· Web viewWord count: 6,960. Journal pages for figures and table: 5. ABSTRACT. Aim: To...

43
1 Article type: Original Article Geographic variation in the evolutionary diversity of tree communities across southern South America Vanessa L Rezende 1,2 ; Kyle G. Dexter 2,3 ; R. Toby Pennington 2 ; Ary T Oliveira-Filho 1 1 Department of Plant Biology, University of Minas Gerais. Belo Horizonte, MG 31270-901, Brazil. 2 Royal Botanic Garden Edinburgh, Edinburgh EH3 5LR, UK. 3 School of GeoSciences, University of Edinburgh, Edinburgh EH9 3FF, UK. Correspondence author: Vanessa Leite Rezende, University of Minas Gerais, Belo Horizonte, MG 31270-901, Brazil. Email: [email protected] Running head: Phylogenetic patterns across southern South America Word count: 6,960 Journal pages for figures and table: 5 ABSTRACT Aim: To determine the principal drivers of variation in the evolutionary diversity of forest tree communities, with a focus on the temperate forests of South America. Location: Forests across southern South America, extending from tropical forests in southern Brazil, to the temperate forests of southern Chile and Argentina. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Transcript of   · Web viewWord count: 6,960. Journal pages for figures and table: 5. ABSTRACT. Aim: To...

1

Article type: Original Article

Geographic variation in the evolutionary diversity of tree communities

across southern South America

Vanessa L Rezende1,2; Kyle G. Dexter2,3; R. Toby Pennington2; Ary T Oliveira-Filho1

1Department of Plant Biology, University of Minas Gerais. Belo Horizonte, MG 31270-901, Brazil.

2Royal Botanic Garden Edinburgh, Edinburgh EH3 5LR, UK. 3School of GeoSciences, University of Edinburgh, Edinburgh EH9 3FF, UK.

Correspondence author: Vanessa Leite Rezende, University of Minas Gerais, Belo Horizonte, MG 31270-901,

Brazil. Email: [email protected]

Running head: Phylogenetic patterns across southern South America

Word count: 6,960

Journal pages for figures and table: 5

ABSTRACT

Aim: To determine the principal drivers of variation in the evolutionary diversity of forest tree

communities, with a focus on the temperate forests of South America.

Location: Forests across southern South America, extending from tropical forests in southern Brazil, to

the temperate forests of southern Chile and Argentina.

Methods: We compiled a database of 742 forest tree community inventories spread over six countries

and 12 biomes, or major vegetation types. In total, these inventories covered 3075 species of shrubs and

trees. We combined this dataset with a temporally calibrated phylogeny that included all species. For

each community we evaluated multiple measures of evolutionary diversity, including phylogenetic

diversity sensu stricto (PD), which is the sum of the branch lengths of a phylogeny that includes all

species in a community, and its equivalent standardised for variation in species richness (ses.PD), which

we refer to as lineage diversity.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

2

Results: We found that biome affiliation is the most important determinant of the evolutionary diversity

of woody plant communities, with climate also showing a significant influence. Communities in wet

evergreen tropical forest have the highest species richness and the highest PD, but the lowest lineage

diversity, while temperate forests in southern South America show the lowest species richness and PD,

but the highest lineage diversity.

Main conclusions: Our results contradict the idea that temperate floras represent recently derived,

evolutionarily poor subsets of tropical floras. Rather, the high lineage diversity we find in temperate

forest communities supports the Austral Conservatism Hypothesis, which states that the flora of

southern South America has evolved independently from the Neotropical Domain over tens of millions

of years. Our identification of the evolutionary distinctness and richness of this flora suggests that it

deserves as much conservation attention as the more species-rich tropical forests of South America and

that southern South American forests should not be lumped into the Neotropical Floristic Province.

Key words: Temperate Forest, Tropical Forest, Biome, Phylogenetic Diversity, Lineage Diversity,

Tropical Conservatism Hypothesis, Austral Conservatism Hypothesis, Latitudinal Gradients, Species

Richness.

INTRODUCTION

The tropical conservatism hypothesis (TCH, Wiens&Donoghue, 2004) has been proposed to

underlie the present-day latitudinal diversity gradient, the pattern whereby species richness is currently

highest in equatorial regions and declines towards the poles. It suggests that species distribution patterns

at global scales, particularly for plants, are largely governed by evolutionarily conserved ancestral

preferences (Wiens & Graham, 2005; Kozak & Wiens, 2010; Romdal et al., 2013). It assumes that most

clades originated in tropical conditions that were widespread from the beginning of the Cretaceous to the

end of the Eocene (Davies et al., 2004; Ruddiman, 2006), and that most species in these clades have at

least partly retained their ancestral physiological tolerances (Ricklefs & Schluter, 1993; Wiens &

Donoghue 2004; Wiens et al., 2010; Jansson et al., 2013). An important component of this hypothesis is

the inability of many lineages to adapt to freezing environments (Condamine et al., 2012), limiting

dispersal and evolution from tropical to temperate regions (Crispet al., 2009). Thus, according to the

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

3

TCH, species richness, evolutionary diversity and lineage age will be high in regions characterized by

warm, non-freezing temperatures, because these conditions match the ancestral niches of many extant

clades, while temperate regions will be occupied by fewer, younger and less diverse clades, because the

transitions to cold, temperate environments have been relatively infrequent and recent (Wiens &

Donoghue 2004).

In support of the TCH, two recent studies have shown that the mean age of angiosperm families

declines away from the equator (Hawkins et al., 2011; Romdal et al., 2013). Focusing on the New

World, Kerkhoff et al. (2014) used a different approach to test the TCH, which consisted of examining

the evolutionary diversity of latitudinal bands (estimated using various metrics of phylogenetic

diversity). Their results supported the TCH in finding that latitudinal bands further away from the

equator have less lineage diversity. In contrast, Boucher-Lalonde et al. (2015) found that once positive

relationships between species richness and current climate are taken into account, the temperature, and

presumably latitude, at which a clade originates has little influence on its species richness, which the

authors take as evidence against the TCH. Other studies have also found results that are contradictory to

patterns expected under the TCH, both at a global scale (Jansson et al., 2013) and in South America

(Segovia et al., 2013; Qian, 2014; Tiede et al., 2016). However, most previous studies have largely

focused on the northern hemisphere when interpreting results (Segovia & Armesto, 2015). Furthermore,

the patterns used to argue for the TCH, for example a decline in family age and evolutionary diversity

away from the equator, are not particularly evident in the southern hemisphere , at least for woody

angiosperm families (see Fig. 4b in Hawkins et al., 2011, and Fig. 1a in Kerkhoff et al., 2014). Here, we

aim to test the TCH in the southern hemisphere, by examining evolutionary diversity in woody plant

communities across southern South America, where support for the TCH has been mixed and uncertain.

The present-day South American flora has complex origins (Pennington & Dick, 2004) and

includes a group of lineages with southern temperate affinities, which have been suggested to have

evolved during and after the breakup of Gondwana, and a group of Neotropical elements which are

largely found in northern South America (Romero, 1986; Villagran & Hinojosa, 1997; Eisenlohr &

Oliveira-filho, 2015). Based on this, Segovia & Armesto (2015) proposed the idea of the Austral

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

4

Conservatism Hypothesis (ACH), that the floras of temperate South America are not simply an

evolutionarily poor subset of tropical clades, but contain many lineages that have diversified outside of

the tropics and retained their ancestral preference for temperate conditions. In order to support the ACH,

Segovia & Armesto (2015) reviewed data showing that the mean age of angiosperm families increases

with altitude and as one moves southward in South America. However, Boucher-Lalonde et al. (2015)

have shown that analyses of mean family ages may be insufficient to test hypotheses around niche

conservatism. As an alternative, studies of the evolutionary diversity of ecological communities can be

used to test the potential role for niche conservatism in determining spatial patterns of biodiversity (c.f.

Kerkhoff et al., 2014) and to distinguish the relative merit of the TCH and the ACH in shaping the South

American flora.

Many metrics have been developed to measure the evolutionary diversity of communities, with

most quantitative metrics being derived from phylogenies (Cadotte et al., 2010). The most basic metric

is to sum the branch lengths of a phylogeny that covers all species in a given community, which we refer

to as phylogenetic diversity sensu stricto (PD). In most empirical datasets, PD is strongly positively

correlated with species richness (SR) (e.g. Forest et al., 2007, Honorio-Coronado et al., 2015). As

species richness is strongly influenced by current climate (Kreft & Jetz, 2007, Boucher-Lalonde et al.,

2015), it is often of interest to assess whether communities show more or less PD than expected given

their SR. One particularly common metric to measure this departure from expectation is the standardised

effect size of PD (ses.PD), which we refer to as lineage diversity (sensu Honorio-Coronado et al., 2015).

Here, in our analyses of the evolutionary diversity of forest tree communities across southern South

America, we focus on measuring phylogenetic diversity sensu stricto and lineage diversity as well as

other commonly employed metrics of phylogenetic diversity.

Most prior studies have used data from herbarium collections, which have many sources of

potential bias when it comes to estimating SR or other metrics of diversity, particularly those that are

strongly correlated with species richness, e.g. PD (Droissart et al., 2012). Previous large-scale studies in

South America have therefore often been restricted to estimating SR, PD or lineage diversity for large

latitudinal bins (e.g. Romdal et al., 2013; Kerkhof et al., 2014), rather than for local communities. By

not analysing local communities, these studies have been unable to adequately assess the potential

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

107

108

109

110

5

influence of environmental factors on SR, PD or lineage diversity (c.f. Boucher-Lalonde et al., 2015).

Those studies that have taken a community approach have had sparse sampling across the southern

hemisphere, particularly in South America (e.g. Kreft & Jetz, 2007; Qian, 2014).

Previous large-scale studies of latitudinal diversity patterns in plants have also focused largely

on angiosperms (e.g. Hawkins et al., 2011, Kerkhoff et al., 2014).Although angiosperms dominate the

majority of terrestrial ecosystems (Lidgard & Crane, 1988; Willis & McElwain, 2002; Crisp & Cook,

2011), in some forests, mainly located at high latitudes or altitudes, gymnosperms dominate (Aerts,

1995; Augusto et al., 2014). Likewise, tree ferns can be another important component of tree

communities, particularly in tropical, wet forests (Smale et al., 1997; Chacón-Labellaet al., 2014). Thus,

to have a comprehensive understanding of PD and lineage diversity in tree communities, it is critical to

include gymnosperm and tree fern lineages, which, being early diverging lineages, may have a large

impact on phylogenetic diversity metrics.

Finally, none of these previous studies in South America have incorporated biome identity into

analyses, which is potentially important as different biomes may have different biogeographic histories

and show different patterns of variation in species richness or phylogenetic diversity (Pennington et al.,

2009; Hughes et al., 2013). For example, the Neotropical seasonally dry tropical forest (SDTF) biome

may not even show the typical latitudinal diversity gradient. Rather, the opposite may hold, as species

richness seems to increase away from the equator (Gentry, 1995; DRYFLOR, 2016).

In this study, we examine spatial patterns of species richness and evolutionary diversity across

the 12 principal woody biomes of southern South America and across a large latitudinal gradient from

15°S to 54°S. We take a community approach, based on 742 species inventories of forest tree

communities, and combine this with a phylogenetic hypothesis for all non-climbing woody plants in

these communities, including gymnosperms and tree ferns. We assess how SR, PD and lineage diversity

vary across climatic environments and biomes. If a tropical origin and phylogenetic conservatism for

tropical environments were the dominant forces shaping large-scale patterns, we would expect tropical

biomes to show higher lineage diversity. Conversely, if the austral conservatism hypothesis holds, then

we should expect comparable or even higher lineage diversity in temperate biomes compared to tropical

ones.

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

6

MATERIALS AND METHODS

Study Area

Our study area spans portions of six countries, Paraguay, Chile, Argentina, Uruguay, Brazil and

Bolivia (Fig. 1), with a latitudinal range from 15° 32’ S to 54° 50’ S and a longitudinal range from 41°

02’ W to 78° 50’ W. We focused on sampling the 12 major forest types, or biomes, in temperate South

America and adjacent subtropical and tropical regions. The biomes were defined following the

classification proposed by Rezende et al. (2016) and comprise: (1) Seasonally Dry Tropical Forest, with

>60% leaf deciduousness in the dry season; (2) Tropical Semideciduous Forest, characterized by 30-

60% leaf deciduousness in the dry season; (3) Wet Evergreen Tropical Forest, with <30% leaf

deciduousness, generally hot and wet climates and associated with the Atlantic coast; (4) Subtropical

Semideciduous Forest, which is similar to Tropical Semideduous, but experiences frost; (5) Araucaria

Mixed Forest, which is similar to Wet Evergreen Tropical Forest, but cooler and characterized by the

presence of Araucaria angustifolia (Araucariaceae); (6) Cloud Dwarf-forest, characteristic of high

altitude areas (above 1500m) with a low canopy (between 3 and 5m in height) and frequent cloud

immersion; (7) Southern Andean Forest, which occurs along the eastern foothills of the Andes in

northern Argentina and southern Bolivia and includes communities variously designated as Tucumano-

Boliviano, Yungas and Chaco Serrano; (8) Gran Chaco, a dry biome which experiences frost and which

encompasses the dry and wet Chaco; (9) Western Pampa Forest, consisting of deciduous or

semideciduous forest with a seasonally cold climatic regime and found along rivers; (10) Eastern Pampa

Forest, which is similar to Western Pampa Forest, but experiences a more maritime climatic regime and

is not specifically associated with rivers; (11) Semi-Arid and Arid Patagonian Forest, which consists of

the Patagonian Temperate Steppe, the Monte Subtropical Semi-deserts and the Espinal region; and (12)

Pacific Forest, that encompasses Mediterranean Chile and Temperate Pacific forests. We combined

these latter two as one biome due to their high floristic similarity to each other relative to other biomes

and the low number of sampling sites in Mediterranean Chilean forests (see Fig. 4 in Rezende et al.,

2016).

139

140

141

142

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

7

Database

We obtained floristic data from the NeoTropTree database (Oliveira-Filho, 2015), which is a

compilation of tree species checklists gathered from the literature with additional occurrence records

obtained from verified herbarium specimens. At present, NeoTropTree contains information on the

composition of tree species for >5,000 georeferenced locations spread across South America. Each site

comprises a 5-km radius circle and contains records of tree species that can be found within this area. In

addition to the presence and absence data on tree species, each site contains descriptive and

environmental data such as biome, altitude, geo-edaphic and climatic variables (see description, history

and NeoTropTree protocol at http://www.icb.ufmg.br/treeatlan). NeoTropTree defines trees as

freestanding plants with stems that can reach over 3m in height. Thus, the NeoTropTree database

includes, for example, palms, tree ferns and bamboos, when they fit the inclusion criteria.

Analysis of Phylogenetic Diversity

We generated an ultrametric, temporally calibrated phylogeny for all genera in the dataset, using

Phylocom (ver 4.2), starting with the following reference tree: R20120829.new (from

http://phylodiversity.net/phylomatic/). We incorporated gymnosperms and tree ferns into this starting

tree based on relationships in Lu et al. (2014) and Quiroga et al. (2015) for gymnosperms and Silvestro

et al. (2015) for tree ferns. All taxa (i.e. species) are first assigned to the clade to which they belong

taxonomically. To temporally calibrate this phylogeny, we assigned fixed ages to 125 of the 820 nodes

in the phylogeny and used the bladj algorithm to constrain the remaining node ages (n= 695). This

algorithm spaces the remaining nodes evenly in time between the nodes with fixed ages, based on their

hierarchical position with respect to each other. We began with a set of ages from Gastauer & Meira-

Neto (2013), and manually modified clade ages based on Magallón et al. (2015) for angiosperms,

Quiroga et al. (2015) and Silvestro et al. (2015) for gymnosperms and Lu et al. (2014) for tree ferns.

For each floristic list, or tree community, in the dataset, we calculated a series of phylogenetic

diversity metrics (sensu Honorio-Coronado et al., 2015): phylogenetic diversity sensu stricto (PD),

which is the total sum of branch lengths in a phylogeny comprising species in the community; the mean

167

168

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

8

pairwise phylogenetic distance between species in the community (MPD); the mean nearest taxon

distance (MNTD), which is the mean phylogenetic distance from each taxon to its closest relative in the

community; and their equivalents, standardized for species richness (ses.PD, ses.MPD, ses.MNTD). The

standardizations were accomplished by randomly drawing the same number of species from the

phylogeny as present in the community, repeating this 999 times, calculating PD, MPD and MNTD for

each randomization, taking the difference between the observed value and the mean of the random

values, and dividing this difference by the standard deviation of values across the randomizations.

We constructed linear models, using a generalized least squares (GLS) approach to account for

spatial autocorrelation, to test the relationship between the phylogenetic metrics and our explanatory

variables. The NeoTropTree database contains climatic data for each site, including the 19 BIOCLIM

variables (Hijmans et al., 2005) and measures of water deficit duration and severity, potential

evapotranspiration (PET), seasonality in water availability(sum of duration of the excess and water

deficit periods), aridity, days of frost and cloud interception (see Neves et al., 2015 for details). We

assessed correlations amongst these climatic variables, and for each statistical model, we included the

uncorrelated variables that had the most explanatory power. Preliminary analyses indicated that latitude

and longitude showed strong relationships with diversity metrics, and we therefore also included these

variables as proxies for climatic variation that may not have been captured by the measured climatic

variables. We ensured that all variance inflation factors (VIFs) were less than five for each explanatory

variable (Quinn & Keough, 2002). We included biome as a separate explanatory variable. Biome

covaries with climate, but we were interested in testing the role of biome alone and to assess if biomes

have explanatory power beyond the climate with which they are associated (and vice versa). Our GLS

approach allowed us to account for spatial autocorrelation when testing the relative influence of biome

versus climate on different diversity metrics. In preliminary analyses, we found an exponential spatial

autocorrelation structure to best fit the data, and we therefore used this structure when generating all

models. Models of the full set of explanatory variables (best climate variables + biome) were compared

to nested subsets where climate or biome were excluded. We used the Akaike Information Criterion

(AIC) to compare models. Lastly, for comparison with previous studies and to explore latitudinal

gradients, we visually explored how species richness and phylogenetic diversity metrics vary with

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

216

217

218

219

220

221

222

9

latitude within and across biomes. We also conducted a principal component analysis to visualize the

relationships amongst climatic variables and with latitude (Fig. 2). All analyses were conducted in the R

Statistical Environment (R Core Working Team, 2015) using the picante (Kembel et al., 2016), vegan

(Oksanen et al., 2016), car (Fox et al., 2016) and nlme (Pinheiro et al., 2016) packages.

RESULTS

In the database, 98.87% of species are angiosperms, 0.69% are gymnosperms and 0.50% are

tree ferns. On average, the biome with the lowest proportion of angiosperms is the Pacific Forest, where

a mean of 92.73% of species per site belong to angiosperms, while the biome with the highest

proportion of angiosperms is the Wet Evergreen Tropical Forest, with a mean of 98.46% species per site

belong to angiosperms per plot (Table 1). When analysing species richness (SR), Wet Evergreen

Tropical Forest has the highest SR, with an average of 323 species per site, whereas the lowest SR was

found in the Pacific Forest (29 species per community; Table 1).

Across all sites, we found phylogenetic diversity sensu stricto (PD) to be strongly correlated

with SR, while the standardised measure of PD (ses.PD), termed lineage diversity, was positively

correlated with mean nearest taxon distance (MNTD) and its standardised equivalent (ses.MNTD) (Fig.

2; see Appendix S1 in Supporting Information). Mean pairwise distance (MPD) and its standardised

equivalent (ses.MPD) are strongly correlated with each other (Fig. 3). MPD and ses.MPD are strongly

influenced by branch lengths at the deepest nodes of the phylogeny, and as a result, these metrics are

often strongly driven by how evenly taxa are divided among major clades (Swenson, 2014). This is the

case in our dataset, where we found that the overwhelming influence on MPD and ses.MPD values was

how evenly taxa are divided amongst tree ferns, gymnosperms and angiosperms, and secondarily

amongst the three major clades of angiosperms: magnoliids, monocots and eudicots. The evenness with

which species in communities are divided amongst these clades explains 96% and 85% of the variation

in MPD and ses.MPD values respectively (see Appendix S2). Thus, following Honorio Coronado et al.

(2015), we focus our results and discussion on PD, lineage diversity, MNTD and ses.MNTD.

We observe similar patterns for PD as we found for SR, where the highest value was found for

communities in Wet Evergreen Tropical Forest (17820.43), while the lowest values were found for

223

224

225

226

227

228

229

230

231

232

233

234

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

10

communities in Semi-Arid and Arid Patagonian Forest (2684.408) and the Pacific Forest (3058.268; see

Appendix S3; Fig. 3). MNTD showed an opposite result to that found for PD and SR. The highest value

was found for the Pacific Forest communities (154.2256), while the Wet Evergreen Tropical Forest had

the lowest value (90.27885). For the standardized metrics (ses.PD and ses.MNTD), the highest values

were also found in the Temperate Pacific Forest communities (0.975112 and -0.21606 respectively).

The best models for all metrics were the full models, including biome and climatic variables

(Table 2). Interestingly though, for all metrics except MNTD, the pure biome model explained the data

better than a model with just climatic data. The best climatic model for PD included the same climatic

variables as the model for SR, which is unsurprising given how tightly correlated it is with SR (see Fig.

S1; Table 2). For the standardised metrics, ses.PD and ses.MNTD, adding the climatic data increased

model performance less than it did for the raw diversity metrics. The high importance found for

temperature-related variables, such as days of frost, relative to precipitation-related variables

corroborates other studies in the extratropical portion of southern South America (Rezende et al., 2015;

Oliveira-Filho et al., 2013; Gonçalves & Souza, 2014).

We then visualized how these different phylogenetic diversity metrics vary with latitude (Fig.3).

As ses.PD and ses.MNTD are strongly correlated (r= 0.844, p<0.001) and thus show similar patterns, the

results for ses.MNTD are not shown. The direction of trends with latitude were similar across biomes,

with a few exceptions. In Cloud Forest and Tropical Semideciduous Forest, ses.PD increased towards

the equator. In all other biomes, ses.PD values increased away from the equator in a southerly direction.

DISCUSSION

In agreement with previous studies (Hawkins, 2001; Turner, 2004; Oliveira-Filho et al., 2013),

we found that the species richness (SR) of tree communities in the southern hemisphere declines as one

moves southwards, away from the equator (Fig. 3A). Similarly, phylogenetic diversity, measured as the

total sum of phylogenetic branch lengths in a community (PD), declines away from the equator (Fig.

3B), which is expected given the tight correlation of SR and PD (see Fig. S1). In contrast, once we

control for the effect of species richness on PD and evaluate phylogenetic, or lineage, diversity in

communities using a standardized metric (ses.PD), we found that lineage diversity increases as one

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

11

moves away from the equator (Fig. 3D). Thus temperate tree communities of southern South America

show greater lineage diversity than expected given their species richness. We found similar results in

analyses restricted to angiosperms (see Appendix S3), which indicates that this result is not due solely to

the increased frequency of gymnosperms in these far southern communities. Rather, in general, the tree

species in these southern communities, both angiosperms and gymnosperms, tend to come from older

lineages (see also Segovia et al., 2013). Overall, our results provide strong support for the Austral

Conservatism Hypothesis (ACH). Our results do not support the idea that southern hemisphere

temperate floras are recently derived, evolutionarily poor subsets of tropical floras, which is an

implication of the Tropical Conservatism Hypothesis (TCH) as it is often interpreted (e.g. Kerkhoff et

al., 2014).

In our analyses, biome was the variable that best explained variation in the phylogenetic

diversity metrics. However, climate is a major factor determining the distribution of terrestrial biomes

(Whittaker, 1975, Lyra et al., 2016), which could suggest that climate is in fact the major driver of

variation in phylogenetic diversity, not biomes per se. Climatic gradients and energetic constraints are

often invoked to explain variation in species richness, with which phylogenetic diversity sensu stricto is

strongly correlated. However, we found higher lineage diversity in colder places, while a hypothesis

based on energetic constraints would predict higher lineage diversity in warmer places. Therefore, we

suggest that the high lineage diversity values found in southern, temperate communities reflects

accumulated lineage diversity, with many deep phylogenetic branches for communities in these southern

biomes, relative to their species richness (Swenson, 2009). Kerkhoff et al. (2014) noted an increase in

lineage diversity from subtropical areas to the poles, but they related this to latitudinal extent patterns,

stating that temperate and boreal areas occupy a larger area than subtropical areas. While this may be the

case in the northern hemisphere, it cannot be an explanation for South America, where tropical and sub-

tropical biomes occupy a much larger area than temperate biomes, and a boreal zone is lacking.

One reason for higher lineage diversity in temperate regions is that close relatives in temperate

forest communities diverged longer ago than close relatives in tropical forest communities, which is

indicated by the higher mean nearest taxon distance (MNTD) values in more southern, colder areas (Fig.

3C). This result may be because the temperate flora has high richness at the family level with low

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

305

306

12

generic and species richness, as well as a high proportion of monotypic genera, with several of these

genera even representing monogeneric families (e.g., Aextoxicaceae, Gomortegaceae, Desfontainiaceae

and Eucryphiaceae; Vilagran & Hinojosa, 1997). Whilst family richness in the tropical biomes is high,

there is also high richness at the genus and species levels.

Variation in diversification rates across biomes could contribute to the observed variation in

evolutionary diversity metrics. Our finding of lower lineage diversity in present-day tropical tree

communities could be due to higher recent diversification rates in the tropics, which could be caused in

turn by higher recent speciation rates in tropical regions relative to temperate regions or higher recent

extinction rates in temperate regions relative to tropical regions. For example, among conifers, total

diversity was higher in South America in the Eocene, with a number of lineages present that are now

restricted to Australasia (Agathis, Papuacedrus, Dacrycarpus) (Wilf et al., 2009; Wilf, 2012; Wilf et al.,

2014), and their extinction may be affecting lineage diversity estimates. However, as with most

phylogenies derived using the Phylocom software (Webb et al., 2008), our phylogeny lacks resolution

amongst species within genera and often lacks resolution amongst genera within families, which

prevents us from being able to use it to estimate diversification, speciation or extinction rates. In

addition, our community surveys do not contain all species in the clades we are sampling and thus our

phylogeny, even if fully resolved, would have real missing data issues when it comes to estimating these

rates. Meanwhile, some recent studies have shown exceptional diversification rates in temperate plant

clades (e.g. Valente et al., 2010), although comparative studies across latitudes have tended to show

higher diversification rates in plants at tropical latitudes (Jansson & Davies, 2008; Svenning et al.,

2008). Diversification and age estimates derived from molecular phylogenies must be interpreted with

caution however, as rates of genetic substitution also vary with latitude (Gillman & Wright, 2014). In

any case, there have been no synthetic studies of diversification rates for tree clades of southern South

America to allow us to assess the importance of variation in diversification rates to the observed

patterns. Instead, we have shown how a community phylogenetic approach can be used to try and

understand variation in the evolutionary diversity of tree communities.

The results for the phylogenetic diversity metrics we evaluated in tree communities are

explicable by the biogeographic history of the flora of southern South America. Overall, our findings

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

328

329

330

331

332

333

334

13

show that the floras of South America’s extratropical biomes are not a narrow, cold-adapted subset of

that of tropical regions, but, as suggested by Segovia & Armesto (2015), they also include diverse

lineages that have a temperature southern hemisphere origin and a history of diversification outside of

the tropics. Southern South American floras comprise considerable numbers of southern temperate

genera such as Nothofagus (~80ma), Podocarpus (~63ma), Aextoxicon, Gomortega (~78ma), Pitavia,

Legrandia, Laureliopsis, Laurelia, Citronella, Cryptocarya, Persea and Drimys that give this vegetation

a markedly different floristic composition from the rest of South America (Zuloaga et al., 2008). There

are also numerous families that are rare in the tropics but much more abundant in southern temperate

latitudes (e.g. Proteaceae, Casuarinaceae, Cunoniaceae, Atherospermataceae and Akaniaceae). Another

conspicuous aspect of these southern floras is the absence of major Neotropical clades (e.g. Meliaceae,

Anacardiaceae, Sapotaceae, Moraceae and Annonaceae). The Pacific forests in this study which show

particularly high lineage diversity are currently isolated from other forests in South America by the

Andes. However, the lineages that are contributing to this lineage diversity predate the rise of the Andes.

Furthermore, other studies beyond South America (Kooyman et al., 2012; Kooyman et al., 2014; Lee et

al., 2016) suggest that high lineage diversity may be a phenomenon common to temperate forests of the

southern hemisphere, e.g.. in Australia and New Zealand, although the exact lineage composition may

differ across continents and islands (Lee et al., 2016). In the northern hemisphere lineage diversity

patterns may differ from those we found in South America; for example, Qian & Ricklefs (2016) found

that in North America, temperate plant communities seem to have been assembled from tropical lowland

floras.

Our results suggest that the distinctive ecology and biogeographical history of each biome could

be a key to understanding the distribution of tree species in southern South America. We have shown

that the forests of southern temperate biomes in South America have a unique and diverse evolutionary

history, which has important implications for biological conservation and in delimiting major floristic

provinces. All of South America is usually included in the Neotropical floristic province (Cox et al.,

2001), but clearly the temperate forests of southern South America should be regarded as distinct. In

terms of conservation, it is imperative to preserve the temperate forests of South America, as well as

tropical forests, given their unexpectedly high and exceptional lineage diversity. Should these temperate

335

336

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

352

353

354

355

356

357

358

359

360

361

362

14

forests be lost, we stand to lose a significant amount of evolutionary history that is not found anywhere

else in South America and perhaps in the world.

ACKNOWLEDGEMENTS

VLR thanks CNPq for supporting a 9-month study period at the Royal Botanic Garden Edinburgh (grant

SWE- 205162/2014-2) and CAPES for the PHD scholarship in UFMG. VLR also thanks the Royal

Botanic Garden Edinburgh for support during the time this research was conducted.

REFERENCES

Aerts, R. (1995) The advantages of being evergreen. Trends in Ecology & Evolution, 10, 402–407.

Augusto, L., Davies, T.J., Delzon, S. & Schrijver, A. De. (2014) The enigma of the rise of angiosperms:

Can we untie the knot? Ecology Letters, 17, 1326–1338.

Bellingham, P.J., Wiser, S.K., Hall, G.M.J., Alley, J.C., Allen, R.B. & Suisted, P.A. (1999) Impacts of

possum browsing on the long-term maintenance of forest biodiversity. Science for Conservation,

103, 1–59.

Boucher-Lalonde, V., De Camargo, R.X., Fortin, J-M., Khair, S., So, R.I., Rivera, H.V., Watson, D.,

Zuloaga, J. & Currie, D.J. (2015) The weakness of evidence supporting tropical niche

conservatism as a main driver of current richness-temperature gradients. Global Ecology and

Biogeography, 24, 795–803.

Cadotte, M.W, Davies, T.J., Regetz, J., Kembel, S.W., Cleland, E. & Oakley, T.H. (2010) Phylogenetic

diversity metrics for ecological communities: Integrating species richness, abundance and

evolutionary history. Ecology Letters, 13, 96–105.

Chacón-Labella, J., De la Cruz, M., Vicuña, R., Tapia, K. & Escudero, A. (2014) Negative density

dependence and environmental heterogeneity effects on tree ferns across succession in a tropical

montane forest. Perspectives in Plant Ecology, Evolution and Systematics, 16, 52–63.

363

364

365

366

367

368

369

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

15

Condamine, F.L., Sperling, F.A.H., Wahlberg, N., Rasplus, J-Y & Kergoat, G.J. (2012) What causes

latitudinal gradients in species diversity? Evolutionary processes and ecological constraints on

swallowtail biodiversity. Ecology Letters, 15, 267–277.

Cox, C.B., Cottage, F. & Close, B. (2001) The Biogeographic Regions Reconsidered. Journal of

Biogegraphy, 28, 511– 523.

Crisp, M.D., Arroyo, M.T.K.,Cook, L.G., Gandolfo, M.A., Jordan, G.J., McGlone, MS.,Weston, P.H.,

Westoby, M. Wilf, P. & Linder, P. (2009) Phylogenetic biome conservatism on a global scale.

Nature, 458, 754–756.

Crisp, M.D. & Cook, L.G. (2011) Cenozoic extinctions account for the low diversity of extant

gymnosperms compared with angiosperms. New Phytologist, 192, 997–1009.

Davies, T.J., Savolainen, V., Chase, M.W., Moat, J. & Barraclough, T.G. (2004) Environmental energy

and evolutionary rates in flowering plants. Proceedings of the Royal Society B: Biological

sciences, 271, 2195–200.

Droissart, V., Hardy, O.J., Sonké, B., Dahdouh-Guebas, F. & Stévart, T. (2012) Subsampling herbarium

collections to assess geographic diversity gradients: A case study with endemic Orchidaceae and

Rubiaceae in Cameroon. Biotropica, 44, 44–52.

Eisenlohr, P.V. & Oliveira-Filho, A.T. (2015) Revisiting patterns of tree species composition and their

driving forces in the Atlantic forests of southeastern Brazil. Biotropica, 47, 689-701.

Forest, F., Grenyer, R., Rouget, M., Davies, T.J., Cowling, R.M., Faith, D.P., Balmford, A., Manninga,

J.C., Provhes, S., van der Bank, M., Reeves, G., Hedderson, T.A.J. & Savolainen, V. (2007)

Preserving the evolutionary potential of floras in biodiversity hotspots. Nature, 445, 757–760.

Fox, J., Weisberg, S., Adler, D., et al. (2014) Package “car”. R package version 2.1-2, companion to

applied regression. Available at: http://CRAN.R-project.org/package=car.

Gastauer, M. & Meira-Neto, J.A.A. (2013) Avoiding inaccuracies in tree calibration and phylogenetic

community analysis using Phylocom 4.2. Ecological Informatics, 15, 85–90.

388

389

390

391

392

393

394

395

396

397

398

399

400

401

402

403

404

405

406

407

408

409

410

411

412

16

Gillman, L.N. & Wright, S.D. (2014) Species richness and evolutionary speed: The influence of

temperature, water and area. Journal of Biogeography, 41, 39–51.

Gentry, A.H. (1995) Diversity and floristic composition of neotropical dry forests. Seasonally dry

tropical forests (ed. by S.H.Bullock, H.A.Mooney and E.Medina), pp. 146–194. Cambridge

University Press, Cambridge.

Gonçalves, E.T. & Souza, A.F. (2014) Floristic variation in ecotonal areas: Patterns, determinants and

biogeographic origins of subtropical forests in South America. Austral Ecology, 39, 122–134.

Hawkins, B.A. (2001) Ecology’s oldest pattern? Endeavour, 25, 133–134.

Hawkins, B. a., Rodríguez, M.Á. & Weller, S.G. (2011) Global angiosperm family richness revisited:

Linking ecology and evolution to climate. Journal of Biogeography, 38, 1253–1266.

Hijmans, R.J., Cameron, S.E., Parra, J.L., Jones, P.G. & Jarvis, A. (2005) Very high resolution

interpolated climate surfaces for global land areas. International Journal of Climatology, 25, 1965–

1978.

Honorio Coronado, E.N., Dexter, K.G., Pennington, R.T., et al. (2015) Phylogenetic diversity of

Amazonian tree communities. Diversity and Distributions, 21, 1295–1307.

Hughes, C.E., Pennington, R.T. & Antonelli, A. (2013) Neotropical Plant Evolution: Assembling the

Big Picture. Botanical Journal of the Linnean Society, 171, 1–18.

Jansson, R., Rodríguez-Castañeda, G. & Harding, L.E. (2013) What Can Multiple Phylogenies Say

About the Latitudinal Diversity Gradient? a New Look At the Tropical Conservatism, Out of the

Tropics, and Diversification Rate Hypotheses. Evolution, 67, 1741–1755.

Jansson, R. & Davies, T.J. (2008) Global variation in diversification rates of flowering plants: Energy

vs. climate change. Ecology Letters, 11, 173–183.

Kembel, S.W., Ackerly, D.D., Blomberg, S.P., Cornwell, W.K., Cowan, P.D., Helmus, M.R., Morlon,

H. & Webb, C.O. (2015) Package “picante”. R tools for integrating phylogenies and ecology. R

package version 1.6-2.

413

414

415

416

417

418

419

420

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

17

Kerkhoff, A.J., Moriarty, P.E. & Weiser, M.D. (2014) The latitudinal species richness gradient in New

World woody angiosperms is consistent with the tropical conservatism hypothesis. Proceedings of

the National Academy of Sciences, 111, 8125–8130.

Kooyman, R., Rossetto, M., Allen, C. & Cornwell, W. (2012) Australian Tropical and Subtropical Rain

Forest Community Assembly: Phylogeny, Functional Biogeography, and Environmental Gradients.

Biotropica, 44, 668–679.

Kooyman, R.M. , Wilf, P., Barreda, V.D., et al. (2014) Paleo-Antarctic rainforest into the modern old

world tropics: The rich past and threatened future of the “southern wet forest survivors.” American

Journal of Botany, 101, 2121–2135.

Kozak, K.H. & Wiens, J.J. (2010) Accelerated rates of climatic-niche evolution underlie rapid species

diversification. Ecology Letters, 13, 1378–1389.

Kreft, H. & Jetz, W. (2007) Global patterns and determinants of vascular plant diversity. Proceedings of

the National Academy of Sciences, 104, 5925–5930.

Lee, D.E., Lee, W.G., Jordan, G.J & Barreda, V.D. (2016) The Cenozoic history of New Zealand

temperate rainforests: comparisons with southern Australia and South America. New Zealand

Journal of Botany, 8643, 1–28.

Lidgard, S. & Crane, P.R. (1988) Quantitative analyses of the early angiosperm radiation. Nature, 331,

344–346.

Lu, Y., Ran, J-H., Guo, D-M., Yang, Z-Y. & Wang, X-Q. (2014) Phylogeny and divergence times of

gymnosperms inferred from single-copy nuclear genes. PloS one, 9, p.e107679.

Lyra, A.D.A., Chou, S.C. & Sampaio, G.D.O. (2016) Sensitivity of the Amazon biome to high

resolution climate change projections. Acta Amazonica, 46, 175–188.

Magallón, S., Gómez-Acevedo, S., Sánchez-Reyes, L.L. & Hernández-Hernández, T. (2015) A

metacalibrated time-tree documents the early rise of flowering plant phylogenetic diversity. New

Phytologist, 207, 437-53.

438

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

454

455

456

457

458

459

460

461

462

18

Neves, D.M., Dexter, K.G., Pennington, R.T., Bueno, M.L. & Oliveira-Filho, A.T. (2015)

Environmental and historical controls of floristic composition across the South American Dry

Diagonal. Journal of Biogeography, 42, 1566–1576.

Oksanen, J., Blanchet, F.G., Kindt, R., Legendre, P., Minchin, P.R., O’Hara, R.B., Simpson, G.L.,

Solymos, P., Stevens, M.H.H. & Wagner, H. (2016) vegan: community ecology package. R

package version 2.0–3. Available at: http://CRAN.R-project.org/package=vegan.

Oliveira-Filho, A.T. (2015) Um Sistema de classificação fisionômico-ecológica da vegetação

Neotropical. Fitossociologia no Brasil: Métodos e estudos de casos, volume 2. (ed. by P.V.

Eisenlohr, J.M. Felfili, M.M.R.F. Melo, L.A. Andrade & J.A.A. Meira-Neto), pp. 452-473. Editora

UFV, Viçosa, Brazil.

Oliveira-Filho, A.T., Budke, J.C., Jarenkow, J. A.,Eisenlohr, P.V. & Neves, D.R.M (2013) Delving into

the variations in tree species composition and richness across South American subtropical Atlantic

and Pampean forests. Journal of Plant Ecology, 8, 242-260.

Pennington, R.T. & Dick, C.W. (2004) The role of immigrants in the assembly of the South American

rainforest tree flora. Philosophical transactions of the Royal Society of London. Series B,

Biological sciences, 359, 1611–1622.

Pennington, R.T., Lavin, M. & Oliveira-Filho, A. (2009) Woody Plant Diversity, Evolution, and

Ecology in the Tropics: Perspectives from Seasonally Dry Tropical Forests. Annual Review of

Ecology, Evolution, and Systematics, 40, 437–457.

Pinheiro J, Bates D, DebRoy S, Sarkar D and R Core Team (2016). nlme: Linear and Nonlinear Mixed

Effects Models. R package version 3.1-128. Avaiable at:

http://CRAN.R-project.org/package=nlme.

Qian, H. (2014) Contrasting relationships between clade age and temperature along latitudinal versus

elevational gradients for woody angiosperms in forests of south america. Journal of Vegetation

Science, 25, 1–8.

463

464

465

466

467

468

469

470

471

472

473

474

475

476

477

478

479

480

481

482

483

484

485

486

487

19

Qian, H. & Ricklefs, R. (2016) Out of the Tropical Lowlands: Latitude versus Elevation. Trends in

Ecology & Evolution, 31, 738–741.

Quinn, G.P. & Keough, M.J. (2002) Experimental Design and Data Analysis for Biologists. Cambridge

University Press.

Quiroga, M.P., Mathiasen, P., Iglesias, A., Mill, R.R. & Premoli, A.C. (2015) Molecular and fossil

evidence disentangle the biogeographical history of Podocarpus, a key genus in plant geography.

Journal of Biogeography, 43, 372–383.

R Core Team (2015) R: a language and environment for statistical computing. Version 3.1.0. R

Foundation for Statistical Computing, Vienna. Available at: http://www.Rproject.org/.

Rezende, V.L., Eisenlohr, P.V., Vibrans, A.C., Oliveira-Filho, A.T. (2015) Humidity, low temperature

extremes, and space influence floristic variation across an insightful gradient in the Subtropical

Atlantic Forest. Plant Ecology, 216, 759–774.

Rezende, V.L., Bueno, M.L. & de Oliveira-Filho, A.T. (2016) Patterns of tree composition in the

southern cone of South America and its relevance to the biogeographic regionalization. Plant

Ecology, 217, 97-110.

Ricklefs, R.E. & Schluter, D. (1993) Species diversity: regional and historical influences. Species

Diversity in Ecological Communities: Historical and Geographical Perspectives (ed. by

R.E.Ricklefs and D. Schluter) pp. 350-363.University of Chicago Press, Chicago.

Romdal, T.S., Araújo, M.B. & Rahbek, C. (2013) Life on a tropical planet: Niche conservatism and the

global diversity gradient. Global Ecology and Biogeography, 22, 344–350.

Ruddiman, W. F. (2007) The early anthropogenic hypothesis: Challenges and responses. Reviews of

Geophysics, 45, RG4001.

Segovia, R.A. & Armesto, J.J. (2015) The Gondwanan legacy in South American biogeography.

Journal of Biogeography, 42, 209–217.

Segovia, R., Hinojosa, L.F., Pérez, M.F. & Hawikins, B.A. (2013) Biogeographic anomalies in the

488

489

490

491

492

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

510

511

512

20

species richness of Chilean forests: Incorporating evolution into a climatic - historic scenario.

Austral Ecology, 38, 905–914.

Silvestro,D., Cascales-Miñana,B., Bacon, C.D., Antonelli, A. (2015) Revisiting the origin and

diversification of vascular plants through a comprehensive Bayesian analysis of the fossil record.

New Phytologist, 207, 425–436.

Smale, M.C, Burns, B.R., Smale, P.N. & Whaley, P.T. (1997) Dynamics of upland

podocarp/broadleaved forest on Mamaku Plateau, central North Island, New Zealand. Journal of

the Royal Society of New Zealand, 27, 513–532.

Svenning, J.C., Borchsenius, F., Bjorholm, S. & Balslev, H. (2008). High tropical net diversification

drives the New World latitudinal gradient in palm (Arecaceae) species richness. Journal of

Biogeography, 35, 394–406.

Swenson, N. (2009) Phylogenetic Resolution and Quantifying the Phylogenetic Diversity and

Dispersion of Communities. PLoS ONE, 4, e4390.

Swenson, N. (2014) Phylogenetic and functional ecology in R. Springer, New York City, New York.

Tiede, Y., Homeier, J., Cumbicus, N., Peña, J., Albrecht, J., Ziegenhagen, B., Bendix, J., Brandl, R. &

Farwig, N. (2015) Phylogenetic niche conservatism does not explain elevational patterns of species

richness, phylodiversity and family age of tree assemblages in Andean rainforest. Erdkunde, 70,

83–106.

Turner, J.R.G. (2004) Explaining the global biodiversity gradient: Energy, area, history and natural

selection. Basic and Applied Ecology, 5, 435–448.

Valente, L.M., Savolainen, V. & Vargas, P. (2010) Unparalleled rates of species diversification in

Europe. Proceedings of the Royal Society of London B, 277, 1489–1496.

Villagran, C. & Hinojosa, L.F. (1997) Historia de los bosques del sur de Sudamérica, II: Análisis

fitogeográfico. Revista Chilena de Historia Natural, 70, 241–267.

Webb, C. O., Ackerly, D. D. & Kembel, S. W. (2008) Phylocom: software for the analysis of

513

514

515

516

517

518

519

520

521

522

523

524

525

526

527

528

529

530

531

532

533

534

535

536

537

21

phylogenetic community structure and trait evolution. Bioinformatics, 24, 2098-2100

Weir, J.T. & Schluter, D.(2007) The latitudinal gradient in recent speciation and extinction rates of birds

and mammals. Science, 315, 1574–1576.

Wiens, J.J., Ackerly, D.D., Allen, A.P., Anacker, B.L., Buckley, L.B., Cornell, H.V., Damschen, E.I.,

Davies, T.J., Grytnes, J-A., Harrison, S.P., Hawkins, B.A.,Holt, R.D., McCain, C.M. & Stephens,

P.R. (2010) Niche conservatism as an emerging principle in ecology and conservation biology.

Ecology Letters, 13, 1310–1324.

Wiens, J.J. & Donoghue, M.J. (2004) Historical biogeography, ecology and species richness. Trends in

Ecology and Evolution, 19, 639–644.

Wiens, J.J. & Graham, C.H. (2005) Niche Conservatism: Integrating Evolution, Ecology, and

Conservation Biology. Annual Review of Ecology, Evolution, and Systematics, 36, 519–539.

Whittaker, R.H. (1975) Communities and ecosystems, 2nd ed.MacMillan, New York.

Wilf, P., Escapa, I.H., Cúneo, N.R., Kooyman, R.M., Johnson, K.R. & Iglesias, A. (2014) First South

American Agathis (Araucariaceae), Eocene of Patagonia. American Journal of Botany, 101, 156–

179.

Wilf, P., Little, S.A., Iglesias,A., Zamaloa, M,C., Gandolfo, M.A., Cúneo, N.R. & Johnson, K.R. (2009)

Papuacedrus (Cupressaceae) in Eocene Patagonia: A new fossil link to Australasian rainforests.

American journal of botany, 96, 2031–47.

Wilf, P. (2012) Rainforest conifers of Eocene Patagonia: Attached cones and foliage of the extant

Southeast Asian and Australasian genus Dacrycarpus (Podocarpaceae). American Journal of

Botany, 99, 562–584.

Willis, K.J. & McElwain, J.C. (2002) The evolution of plants. Oxford University Press, Oxford, UK.

Zuloaga, F.O., Morrone, O. & Beltrano, M.J. (2008) Catálogo de las plantas vasculares del Cono Sur.

Monographs in Systematic Botany from the Missouri Botanical Garden, 107, 1–161.

538

539

540

541

542

543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

22

SUPPORTING INFORMATION

Additional Supporting Information may be found in the online version of this article:

Appendix S1 Pairwise correlations between species richness and all phylogenetic diversity metrics.

Appendix S2 MPD and ses.MPD correlations with lineage composition.

Appendix S3 Phylogenetic results for angiosperm.

BIOSKETCH

Vanessa L Rezende is a Brazilian Research Fellow at the University of Minas Gerais, Brazil. She is

interested in subtropical ecology and evolutionary biology of plants, with an emphasis on niche

evolution of southern South America.

Author contributions: V.L.R., K.G.D., R.T.P. and A.O.F. designed the paper; V.L.R., A.O.F. and K.G.D.

assembled the database; V.L.R. and K.G.D. analysed the data; V.L.R., K.G.D., R.T.P. and A.O.F. led

the writing. All authors read and approved the final manuscript.

TABLES

Table 1. Sampling and community composition across biomes, showing proportional representation of

major clades and values of species richness (SR)

Biome No. of

sites

Mean (and

range) for no. of

species per site

Mean(and range)

for no. of

angiosperm

species per site

Mean(and

range) for no. of

gymnosperm

species per site

Mean(and

range) for no. of

tree fern species

per site

Tropical

Semid.* Forest

155 201 (38 – 507) 199 (38-505) 0 2 (0-8)

Seasonally Dry 74 96 (59 – 141) 96 (59-141) 0 0

562

563

564

565

566

567

568

569

570

571

572

573

574

575

576

577

578

23

Tropical Forest

Wet Evergreen

Tropical Forest

95 323 (49 – 746) 318 (49-736) 1 (0-2) 4 (0-12)

Cloud Forest 20 233 (79 – 376) 225 (74-365) 2 (0-3) 6 (0-11)

Subtropical

Semid.* Forest

41 179 (60 – 343) 178 (60-343) 0 1 (0-4)

Araucaria

Mixed Forest

43 210 (102 – 408) 204 (99-399) 2 (0-3) 3 (1-6)

Southern

Andean Forest

37 106 (36 – 226) 104 (36-223) 1 (0-3) 1 (0-4)

Gran Chaco 81 53 (21 – 236) 53 (21-236) 0 0

Western Pampa

Forest

68 75 (23 – 197) 75 (23-196) 0 0

Eastern Pampa

Forest

63 84 (20 – 194) 83 (20-192) 0.3 (0-2) 0.8 (0-4)

Semi-Arid and

Arid Patagonian

Forest

15 34 (9 – 64) 34 (9-64) 0 0

Pacific Forest 50 30 (10 – 46) 28 (10-43) 2 (0-5) 0

*Semid. = Semideciduous

Table 2. delta AIC values for different models to explain variation in species richness and phylogenetic

diversity metrics. All delta AIC values are scaled with respect to the best model for a given metric.

Potential.ET: Potential Evapotranspiration; PrecWetP: Precipitation of wettest month; PrecAnn: Annual

precipitation; TempSeas: Temperature seasonality; TempAnn: Annual mean temperature. TempMin:

Minimum Temperature of coldest month; Isotherm: Isothermality; DaysFrost: Days of frost.

Metrics SR PD MNTD ses.PD ses.MNTD

Biome and Climate 0 0 0 0 0

579

580

581

582

583

584

585

24

Just Biome, no Climate -189 -188 -115 -18 -21

Just Climate, no Biome -218 -262 -83 -107 -93

Climatic variables included TempMin/

Isotherm/

TempSeas/

PrecAnn

TempMin/

Isotherm/

TempSeas/

PrecAnn

TempMin/

TempSeas

DaysFrost TempAnn/

PrecWetP/

PET

LIST OF FIGURE LEGENDS

Figure 1. Tree communities extracted from the NeotropTree database and used in this study. The

symbols correspond to the biomes. Each tree community represents an inventory of non-climbing woody

plants that reach ≥3 m in height and are derived from published and unpublished (grey) literature as well

as herbarium records. See main text for further details.

Figure 2. First two principal components from a principal component analysis of all climatic variables.

Potential.ET: Potential Evapotranspiration; PrecWetP: Precipitation of wettest month; PrecAnn: Annual

precipitation; TempSeas: Temperature seasonality; TempAnnRng: Temperature annual range.

HyperSeas: Seasonality in water availability; TempDayRng: Temperature day range; PrecDryP:

Precipitation of driest month; TempMax: Max temperature of warmest month; WaterDefSev: Water

deficit severity; WaterDefDur: Water deficit duration; TempMin: Minimum Temperature of coldest

month; Isotherm: Isothermality; CloudItcp: Cloud interception; DaysFrost: Days of frost. The

relationship of latitude to these first two principal components is shown with contour (red) lines.

Figure 3. Variation with latitudinal across southern South America for (a) species richness, (b)

phylogenetic diversity sensu stricto, (c) mean nearest taxon distance and (d) phylogenetic diversity

sensu stricto standardised for variation in species richness (ses.PD). Each point represents the estimated

value for a single tree community. See the legend in Figure 1 for explanation of colour codes for biomes.

586

587

588

589

590

591

592

593

594

595

596

597

598

599

600

601

602

603

604

25

Figure 1. Tree communities extracted from the NeotropTree database and used in this study. The

symbols correspond to the biomes. Each tree community represents an inventory of non-climbing woody

plants that reach ≥3 m in height and are derived from published and unpublished (grey) literature as well

as herbarium records. See main text for further details.

605

606

607

608

609

610

611

612

613

26

Figure 2. First two principal components from a principal component analysis of all climatic variables.

Potential.ET: Potential Evapotranspiration; PrecWetP: Precipitation of wettest month; PrecAnn: Annual

precipitation; TempSeas: Temperature seasonality; TempAnnRng: Temperature annual range.

HyperSeas: Seasonality in water availability; TempDayRng: Temperature day range; PrecDryP:

Precipitation of driest month; TempMax: Max temperature of warmest month; WaterDefSev: Water

deficit severity; WaterDefDur: Water deficit duration; TempMin: Minimum Temperature of coldest

month; Isotherm: Isothermality; CloudItcp: Cloud interception; DaysFrost: Days of frost. The

relationship of latitude to these first two principal components is shown with contour (red) lines.

Figure 3. Variation with latitudinal across southern South America for (a) species richness, (b)

phylogenetic diversity sensu stricto, (c) mean nearest taxon distance and (d) phylogenetic diversity

sensu stricto standardised for variation in species richness (ses.PD). Each point represents the estimated

value for a single tree community. See the legend in Figure 1 for explanation of colour codes for biomes.

614

615

616

617

618

619

620

621

622

623

624

625

626

627

628