Interspecific phylogenetic analysis enhances intraspecific phylogeographical inference: a case study in Pinus lambertiana

Share Embed


Descripción

1

Molecular Ecology In press.

2 3

Interspecific phylogenetic analysis enhances intraspecific phylogeographic inference: A

4

case study in Pinus lambertiana

5

AARON LISTON*, MARIAH PARKER-DEFENIKS*, JOHN V. SYRING*,‡, ANN

6

WILLYARD*,§ , and RICHARD CRONN†

7 8

*

9

USA, †Pacific Northwest Research Station, USDA Forest Service, 3200 SW Jefferson Way,

10

Corvallis, Oregon 97331,USA, ‡Current Address: Department of Biological and Physical

11

Sciences, Montana State University – Billings, Billings, Montana 59101, USA, §Current

12

Address:Department of Biology, University of South Dakota, Vermillion, SD 57069 USA

Department of Botany and Plant Pathology, Oregon State University, Corvallis, Oregon 97331,

13 14 15

Keywords: chloroplast introgression, Cronartium ribicola, Pinus lambertiana,Pinus subsect.

16

Strobus, phylogeography, white pine blister rust resistance

17 18 19

Correspondence: A. Liston. Department of Botany and Plant Pathology, 2082 Cordley Hall,

20

Oregon State University, Corvallis, Oregon 97331, USA. Fax: 1 541 737 3573. E-mail:

21

[email protected]

22 23 24

Running title: Phylogeny and phylogeography

1

25

Abstract

26

Pinus lambertiana (sugar pine) is an economically and ecologically important conifer with a

27

1600 km latitudinal range extending from Oregon, USA to northern Baja California, Mexico.

28

Like all North American white pines (subsect. Strobus), sugar pine is highly susceptible to white

29

pine blister rust, a disease caused by the fungus Cronartium ribicola. We conducted a chloroplast

30

DNA survey of Pinus subsect. Strobus with comprehensive geographic sampling of P.

31

lambertiana. Sequence analysis of 12 sugar pine individuals revealed strong geographic

32

differentiation for two chloroplast haplotypes. A diagnostic restriction site survey of an

33

additional 72 individuals demarcated a narrow 150 km contact zone in northeastern California. In

34

the contact zone, maternal (megagametophtye) and paternal (embryo) haplotypes were identified

35

in 31 single seeds, demonstrating bidirectional pollen flow extending beyond the range of

36

maternal haplotypes. The frequencies of the Cr1 allele for white pine blister rust major gene

37

resistance, previously determined for 41 seed zones, differ significantly among seed zones that

38

are fixed for the alternate haplotypes, or contain a mixture of both haplotypes. Interspecific

39

phylogenetic analysis reveals that the northern sugar pine haplotype belongs to a clade that

40

includes P. albicaulis (whitebark pine) and all of the East Asian white pines. Furthermore, there

41

is little cpDNA divergence between northern sugar pine and whitebark pine (dS = 0.00058).

42

These results are consistent with a Pleistocene migration of whitebark pine into North America

43

and subsequent chloroplast introgression from whitebark pine to sugar pine. This study

44

demonstrates the importance of placing phylogeographic results in a broader phylogentic

45

context.

2

46

Introduction

47 48

Pinus has been described as “the most economically and ecologically significant tree genus in

49

the world” (Richardson and Rundel 1998). Support for this claim is found in the large number of

50

population genetic studies conducted in pine species. Ledig (1998) summarized genic diversity

51

statistics (primarily from isozymes) for 51 of the ca. 110 species of pine. Over the last 15 years,

52

at least 26 species of pine have been evaluated for among population DNA variation (tabulated

53

in Petit et al. 2005, see also Chiang et al. 2006; Navascués et al. 2006). The focus of many of

54

these studies is phylogeographic inference using chloroplast DNA. Despite the prevalence of

55

interspecific hybridization in pines (Ledig 1998) few of these studies sample other related

56

species, and thus cannot place the within species genetic variation in a broader phylogenetic

57

context. This study of Pinus lambertiana, sugar pine, demonstrates how resolution of

58

interspecific phylogeny can have a profound impact on the interpretation of intraspecific

59

phylogeographic results. Just as studies can be enhanced by “putting the geography into

60

phylogeography” (Kidd and Ritchie 2006) it is imperative to incorporate a broad phylogenetic

61

perspective as well.

62 63

Pinus lambertiana is one of the ca. 20 species of subsection Strobus (Gernandt et al. 2005;

64

Syring et al. 2007). This clade is known by the common name “white pines” and is distributed

65

discontinuously throughout the Northern Hemisphere (Table 1). The monophyly of subsect.

66

Strobus is strongly supported by chloroplast sequences (Gernandt et al. 2005; Eckert and Hall

67

2006) and nuclear ribosomal DNA (Liston et al. 1999) and moderately supported by a low copy

68

nuclear locus (Syring et al. 2007). The 20 species share a similar vegetative morphology (five

69

relatively narrow and strongly amphistomatic needles per fascicle) but differ dramatically in

70

ovulate cone size (from 5 cm in P. pumila to 60 cm in P. lambertiana) and shape. The five

71

species (P. albicaulis, P. cembra, P. koraiensis, P. pumila, P. sibirica) with indehiscent “closed”

72

cones adapted for bird dispersal were traditionally treated as “subsect. Cembrae”, or stone pines.

73 74

Like all North American members of subsect. Strobus, sugar pine is highly susceptible to white

75

pine blister rust, a disease caused by the heterocyclic rust fungus Cronartium ribicola. This

76

pathogen is native to Asia (Kinloch 2003) and was accidentally introduced to North America in

3

77

the eastern U.S. and British Columbia in the early 20th century (Mielke 1943; Scharpf 1993). It

78

has subsequently spread to all species of subsect. Strobus that occur in the United States and

79

Canada (Scharpf 1993; Kinloch 2003). The disease results in cankers that girdle the main stem

80

and kill infected seedlings and trees (Kinloch and Scheuner 2004). While a mapped locus (Cr1)

81

that confers qualitative (major gene) resistance has been identified in sugar pine (Kinloch 1992,

82

2003; Devey et al. 1995), its frequency in populations is typically low. Cr1 frequency varies

83

from less than 10% in the southern part of the range of P. lambertiana to near absence in the

84

north (Kinloch 1992, summarized in our Table 2) . Attempts to increase the frequency of white

85

pine blister rust resistance have prompted federal (USDA Forest Service, US Bureau of Land

86

Management) and private agencies to make extensive seed collections for this species, and to

87

initiate large-scale screening programs.

88 89

We used DNA sequences of two chloroplast loci to conduct a phylogenetic analysis of North

90

American and Eurasian members of Pinus subsect. Strobus, in concert with a phylogeographic

91

survey of P. lambertiana. Our study documents significant cpDNA divergence between two

92

major haplotypes in P. lambertiana. The distribution of these haplotypes is concordant with the

93

geographic distribution of white pine blister rust major gene resistance. A narrow contact zone

94

between haplotypes, and limited divergence within each haplotype, strongly suggests that

95

secondary contact between two cytoplasmically divergent groups has occurred in the

96

(evolutionarily) recent past. Our phylogenetic analysis provides evidence that the nothern

97

populations of Pinus lambertiana may have obtained their chloroplast via introgression from P.

98

albicaulis, whitebark pine. The integration of phylogenetic and phylogeographic approaches

99

allowed us to recover this unexpected evolutionary history.

100 101 102

Materials and Methods

103

Organismal sampling and DNA genotyping

104

Pinus lambertiana Douglas is an economically and ecologically important conifer with a 1600

105

km latitudinal range extending from Oregon, USA to northern Baja California, Mexico. Eighty-

106

four individuals representing the geographic range of this species were included in this analysis.

107

Nineteen additional species from Pinus subsect. Strobus were sampled for the phylogenetic

4

108

analysis (Table 1), and Pinus gerardiana from subsect. Gerardianae was used as the outgroup.

109

DNA was extracted from the haploid megagametophyte of individual seeds as described in

110

Syring et al. (2007). PCR amplification followed Gernandt et al. (2005). Approximately 90% of

111

the chloroplast matK open reading frame (1404 bp) and ca. 150 bp of the 3' trnK (UUU) intron

112

(see Hausner et al. 2005 for a recent review) were amplified using primers matK1F (Wang et al.

113

1999) and ORF515-900F (Gadek et al. 2000). The chloroplast trnG (UCC) intron (ca. 780 bp)

114

was amplified using the primers 3' trnG and 5' trnG2G (Shaw et al. 2005). For divergence time

115

estimates between P. albicaulis and P. lambertiana (described below), three additional loci were

116

added to the cpDNA data set for these species. These include new sequences for the trnL-trnF

117

intergenic region (including trnL exon 1 and its intron) and the rpl16 intron (Shaw et al. 2005),

118

as well as previously published rbcL sequences (Gernandt et al. 2005). Predicted amplicon

119

lengths were based on the Pinus koraiensis chloroplast genome (Noh et al. unpublished,

120

AY228468). Uncloned PCR products were submitted to High-Throughput Sequencing Solutions

121

(University of Washington) for ExoSAP purification and automated capillary sequencing.

122

Electropherograms were examined and aligned with BioEdit 7.0.5.2 (Hall 1999). All new

123

sequences are deposited in GenBank under the accessions EF546699-EF546759.

124 125

Chloroplast matK and trnG sequences from 12 individuals of P. lambertiana identified two

126

divergent haplotypes (see results) abbreviated N (for North) and S (for South). Using methods

127

described in Liston (1992), we screened 72 individuals for an AluI restriction site that

128

differentiates these two haplotypes. This assay is diagnostic for a C/T polymorphism at position

129

1352 in matK. Since chloroplasts are paternally inherited in Pinus (Neale and Sederoff 1989), the

130

AluI polymorphism was also used to determine maternal versus pollen cpDNA haplotype by

131

extracting DNA from megagametophyte (maternal) and embryo (paternal) tissue in 31 individual

132

seeds.

133 134

Phylogenetic, phylogeographic and statistical analyses

135

Phylogenetic analyses and constraint tests were conducted with PAUP* 4.0b10 (Swofford 2002)

136

following the procedures outlined in Syring et al. (2005). Indels were added to the parsimony

137

data matrix as binary characters. Parsimony analysis was conducted with a heuristic search

138

strategy of 1000 random addition sequences and TBR swapping. Branch support was assessed

5

139

using 2000 bootstrap replicates. Indels and duplicate sequences were excluded, and the most

140

appropriate likelihood model was selected using the method of Posada and Crandall (1998) and

141

AIC scores at the FindModel website (http://hcv.lanl.gov/content/hcv-

142

db/findmodel/findmodel.html). To evaluate whether sequences diverged at clock-like rates,

143

maximum likelihood trees were estimated using the GTR+ γ model as implemented in PAUP*

144

4.0b10 (Swofford 2002). Likelihood scores obtained with and without a molecular clock

145

constraint were evaluated using the likelihood ratio test (LRT) of Muse and Weir (1992). Under

146

assumptions of a molecular clock, the divergence time(Tdiv) between two groups of sequences is

147

approximately Tdiv = dS / 2μ where dS is the average pairwise distance among sequences at

148

presumably neutral (synonymous and non-coding) sites and μ is the neutral mutation rate.

149

Estimates of dS were calculated with DnaSP 4.10.9 (Rozas et al. 2003, note that the program

150

uses the abbreviation Ks). The estimate of μ for Pinus cpDNA (0.22 ×10-9 silent substitutions

151

site-1 year-1; standard error = 0.55 ×10-10) is based on an 85 million years ago (mya) divergence

152

time between the two subgenera of Pinus; see Willyard et al. (2007) for details. Divergence

153

times estimated here are reported with ± one standard error.

154 155

To investigate whether the white pine blister rust major gene resistance allele (Cr1) frequencies

156

were different among chloroplast haplotype classes, 41 tree seed zones used in blister rust

157

screening (Table 2; Kinloch 1992) were classified as either fixed for the N haplotype, fixed for

158

the S haplotype, or polymorphic for the two haplotypes. One-way ANOVA was used to examine

159

variance partitioning and to test the hypothesis that means were equivalent among these three

160

groups. Given the large number of zeros present in estimated Cr1 allele frequencies (especially

161

in the northern part of the range), we also estimated means and confidence intervals for allele

162

frequencies in the N, S, and mixed haplotype groups using 10,000 nonparametric bootstrap

163

resamplings. Statistical and nonparametric bootstraps were performed using PopTools version

164

2.7 (Hood 2006). For mapping, the program TRS2LL (Wefald 2001) was used to convert

165

township/range/section localities to latitude and longitude.

166 167

Results

168

Phylogeographic haplotype variation in sugar pine

6

169

Sequences of cpDNA matK and trnG intron in 12 Pinus lambertiana individuals revealed two

170

predominant haplotypes with fixed differences at 10 sites (seven in matK, including three amino

171

acid replacements; three in the trnG intron). One additional haplotype was observed in the Baja

172

California individual (an autapomorphic matK replacement substitution). The sequence results

173

combined with AluI restriction digest assays for 72 individuals demonstrated an abrupt transition

174

in the distribution of the two predominant haplotypes (Fig. 1A). Plants from Oregon and

175

northwestern California (Klamath mountains and North Coast range) were fixed for a common

176

haplotype “N” (thymine at position 1352 of matK), while plants from the Sierra Nevada and

177

Transverse and Peninsular ranges in California were fixed for the alternate haplotype “S”

178

(cytosine at position 1352). The contact zone between the N and S haplotypes occurs near

179

latitude 40˚ 30' N in northeastern California, and the zone of polymorphism is remarkably well-

180

defined, spanning less than 150 km of the ca. 1600 km latitudinal range of this species (Figs. 1,

181

2). Megagametophyte and embryo comparisons in 31 individuals from the contact zone revealed

182

that 12 (39%) seeds contained different maternal and paternal haplotypes, indicating that

183

seedlings are frequently sired by pollen parents with different haplotypes than the ovulate parent.

184

Embryos with the S haplotype were found in seeds from eight N haplotype maternal trees

185

distributed in seed zones 522, 523, 732 and 771 (Fig. 2). The converse pattern was also observed

186

as four S haplotype maternal plants from seed zones 522 and 731 produced seed containing N

187

haplotype embryos (Fig. 2).

188 189

At least one sugar pine individual was assayed for the S vs. N haplotype in 35 of the 41 seed

190

zones sampled by Kinloch (1992) in his survey of Cr1 allele frequencies (the factor conferring

191

major gene resistance to white pine blister rust; Table 2), with intensive sampling in the contact

192

zone (Fig. 2). The haplotype for six unassayed seed zones (one in northwestern California and

193

five in Oregon) was inferred to be N based on results from adjacent seed zones. The Baja

194

California haplotype S' (differing from S by one substitution, Fig. 3) was grouped with S for this

195

analysis. The N, S, and polymorphic seed zones showed significantly different Cr1 frequencies

196

by one-way ANOVA (F = 32.3, P = 7.6e-9). The N haplotype seed zones had the lowest Cr1

197

allele frequency (0.0032 + 0.0038; n = 23), mixed haplotype seed zones had intermediate Cr1

198

frequencies (0.0380 + 0.0185; n = 3), and S haplotype seed zones had the highest Cr1

7

199

frequencies (0.0484 + 0.0260; n = 15). Nonparametric bootstrapping resulted in the same mean

200

allele frequencies and confidence intervals.

201 202

Phylogenetic relationships and variation in sugar pine and other white pine species. Alignment

203

of matK and trnG intron sequences required one 6 bp indel in the matK ORF and four indels (1,

204

4, 10 and 15 bp) in the trnG intron. Three of the indels were confined to P. parviflora, the 15 bp

205

trnG intron indel was shared by P. ayacahuite, P. flexilis, P. strobiformis and P. peuce, and the 1

206

bp indel was autapomorphic in P. armandii ‘b’. Combined phylogenetic analysis of matK (1545

207

bp), trnG intron (746 bp) and the five indels resulted in two most parsimonious trees of length 57

208

with a consistency index of 0.91 (Fig. 3). The two trees differ in the resolution of P. monticola as

209

a grade (shown) or clade (not shown). All Asian species of subsect. Strobus form a strongly

210

supported clade that includes the North American P. albicaulis and the P. lambertiana N

211

haplotype. At these two loci, no sequence divergence was found among northern P. lambertiana,

212

P. albicaulis and representatives of six Asian species (P. dalatensis, P. koraiensis, P. pumila, P.

213

sibirica, P. wallichiana, P. kwangtungensis) and P. cembra from Romania. The P. lambertiana S

214

haplotypes and the remaining North American white pine species form a grade, with P.

215

monticola in a strongly supported sister position to the Eurasian clade. Constraining P.

216

lambertiana samples to monophyly results in trees of 66 steps, which is significantly longer

217

based on both Templeton (p = 0.0039) and Kishino-Hasegawa tests (p = 0.0027).

218 219

Maximum likelihood trees obtained with and without a molecular clock were topologically

220

identical to one of the most parsimonious trees and were not significantly different from each

221

other based on the LRT (ΔlnL = 10.44013, χ2 = 20.88, d.f. = 19, p = 0.34). This suggests that the

222

sequences are diverging at equivalent rates, and a simple molecular clock calculation can be

223

applied. The mean dS between P. monticola and the Asian clade is 0.00232, resulting in an

224

estimated divergence of 5.3 ± 1.3 mya (Late Miocene – Early Pliocene). Sequences of 5799 bp

225

of cpDNA (adding rbcL, rpl16 and trnL-F to the matK and trnG used in the phylogenetic

226

analysis) revealed a single substitution in trnL-F between a P. albicaulis accession (Washington

227

state) and an N haplotype P. lambertiana (dS = 0.00058), resulting in an estimated divergence of

228

0.6 ± 0.15 mya (Pleistocene). Sequences of trnL-F from an additional five accessions of P.

229

albicaulis and three of N haplotype P. lambertiana (Table 1) confirmed that this is a fixed

8

230

difference between these two taxa. The estimated divergence time between the two P.

231

lambertiana chloroplast haplotypes is 15.5 ± 3.9 mya (dS = 0.00681) and between the S

232

haplotype of P. lambertiana and the other North American white pines is 9.0 ± 2.25 mya (dS =

233

0.00396).

234 235

Discussion

236

The pattern of genetic subdivision in sugar pine

237

Previous accounts of sugar pine have described neither morphological nor ecological differences

238

that can be associated with the phylogeographic pattern observed here. In fact, Mirov (1967)

239

described Pinus lambertiana as “rather stable morphologically” (p. 142) and he considered it to

240

be a prime example of a “good species” that is “clearly delimited [and] can be identified without

241

difficulty” (p. 531). The southern limit of the contact zone (Fig. 1A) does coincide with the

242

interface of the Cascade and Sierra Nevada mountain ranges, characterized by relatively recent

243

volcanics and predominantly metamorphics (with granitic intrusions and volcanics), respectively

244

(Hickman 1993). Although this geological transition is used to demarcate two floristic

245

subregions, there is no apparent vegetational break between the forests of the Cascades and

246

northern Sierra Nevada (Hickman 1993).

247 248

In her PhD thesis, Martinson (1997) conducted an analysis of allozyme data collected by Conkle

249

(unpublished; 1996). Forty populations and 400 individuals of P. lambertiana were assayed at 30

250

allozyme loci. Average heterozygosity was 0.22, and no geographic region showed reduced

251

genetic diversity. Clustering of genetic distances separated populations from Oregon and

252

northwestern California from populations in the Sierra Nevada. Unfortunately, no populations

253

were sampled in the chloroplast haplotype contact zone in northeastern California. Thus,

254

although the reported allozyme differentiation may correspond to the chloroplast haplotype

255

distribution, it will require allozyme sampling in the contact zone to confirm this. The allozyme

256

data also separated the Sierra Nevada populations from those sampled in southern California and

257

Baja California. This division is not apparent in our data set. However, a unique substitution was

258

found in the matK sequence of the Baja California individual. This evidence for genetic isolation

259

is consistent with the geographic isolation of this disjunct population (Fig. 1A, B).

260

9

261

The narrow transition zone between the N and S haplotypes of P. lambertiana suggests that these

262

populations have only recently come into contact. This would be consistent with a Holocene

263

range expansion from separate northern and southern refugia. There is abundant evidence from

264

the pollen record for post-glacial movement of pines (Mohr et al. 2000; Thompson and Anderson

265

2000). Unfortunately, individual Pinus species cannot be identified from pollen. Narrow

266

haplotype (chloroplast or mitochondrial) transition zones observed in other western North

267

American plants (Soltis et al. 1997; Aagaard et al. 1998; Latta and Mitton 1999; Johansen and

268

Latta 2003) have also been attributed to post-glacial contact of previously separated populations.

269 270

The examination of maternal and paternal haplotypes offers insight into the dynamics of

271

dispersal at the contact zone (Fig. 2). The two P. lambertiana trees sampled in the southern part

272

of seed zone 731 represent a population that is isolated on Happy Camp mountain, Modoc

273

County. These trees have the S haplotype, and are presumed to have colonized this location by

274

long distance seed dispersal. The closest sampled potential source is ca. 90 km away. The large

275

(228 ± 40 mg) and “flimsy” winged seeds of P. lambertiana are “seldom dispersed far by wind”,

276

but caching by Steller’s jays (Thayer and Vander Wall 2005) and Clark’s nutcrackers (D.

277

Tomback, pers. comm.) can potentially lead to long-distance dispersal. No other example of a

278

disjunct haplotype was observed. Embryos possessing the S haplotype occur up to 25 km north

279

of the northernmost potential source trees, indicating that pollen flow advances ahead of seed

280

dispersal. The two Happy Camp mountain trees whose megagametophytes carry the S haplotype

281

have apparently been pollinated by trees of the N haplotype. Comparison of chloroplast and

282

mitochondrial haplotypes in a Pinus ponderosa contact zone in western Montana has found a

283

similar pattern of more extensive pollen flow and rare long distance seed dispersal (Latta and

284

Mitton 1993; Johansen and Latta 2003).

285 286

One of the most surprising results of our study is the concordance between the distribution of the

287

two cpDNA haplotypes and the relative frequency of a white pine blister rust major gene

288

resistance locus (Cr1) in sugar pine. Kinloch (1992) described the pattern of Cr1 frequency as a

289

cline. In contrast, the significant differences in Cr1 frequency observed among the three

290

haplotype groups (N, S, and mixed) suggests that the gene frequency does not change in a

291

gradual manner, but rather shows the same abrupt transition as observed in the chloroplast. There

10

292

is no evidence for a causal link between the cpDNA haplotype and Cr1 distribution patterns. It is

293

well-established that resistance shows nuclear inheritance (Kinloch 1992), and the highest

294

frequencies of Cr1 (4 – 9% per seed zone) are far lower than the frequency of the S haplotype.

295

Furthermore, examination of three resistant trees (heterozygous for the dominant Cr1 allele; J.

296

Gleason, unpublished data) in the contact zone found both N and S haplotypes (Fig. 2). Note that

297

all other genotyped individuals in the contact zone are non-resistant (J. Gleason, unpublished

298

data). We predict that the concordance between the Cr1 and chloroplast haplotype frequencies

299

reflects a common history of genetic isolation, followed by recent migration and contact (see

300

below).

301 302

Evidence for chloroplast introgression

303

The chloroplast haplotypes of P. lambertiana resolve in two different clades in the phylogenetic

304

analysis of Pinus subsect. Strobus, one comprised of five other North American species and the

305

other encompassing 12 Eurasian species and the North American P. albicaulis (whitebark pine).

306

Two biological processes could explain these results: incomplete lineage sorting of an ancestral

307

polymorphism, or chloroplast introgression. Incomplete lineage sorting has been determined to

308

be the most probable source of widespread allelic non-monophyly at nuclear loci in species of

309

Pinus subgenus Strobus (Syring et al. 2007). It has also been considered a potential cause of

310

similar patterns observed in chloroplast studies in other plant species (Tsitrone et al. 2003).

311

However, the stochastic process of incomplete lineage sorting is not expected to show the strong

312

geographic partitioning observed for the two chloroplast haplotypes. On the other hand, if the

313

two subgroups of sugar pine have been separated since the Miocene (15.1 ± 3.8 mya) and each

314

retained a different haplotype, one might expect to find morphological divergence between (and

315

sequence variation within) the two groups, particularly since this same interval has apparently

316

been accompanied by multiple speciation events in these lineages. Although some sequence

317

divergence was found in the S haplotype clade (in the geographically isolated Baja California

318

population), none was found in the N haplotype clade. The amount of sequence divergence

319

between the S haplotype of P. lambertiana and the other North American white pines is

320

consistent with genetic isolation since the Miocene. In contrast, the high sequence similarity

321

between the N haplotype of sugar pine and the Asian clade is suggestive of a much more recent

322

shared plastid ancestry.

11

323 324

To account for this unexpected genetic similarity, we suggest that the N haplotype of P.

325

lambertiana may have its origin in a chloroplast introgression event involving P. albicaulis.

326

Introgression-mediated chloroplast transfer has been named “chloroplast capture” (Rieseberg and

327

Soltis 1991; Tsitrone et al. 2003) and has been offered as an explanation for similar patterns of

328

cytonuclear incongruence observed in many plant genera (reviewed in Rieseberg et al. 1996;

329

Wendel and Doyle 1998). Tsitrone et al. (2003) have modeled the process under the assumption

330

of maternal chloroplast inheritance and they determined conditions likely to promote its

331

occurence. A key aspect of their model is the observation that cytonuclear incompatibility often

332

results in full or partial male steritily and thus can increase maternal fitness through enhanced

333

seed production. Although they do not explicitly model paternal inheritance (the situation in

334

Pinaceae), they suggest that chloroplast introgression should be less common here, since

335

cytonuclear interactions typically reduce male fitness. Theoretically, chloroplast substitution

336

could result in an advantage in male function, but this situation has apparently not been

337

documented (Tsitrone et al. 2003). However, a pattern consistent with chloroplast introgression

338

has been observed in other species of Pinaceae, e.g. Pinus montezumae (Matos and Schaal 2000),

339

Pinus muricata (Hong et al. 2003) and Larix sibirica (Wei and Wang 2003).

340 341

Petit et al. (2003) offer “pollen swamping” as an alternative explanation for the lack of

342

cytoplasmic (cpDNA and mtDNA) differentiation between Quercus petraea and Q. robur, two

343

sympatric oaks that are consistently differentiated at nuclear markers. In their scenario, Q. robur

344

seed disperses into new habitats which are subsequently colonized by Q. petraea via pollen flow,

345

resulting in F1 hybrids. Asymmetric introgression and strong selection for the Q. petraea

346

phenotype results in mixed populations that share a single cytoplasm. Petit et al. invoke the fact

347

that the seeds of Q. robur are better adapted to bird-dispersal than Q. petraea, and thus are more

348

likely to establish through long-distance dispersal. A similar relationships exists between P.

349

albicaulis (dispersed mainly by the pine seed specialist, Clark’s nutcracker, Tomback 2005) and

350

P. lambertiana (dispersed to a limited extent by wind and primarily by generalist Steller’s jays

351

and yellow pine chipmunks, Thayer and Vander Wall 2005). In both cases, the better disperser is

352

thought to have contributed its chloroplast to the other species. An important caveat is that the

12

353

cytoplasm is maternally inherited in oaks, and thus carried by the seed, and not the pollen as in

354

pines.

355 356 357

If chloroplast introgression is responsible for this pattern, how does one account for the

358

otherwise “Eurasian” haplotype in two species of North American subsect. Strobus? The

359

cpDNA-based phylogeny (Fig. 3) requires two dispersal events between western North America

360

and Asia. The disjunction between P. monticola and the Eurasian clade can be dated to the late

361

Miocene or early Pliocene. This timing is consistent with estimates of 2.6 – 16.7 mya from

362

eleven other eastern Asian / western North American plant disjunctions (Zhu et al. 2006; Zhang

363

et al. 2007). It is noteworthy that well-preserved Pliocene and Pleistocene fossils of P. monticola

364

have been collected in northeastern Siberia and Alaska (reviewed in Bingham et al. 1974).

365

Following diversification of the Eurasian white pines, and origin of the “closed cone”

366

morphology characteristic of stone pines, we propose that the ancestor of P. albicaulis dispersed

367

from Asia to North America via Beringia, presumably during the Pleistocene. Beringia was

368

largely unglaciated during the Pleistocene, and is known to have served as a refugium for trees

369

and shrubs, including Pinus, through the late glacial maximum (Brubaker et al. 2005).

370

Krutovskii et al. (1995) proposed a similar scenario based on allozyme and chloroplast

371

restriction fragment analysis of the “subsect. Cembrae” pines, but placed the migration at an

372

earlier period (Pliocene).

373 374

Evidence from mtDNA haplotypes has been used to infer the existence of three late Pleistocene

375

refugia for P. albicaulis (Richardson et al. 2002): western Wyoming, western Idaho and the

376

southern Cascades of Oregon. The southern Cascades currently support large populations of P.

377

lambertiana, and we propose that this region could also have served as a northern glacial

378

refugium for sugar pine, or possibly further west in the Klamath/Siskiyou Mts. (a region with

379

several paleo-endemic conifers, e.g. Picea breweriana and Chamaecyparis lawsoniana). Sugar

380

pine is common here, but whitebark pine has only recently been discovered in a small population

381

on Mt. Ashland (nine individuals; Murray 2005). Regardless of their current geographic

382

distributions, sympatry in a glacial refugium could have provided the opportunity for the

383

northern populations of sugar pine to acquire the chloroplast of whitebark pine. Although P.

13

384

albicaulis generally occurs at higher elevations than P. lambertiana, the two are partly sympatric

385

in northern California and southern Oregon (Fig. 1B). There is also evidence that P. lambertiana

386

formerly occurred at higher elevations than its current distribution (May 1974; Millar et al.

387

2006).

388 389

Two factors are required for chloroplast introgression to occur; sympatry and reproductive

390

compatibility. While many interspecific crosses have been conducted in subsect. Strobus, there is

391

no record of attempts to cross P. lambertiana and P. albicaulis (Critchfield 1986, R. Sniezko,

392

pers. comm.). Critchfield and Kinloch (1986) do, however, document interspecific hybridization

393

between P. lambertiana and Asian members of subsect. Strobus, namely P. armandii and P.

394

koraiensis. Seed set in these artificial crosses averaged 2.1% and 0.2% viable seed per cone,

395

respectively. In contrast, P. lambertiana is apparently intersterile with all other North American

396

white pines (Critchfield 1986; Fernando et al. 2005). No naturally occurring hybrids of P.

397

lambertiana have been recorded (Mirov 1967; Critchfield 1986; R. Sniezko, pers. comm.).

398 399

Our results explain why Pinus lambertiana was resolved in conflicting positions in recent

400

chloroplast sequence-based phylogenetic analyses of pines. Gernandt et al. (2005) sampled P.

401

lambertiana in Oregon (a region fixed for the N haplotype), while Eckert and Hall (2006)

402

sampled an individual from southern California (a region fixed for the S haplotype). Each study

403

placed P. lambertiana in a position that is consistent with the resolution of the respective

404

haplotypes in our study (Fig. 3). This demonstrates the importance of sampling multiple

405

individuals per species in phylogenetic analyses of closely related species (see also Syring et al.

406

2007).

407 408

The results reported here provide the first phylogeographic hypothesis for an ecologically and

409

economically important conifer, Pinus lambertiana. By placing the intraspecific results within a

410

broader phylogenetic context, further insights were gained into the evolutionary history of this

411

species. The novel hypotheses of two Pleistocene refugia for P. lambertiana and chloroplast

412

introgression with P. albicaulis can be tested with additional molecular markers, in particular

413

nuclear and mitochondrial loci. The observation that the two chloroplast haplotypes demarcate

414

population groups that differ in their vulnerability to white pine blister rust is also a significant

14

415

result that merits further attention. Beyond sugar pine, this study demonstrates the value of

416

including an interspecific phylogenetic component in phylogeographic research. Without this

417

broader perspective, the antiquity of the haplotype groups would remain unknown, as would the

418

unexpected, and potentially reticulate, history of these species.

419 420

Acknowledgements

421

We are indebted to John Gleason (USFS, Placerville Nursery and Disease Resistance Program),

422

Jerry Berdeen and Richard Sniezko (USFS, Dorena Genetic Resource Center) and David

423

Johnson (USFS, Institute of Forest Genetics) for supplying sugar pine seeds and Roman

424

Businský for providing his collections of Asian species. We thank David Gernandt, Bohun

425

Kinloch, Todd Ott, Paul Severns, Richard Sniezko, and Diana Tomback for constructive

426

comments on the manuscript. This research was funded by National Science Foundation grants

427

DEB 0317103 and ATOL 0629508, an NSF Research Experience for Undergraduates

428

suppplement, and the Pacific Northwest Research Station, USDA Forest Service.

429 430

References

431

Aagaard JE, Krutovskii KV, Strauss SH (1998) RAPD markers of mitochondrial origin exhibit

432

lower population diversity and higher differentiation than RAPDs of nuclear origin in

433

Douglas fir. Molecular Ecology, 7, 801-812.

434 435 436

Bingham RT, Hoff RJ, Steinhoff RJ (1972) Genetics of Western White Pine, USDA Forest Service Research Paper WO-12, USDA Forest Service, Washington, DC. Brubaker LB, Anderson PM, Edwards ME, Lozhkin AV (2005) Beringia as a glacial refugium

437

for boreal trees and shrubs: new perspectives from mapped pollen data. Journal of

438

Biogeography, 32, 833-848.

439

Chiang YC, Hung KH, Schaal BA et al. (2006) Contrasting phylogeographical patterns between

440

mainland and island taxa of the Pinus luchuensis complex. Molecular Ecology, 15, 765-

441

779.

442

Conkle MT (1996) Patterns of variation in isozymes of sugar pine. In: Sugar Pine: Status,

443

Values, and Roles in Ecosystems: Proceedings of a Symposium presented by the

444

California Sugar Pine Management Committee (eds. Kinloch BB, Marosy M, Huddleston

15

445

ME), p. 99. University of California Division of Agriculture and Natural Resources,

446

Davis, CA.

447 448

Critchfield WB (1986) Hybridization and classification of the white pines (Pinus section Strobus). Taxon, 35, 647-656.

449

Critchfield WB, Kinloch BB (1986) Sugar pine and its hybrids. Silvae Genetica, 35, 138-145.

450

Critchfield WB, Little EL, Jr. (1966) Geographic Distribution of the Pines of the World, Forest

451 452

Service Misc. Publication 991, USDA Forest Service, Washington, DC. Devey ME, Delfino-Mix A, Kinloch BB, Neale DB (1995) Random amplified polymorphic

453

DNA markers tightly linked to a gene for resistance to white pine blister rust in sugar

454

pine. Proceedings of the National Academy of Sciences, 92, 2066-2070.

455

Eckert AJ, Hall BD (2006) Phylogeny, historical biogeography, and patterns of diversification

456

for Pinus (Pinaceae): phylogenetic tests of fossil-based hypotheses. Molecular

457

Phylogenetics and Evolution, 40, 166-182.

458

Fernando DD, Long SM, Sniezko RA (2005) Sexual reproduction and crossing barriers in white

459

pines: the case between Pinus lambertiana (sugar pine) and P. monticola (western white

460

pine). Tree Genetics & Genomes, 1, 143-150.

461

Gadek PA, Alpers DL, Heslewood MM, Quinn CJ (2000) Relationships within Cupressaceae

462

sensu lato: a combined morphological and molecular approach. American Journal of

463

Botany, 87, 1044-1057.

464 465 466 467

Gernandt DS, Gaeda López G, Ortiz García S, Liston A (2005) Phylogeny and classification of Pinus. Taxon, 54, 29-42. Hall TA (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series, 41, 95-98.

468

Hausner G, Olson R, Simon D, et al. (2006) Origin and evolution of the chloroplast trnK (matK)

469

intron: A model for evolution of group II intron RNA structures. Molecular Biology and

470

Evolution, 23, 380-391.

471

Hickman JC (1993) Geographic subdivisions of California. In: The Jepson Manual: Higher

472

Plants of California (ed. Hickman JC), pp. 37-48. University of California, Berkeley,

473

CA.

16

474

Hong Y-P, Krupkin AB, Strauss SH (1993) Chloroplast DNA transgresses species boundaries

475

and evolves at variable rates in the California closed-cone pines (P. radiata, P. muricata,

476

and P. attenuata). Molecular Phylogenetics and Evolution, 2, 322-329.

477 478 479 480 481 482 483 484

Hood GM (2006) PopTools version 2.7.5. CSIRO, Canberra, Australia, http://www.cse.csiro.au/poptools/. Johansen AD, Latta RG (2003) Mitochondrial haplotype distribution, seed dispersal and patterns of postglacial expansion of ponderosa pine. Molecular Ecology, 12, 293-298. Kidd DM, Ritchie MG (2006) Phylogeographic information systems: putting the geography into phylogeography. Journal of Biogeography, 33, 1851-1865. Kinloch BB (2003) White pine blister rust in North America: past and prognosis. Phytopathology, 93, 1044-1047.

485

Kinloch BB (1992) Distribution and frequency of a gene for resistance to white pine blister rust

486

in natural populations of sugar pine. Canadian Journal of Botany, 70, 1319-1323.

487

Kinloch BB, Scheuner WH (2004) Pinus lambertiana Dougl., Sugar Pine, USDA Forest Service,

488

Washington, DC.

489

http://na.fs.fed.us/spfo/pubs/silvics_manual/Volume_1/vol1_Table_of_contents.htm.

490

Krutovskii KV, Politov DV, Altukhov YP (1995) Isozyme study of population genetic structure,

491

mating system and phylogenetic relationships of the five stone pine species (subsection

492

Cembrae, section Strobi, subgenus Strobus). In: Population genetics and genetic

493

conservation of forest trees (eds. Baradat P, Adams WT, Müller-Starck G), pp. 279-304.

494

Academic Publishing, Amsterdam.

495 496

Latta RG, Mitton JB (1999) Historical separation and present gene flow through a zone of secondary contact in ponderosa pine. Evolution, 53, 769-776.

497

Ledig FT (1998) Genetic variation in Pinus. In: Ecology and biogeography of Pinus (ed.

498

Richardson DM), pp. 251-280. Cambridge University Press, Cambridge.

499

Liston A (1992) Variation in the chloroplast genes rpoC1 and rpoC2 of the genus Astragalus

500

(Fabaceae): Evidence from restriction site mapping of a PCR-amplified fragment.

501

American Journal of Botany, 79, 953-961.

502

Liston A, Robinson WA, Piñero D, Alvarez-Buylla ER (1999) Phylogenetics of Pinus (Pinaceae)

503

based on nuclear ribosomal DNA internal transcribed spacer sequences. Molecular

504

Phylogenetics and Evolution, 11, 95-109.

17

505

Martinson SR (1997) Association among geographic, allozyme, and growth variables for sugar

506

pine (Pinus lambertiana Dougl.) in southwest Oregon and throughout the species range.

507

PhD thesis, North Carolina State University.

508

Matos JA, Schaal BA (2000) Chloroplast evolution in the Pinus montezumae complex: A

509

coalescent approach to hybridization. Evolution 54, 1218-1233.

510

May RH (1974) An observation of some sugar pine relicts. Madroño, 22, 276.

511

Mielke JL (1943) White Pine Blister Rust in Western North America, Bulletin No. 52, Yale

512 513

University School of Forestry, New Haven. Millar CI, King JC, Westfall RD, Alden HA, Delany DL (2006) Late Holocene forest dynamics,

514

volcanism, and climate change at Whitewing Mountain and San Joaquin Ridge, Mono

515

County, Sierra Nevada, CA, USA. Quaternary Research, 66, 273-287.

516

Mirov NT (1967) The genus Pinus, The Ronald Press Company, New York.

517

Mohr JA, Whitlock C, Skinner CN (2000) Postglacial vegetation and fire history, eastern

518 519 520

Klamath mountains, California, USA. The Holocene, 10, 587-601. Murray M (2005) Our threatened timberlines: the plight of whitebark pine ecosystems. Kalmiopsis, 12, 25-29.

521

Muse SV, Weir BS (1992) Testing for equality of evolutionary rates. Genetics, 132, 269-276.

522

Navascues M, Vaxevanidou Z, Gonzalez-Martinez SC, et al. (2006) Chloroplast microsatellites

523

reveal colonization and metapopulation dynamics in the Canary Island pine. Molecular

524

Ecology, 15, 2691-2698.

525 526 527

Neale DB, Sederoff RR (1989) Paternal inheritance of chloroplast DNA and maternal inheritance of mitochondrial DNA in loblolly pine. Theoretical and Applied Genetics, 77, 212-216. Petit RJ, Duminil J, Fineschi S, et al. (2005) Comparative organization of chloroplast,

528

mitochondrial and nuclear diversity in plant populations. Molecular Ecology, 14, 689-

529

701.

530 531 532 533

Petit RJ, Bodénès C, Ducousso A, Roussel G, Kremer A (2003) Hybridization as a mechanism of invasion in oaks. New Phytologist 161, 151-164. Posada D, Crandall KA (1998) MODELTEST: testing the model of DNA substitution. Bioinformatics, 14, 817-818.

18

534

Richardson BA, Brunsfeld SJ, Klopfenstein NB (2002) DNA from bird-dispersed seed and wind-

535

disseminated pollen provides insights into postglacial colonization and population genetic

536

structure of whitebark pine (Pinus albicaulis). Molecular Ecology, 11, 215-227.

537

Richardson DM, Rundel PW (1998) Ecology and biogeography of Pinus: an introduction. In:

538

Ecology and biogeography of Pinus (ed. Richardson DM), pp. 3-46. Cambridge

539

University Press, Cambridge.

540 541 542 543

Rieseberg LH, Soltis DE (1991) Phylogenetic consequences of cytoplasmic gene flow in plants. Evolutionary Trends in Plants, 5, 65-84. Rieseberg LH, Whitton J, Linder CR (1996) Molecular marker incongruence in plants hybrid zones and phylogenetic trees. Acta Botanica Neerlandica, 45, 243-262.

544

Rozas J, Sánchez-DelBarrio JC, Messeguer X, Rozas R (2003) DnaSP, DNA polymorphism

545

analyses by the coalescent and other methods. Bioinformatics, 19, 2496-2497.

546

Scharpf RF (1993) Diseases of Pacific Coast Conifers, Handbook 521, USDA Forest Service,

547 548

Washington D.C. Shaw J, Lickey EB, Beck JT, et al. (2005) The tortoise and the hare II: relative utility of 21

549

noncoding chloroplast DNA sequences for phylogenetic analysis. American Journal of

550

Botany, 92, 142-166.

551

Soltis DE, Gitzendanner MA, Strenge DD, Soltis PS (1997) Chloroplast DNA intraspecific

552

phylogeography of plants from the Pacific Northwest of North America. Plant

553

Systematics and Evolution, 206, 353-373.

554 555 556 557 558

Swofford DL (2002) PAUP*, Phylogenetic Analysis Using Parsimony (*and Other Methods), Version 4, Sinauer Associates, Sunderland, Massachusetts. Syring J, Farrell K, Businský R, Cronn R, Liston A (2007) Widespread genealogical nonmonophyly in species of Pinus subgenus Strobus. Systematic Biology, 56, 163-181. Syring J, Willyard A, Cronn R, Liston A (2005) Evolutionary relationships among Pinus

559

(Pinaceae) subsections inferred from multiple low-copy nuclear loci. American Journal

560

of Botany, 92, 2086-2100.

561

Thayer TC, Vander Wall SB (2005) Interactions between Steller’s jays and yellow pine

562

chipmunks over scatter-hoarded sugar pine seeds. Journal of Animal Ecology, 74, 364-

563

375.

19

564

Thompson RS, Anderson KH (2000) Biomes of western North America at 18,000, 6000 and 0

565

14

566

27, 555-584.

567

C yr bp reconstructed from pollen and packrat midden data. Journal of Biogeography,

Tomback DF (2005) The impact of seed dispersal by Clark's nutcracker on whitebark pine:

568

multi-scale perspective on a high mountain mutualism. In: Mountain ecosystems: studies

569

in treeline ecology (eds. Broll G, Keplin B), pp. 181-201. Springer, Berlin.

570 571

Tsitrone A, Kirkpatrick M, Levin DA (2003) A model for chloroplast capture. Evolution, 57, 1776-1782.

572

Wang XR, Tsumura Y, Yoshimaru H, Nagasaka K, Szmidt AE (1999) Phylogenetic relationships

573

of Eurasian pines (Pinus, Pinaceae) based on chloroplast rbcL, matK, rpl20-rps18 spacer,

574

and trnV intron sequences. American Journal of Botany, 86, 1742-1753.

575

Wefald M (2001) TRS2LL, http://www.geocities.com/jeremiahobrien/trs2ll.html.

576

Wei XX, Wang XQ (2003) Phylogenetic split of Larix: evidence from paternally inherited

577 578

cpDNA trnT- trnF region. Plant Systematics and Evolution, 239, 67-77. Wendel JF, Doyle JJ (1998) Phylogenetic incongruence: window into genome history and

579

molecular evolution. In: Molecular Systematics of Plants: DNA Sequencing (eds. Soltis

580

DE, Soltis PS, Doyle JJ), pp. 265-296. Kluwer, Boston.

581

Willyard A, Syring J, Gernandt DS, Liston A, Cronn R (2007) Fossil calibration of molecular

582

divergence infers a moderate mutation rate and recent radiations for Pinus. Molecular

583

Biology and Evolution, 24, 90-101.

584

Zhang M-L, Uhink CH, Kadereit JW (2007) Phylogeny and biogeography of

585

Epimedium/Vancouveria (Berberidaceae): Western North American - East Asian

586

disjunctions, the origin of European mountain plant taxa, and East Asian species

587

diversity. Systematic Botany 32, 81-92.

588

Zhu Y-P, Wen J, Zhang Z-Y, Chen Z-D (2006) Evolutionary relationships and diversification of

589

Stachyuraceae based on sequences of four chloroplast markers and the nuclear ribosomal

590

ITS region. Taxon, 55, 931-940.

20

591

Table 1 Geographic origin and Genbank accessions for samples sequenced for the cpDNA matK

592

and trnG intron loci.

593 Species

Haplo-

Country

Administrative Unit

type

matK

trnG

accession

accession

EF546699

EF546730

albicaulis *

U.S.

California: Mono Co.

albicaulis *

U.S.

California: Siskiyou Co.

"

"

*

U.S.

Montana

"

"

albicaulis *

U.S.

Oregon

"

"

albicaulis *, ‡

U.S.

Washington

"

"

albicaulis *

U.S.

Wyoming

"

"

albicaulis

armandii †

a

China

Anhui

EF546700

EF546731

armandii †

b

Taiwan

Kaohsiung

EF546701

"

ayacahuite

Honduras

La Paz

EF546702

EF546732

ayacahuite

Mexico

Mexico

"

"

ayacahuite

Mexico

Michoacan

"

"

bhutanica

India

West Kameng

EF546703

EF546733

cembra

a

Austria

EF546704

EF546734

cembra

b

Switzerland

EF546705

"

cembra

c

Romania

EF546706

"

chiapensis

Guatemala

EF546707

EF546735

chiapensis

Mexico

Chiapas

"

"

chiapensis

Mexico

Guerrero

"

"

dalatensis †

Vietnam

Kon Tum

EF546708

EF546736

flexilis

a

U.S.

California

EF546709

EF546737

flexilis

b

U.S.

Colorado

EF546710

"

flexilis

c

Canada

Alberta

EF546711

"

koraiensis

Russia

EF546712

EF546738

koraiensis

Japan

"

"

koraiensis

S. Korea

"

"

kwangtungensis

China

EF546713

EF546739

21

lambertiana *

N

U.S.

California: seedzone 091

EF546715

EF546741

lambertiana

N

U.S.

California: seedzone 372

"

"

N

U.S.

California: seedzone 516

"

"

N

U.S.

California: seedzone 732

"

"

lambertiana *

N

U.S.

Oregon: seedzone 472

"

"

lambertiana

N

U.S.

Oregon: seedzone 731

"

"

lambertiana

S

U.S.

California: seedzone 120

EF546714

EF546740

lambertiana

S

U.S.

California: seedzone 526

"

"

lambertiana *, ‡

S

U.S.

California: seedzone 731

"

"

lambertiana

S

U.S.

California: seedzone 992

"

"

lambertiana

S

U.S.

California: seedzone 994

"

"

lambertiana

S'

Mexico

Baja California

EF546716

"

monticola

a

U.S.

Oregon

EF546717

EF546742

monticola

b

Canada

British Columbia

EF546718

EF546743

monticola

b

U.S.

California

"

"

EF546719

EF546744

EF546720

EF546745

"

"

EF546721

EF546746

lambertiana * lambertiana

*, ‡

morrisonicola

Taiwan

parviflora

Japan

Hokkaido

parviflora

Japan

Honshu

peuce

Bulgaria

pumila

Japan

Hokkaido

EF546722

EF546747

pumila

Japan

Hokkaido

"

"

pumila

unknown

"

"

sibirica

a

Russia

Krasnoyarsk Krai

EF546723

EF546748

sibirica

b

Russia

Kemorovo

EF546724

"

strobiformis

a

Mexico

Coahuila

EF546725

EF546749

strobiformis

b

Mexico

Durango

EF546726

"

strobiformis

b

U.S.

Texas

"

"

strobus

U.S.

Minnesota

EF546727

EF546750

strobus

Canada

Newfoundland

"

"

strobus

U.S.

North Carolina

"

"

Pakistan

Punjab

EF546728

EF546751

wallichiana †

a

22

wallichiana gerardiana †

b

Nepal

Karnali

EF546729

"

Pakistan

Gilgit

AY115801

EF546752

594 595

Haplotypes N, S and S' for P. lambertiana are described in the text. In other species, the letters a,

596

b and c were applied as needed for multiple haplotypes within a species. Voucher specimens are

597

deposited at Oregon State University (OSC), unless indicated otherwise.

598

*

599

Genbank accessions EF546757 - EF546759.

600



601

Gardening, Průhonice, Czech Republic (RILOG).

Sequenced for trnL-F, GenBank accessions EF546753 - EF546756. ‡ Sequenced for rpl16, Voucher deposited at the Silva Tarouca Research Institute for Landscape and Ornamental

602 603

23

604

Table 2 The frequency of the white pine blister rust resistance gene Cr1 in sugar pine seed zones

605

from the region of haplotype contact, compared to the number of southern and northern

606

haplotypes. Data in columns 2 and 3 are from the 41 seed zones sampled by Kinloch (1992).

607 seed

trees / seeds

zone

sampled

frequency of

haplo-

N

S

frequency

major gene

type

haplo-

haplo-

of S

resistance

group

type

type

haplotype

Oregon: Coast ranges and Cascades 090

10 / 74

0.0000

N

452

N

1

0.00

472

N

1

0.00

1

0.00

491

24 / 430

0.0000

N

492

7 / 231

0.0000

N

501

8 / 237

0.0127

N

502

18 / 379

0.0053

N

511

12 / 204

0.0049

N

1

0.00

512

9 / 205

0.0000

N

1

0.00

681

N

1

0.00

701

N

1

0.00

702

N

1

0.00

N

1

0.00

703

8 / 157

0.0000

721

N

Northwest California: Coast ranges 081

2 / 15

0.0000

N

2

0.00

091

36 / 1763

0.0085

N

2

0.00

095

8 / 136

0.00

N

301

162 / 4752

0.0061

N

1

0.00

302

5 / 137

0.00

N

1

0.00

N

1

0.00

303 311

28 / 683

0.0073

N

1

0.00

321

141 / 5197

0.0073

N

1

0.00

24

322

5 / 302

0.00

N

1

0.00

331

3 / 44

0.00

N

1

0.00

N

1

0.00

332 340

33 / 2030

0.0030

N

1

0.00

371

4 / 190

0.0053

N

1

0.00

372

101 / 4065

0.0096

N

1

0.00

Northeast California: Cascades 516

8 / 265

0.00

N

1

0.00

521

12 / 779

0.0026

N

3

0.00

522

264 / 15815

0.0192

mixed

4

3

0.43

523

42 / 1869

0.0316

mixed

1

3

0.75

mixed

1

2

0.67

4

0.33

731 732

7 / 174

0.0632

mixed

8

741

9 / 538

0.00

N

2

0.00

742

N

4

0.00

771

mixed

1

2

0.67

Eastern California: Sierra Nevada 524

85 / 3141

0.0274

S

3

1.00

525

116 / 5099

0.0322

S

2

1.00

526

222 / 7714

0.0460

S

2

1.00

531

91 / 2976

0.0339

S

1

1.00

532

16 / 941

0.0659

S

1

1.00

533

16 / 635

0.0819

S

1

1.00

534

124 / 2668

0.0727

S

1

1.00

540

44 / 1077

0.0706

S

1

1.00

772

14 / 222

0.0180

S

3

1.00

S

1

1.00

California: Central Coast range 120

93 / 2133

0.0886

Southern California: Transverse ranges 992

8 / 644

0.0699

S

1

1.00

993

29 / 463

0.0324

S

1

1.00

25

994

31 / 337

0.0297

S

1

1.00

997

32 / 517

0.0561

S

1

1.00

S

1

1.00

Baja California: Sierra San Pedro Martír 12 / 254

0.00

608 609

26

610

Fig. 1A Sampled individuals of P. lambertiana in western North America. Yellow squares

611

represent the S haplotype and blue squares represent the N haplotype. Forest tree seed zones

612

sampled by Kinloch (1992) and / or this study are shaded according to the frequency of the Cr1

613

allele (Table 2), ranging from 0% (no shading) to 8.9% (dark gray, seed zone 120). Seed zones

614

outlined in red were not sampled for Cr1. The haplotype contact zone (green rectangle) is shown

615

in more detail in Figure 2.

616 617

Fig. 1B Approximate geographic distribution of Pinus lambertiana (dark gray) and P. albicaulis

618

(light gray) from Critchfield and Little (1966). Red triangles represent P. albicaulis samples used

619

in this study.

620 621

Fig. 2 Detail of the contact zone between the S (yellow) and N (blue) cpDNA haplotypes of P.

622

lambertiana. Squares represent the maternal (megagametophyte) haplotype and circles represent

623

the paternal (embryo) haplotype. Asterisks denote white pine blister rust resistant individuals.

624 625

Fig. 3 One of two most parsimonious trees estimated from cpDNA matK and trnG intron

626

sequences. Bootstrap values are shown below the branches. Tree length = 57, consistency index

627

= 0.91, retention index = 0.97. When applicable, haplotypes (see Table 2) and the number of

628

individuals that share a particular sequence and haplotypes follow the species names.

629 630

Author Information

631 632 633 634 635 636 637

Aaron Liston and Rich Cronn collaboratively study the systematics, population genetics, and evolution of pines. John Syring and Ann Willyard are former PhD students of Cronn and Liston. Dr. Syring is now an Assistant Professor of Plant Systematics at Montana State UniversityBillings, and Dr. Willyard is a post-doc at the University of South Dakota. Mariah ParkerdeFeniks is a Sociology major at Oregon State University, pursuing a career in Criminal Justice and Social Research.

27

638

Figure 1

639

28

640

Figure 2

641

29

642

Figure 3

643

30

Lihat lebih banyak...

Comentarios

Copyright © 2017 DATOSPDF Inc.