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

advertisement
Molecular Ecology (2007) 16, 3926– 3937
doi: 10.1111/j.1365-294X.2007.03461.x
Interspecific phylogenetic analysis enhances intraspecific
phylogeographical inference: a case study in Pinus
lambertiana
Blackwell Publishing Ltd
A A R O N L I S T O N ,* M A R I A H P A R K E R - D E F E N I K S ,* J O H N V . S Y R I N G ,*‡ A N N W I L L Y A R D ,*§
and R I C H A R D C R O N N †
*Department of Botany and Plant Pathology, 2082 Cordley Hall, Oregon State University, Corvallis, Oregon 97331, USA, †Pacific Northwest Research Station, USDA Forest Service, 3200 SW Jefferson Way, Corvallis, Oregon 97331, USA
Abstract
Pinus lambertiana (sugar pine) is an economically and ecologically important conifer with
a 1600-km latitudinal range extending from Oregon, USA, to northern Baja California,
Mexico. Like all North American white pines (subsect. Strobus), sugar pine is highly
susceptible to white pine blister rust, a disease caused by the fungus Cronartium ribicola.
We conducted a chloroplast DNA (cpDNA) survey of Pinus subsect. Strobus with com­
prehensive geographical sampling of P. lambertiana. Sequence analysis of 12 sugar pine
individuals revealed strong geographical differentiation for two chloroplast haplotypes.
A diagnostic restriction site survey of an additional 72 individuals demarcated a narrow
150-km contact zone in northeastern California. In the contact zone, maternal (megagame­
tophtye) and paternal (embryo) haplotypes were identified in 31 single seeds, demonstrating
bidirectional pollen flow extending beyond the range of maternal haplotypes. The
frequencies of the Cr1 allele for white pine blister rust major gene resistance, previously
determined for 41 seed zones, differ significantly among seed zones that are fixed for the
alternate haplotypes, or contain a mixture of both haplotypes. Interspecific phylogenetic
analysis reveals that the northern sugar pine haplotype belongs to a clade that includes
Pinus albicaulis (whitebark pine) and all of the East Asian white pines. Furthermore, there
is little cpDNA divergence between northern sugar pine and whitebark pine (dS = 0.00058).
These results are consistent with a Pleistocene migration of whitebark pine into North
America and subsequent chloroplast introgression from whitebark pine to sugar pine.
This study demonstrates the importance of placing phylogeographical results in a broader
phylogenetic context.
Keywords: chloroplast introgression, Cronartium ribicola, phylogeography, Pinus lambertiana, Pinus
subsect. Strobus, white pine blister rust resistance
Received 13 January 2007; revision received 17 April 2007; accepted 11 June 2007
Introduction
Pinus has been described as ‘the most economically and
ecologically significant tree genus in the world’ (Richardson
& Rundel 1998). Support for this claim is found in the large
Correspondence: A. Liston, Fax: 1 541 737 3573.
E-mail: listona@science.oregonstate.edu
§Present address: Department of Biology, University of South
Dakota, Vermillion, SD 57069, USA
‡Present address: Department of Biological and Physical Sciences,
Montana State University–Billings, Billings, MT 59101, USA
number of population genetic studies conducted in pine
species. Ledig (1998) summarized genic diversity statistics
(primarily from isozymes) for 51 of the c. 110 species of
pine. Over the last 15 years, at least 26 species of pine have
been evaluated for among-population DNA variation
(tabulated in Petit et al. 2005; see also Chiang et al. 2006;
Navascues et al. 2006). The focus of many of these studies
is phylogeographical inference using chloroplast DNA
(cpDNA). Despite the prevalence of interspecific
hybridization in pines (Ledig 1998), few of these studies
sample other related species, and thus cannot place the
within-species genetic variation in a broader phylogenetic
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
P H Y L O G E N Y A N D P H Y L O G E O G R A P H Y 3927
context. This study of Pinus lambertiana, sugar pine,
demonstrates how resolution of interspecific phylogeny can
have a profound impact on the interpretation of intraspecific
phylogeographical results. Just as studies can be enhanced
by ‘putting the geography into phylogeography’ (Kidd &
Ritchie 2006), it is imperative to incorporate a broad
phylogenetic perspective as well.
Pinus lambertiana is one of the c. 20 species of subsection
Strobus (Gernandt et al. 2005; Syring et al. 2007). This clade is
known by the common name ‘white pines’ and is distributed
discontinuously throughout the Northern Hemisphere
(Table 1). The monophyly of subsect. Strobus is strongly
supported by chloroplast sequences (Gernandt et al. 2005;
Eckert & Hall 2006) and nuclear ribosomal DNA (Liston
et al. 1999) and moderately supported by a low copy
nuclear locus (Syring et al. 2007). The 20 species share a
similar vegetative morphology (five relatively narrow and
strongly amphistomatic needles per fascicle) but differ
dramatically in ovulate cone size (from 5 cm in Pinus pumila
to 60 cm in P. lambertiana) and shape. The five Pinus species
(P. albicaulis, P. cembra, P. koraiensis, P. pumila, P. sibirica) with
indehiscent ‘closed’ cones adapted for bird dispersal were
traditionally treated as ‘subsect. Cembrae’, or stone pines.
Like all North American members of subsect. Strobus,
sugar pine is highly susceptible to white pine blister rust, a
disease caused by the heterocyclic rust fungus Cronartium
ribicola. This pathogen is native to Asia (Kinloch 2003) and
was accidentally introduced to North America in the eastern
United States and British Columbia in the early 20th
century (Mielke 1943; Scharpf 1993). It has subsequently
spread to all species of subsect. Strobus that occur in the
United States and Canada (Scharpf 1993; Kinloch 2003).
The disease results in cankers that girdle the main stem
and kill infected seedlings and trees (Kinloch & Scheuner
2004). While a mapped locus (Cr1) that confers qualitative
(major gene) resistance has been identified in sugar pine
(Kinloch 1992, 2003; Devey et al. 1995), its frequency in
populations is typically low. Cr1 frequency varies from less
than 10% in the southern part of the range of P. lambertiana
to near absence in the north (Kinloch 1992; summarized in
our Table 2). Attempts to increase the frequency of white
pine blister rust resistance have prompted federal (USDA
Forest Service, US Bureau of Land Management) and private
agencies to make extensive seed collections for this species,
and to initiate large-scale screening programmes.
We used DNA sequences of two chloroplast loci to
conduct a phylogenetic analysis of North American and
Eurasian members of Pinus subsect. Strobus, in concert
with a phylogeographical survey of P. lambertiana. Our
study documents significant cpDNA divergence between
two major haplotypes in P. lambertiana. The distribution of
these haplotypes is concordant with the geographical
distribution of white pine blister rust major gene resistance.
A narrow contact zone between haplotypes, and limited
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
divergence within each haplotype, strongly suggests that
secondary contact between two cytoplasmically divergent
groups has occurred in the (evolutionarily) recent past.
Our phylogenetic analysis provides evidence that the
northern populations of P. lambertiana may have obtained
their chloroplast via introgression from P. albicaulis,
whitebark pine. The integration of phylogenetic and phyl­
ogeographical approaches allowed us to recover this unex­
pected evolutionary history.
Materials and methods
Organismal sampling and DNA genotyping
Pinus lambertiana Douglas is an economically and eco­
logically important conifer with a 1600-km latitudinal
range extending from Oregon, USA to northern Baja
California, Mexico. Eighty-four individuals representing
the geographical range of this species were included in this
analysis. Nineteen additional species from Pinus subsect.
Strobus were sampled for the phylogenetic analysis
(Table 1), and Pinus gerardiana from subsect. Gerardianae
was used as the outgroup. DNA was extracted from the
haploid megagametophyte of individual seeds as described
in Syring et al. (2007). Polymerase chain reaction (PCR)
amplification followed Gernandt et al. (2005). Approximately
90% of the chloroplast matK open reading frame (1404 bp)
and c. 150 bp of the 3′ trnK (UUU) intron (see Hausner et al.
2006 for a recent review) were amplified using primers
matK1F (Wang et al. 1999) and ORF515–900F (Gadek et al.
2000). The chloroplast trnG (UCC) intron (c. 780 bp) was
amplified using the primers 3′ trnG and 5′ trnG2G (Shaw
et al. 2005). For divergence time estimates between Pinus
albicaulis and P. lambertiana (described below), three
additional loci were added to the cpDNA data set for these
species. These include new sequences for the trnL-trnF
intergenic region (including trnL exon 1 and its intron) and
the rpl16 intron (Shaw et al. 2005), as well as previously
published rbcL sequences (Gernandt et al. 2005). Predicted
amplicon lengths were based on the Pinus koraiensis
chloroplast genome (EW Noh, JS Lee, YI Choi, MS Han, YS
Yi, and SU Han, unpublished, AY228468). Uncloned PCR
products were submitted to High-Throughput Sequencing
Solutions (University of Washington) for ExoSAP purification
and automated capillary sequencing. Electropherograms
were examined and aligned with bioedit 7.0.5.2 (Hall
1999). All new sequences are deposited in GenBank under
the Accessions nos EF546699–EF546759.
Chloroplast matK and trnG sequences from 12 individuals
of P. lambertiana identified two divergent haplotypes
(see results) abbreviated N (for North) and S (for South).
Using methods described in Liston (1992), we screened 72
individuals for an AluI restriction site that differentiates
these two haplotypes. This assay is diagnostic for a C/T
3928 A . L I S T O N E T A L .
Table 1 Geographical origin and GenBank Accessions for samples sequenced for the cpDNA matK and trnG intron loci
Pinus species
albicaulis*
albicaulis*
albicaulis*
albicaulis*
albicaulis*‡
albicaulis*
armandii†
armandii†
ayacahuite
ayacahuite
ayacahuite
bhutanica
cembra
cembra
cembra
chiapensis
chiapensis
chiapensis
dalatensis†
flexilis
flexilis
flexilis
koraiensis
koraiensis
koraiensis
kwangtungensis
lambertiana*
lambertiana
lambertiana*
lambertiana*‡
lambertiana*
lambertiana
lambertiana
lambertiana
lambertiana*‡
lambertiana
lambertiana
lambertiana
monticola
monticola
monticola
morrisonicola
parviflora
parviflora
peuce
pumila
pumila
pumila
sibirica
sibirica
strobiformis
strobiformis
strobiformis
strobus
strobus
strobus
wallichiana†
wallichiana
gerardiana†
Haplotype
a
b
a
b
c
a
b
c
N
N
N
N
N
N
S
S
S
S
S
S′
a
b
b
a
b
a
b
b
a
b
Country
Administrative unit
matK accession
trnG accession
USA
USA
USA
USA
USA
USA
China
Taiwan
Honduras
Mexico
Mexico
India
Austria
Switzerland
Romania
Guatemala
Mexico
Mexico
Vietnam
USA
USA
Canada
Russia
Japan
South Korea
China
USA
USA
USA
USA
USA
USA
USA
USA
USA
USA
USA
Mexico
USA
Canada
USA
Taiwan
Japan
Japan
Bulgaria
Japan
Japan
unknown
Russia
Russia
Mexico
Mexico
USA
USA
Canada
USA
Pakistan
Nepal
Pakistan
California: Mono Co.
California: Siskiyou Co.
Montana
Oregon
Washington
Wyoming
Anhui
Kaohsiung
La Paz
Mexico
Michoacan
West Kameng
EF546699
″
″
″
″
″
EF546700
EF546701
EF546702
″
″
EF546703
EF546704
EF546705
EF546706
EF546707
″
″
EF546708
EF546709
EF546710
EF546711
EF546712
″
″
EF546713
EF546715
″
″
″
″
″
EF546714
″
″
″
″
EF546716
EF546717
EF546718
″
EF546719
EF546720
″
EF546721
EF546722
″
″
EF546723
EF546724
EF546725
EF546726
″
EF546727
″
″
EF546728
EF546729
AY115801
EF546730
″
″
″
″
″
EF546731
″
EF546732
″
″
EF546733
EF546734
″
″
EF546735
″
″
EF546736
EF546737
″
″
EF546738
″
″
EF546739
EF546741
″
″
″
″
″
EF546740
″
″
″
″
″
EF546742
EF546743
″
EF546744
EF546745
″
EF546746
EF546747
″
″
EF546748
″
EF546749
″
″
EF546750
″
″
EF546751
″
EF546752
Chiapas
Guerrero
Kon Tum
California
Colorado
Alberta
California: seed zone 091
California: seed zone 372
California: seed zone 516
California: seed zone 732
Oregon: seed zone 472
Oregon: seed zone 731
California: seed zone 120
California: seed zone 526
California: seed zone 731
California: seed zone 992
California: seed zone 994
Baja California
Oregon
British Columbia
California
Hokkaido
Honshu
Hokkaido
Hokkaido
Krasnoyarsk Krai
Kemorovo
Coahuila
Durango
Texas
Minnesota
Newfoundland
North Carolina
Punjab
Karnali
Gilgit
Haplotypes N, S and S’ for P. lambertiana are described in the text. In other species, the letters a, b and c were applied as needed for multiple
haplotypes within a species. Voucher specimens are deposited at Oregon State University (OSC), unless indicated otherwise; *sequenced
for trnL-F, GenBank Accession nos EF546753 –EF546756; †Voucher deposited at the Silva Tarouca Research Institute for Landscape and
Ornamental Gardening, Prùhonice, Czech Republic (RILOG); ‡ sequenced for rpl16, GenBank Accession nos EF546757–EF546759.
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
P H Y L O G E N Y A N D P H Y L O G E O G R A P H Y 3929
Table 2 The frequency of the white pine blister rust resistance gene Cr1 in sugar pine seed zones, compared to the number of southern and
northern haplotypes. Data in columns 2 and 3 are from the 41 seed zones sampled by Kinloch (1992)
Seed zone
Trees/seeds
sampled
Frequency of major
gene resistance
Oregon: coast ranges and Cascades
090
10/74
0.0000
452
472
491
24/430
0.0000
492
7/231
0.0000
501
8/237
0.0127
502
18/379
0.0053
511
12/204
0.0049
512
9/205
0.0000
681
701
702
703
8/157
0.0000
721
Northwest California: coast ranges
081
2/15
0.0000
091
36/1763
0.0085
095
8/136
0.00
301
162/4752
0.0061
302
5/137
0.00
303
311
28/683
0.0073
321
141/5197
0.0073
322
5/302
0.00
331
3/44
0.00
332
340
33/2030
0.0030
371
4/190
0.0053
372
101/4065
0.0096
Northeast California: Cascades
516
8/265
0.00
521
12/779
0.0026
522
264/15815
0.0192
523
42/1869
0.0316
731
732
7/174
0.0632
741
9/538
0.00
742
771
Eastern California: Sierra Nevada
524
85/3141
0.0274
525
116/5099
0.0322
526
222/7714
0.0460
531
91/2976
0.0339
532
16/941
0.0659
533
16/635
0.0819
534
124/2668
0.0727
540
44/1077
0.0706
772
14/222
0.0180
California: central coast range
120
93/2133
0.0886
Southern California: transverse ranges
992
8/644
0.0699
993
29/463
0.0324
994
31/337
0.0297
997
32/517
0.0561
Baja California: Sierra San Pedro Martír
12/254
0.00
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
Haplotype
group
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N haplotype
S haplotype
Frequency of
S haplotype
1
1
1
0.00
0.00
0.00
1
1
1
1
1
0.00
0.00
0.00
0.00
0.00
1
0.00
N
N
N
N
N
N
N
N
N
N
N
N
N
N
2
2
0.00
0.00
1
1
1
1
1
1
1
1
1
1
1
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
N
N
Mixed
Mixed
Mixed
Mixed
N
N
Mixed
1
3
4
1
1
8
2
4
1
2
0.00
0.00
0.43
0.75
0.67
0.33
0.00
0.00
0.67
S
S
S
S
S
S
S
S
S
3
2
2
1
1
1
1
1
3
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
S
1
1.00
S
S
S
S
1
1
1
1
1.00
1.00
1.00
1.00
S
1
3
3
2
4
1.00
3930 A . L I S T O N E T A L .
polymorphism at position 1352 in matK. Since chloroplasts
are paternally inherited in Pinus (Neale & Sederoff 1989),
the AluI polymorphism was also used to determine mater­
nal vs. pollen cpDNA haplotype by extracting DNA from
megagametophyte (maternal) and embryo (paternal)
tissue in 31 individual seeds.
nonparametric bootstraps were performed using poptools
version 2.7 (Hood 2006). For mapping, the program trs2ll
(Wefald 2001) was used to convert township/range/section
localities to latitude and longitude.
Phylogenetic, phylogeographical and statistical analyses
Phylogeographical haplotype variation in sugar pine
Phylogenetic analyses and constraint tests were conducted
with paup* 4.0b10 (Swofford 2002) following the procedures
outlined in Syring et al. (2005). Indels were added to the
parsimony data matrix as binary characters. Parsimony
analysis was conducted with a heuristic search strategy
of 1000 random addition sequences and tree-bisection–
reconnection swapping. Branch support was assessed
using 2000 bootstrap replicates. Indels and duplicate
sequences were excluded, and the most appropriate
likelihood model was selected using the method of
Posada & Crandall (1998) and AIC scores at the findmodel
website (http://hcv.lanl.gov/content/hcv-db/findmodel/
findmodel.html). To evaluate whether sequences diverged
at clock-like rates, maximum-likelihood trees were estimated
using the GTR + γ model as implemented in paup* 4.0b10
(Swofford 2002). Likelihood scores obtained with and
without a molecular clock constraint were evaluated using
the likelihood ratio test (LRT) of Muse & Weir (1992).
Under assumptions of a molecular clock, the divergence
time (Tdiv) between two groups of sequences is
approximately Tdiv = dS/2µ, where dS is the average
pairwise distance among sequences at presumably neutral
(synonymous and noncoding) sites and µ is the neutral
mutation rate. Estimates of dS were calculated with dnasp
4.10.9 (Rozas et al. 2003; note that the program uses the
abbreviation Ks). The estimate of µ for Pinus cpDNA
(0.22 × 10–9 silent substitutions site–1 years–1; standard
error = 0.55 × 10–10) is based on a divergence time of 85
million years between the two subgenera of Pinus; see
Willyard et al. (2007) for details. Divergence times estimated
here are reported with ± one standard error.
To investigate whether the white pine blister rust major
gene resistance allele (Cr1) frequencies were different
among chloroplast haplotype classes, 41 tree seed zones
used in blister rust screening (Table 2; Kinloch 1992) were
classified as either fixed for the N haplotype, fixed for the
S haplotype, or polymorphic for the two haplotypes. Oneway analysis of variance (anova) was used to examine
variance partitioning and to test the hypothesis that means
were equivalent among these three groups. Given the large
number of zeros present in estimated Cr1 allele frequencies
(especially in the northern part of the range), we also esti­
mated means and confidence intervals for allele frequencies
in the N, S, and mixed haplotype groups using 10 000
nonparametric bootstrap resamplings. Statistical and
Sequences of cpDNA matK and trnG intron in 12 Pinus
lambertiana individuals revealed two predominant
haplotypes with fixed differences at 10 sites (seven in matK,
including three amino acid replacements; three in the trnG
intron). One additional haplotype was observed in the Baja
California individual (an autapomorphic matK replacement
substitution). The sequence results combined with AluI
restriction digest assays for 72 individuals demonstrated
an abrupt transition in the distribution of the two
predominant haplotypes (Fig. 1A). Plants from Oregon
and northwestern California (Klamath mountains and
North Coast range) were fixed for a common haplotype ‘N’
(thymine at position 1352 of matK), while plants from
the Sierra Nevada and Transverse and Peninsular ranges
in California were fixed for the alternate haplotype ‘S’
(cytosine at position 1352). The contact zone between the
N and S haplotypes occurs near latitude 40°30′N in
northeastern California, and the zone of polymorphism is
remarkably well-defined, spanning less than 150 km of the
c. 1600-km latitudinal range of this species (Figs 1 and 2).
Megagametophyte and embryo comparisons in 31
individuals from the contact zone revealed that 12 (39%)
seeds contained different maternal and paternal haplotypes,
indicating that seedlings are frequently sired by pollen
parents with different haplotypes than the ovulate parent.
Embryos with the S haplotype were found in seeds from
eight N haplotype maternal trees distributed in seed zones
522, 523, 732 and 771 (Fig. 2). The converse pattern was also
observed as four S haplotype maternal plants from seed
zones 522 and 731 produced seed containing N haplotype
embryos (Fig. 2).
At least one sugar pine individual was assayed for the S
vs. N haplotype in 35 of the 41 seed zones sampled by
Kinloch (1992) in his survey of Cr1 allele frequencies (the
factor conferring major gene resistance to white pine
blister rust; Table 2), with intensive sampling in the contact
zone (Fig. 2). The haplotype for six unassayed seed zones
(one in northwestern California and five in Oregon) was
inferred to be N based on results from adjacent seed zones.
The Baja California haplotype S′ (differing from S by one
substitution, Fig. 3) was grouped with S for this analysis.
The N, S, and polymorphic seed zones showed significantly
different Cr1 frequencies by one-way anova (F = 32.3,
P = 7.6 × 10 – 9). The N haplotype seed zones had the lowest
Cr1 allele frequency (0.0032 ± 0.0038; n = 23), mixed
Results
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
P H Y L O G E N Y A N D P H Y L O G E O G R A P H Y 3931
Fig. 1 (A) Sampled individuals of Pinus
lambertiana in western North America.
Yellow squares represent the S haplotype
and blue squares represent the N haplotype.
Forest tree seed zones sampled by Kinloch
(1992) and/or this study are shaded
according to the frequency of the Cr1 allele
(Table 2), ranging from 0% (no shading) to
8.9% (dark grey, seed zone 120). Seed zones
outlined in red were not sampled for Cr1.
The haplotype contact zone (green rectangle)
is shown in more detail in Fig. 2. (B)
Approximate geographical distribution of
P. lambertiana (dark grey) and P. albicaulis
(light grey) from Critchfield & Little (1966).
Red triangles represent P. albicaulis samples
used in this study.
haplotype seed zones had intermediate Cr1 frequencies
(0.0380 ± 0.0185; n = 3), and S haplotype seed zones had
the highest Cr1 frequencies (0.0484 ± 0.0260; n = 15).
Nonparametric bootstrapping resulted in the same mean
allele frequencies and confidence intervals.
Phylogenetic relationships and variation in sugar pine and other
white pine species. Alignment of matK and trnG intron
sequences required one 6-bp indel in the matK ORF and
four indels (1, 4, 10 and 15 bp) in the trnG intron. Three of
the indels were confined to Pinus parviflora, the 15-bp
trnG intron indel was shared by P. ayacahuite, P. flexilis,
P. strobiformis and P. peuce, and the 1-bp indel was
autapomorphic in P. armandii ‘b’. Combined phylogenetic
analysis of matK (1545 bp), trnG intron (746 bp) and the five
indels resulted in two most parsimonious trees of length 57
with a consistency index of 0.91 (Fig. 3). The two trees
differ in the resolution of Pinus monticola as a grade
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
(shown) or clade (not shown). All Asian species of subsect.
Strobus form a strongly supported clade that includes the
North American Pinus albicaulis and the P. lambertiana
N haplotype. At these two loci, no sequence divergence
was found among northern P. lambertiana, P. albicaulis and
representatives of six Asian Pinus species (P. dalatensis,
P. koraiensis, P. pumila, P. sibirica, P. wallichiana, P. kwan­
gtungensis) and P. cembra from Romania. The P. lambertiana
S haplotypes and the remaining North American white
pine species form a grade, with P. monticola in a strongly
supported sister position to the Eurasian clade.
Constraining P. lambertiana samples to monophyly results
in trees of 66 steps, which is significantly longer based on
both Templeton (P = 0.0039) and Kishino –Hasegawa tests
(P = 0.0027).
Maximum-likelihood trees obtained with and without
a molecular clock were topologically identical to one of
the most parsimonious trees and were not significantly
3932 A . L I S T O N E T A L .
Fig. 2 Detail of the contact zone between
the S (yellow) and N (blue) cpDNA
haplotypes of Pinus lambertiana. Squares
represent the maternal (megagametophyte)
haplotype and circles represent the pater­
nal (embryo) haplotype. Asterisks denote
white pine blister rust resistant individuals.
different from each other based on the LRT (∆ln L =
10.44013, χ2 = 20.88, d.f. = 19, P = 0.34). This suggests that the
sequences are diverging at equivalent rates, and a simple
molecular clock calculation can be applied. The mean dS
between P. monticola and the Asian clade is 0.00232, resulting
in an estimated divergence of 5.3 ± 1.3 million years ago
(Ma; Late Miocene–Early Pliocene). Sequences of 5799 bp
of cpDNA (adding rbcL, rpl16 and trnL-F to the matK and
trnG used in the phylogenetic analysis) revealed a single
substitution in trnL-F between a P. albicaulis accession
(Washington state) and an N haplotype P. lambertiana
(dS = 0.00058), resulting in an estimated divergence of
0.6 ± 0.15 Ma (Pleistocene). Sequences of trnL-F from an
additional five accessions of P. albicaulis and three of N
haplotype P. lambertiana (Table 1) confirmed that this is a
fixed difference between these two taxa. The estimated
divergence time between the two P. lambertiana chloroplast
haplotypes is 15.5 ± 3.9 Ma (dS = 0.00681) and between the
S haplotype of P. lambertiana and the other North American
white pines is 9.0 ± 2.25 Ma (dS = 0.00396).
Discussion
The pattern of genetic subdivision in sugar pine
Previous accounts of sugar pine have described neither
morphological nor ecological differences that can be
associated with the phylogeographical pattern observed
here. In fact, Mirov (1967) described Pinus lambertiana as
‘rather stable morphologically’ (p. 142) and he considered
it to be a prime example of a ‘good species’ that is ‘clearly
delimited [and] can be identified without difficulty’
(p. 531). The southern limit of the contact zone (Fig. 1A)
does coincide with the interface of the Cascade and Sierra
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
P H Y L O G E N Y A N D P H Y L O G E O G R A P H Y 3933
Fig. 3 One of two most parsimonious trees estimated from
cpDNA matK and trnG intron sequences. Bootstrap values are
shown below the branches. Tree length = 57, consistency
index = 0.91, retention index = 0.97. When applicable, haplotypes
(see Table 2) and the number of individuals that share a particular
sequence and haplotypes follow the species names.
Nevada mountain ranges, characterized by relatively recent
volcanic activities and predominantly metamorphics (with
granitic intrusions and volcanic activities), respectively
(Hickman 1993). Although this geological transition is
used to demarcate two floristic subregions, there is no
apparent vegetational break between the forests of the
Cascades and northern Sierra Nevada (Hickman 1993).
In her PhD thesis, Martinson (1997) conducted an analysis
of allozyme data collected by Conkle (1996, abstract only).
Forty populations and 400 individuals of P. lambertiana
were assayed at 30 allozyme loci. Average heterozygosity
was 0.22, and no geographical region showed reduced
genetic diversity. Clustering of genetic distances separated
populations from Oregon and northwestern California
from populations in the Sierra Nevada. Unfortunately, no
populations were sampled in the chloroplast haplotype
contact zone in northeastern California. Thus, although
the reported allozyme differentiation may correspond to
the chloroplast haplotype distribution, it will require
allozyme sampling in the contact zone to confirm this.
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
The allozyme data also separated the Sierra Nevada
populations from those sampled in southern California and
Baja California. This division is not apparent in our data
set. However, a unique substitution was found in the matK
sequence of the Baja California individual. This evidence
for genetic isolation is consistent with the geographical
isolation of this disjunct population (Fig. 1A,B).
The narrow transition zone between the N and S haplo­
types of P. lambertiana suggests that these populations have
only recently come into contact. This would be consistent
with a Holocene range expansion from separate northern
and southern refugia. There is abundant evidence from the
pollen record for postglacial movement of pines (Mohr
et al. 2000; Thompson & Anderson 2000). Unfortunately,
individual Pinus species cannot be identified from pollen.
Narrow haplotype (chloroplast or mitochondrial) transition
zones observed in other western North American plants
(Soltis et al. 1997; Aagaard et al. 1998; Latta & Mitton 1999;
Johansen & Latta 2003) have also been attributed to postglacial contact of previously separated populations.
The examination of maternal and paternal haplotypes
offers insight into the dynamics of dispersal at the contact
zone (Fig. 2). The two P. lambertiana trees sampled in the
southern part of seed zone 731 represent a population that
is isolated on Happy Camp mountain, Modoc County.
These trees have the S haplotype, and are presumed to
have colonized this location by long-distance seed dispersal.
The closest sampled potential source is c. 90 km away.
The large (228 ± 40 mg) and ‘flimsy’ winged seeds of P.
lambertiana are ‘seldom dispersed far by wind’, but caching
by Steller’s jays (Thayer & Vander Wall 2005) and Clark’s
nutcrackers (D. Tomback, personal communication) can
potentially lead to long-distance dispersal. No other example
of a disjunct haplotype was observed. Embryos possessing
the S haplotype occur up to 25 km north of the northern­
most potential source trees, indicating that pollen flow
advances ahead of seed dispersal. The two Happy Camp
mountain trees whose megagametophytes carry the S
haplotype have apparently been pollinated by trees of the
N haplotype. Comparison of chloroplast and mitochondrial
haplotypes in a Pinus ponderosa contact zone in western
Montana has found a similar pattern of more extensive
pollen flow and rare long-distance seed dispersal (Latta &
Mitton 1999; Johansen & Latta 2003).
One of the most surprising results of our study is the
concordance between the distribution of the two cpDNA
haplotypes and the relative frequency of a white pine blister
rust major gene resistance locus (Cr1) in sugar pine. Kinloch
(1992) described the pattern of Cr1 frequency as a cline.
In contrast, the significant differences in Cr1 frequency
observed among the three haplotype groups (N, S, and
mixed) suggests that the gene frequency does not change
in a gradual manner, but rather shows the same abrupt
transition as observed in the chloroplast. There is no
3934 A . L I S T O N E T A L .
evidence for a causal link between the cpDNA haplotype
and Cr1 distribution patterns. It is well-established that
resistance shows nuclear inheritance (Kinloch 1992), and
the highest frequencies of Cr1 (4 – 9% per seed zone) are far
lower than the frequency of the S haplotype. Furthermore,
examination of three resistant trees (heterozygous for the
dominant Cr1 allele; J. Gleason, unpublished data) in
the contact zone found both N and S haplotypes (Fig. 2).
Note that all other genotyped individuals in the contact
zone are nonresistant ( J. Gleason, unpublished data).
We predict that the concordance between the Cr1 and
chloroplast haplotype frequencies reflects a common
history of genetic isolation, followed by recent migration
and contact (see below).
Evidence for chloroplast introgression
The chloroplast haplotypes of P. lambertiana resolve in two
different clades in the phylogenetic analysis of Pinus
subsect. Strobus, one comprised of five other North
American species and the other encompassing 12 Eurasian
species and the North American Pinus albicaulis (whitebark
pine). Two biological processes could explain these results:
incomplete lineage sorting of an ancestral polymorphism,
or chloroplast introgression. Incomplete lineage sorting
has been determined to be the most probable source of
widespread allelic nonmonophyly at nuclear loci in species
of Pinus subgenus Strobus (Syring et al. 2007). It has also
been considered a potential cause of similar patterns
observed in chloroplast studies in other plant species
(Tsitrone et al. 2003). However, the stochastic process of
incomplete lineage sorting is not expected to show the
strong geographical partitioning observed for the two
chloroplast haplotypes. On the other hand, if the two
subgroups of sugar pine have been separated since the
Miocene (15.1 ± 3.8 Ma) and each retained a different
haplotype, one might expect to find morphological
divergence between (and sequence variation within) the
two groups, particularly since this same interval has
apparently been accompanied by multiple speciation events
in these lineages. Although some sequence divergence was
found in the S haplotype clade (in the geographically
isolated Baja California population), none was found in the
N haplotype clade. The amount of sequence divergence
between the S haplotype of P. lambertiana and the other
North American white pines is consistent with genetic
isolation since the Miocene. In contrast, the high sequence
similarity between the N haplotype of sugar pine and the
Asian clade is suggestive of a much more recent shared
plastid ancestry.
To account for this unexpected genetic similarity, we
suggest that the N haplotype of P. lambertiana may have
its origin in a chloroplast introgression event involving P.
albicaulis. Introgression-mediated chloroplast transfer has
been named ‘chloroplast capture’ (Rieseberg & Soltis 1991;
Tsitrone et al. 2003) and has been offered as an explanation
for similar patterns of cytonuclear incongruence observed
in many plant genera (reviewed in Rieseberg et al. 1996;
Wendel & Doyle 1998). Tsitrone et al. (2003) have modelled
the process under the assumption of maternal chloroplast
inheritance and they determined conditions likely to
promote its occurrence. A key aspect of their model is the
observation that cytonuclear incompatibility often results
in full or partial male sterility and thus can increase maternal
fitness through enhanced seed production. Although they
do not explicitly model paternal inheritance (the situation
in Pinaceae), they suggest that chloroplast introgression
should be less common here, since cytonuclear interactions
typically reduce male fitness. Theoretically, chloroplast
substitution could result in an advantage in male function,
but this situation has apparently not been documented
(Tsitrone et al. 2003). However, a pattern consistent with
chloroplast introgression has been observed in other
species of Pinaceae, e.g. Pinus montezumae (Matos & Schaal
2000), Pinus muricata (Hong et al. 2003) and Larix sibirica
(Wei & Wang 2003).
Petit et al. (2003) offer ‘pollen swamping’ as an alternative
explanation for the lack of cytoplasmic (cpDNA and mito­
chondrial DNA) differentiation between Quercus petraea
and Quercus robur, two sympatric oaks that are consistently
differentiated at nuclear markers. In their scenario, Q. robur
seed disperses into new habitats which are subsequently
colonized by Q. petraea via pollen flow, resulting in F1
hybrids. Asymmetric introgression and strong selection
for the Q. petraea phenotype results in mixed populations
that share a single cytoplasm. Petit et al. (2003) invoke the
fact that the seeds of Q. robur are better adapted to birddispersal than Q. petraea, and thus are more likely to establish
through long-distance dispersal. A similar relationships
exists between P. albicaulis (dispersed mainly by the pine
seed specialist, Clark’s nutcracker, Tomback 2005) and
P. lambertiana (dispersed to a limited extent by wind and
primarily by generalist Steller’s jays and yellow pine
chipmunks, Thayer & Vander Wall 2005). In both cases,
the better disperser is thought to have contributed its
chloroplast to the other species. An important caveat is
that the cytoplasm is maternally inherited in oaks, and thus
carried by the seed and not by the pollen as in pines.
If chloroplast introgression is responsible for this pattern,
how does one account for the otherwise ‘Eurasian’ haplotype
in two species of North American subsect. Strobus? The
cpDNA-based phylogeny (Fig. 3) requires two dispersal
events between western North America and Asia. The
disjunction between P. monticola and the Eurasian clade
can be dated to the Late Miocene or Early Pliocene. This
timing is consistent with estimates of 2.6 –16.7 Ma from
11 other eastern Asian/western North American plant
disjunctions (Zhu et al. 2006; Zhang et al. 2007). It is
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
P H Y L O G E N Y A N D P H Y L O G E O G R A P H Y 3935
noteworthy that well-preserved Pliocene and Pleistocene
fossils of P. monticola have been collected in northeastern
Siberia and Alaska (reviewed in Bingham et al. 1972).
Following diversification of the Eurasian white pines, and
origin of the ‘closed cone’ morphology characteristic of
stone pines, we propose that the ancestor of P. albicaulis
dispersed from Asia to North America via Beringia,
presumably during the Pleistocene. Beringia was largely
unglaciated during the Pleistocene, and is known to have
served as a refugium for trees and shrubs, including
Pinus, through the late glacial maximum (Brubaker et al.
2005). Krutovskii et al. (1995) proposed a similar scenario
based on allozyme and chloroplast restriction fragment
analysis of the ‘subsect. Cembrae’ pines, but placed the
migration at an earlier period (Pliocene).
Evidence from mitochondrial DNA haplotypes has been
used to infer the existence of three Late Pleistocene refugia
for P. albicaulis (Richardson et al. 2002): western Wyoming,
western Idaho and the southern Cascades of Oregon. The
southern Cascades currently support large populations of
P. lambertiana, and we propose that this region could also
have served as a northern glacial refugium for sugar pine,
or possibly farther west in the Klamath/Siskiyou Mountains
(a region with several palaeo-endemic conifers, e.g. Picea
breweriana and Chamaecyparis lawsoniana). Sugar pine is
common here, but whitebark pine has only recently been
discovered in a small population on Mount Ashland (nine
individuals; Murray 2005). Regardless of their current
geographical distributions, sympatry in a glacial refugium
could have provided the opportunity for the northern
populations of sugar pine to acquire the chloroplast of
whitebark pine. Although P. albicaulis generally occurs at
higher altitudes than P. lambertiana, the two are partly
sympatric in northern California and southern Oregon
(Fig. 1B). There is also evidence that P. lambertiana formerly
occurred at higher altitudes than its current distribution
(May 1974; Millar et al. 2006).
Two factors are required for chloroplast introgression to
occur; sympatry and reproductive compatibility. While
many interspecific crosses have been conducted in subsect.
Strobus, there is no record of attempts to cross P. lambertiana
and P. albicaulis (Critchfield 1986; R. Sniezko, personal
communication). Critchfield & Kinloch (1986) do, however,
document interspecific hybridization between P. lambertiana
and Asian members of subsect. Strobus, namely P. armandii
and P. koraiensis. Seed set in these artificial crosses averaged
2.1% and 0.2% viable seed per cone, respectively. In
contrast, P. lambertiana is apparently intersterile with all
other North American white pines (Critchfield 1986;
Fernando et al. 2005). No naturally occurring hybrids of
P. lambertiana have been recorded (Mirov 1967; Critchfield
1986; R. Sniezko, personal communication).
Our results explain why P. lambertiana was resolved in
conflicting positions in recent chloroplast sequence-based
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
phylogenetic analyses of pines. Gernandt et al. (2005)
sampled P. lambertiana in Oregon (a region fixed for the N
haplotype), while Eckert & Hall (2006) sampled an individual
from southern California (a region fixed for the S haplo­
type). Each study placed P. lambertiana in a position that is
consistent with the resolution of the respective haplotypes
in our study (Fig. 3). This demonstrates the importance of
sampling multiple individuals per species in phylogenetic
analyses of closely related species (see also Syring et al.
2007).
The results reported here provide the first phylogeo­
graphical hypothesis for an ecologically and economically
important conifer, P. lambertiana. By placing the intraspecific
results within a broader phylogenetic context, further
insights were gained into the evolutionary history of this
species. The novel hypotheses of two Pleistocene refugia
for P. lambertiana and chloroplast introgression with P.
albicaulis can be tested with additional molecular markers,
in particular nuclear and mitochondrial loci. The observation
that the two chloroplast haplotypes demarcate population
groups that differ in their vulnerability to white pine blister
rust is also a significant result that merits further attention.
Beyond sugar pine, this study demonstrates the value of
including an interspecific phylogenetic component in
phylogeographical research. Without this broader per­
spective, the antiquity of the haplotype groups would
remain unknown, as would the unexpected, and potentially
reticulate, history of these species.
Acknowledgements
We are indebted to John Gleason (USFS, Placerville Nursery and
Disease Resistance Program), Jerry Berdeen and Richard Sniezko
(USFS, Dorena Genetic Resource Center) and David Johnson
(USFS, Institute of Forest Genetics) for supplying sugar pine seeds
and Roman Businsky for providing his collections of Asian
species. We thank David Gernandt, Bohun Kinloch, Todd Ott,
Paul Severns, Richard Sniezko, and Diana Tomback for constructive
comments on the manuscript. This research was funded by
National Science Foundation grants DEB 0317103 and ATOL
0629508, an NSF Research Experience for Undergraduates
supplement, and the Pacific Northwest Research Station,
USDA Forest Service.
References
Aagaard JE, Krutovskii KV, Strauss SH (1998) RAPD markers of
mitochondrial origin exhibit lower population diversity and
higher differentiation than RAPDs of nuclear origin in Douglas
fir. Molecular Ecology, 7, 801–812.
Bingham RT, Hoff RJ, Steinhoff RJ (1972) Genetics of Western White
Pine, USDA Forest Service Research Paper WO-12. USDA Forest
Service, Washington, D.C.
Brubaker LB, Anderson PM, Edwards ME, Lozhkin AV (2005)
Beringia as a glacial refugium for boreal trees and shrubs: new
perspectives from mapped pollen data. Journal of Biogeography,
32, 833–848.
3936 A . L I S T O N E T A L .
Chiang YC, Hung KH, Schaal BA et al. (2006) Contrasting phylo­
geographical patterns between mainland and island taxa of the
Pinus luchuensis complex. Molecular Ecology, 15, 765 – 779.
Conkle MT (1996) Patterns of variation in isozymes of sugar pine.
In: Sugar Pine: Status, Values, and Roles in Ecosystems: Proceedings
of a Symposium Presented by the California Sugar Pine Management
Committee (eds Kinloch BB, Marosy M, Huddleston ME), p. 99.
University of California Division of Agriculture and Natural
Resources, Davis, California.
Critchfield WB (1986) Hybridization and classification of the
white pines (Pinus section Strobus). Taxon, 35, 647 – 656.
Critchfield WB, Kinloch BB (1986) Sugar pine and its hybrids.
Silvae Genetica, 35, 138 – 145.
Critchfield WB, Little EL Jr (1966) Geographic Distribution of the
Pines of the World, Forest Service Miscellaneous Publication 991.
USDA Forest Service, Washington, D.C.
Devey ME, Delfino-Mix A, Kinloch BB, Neale DB (1995) Random
amplified polymorphic DNA markers tightly linked to a gene
for resistance to white pine blister rust in sugar pine. Proceedings
of the National Academy of Sciences, USA, 92, 2066 – 2070.
Eckert AJ, Hall BD (2006) Phylogeny, historical biogeography, and
patterns of diversification for Pinus (Pinaceae): phylogenetic
tests of fossil-based hypotheses. Molecular Phylogenetics and
Evolution, 40, 166– 182.
Fernando DD, Long SM, Sniezko RA (2005) Sexual reproduction
and crossing barriers in white pines: the case between Pinus
lambertiana (sugar pine) and P. monticola (western white pine).
Tree Genetics and Genomes, 1, 143 – 150.
Gadek PA, Alpers DL, Heslewood MM, Quinn CJ (2000)
Relationships within Cupressaceae sensu lato: a combined
morphological and molecular approach. American Journal of
Botany, 87, 1044–1057.
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.
Hausner G, Olson R, Simon D et al. (2006) Origin and evolution of
the chloroplast trnK (matK) intron: a model for evolution of
group II intron RNA structures. Molecular Biology and Evolution,
23, 380–391.
Hickman JC (1993) Geographic subdivisions of California. In: The
Jepson Manual: Higher Plants of California (ed. Hickman JC),
pp. 37–48. University of California, Berkeley, California.
Hong Y-P, Krupkin AB, Strauss SH (1993) Chloroplast DNA
transgresses species boundaries and evolves at variable rates in
the California closed-cone pines (P. radiata, P. muricata, and
P. attenuata). Molecular Phylogenetics and Evolution, 2, 322–329.
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 (1992) Distribution and frequency of a gene for
resistance to white pine blister rust in natural populations of
sugar pine. Canadian Journal of Botany, 70, 1319 – 1323.
Kinloch BB (2003) White pine blister rust in North America: past
and prognosis. Phytopathology, 93, 1044 – 1047.
Kinloch BB, Scheuner WH (2004) Pinus lambertiana Dougl., Sugar
Pine. USDA Forest Service, Washington, D.C. http://na.fs.fed.us/
spfo/pubs/silvics_manual/Volume_1/vol1_Table_of_contents.htm.
Krutovskii KV, Politov DV, Altukhov YP (1995) Isozyme study of
population genetic structure, mating system and phylogenetic
relationships of the five stone pine species (subsection Cembrae,
section Strobi, subgenus Strobus). In: Population Genetics and
Genetic Conservation of Forest Trees (eds Baradat P, Adams WT,
Müller-Starck G), pp. 279–304. Academic Publishing, Amsterdam.
Latta RG, Mitton JB (1999) Historical separation and present gene
flow through a zone of secondary contact in ponderosa pine.
Evolution, 53, 769–776.
Ledig FT (1998) Genetic variation in Pinus. In: Ecology and Biogeog­
raphy of Pinus (ed. Richardson DM), pp. 251–280. Cambridge
University Press, Cambridge, UK.
Liston A (1992) Variation in the chloroplast genes rpoC1 and rpoC2
of the genus Astragalus (Fabaceae): evidence from restriction site
mapping of a PCR-amplified fragment. American Journal of
Botany, 79, 953–961.
Liston A, Robinson WA, Piñero D, Alvarez-Buylla ER (1999)
Phylogenetics of Pinus (Pinaceae) based on nuclear ribosomal
DNA internal transcribed spacer sequences. Molecular Phylo­
genetics and Evolution, 11, 95–109.
Martinson SR (1997) Association among geographic, allozyme, and
growth variables for sugar pine (Pinus lambertiana Dougl.) in
southwest Oregon and throughout the species range. PhD Thesis,
North Carolina State University.
Matos JA, Schaal BA (2000) Chloroplast evolution in the Pinus
montezumae complex: a coalescent approach to hybridization.
Evolution, 54, 1218–1233.
May RH (1974) An observation of some sugar pine relicts.
Madroño, 22, 276.
Mielke JL (1943) White Pine Blister Rust in Western North America,
Bulletin no. 52, Yale University School of Forestry, New Haven,
Connecticut.
Millar CI, King JC, Westfall RD, Alden HA, Delany DL (2006) Late
Holocene forest dynamics, volcanism, and climate change at
Whitewing Mountain and San Joaquin Ridge, Mono County,
Sierra Nevada, CA, USA. Quaternary Research, 66, 273–287.
Mirov NT (1967) The Genus Pinus. The Ronald Press Co., New
York.
Mohr JA, Whitlock C, Skinner CN (2000) Postglacial vegetation
and fire history, eastern Klamath mountains, California, USA.
Holocene, 10, 587–601.
Murray M (2005) Our threatened timberlines: the plight of
whitebark pine ecosystems. Kalmiopsis, 12, 25–29.
Muse SV, Weir BS (1992) Testing for equality of evolutionary rates.
Genetics, 132, 269–276.
Navascues M, Vaxevanidou Z, Gonzalez-Martinez SC et al. (2006)
Chloroplast microsatellites reveal colonization and metapopu­
lation dynamics in the Canary Island pine. Molecular Ecology, 15,
2691–2698.
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, Bodénès C, Ducousso A, Roussel G, Kremer A (2003)
Hybridization as a mechanism of invasion in oaks. New Phytologist,
161, 151–164.
Petit RJ, Duminil J, Fineschi S et al. (2005) Comparative organization
of chloroplast, mitochondrial and nuclear diversity in plant
populations. Molecular Ecology, 14, 689–701.
Posada D, Crandall KA (1998) modeltest: testing the model of
DNA substitution. Bioinformatics, 14, 817–818.
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
P H Y L O G E N Y A N D P H Y L O G E O G R A P H Y 3937
Richardson BA, Brunsfeld SJ, Klopfenstein NB (2002) DNA from
bird-dispersed seed and wind-disseminated pollen provides
insights into postglacial colonization and population genetic
structure of whitebark pine (Pinus albicaulis). Molecular Ecology,
11, 215–227.
Richardson DM, Rundel PW (1998) Ecology and biogeography of
Pinus: an introduction. In: Ecology and Biogeography of Pinus
(ed. Richardson DM), pp. 3 – 46. Cambridge University Press,
Cambridge, UK.
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.
Rozas J, Sánchez-De I, Barrio JC, Messeguer X, Rozas R (2003)
dnasp, DNA polymorphism analyses by the coalescent and
other methods. Bioinformatics, 19, 2496 – 2497.
Scharpf RF (1993) Diseases of Pacific Coast Conifers, Handbook 521.
USDA Forest Service, Washington D.C.
Shaw J, Lickey EB, Beck JT et al. (2005) The tortoise and the hare II:
relative utility of 21 noncoding chloroplast DNA sequences for
phylogenetic analysis. American Journal of Botany, 92, 142–166.
Soltis DE, Gitzendanner MA, Strenge DD, Soltis PS (1997) Chloro­
plast DNA intraspecific phylogeography of plants from the
Pacific Northwest of North America. Plant Systematics and
Evolution, 206, 353– 373.
Swofford DL (2002) PAUP*, Phylogenetic Analysis Using Parsimony
(*and Other Methods), Version 4. Sinauer Associates, Sunderland,
Massachusetts.
Syring J, Farrell K, Businsky R, Cronn R, Liston A (2007) Wide­
spread 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 (Pinaceae) subsections inferred from
multiple low-copy nuclear loci. American Journal of Botany, 92,
2086–2100.
Thayer TC, Vander Wall SB (2005) Interactions between Steller’s
jays and yellow pine chipmunks over scatter-hoarded sugar
pine seeds. Journal of Animal Ecology, 74, 364 – 375.
Thompson RS, Anderson KH (2000) Biomes of western North
America at 18 000, 6000 and 0 14C yr bp reconstructed from
pollen and packrat midden data. Journal of Biogeography, 27,
555–584.
© 2007 The Authors
Journal compilation © 2007 Blackwell Publishing Ltd
Tomback DF (2005) The impact of seed dispersal by Clark’s
nutcracker on whitebark pine: multi-scale perspective on a high
mountain mutualism. In: Mountain Ecosystems: Studies in Treeline
Ecology (eds Broll G, Keplin B), pp. 181–201. Springer, Berlin.
Tsitrone A, Kirkpatrick M, Levin DA (2003) A model for chloro­
plast capture. Evolution, 57, 1776–1782.
Wang XR, Tsumura Y, Yoshimaru H, Nagasaka K, Szmidt AE
(1999) Phylogenetic relationships of Eurasian pines (Pinus,
Pinaceae) based on chloroplast rbcL, matK, rpl20-rps18 spacer,
and trnV intron sequences. American Journal of Botany, 86, 1742 –
1753.
Wefald M (2001) TRS2LL, http://www.geocities.com/jeremiahobrien/
trs2ll.html.
Wei XX, Wang XQ (2003) Phylogenetic split of Larix: evidence
from paternally inherited cpDNA trnT-trnF region. Plant
Systematics and Evolution, 239, 67–77.
Wendel JF, Doyle JJ (1998) Phylogenetic incongruence: window
into genome history and molecular evolution. In: Molecular
Systematics of Plants: DNA Sequencing (eds Soltis DE, Soltis PS,
Doyle JJ), pp. 265–296. Kluwer, Boston, Massachusetts.
Willyard A, Syring J, Gernandt DS, Liston A, Cronn R (2007) Fossil
calibration of molecular divergence infers a moderate mutation
rate and recent radiations for Pinus. Molecular Biology and Evolution,
24, 90–101.
Zhang M-L, Uhink CH, Kadereit JW (2007) Phylogeny and bio­
geography of Epimedium/Vancouveria (Berberidaceae): western
North American–East Asian disjunctions, the origin of Euro­
pean mountain plant taxa, and East Asian species diversity.
Systematic Botany, 32, 81–92.
Zhu Y-P, Wen J, Zhang Z-Y, Chen Z-D (2006) Evolutionary
relationships and diversification of Stachyuraceae based on
sequences of four chloroplast markers and the nuclear ribosomal
ITS region. Taxon, 55, 931–940.
Aaron Liston and Rich Cronn collaboratively study the system­
atics, 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 University-Billings, and Dr Willyard is a postdoc
at the University of South Dakota. Mariah Parker-deFeniks is a
Sociology major at Oregon State University, pursuing a career in
Criminal Justice and Social Research.
Download