the Guiana Shield: implications for the delimitation of Amazonian areas of endemism Boundaries of the Guianan area of endemism. Journal of biogeography 38:681-696.

Nores, M. 1999. An alternative hypothesis for the origin of Amazonian bird diversity. Journal of Biogeography 26:475-485.

Nores, M. 2000. Species richness in the Amazonian bird fauna from an evolutionary perspective. Pages 419-430.

Nur, N., G. R. Geupel, and G. Ballard. 2004. Estimates of adult survival, capture probability, and recapture probability: evaluating and validating constant-effort mist netting. Studies in Avian Biology 29:63-70.

Olson, D. M., E. Dinerstein, E. D. Wikramanayake, N. D. Burgess, G. V. N. Powell, E. C. Underwood, J. A. D'Amico, I. Itoua, H. E. Strand, J. C. Morrison, C. J. Loucks, T. F. Allnutt, T. H. Ricketts, Y. Kura, J. F. Lamoreux, W. W. Wettengel, P. Hedao, and K. R. Kassem. 2001. Terrestrial ecoregions of the worlds: A new map of life on Earth. Bioscience 51:933-938.

Orrock, J. L., and J. I. Watling. 2010. Local community size mediates ecological drift and competition in metacommunities.

Oswald, J. A., J. G. Burleigh, D. W. Steadman, S. K. Robinson, and A. W. Kratter. 2016. Historical climatic variability and geographical barriers as drivers of community composition in a biodiversity hotspot. Journal of Biogeography 43:123-133.

Pearman, P. B. 2002. The scale of community structure: Habitat variation and avian guilds in tropical forest understory. Ecological Monographs 72:19-39.

Peixoto, F. P., P. H. P. Braga, M. V. Cianciaruso, J. A. F. Diniz, and D. Brito. 2014. Global patterns of phylogenetic beta diversity components in bats. Journal of Biogeography 41:762-772.

Pennington, R. T. 2009. Diversification of the Amazonian Flora and its Relation to key Geological and Environmental Events: A Molecular Perspective. Pages 373-385.

Pianka, E. R. 1966. Latitudinal Gradients in Species Diversity: A Review of Concepts. The American Naturalist 100:33-46.

Pitman, N. C. A., H. Mogollon, N. Davila, M. Rios, R. Garcia-Villacorta, J. Guevara, T. R. Baker, A. Monteagudo, O. L. Phillips, R. Vasquez-Martinez, M. Ahuite, M. Aulestia, D. Cardenas, C. E. Ceron, P. A. Loizeau, D. A. Neill, P. Nunez V, W. A. Palacios, R. Spichiger, and E. Valderrama. 2008. Tree community change across

Page 161: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

161

700 km of lowland Amazonian forest from the Andean foothills to Brazil. Biotropica 40:525-535.

Poletto, F., and A. Aleixo. 2005. Biogeographic implications of new avian records from a patch of white-sand forest in southwestern Brazilian Amazonia. Revista Brasileira De Zoologia 22:1196-1200.

Pomara, L. Y., K. Ruokolainen, H. Tuomisto, and K. R. Young. 2012. Avian Composition Co-varies with Floristic Composition and Soil Nutrient Concentration in Amazonian Upland Forests. Biotropica 44:545-553.

Poulsen, A. D., H. Tuomisto, and H. Balslev. 2006. Edaphic and floristic variation within a 1-ha plot of lowland Amazonian rain forest. Biotropica 38:468-478.

Prance, G. T. 1996. Islands in Amazonia. Philosophical Transactions of the Royal Society of London Series B-Biological Sciences 351:823-833.

Pritchard, J. K., M. Stephens, and P. Donnelly. 2000. Inference of population structure using multilocus genotype data. Genetics 155:945-959.

Puhakka, M., R. Kalliola, M. Rajasilta, and J. Salo. 1992. River types, site evolution and successional vegetation patterns in peruvian Amazonia. Journal of Biogeography 19:651-665.

Remsen Jr, J., and D. A. Good. 1996. Misuse of data from mist-net captures to assess relative abundance in bird populations. The Auk:381-398.

Remsen, J. V., and T. A. Parker. 1983. Contribution of river-created habitats to bird species richness in amazonia. Biotropica 15:223-231.

Remsen, J. V. J., and S. K. Robinson. 1990. A classification scheme for foraging behavior of birds in terrestrial habitats. Morrison, M. L., Et Al. (Ed.). Studies in Avian Biology, No. 13. Avian Foraging: Theory, Methodology, and Applications; International Symposium, Asilomar, California, USA, December 18-19, 1988. X+515p. Cooper Ornithological Society: Los Angeles, California, USA. Illus. Paper:144-160.

Ricklefs, R. E. 2008. Disintegration of the Ecological Community. American Naturalist 172:741-750.

Ricklefs, R. E. 2011. A biogeographical perspective on ecological systems: some personal reflections. Journal of Biogeography 38:2045-2056.

Rimmer, C. C., and K. P. McFarland. 2001. Known breeding and wintering sites of a Bicknell's Thrush. Wilson Bulletin 113:234-236.

Page 162: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

162

Robinson, S. K., and R. T. Holmes. 1982. Foraging behavior of forest birds - The relationships among search tactics, diet, and habitat structure. Ecology 63:1918-1931.

Robinson, S. K., and J. Terborgh. 1990. Bird communities of the Cocha Cashu Biological Station in Amazonian Peru. Four Neotropical rainforests.:199-216.

Robinson, S. K., and J. Terborgh. 1995. Interspecific aggression and habitat selection by Amazonian birds. Journal of Animal Ecology 64:1-11.

Robinson, W. D., J. D. Brawn, and S. K. Robinson. 2000. Forest bird community structure in central Panama: Influence of spatial scale and biogeography. Ecological Monographs 70:209-235.

Rosenzweig, M. L. 1987. Habitat selection as a source of biological diversity. Evolutionary Ecology 1:315-330.

Rosenzweig, M. L. 1995. Species diversity in space and time. Cambridge University Press, Cambridge ; New York.

Ruegg, K. C. 2002. Not as the crow flies: a historical explanation for circuitous migration in Swainson's thrush (Catharus ustulatus). Proceedings of the Royal Society. B, Biological sciences 269:1375-1381.

Ruokolainen, K., A. Linna, and H. Tuomisto. 1997. Use of Melastomataceae and pteridophytes for revealing phytogeographical patterns in Amazonian rain forests. Journal of Tropical Ecology 13:243-256.

Ruokolainen, K., H. Tuomisto, and R. Kalliola. 2005. Landscape heterogeneity and species diversity in Amazonia. Tropical rainforests: past, present and future:251-270.

Ruokolainen, K., H. Tuomisto, M. J. Macia, M. A. Higgins, and M. Yli-Halla. 2007. Are floristic and edaphic patterns in Amazonian rain forests congruent for trees, pteridophytes and Melastomataceae? Journal of Tropical Ecology 23:13-25.

Saaksjarvi, I. E., S. Haataja, S. Neuvonen, I. D. Gauld, R. Jussila, J. Salo, and A. M. Burgos. 2004. High local species richness of parasitic wasps (Hymenoptera : Ichneumonidae; Pimplinae and Rhyssinae) from the lowland rainforests of Peruvian Amazonia. Ecological Entomology 29:735-743.

Saaksjarvi, I. E., K. Ruokolainen, H. Tuomisto, S. Haataja, P. V. A. Fine, G. Cardenas, I. Mesones, and V. Vargas. 2006. Comparing composition and diversity of parasitoid wasps and plants in an Amazonian rain-forest mosaic. Journal of Tropical Ecology 22:167-176.

Page 163: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

163

Salo, M., and A. Pyhala. 2007. Exploring the gap between conservation science and protected area establishment in the Allpahuayo-Mishana National Reserve (Peruvian Amazonia). Environmental Conservation 34:23-32.

Sanderson, M. J. 2003. r8s: inferring absolute rates of molecular evolution and divergence times in the absence of a molecular clock. Bioinformatics 19:301-302.

Santana, C. R., and L. dos Anjos. 2010. On the association of birds to bamboo stands in Southern brazilian Atlantic Forest. Biota Neotropica 10:39-44.

Schulenberg, T. S., D. F. Stotz, D. F. Lane, J. P. O'Neill, and T. A. I. Parker. 2007. Birds of Peru. Helm Field Guides. Birds of Peru. [Helm Field Guides.]:1-656.

Slatkin, M. 1993. Isolation by Distance in Equilibrium and Non-Equilibrium Populations. Evolution 47:264-279.

Stamatakis, A. 2006. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22:2688-2690.

Steege, H. T., D. Sabatier, H. Castellanos, T. Van Andel, J. Duivenvoorden, A. A. De Oliveira, R. Ek, R. Lilwah, P. Maas, and S. Mori. 2000. An analysis of the floristic composition and diversity of Amazonian forests including those of the Guiana Shield. Journal of Tropical Ecology 16:801-828.

Stegen, J. C. 2013. Stochastic and deterministic drivers of spatial and temporal turnover in breeding bird communities Drivers of spatial and temporal turnover. Global ecology and biogeography 22:202-212.

Stiles, F. G. 1975. Ecology, flowering phenology, and hummingbird pollination of some costa rican Heliconia species. Ecology 56:285-301.

Storch, D., R. G. Davies, S. Zajicek, C. D. L. Orme, V. Olson, G. H. Thomas, T. S. Ding, P. C. Rasmussen, R. S. Ridgely, P. M. Bennett, T. M. Blackburn, I. P. F. Owens, and K. J. Gaston. 2006. Energy, range dynamics and global species richness patterns: reconciling mid-domain effects and environmental determinants of avian diversity. Ecology Letters 9:1308-1320.

Stotz. 1996. Neotropical birds : ecology and conservation. Pages xx, 478 p.

Stotz, D. F., J. W. Fitzpatrick, T. Parker III, A., and D. K. Moskowitz. 1996. Neotropical birds : ecology and conservation. Pages xx, 478 p.

Stropp, J., P. Van der Sleen, P. A. Assuncao, A. L. da Silva, and H. Ter Steege. 2011. Tree communities of white-sand and terra-firme forests of the upper Rio Negro. Acta Amazonica 41:521-543.

Page 164: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

164

Svenning, J. C. 2001. On the role of microenvironmental heterogeneity in the ecology and diversification of neotropical rain-forest palms (Arecaceae). Botanical Review 67:1-53.

Svenning, J. C., B. M. J. Engelbrecht, D. A. Kinner, T. A. Kursar, R. F. Stallard, and S. J. Wright. 2006. The relative roles of environment, history and local dispersal in controlling the distributions of common tree and shrub species in a tropical forest landscape, Panama. Journal of Tropical Ecology 22:575-586.

Swenson, N. G., J. C. Stegen, S. J. Davies, D. L. Erickson, J. Forero-Montana, A. H. Hurlbert, W. J. Kress, J. Thompson, M. Uriarte, S. J. Wright, and J. K. Zimmerman. 2012. Temporal turnover in the composition of tropical tree communities: functional determinism and phylogenetic stochasticity. Ecology 93:490-499.

Tello, J. G., R. G. Moyle, D. J. Marchese, and J. Cracraft. 2009. Phylogeny and phylogenetic classification of the tyrant flycatchers, cotingas, manakins, and their allies (Aves: Tyrannides). Cladistics 25:429-467.

Tene Fossog, B. 2015. Habitat segregation and ecological character displacement in cryptic African malaria mosquitoes. Evolutionary applications 8:326-345.

ter Steege, H., N. C. A. Pitman, D. Sabatier, C. Baraloto, R. P. Salomao, J. E. Guevara, O. L. Phillips, C. V. Castilho, W. E. Magnusson, J.-F. Molino, A. Monteagudo, P. Nunez Vargas, J. C. Montero, T. R. Feldpausch, E. N. H. Coronado, T. J. Killeen, B. Mostacedo, R. Vasquez, R. L. Assis, J. Terborgh, F. Wittmann, A. Andrade, W. F. Laurance, S. G. W. Laurance, B. S. Marimon, B.-H. Marimon, Jr., I. C. Guimaraes Vieira, I. L. Amaral, R. Brienen, H. Castellanos, D. Cardenas Lopez, J. F. Duivenvoorden, H. F. Mogollon, F. D. d. A. Matos, N. Davila, R. Garcia-Villacorta, P. R. Stevenson Diaz, F. Costa, T. Emilio, C. Levis, J. Schietti, P. Souza, A. Alonso, F. Dallmeier, A. J. D. Montoya, M. T. Fernandez Piedade, A. Araujo-Murakami, L. Arroyo, R. Gribel, P. V. A. Fine, C. A. Peres, M. Toledo, G. A. Aymard C, T. R. Baker, C. Ceron, J. Engel, T. W. Henkel, P. Maas, P. Petronelli, J. Stropp, C. E. Zartman, D. Daly, D. Neill, M. Silveira, M. R. Paredes, J. Chave, D. d. A. Lima Filho, P. M. Jorgensen, A. Fuentes, J. Schongart, F. Cornejo Valverde, A. Di Fiore, E. M. Jimenez, M. C. Penuela Mora, J. F. Phillips, G. Rivas, T. R. van Andel, P. von Hildebrand, B. Hoffman, E. L. Zent, Y. Malhi, A. Prieto, A. Rudas, A. R. Ruschell, N. Silva, V. Vos, S. Zent, A. A. Oliveira, A. C. Schutz, T. Gonzales, M. Trindade Nascimento, H. Ramirez-Angulo, R. Sierra, M. Tirado, M. N. Umana Medina, G. van der Heijden, C. I. A. Vela, E. Vilanova Torre, C. Vriesendorp, O. Wang, K. R. Young, C. Baider, H. Balslev, C. Ferreira, I. Mesones, A. Torres-Lezama, L. E. Urrego Giraldo, R. Zagt, M. N. Alexiades, L. Hernandez, I. Huamantupa-Chuquimaco, W. Milliken, W. Palacios Cuenca, D. Pauletto, E. Valderrama Sandoval, L. Valenzuela Gamarra, K. G. Dexter, K. Feeley, G. Lopez-Gonzalez, and M. R. Silman. 2013. Hyperdominance in the Amazonian tree flora. Science (New York, N.Y.) 342:1243092-1243092.

Page 165: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

165

Terborgh, J., S. K. Robinson, T. A. Parker, C. A. Munn, and N. Pierpont. 1990. Structure and organization of an Amazonian forest bird community. Ecological Monographs 60:213-238.

Thessler, S., K. Ruokolainen, H. Tuomisto, and E. Tomppo. 2005. Mapping gradual landscape-scale floristic changes in Amazonian primary rain forests by combining ordination and remote sensing. Global Ecology and Biogeography 14:315-325.

Thiollay, J.-M. 1994. Structure, density and rarity in an Amazonian rainforest bird community. Journal of tropical ecology 10:449.

Thomas, C. D., and W. E. Kunin. 1999. The spatial structure of populations. Journal of Animal Ecology 68:647-657.

Tuomisto, H. 2006. Edaphic niche differentiation among Polybotrya ferns in western Amazonia: implications for coexistence and speciation. Ecography 29:273-284.

Tuomisto, H. 2007. Interpreting the biogeography of South America. Journal of Biogeography 34:1294-1295.

Tuomisto, H. 2010a. A diversity of beta diversities: straightening up a concept gone awry. Part 1. Defining beta diversity as a function of alpha and gamma diversity. Ecography 33:2-22.

Tuomisto, H. 2010b. A diversity of beta diversities: straightening up a concept gone awry. Part 2. Quantifying beta diversity and related phenomena. Ecography 33:23-45.

Tuomisto, H., and A. D. Poulsen. 1996. Influence of edaphic specialization on pteridophyte distribution in neotropical rain forests. Journal of Biogeography 23:283-293.

Tuomisto, H., and K. Ruokolainen. 1997. The role of ecological knowledge in explaining biogeography and biodiversity in Amazonia. Biodiversity and Conservation 6:347-357.

Tuomisto, H., K. Ruokolainen, R. Kalliola, A. Linna, W. Danjoy, and Z. Rodriguez. 1995. Dissecting Amazonian Biodiversity. Science 269:63-66.

Tuomisto, H., K. Ruokolainen, A. D. Poulsen, R. C. Moran, C. Quintana, G. Canas, and J. Celi. 2002. Distribution and diversity of pteridophytes and Melastomataceae along edaphic gradients in Yasuni National Park, Ecuadorian Amazonia. Biotropica 34:516-533.

Tuomisto, H., K. Ruokolainen, and M. Yli-Halla. 2003. Dispersal, environment, and floristic variation of western Amazonian forests. Science 299:241-244.

Page 166: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

166

Ungvari-Martin, J., C. M. Heckscher, and K. A. Hobson. 2016. Inter-annual site fidelity and breeding origins of Gray-cheeked Thrushes in

white sand forests of the Peruvian Amazon. Journal of Field Ornithology 87:55-64.

Urban, M. C., M. A. Leibold, P. Amarasekare, L. De Meester, R. Gomulkiewicz, M. E. Hochberg, C. A. Klausmeier, N. Loeuille, C. de Mazancourt, J. Norberg, J. H. Pantel, S. Y. Strauss, M. Vellend, and M. J. Wade. 2008. The evolutionary ecology of metacommunities. Trends in Ecology & Evolution 23:311-317.

Van Houtan, K. S., S. L. Pimm, J. M. Halley, R. O. Bierregaard, and T. E. Lovejoy. 2007. Dispersal of Amazonian birds in continuous and fragmented forest. Ecology Letters 10:219-229.

Veen, T. 2010. Temporal differences in food abundance promote coexistence between two congeneric passerines. Oecologia 162:873-884.

Vellend, M. 2010. Conceptual synthesis in community ecology. Quarterly Review of Biology 85:183-206.

Vidal, M. M., E. Hasui, M. A. Pizo, J. Y. Tamashiro, W. R. Silva, and P. R. Guimaraes, Jr. 2014. Frugivores at higher risk of extinction are the key elements of a mutualistic network. Ecology 95:3440-3447.

Vormisto, J., O. L. Phillips, K. Ruokolainen, H. Tuomisto, and R. Vasquez. 2000. A comparison of fine-scale distribution patterns of four plant groups in an Amazonian rainforest. Ecography 23:349-359.

Wang, X. G., N. G. Swenson, T. Wiegand, A. Wolf, R. Howe, F. Lin, J. Ye, Z. Q. Yuan, S. Shi, X. J. Bai, D. L. Xing, and Z. Q. Hao. 2013. Phylogenetic and functional diversity area relationships in two temperate forests. Ecography 36:883-893.

Warkentin, I. G. 1996. The conservation implications of site fidelity: A case study involving nearctic-neotropical migrant songbirds wintering in a Costa Rican mangrove. Biological conservation 77:143-150.

Wassenaar, L. I., and K. A. Hobson. 2001. A stable-isotope approach to delineate geographical catchment areas of avian migration monitoring stations in North America. Environmental Science & Technology 35:1845-1850.

Webb, C. O., D. D. Ackerly, M. A. McPeek, and M. J. Donoghue. 2002. Phylogenies and community ecology. Annual Review of Ecology and Systematics 33:475-505.

Weinstein, B. G., B. Tinoco, J. L. Parra, L. M. Brown, J. A. McGuire, F. G. Stiles, and C. H. Graham. 2014. Taxonomic, Phylogenetic, and Trait Beta Diversity in South American Hummingbirds. American Naturalist 184:211-224.

Willis, E. O., and Y. Oniki. 1992. Birds and army ants. Boletim do Museu Paraense Emilio Goeldi Serie Zoologia 8:123-150.

Page 167: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

167

Wilman, H., J. Belmaker, J. Simpson, C. de la Rosa, M. M. Rivadeneira, and W. Jetz. 2014. EltonTraits 1.0: Species-level foraging attributes of the world's birds and mammals. Ecology:2027-2027.

Wright, S. 1943. Isolation by distance. Genetics (Austin) 28:114.

Wunderle, J. M., and S. C. Latta. 2000. Winter site fidelity of nearctic migrants in shade coffee plantations of different sizes in the Dominican Republic. Auk 117:596-614.

Wunderle, J. M., M. R. Willig, and L. M. P. Henriques. 2005. Avian distribution in treefall gaps and understorey of terra firme forest in the lowland Amazon. Ibis 147:109-129.

Page 168: © 2016 Judit Ungvari-Martin - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/11/00001/UNGVARI... · 2016-09-01 · judit ungvari-martin a dissertation presented to the

168

BIOGRAPHICAL SKETCH

A biographical sketch is required of all candidates and it typically includes the

educational background of the candidate, but here it is partially replaced by the life story

of Judit Ungvari-Martin. She was born in Hungary, behind the iron curtain. She

remembers the first time she ate a banana in the summer, soon after the fall of

communism in the country. She was somewhat of a black sheep in the family for always

choosing outdoor activities, and anything involving animals or the family garden for

recreational activities. Despite her liking of all things biology and thinking about the

veterinary career, she attended law school 1998-1999 in Miskolc, Hungary. Very soon

she discovered that law is not her calling, and she ran off to volunteer on an Israeli

kibbutz where she worked in citrus groves, in the communal kitchen, in a plastic factory

and with toddlers in a preschool. She attended the Technion University and the

University of Haifa studying natural sciences and Hebrew. She moved to the United

States in 2003, got two BS degrees in 2007 in zoology and wildlife ecology and

conservation. She fell in love with working with birds while completing her

undergraduate Honors thesis with Gustavo Londoño. She fell in love with the tropics

after spending 6 month in Peru volunteering on a field project. Her dissertation research

is a product of 22 months in the field amongst amazing local people, volunteers from all

over the world, and all things that the Amazon can offer. Judit is committed to

enhancing the scientific capacity through direct interactions with students. She has

trained 44 undergraduate field assistants, advised 6 students through their honors

thesis projects. When not running around in the tropics, Judit enjoys life in places where

the weather is mild, cooking, baking, sharing life with family, friends, and pets. She is a

long time foster for pet rescue organizations.