<<

Molecular Phylogenetics and Evolution 69 (2013) 717–727

Contents lists available at SciVerse ScienceDirect

Molecular Phylogenetics and Evolution

journal homepage: www.elsevier.com/locate/ympev

A new phylogeny for the Picea from plastid, mitochondrial, and nuclear sequences ⇑ Jared D. Lockwood a,1, Jelena M. Aleksic´ b,1, Jiabin Zou c,1, Jing Wang c, Jianquan Liu c, Susanne S. Renner a, a Systematic and Mycology, University of Munich (LMU), Menzinger Strasse 67, 80638 Munich, Germany b University of Belgrade, Institute of Molecular Genetics and Genetic Engineering, 11010 Belgrade, Serbia c State Key Laboratory of Grassland Agro-ecosystem, School of Life Sciences, Lanzhou University, Lanzhou, Gansu 730000, article info abstract

Article history: Studies over the past ten years have shown that the crown groups of most genera are only about Received 28 March 2013 15–25 Ma old. The genus Picea (, ), with around 35 , appears to be no exception. In Revised 21 June 2013 addition, molecular studies of co-existing species have demonstrated frequent introgression. Per- Accepted 5 July 2013 haps not surprisingly therefore previous phylogenetic studies of species relationships in Picea, based Available online 18 July 2013 mostly on plastid sequences, suffered from poor statistical support. We therefore generated mitochon- drial, nuclear, and further plastid DNA sequences from carefully sourced material, striking a balance Keywords: between alignability with outgroups and phylogenetic signal content. Motif duplications in mitochon- Molecular clocks drial introns were treated as characters in a stochastic Dollo model; molecular clock models were cali- Historical biogeography Mitochondrial nad introns brated with ; and ancestral ranges were inferred under maximum likelihood. In agreement with Stochastic Dollo model previous findings, Picea diverged from its sister clade 180 million years ago (Ma), and the most recent Secondary structure-based alignment common ancestor of today’s spruces dates to 28 Ma. Different from previous analyses though, we find North American spruces a large Asian clade, an American clade, and a Eurasian clade. Two expansions occurred from to North calibrations America and several between Asia and . Chinese P. brachytyla, American P. engelmannii, and Nor- way spruce, P. abies, are not monophyletic, and North America has ten, not eight species. Divergence times imply that refugia are unlikely to be the full explanation for the relationships between the European species and their East Asian relatives. Thus, northern spruce may be part of an Asian species complex that diverged from the southern Norway spruce lineage in the Upper Miocene, some 6 Ma, which can explain the deep genetic gap noted in phylogeographic studies of Norway spruce. The large effective population sizes of spruces, and incomplete lineage sorting during speciation, mean that the interspecific relationships within each of the geographic clades require further studies, especially based on genomic information and population genetic data. Ó 2013 Elsevier Inc. All rights reserved.

1. Introduction young, with the largest genera, Abies, Cupressus, Picea, Pinus, and Juniperus, having median node ages under 3.5 Mya (Leslie et al., Gymnosperms comprise just over 1000 species in 85 genera, 2012). with making up 600 species and 68 genera (Christenhusz Interspecific relationships in young radiations are usually diffi- et al., 2011). While the gymnosperm fossil record extends back to cult to resolve due to the incomplete lineage sorting, and an addi- the , research over the past 10 years has made clear that the tional challenge in many conifer groups is introgression of most recent common ancestors of most gymnosperm genera date organelle genes following hybridization between co-occurring spe- to just 15–25 Ma (Treutlein and Wink, 2002; Won and Renner, cies. For spruces (Picea), the focal group here, examples are P. abies 2006; Eckert and Hall, 2006; Willyard et al., 2007; Knapp et al., and P. obovata in the Ural Mountains (Krutovskii and Bergmann, 2007; Qiao et al., 2007; Gernandt et al., 2008; Ickert-Bond et al., 1995), P. mariana and P. rubens in eastern North America (Perron 2009; Wei et al., 2010; Lin et al., 2010; Nagalingum et al., 2011; and Bousquet, 1997; Perron et al., 2000; Jaramillo-Correa et al., Mao et al., 2010, 2012; Leslie et al., 2012). Northern hemisphere 2003), P. glauca, P. engelmannii, and P. sitchensis in western North conifer clades (Pinaceae, Cupressoideae) tend to be especially America (Sutton et al., 1991; Rajora and Dancik, 2000), and P.liki- angensis, P. purpurea, and P. wilsonii from the Qinghai–Tibetan Pla- teau in China (Li et al., 2010; Du et al., 2011; Zou et al., 2012). ⇑ Corresponding author. Nevertheless, phylogenetic studies of the genus Picea have relied E-mail address: [email protected] (S.S. Renner). heavily on plastid DNA. The first such study used RFLP data and 1 These authors contributed equally to the work.

1055-7903/$ - see front matter Ó 2013 Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.ympev.2013.07.004 718 J.D. Lockwood et al. / Molecular Phylogenetics and Evolution 69 (2013) 717–727 inferred a North American origin because three North American are shown in Fig. 1). Farjón (1990) considers the American P. mar- species (P. engelmannii, P. glauca, P. mexicana) branched off basally tinezii a of P. chihuahuana; to test this, we included se- in the cladogram (Sigurgeirsson and Szmidt, 1993). A second study quences representing both names, using topotypical material. We relied on plastid loci and mitochondrial intron haplotypes (Ran also included both of P. engelmannii, the typical subspe- et al., 2006), as did a third, which however, focused more on cies and subsp. mexicana (of Schmidt, 1988; Farjón, 1990). The five incongruence (Bouillé et al., 2011). In both these latter studies, Japanese endemics were all sampled, P. alcoquiana (P. bicolor (Max- plastid data yielded a topology in which the western American P. im.) Mayr), P. glehnii, P. koyamae, P. maximowiczii, and P. torano (P. breweriana and P. sitchensis formed a basal grade. This caused polita (Siebold and Zucc.) Carriere). As outgroups, we included rep- Ran et al. (2006) to suggest an origin of Picea in North America. resentatives of the other 10 genera of Pinaceae (a family of c. 215 They also inferred two migrations from North America to Asia, species), namely one species each of Abies (fir), , (ce- one from Asia to North America, one from North America to Eur- dar), , Larix (), (Bristlecone hemlock), ope, and one from Asia to Europe. By contrast, a pre-molecular (Golden larch), (Douglas fir), and study suggested that Picea originated in Asia and expanded to (hemlock), and three species from the two subgenera (Pinus and North America from there (Wright, 1955). None of the molecular Strobus)ofPinus. studies had statistical support for deeper relationships in Picea. A problem in previous studies seems to have been misidentified To better resolve relationships among spruces we decided to fo- material or contaminated DNA, which would explain certain con- cus on mitochondrial (mt) introns and, if appropriate, to combine tradictory findings. For example, some nad1 intron sequences la- mt data with plastid and nuclear data. Among the most informa- beled as representing one species differ strikingly from each tive regions in conifers are group II introns in the mitochondrial other. Thus, the nad1 sequences produced by Sperisen genes that code for NADH dehydrogenase subunits, specifically the et al. (2001) and Tollefsrud et al. (2008) from 369 natural popula- first intron of nad5 and the second intron of nad1 (Grivet et al., tions across Europe do not match the P. abies nad1 sequences sub- 1999; Gugerli et al., 2001a, 2001b; Sperisen et al., 2001; Ran mitted to GenBank by Bouillé et al. (2011; accessions EF440449, et al., 2006; Meng et al., 2007; Bouillé et al., 2011). Using structural EF440450, EF440451). Similarly, sequences in GenBank of P. asper- criteria from the folding of group II introns (Kelchner, 2000), an ata, P. crassifolia, P. jezoensis, P. omorika, and P. wilsonii do not alignment of 17 species of Picea revealed two indels of up to match newly generated sequences from documented mate- 1500 base pairs (bp) and several complex motifs of about 30 bp rial. We have excluded all such doubtful sequences. in domain IV of nad1 intron 2 that were conserved at the level of species groups (Aleksic´, 2008; Aleksic´ and Geburek, 2010, in press). 2.2. Gene sequencing and alignment This suggested that these two introns would be phylogenetically useful as long as alignment homology was maintained. We sequenced mitochondrial introns, several relatively con- We here study phylogenetic relationships in Picea based on served plastid gene regions, and a nuclear gene. Note that spruce DNA sequences from all but one of its ca. 35 species (Farjón, plastid DNA is paternally inherited (Sutton et al., 1991; Mogensen, 1990, 2001), with a few problematic species represented by multi- 1996; Grivet et al., 1999). From the mitochondrial genome, we ple individuals. Picea ranges from the Circle to Mexico and sequenced nad5 intron 1 and nad1 intron 2; from the chloroplast (Farjón, 2001; our Fig. 1), with America harboring a sup- genome, the rbcL gene and the matK gene. We also sequenced three posed eight species, Europe (including Turkey and the Caucasus re- plastid intergenic spacers, viz. trnL-trnF, trnH-psbA, and trnS-trnG gion) to the Ural Mountains three, and Asia c. 23 (Farjón, 2001; the (Ran et al., 2006; Meng et al., 2007; Bouillé et al., 2011; Du et al., lists 16 native species, seven of them endemic; Fu 2009, 2011). All data matrices have the same taxonomic represen- et al., 1999). Of special interest in terms of economic importance tation except for the outgroup genera Abies and Keteleeria, in which is Norway spruce, Picea abies. Phylogeographic studies of this spe- sequences from congeneric species were combined to represent cies have consistently revealed a genetic gap between its northern the genus (see Table S1 in Supporting Information). From the nu- and southern European populations, coinciding with the middle clear genome, we sequenced the gene encoding 4-coumarate: Polish disjunction (Schmidt-Vogt, 1974, 1977; Lagercrantz and Ry- coenzyme A ligase (4CL) in the lignin biosynthetic pathway. Partial man, 1990; Sperisen et al., 2001; Collignon et al., 2002; Tollefsrud 4CL (600 bp) sequences had already been obtained for 400+ indi- et al., 2008; our Fig. 1), which has been attributed to Holocene viduals from several Chinese species as part of phylogeographic range changes one or several millennia ago (Latałowa and van work (Li et al., 2010), and 2000 bp sequences were available from der Knaap, 2006). We thus aimed to infer (i) the age and relation- 117 individuals of most Picea (J.Q. Liu and collaborators, unpub- ships of Norway spruce, (ii) the relationship among the American, lished data). Of the 117 sequences, 53 contained no ambiguous Asian, and European spruces, and (iii) possible topological conflict base calls, while the remaining contained a few ambiguous base between plastid, nuclear, and mitochondrial data in a presumably calls. For P. asperata, P. crassifolia, P. koraiensis, P. meyeri, P. retro- young conifer radiation (Bouillé and Bousquet, 2005; Crisp and flexa, and P. schrenkiana, sequences with ambiguous base calls Cook, 2011), an assumption we test with strict and relaxed clock were identified as either the paternal or maternal allele by compar- models, calibrated with various Pinaceae fossils. ing them to 4CL sequences from haploid megagametophyte tissue. For the Japanese species P. alcoquiana we only obtained 816 bp of 4CL. 2. Materials and methods Total genomic DNA was extracted from silica-dried needles using a commercial plant DNA extraction kit (NucleoSpin, Mache- 2.1. Taxon sampling rey-Nagel, Düren, Germany). Primers used for polymerase chain reactions (PCRs) and sequencing are listed in Table S2. PCR prod- We sampled 47 individuals representing all commonly recog- ucts were purified with the ExoSAP or FastAP clean-up kits (Fer- nized species and subspecies (Farjón, 1990, 2001). Not sampled is mentas, St. Leon-Rot, Germany), and sequencing relied on Big Picea aurantiaca, an endangered species endemic to West , Dye Terminator kits (Applied Biosystems, Foster City, CA, USA) China, which has been treated as a variety of P. asperata (Fu et al., and an ABI 3130 automated sequencer. 1999; Eckenwalder, 2009). To represent P.abies, we sampled three A total of 350 sequences were generated and submitted to Gen- individuals from the Baltico–Nordic domain, one from the Hercy- Bank. Table S1 provides species names with authorities, voucher no-Carpathian domain, and four from the Alpine domain (domains information, GenBank accession numbers, and the general J.D. Lockwood et al. / Molecular Phylogenetics and Evolution 69 (2013) 717–727 719

P. glauca P. mariana P. sitchensis

P. breweriana P. rubens P. engelmannii P. pungens P. mexicana

P. chihuahuana P. martinezii

Picea abies Baltico-Nordic domain

Picea abies Hercyno-Carpathian P. abies P. obovata and Alpine domains P. jezoensis

P. abies P. schrenkiana P. koraiensis P. wilsonii P. omorika P. crassifolia P. glehnii P. orientalis P. purpurea P. meyeri P. asperata P. alcoquiana P. retroflexa P. neoveitchii P. koyamae P. smithiana P. brachytyla P. maximowiczii P. likiangensis P. torano

P. spinulosa P. farreri P. morrisonicola

Fig. 1. Map of the Northern Hemisphere in equal-area projection (Mollweide) showing the distribution of Picea based on Farjón (1990) and eFloras (2008). Black dashed lines represent the Tropic of Cancer and the Arctic Circle. geographic distribution of each species. Alignment was done man- fit the plastid matrix, the TPM3uf + C model the nuclear matrix, ually, and for the group II mitochondrial introns nad1 and nad5, and the SYM + I model the mitochondrial matrix. In all cases, the conserved domains I, II, III, V, and VI and the variable domain IV model with the highest likelihood score was the general time were aligned according to structural criteria (Fig. 2). The combined reversible (GTR) model, and we opted for GTR + C with four cate- alignment contained approximately 50% empty cells, mostly from gories of rate heterogeneity for the final analyses. were outgroup nad1 and 4CL sequences. rooted on the Abietoideae subfamily of Pinaceae (Abies, Cedrus, Keteleeria, Nothotsuga, Pseudolarix, Tsuga), which is the sister group 2.3. Phylogenetic analyses of the Pinoideae (Cathaya, Larix, Picea, Pinus, Pseudotsuga; Gernandt et al., 2008). Phylogenetic trees were obtained by Bayesian inference (BI) in Bayesian analyses (using BEAST) used a Yule tree prior, several BEAST v. 1.7.2 (Drummond et al., 2012) and maximum likelihood Monte Carlo Markov chains (MCMC) of 60 million generations (ML) in RAxML (Stamatakis, 2006) with the raxmlGUI v. 1.1 (Silv- each, with parameters sampled every 10,000 generations. To take estro and Michalak, 2011). Best-fitting models of sequence evolu- advantage of informative insertions and deletions (indels) in the tion were estimated using jModelTest v. 0.1.1 (Posada, 2008). mitochondrial introns, we coded indels as characters using simple Based on the Akaike information criterion, the TIM1 + I + C model indel coding (SIC; Simmons and Ochoterena, 2000) as imple- (I: invariant sites; C: gamma-distributed rate heterogeneity) best mented in SeqState (Müller, 2005). Simple indel coding was 720 J.D. Lockwood et al. / Molecular Phylogenetics and Evolution 69 (2013) 717–727

(a) IV A B

Variable 493 bp 1333 bp Section Domain

B-b G1-1 G1-2 A1 A2 A3 A4 A5 A6A4 A7 A8 A7 B-a B-c

414 bp 76 bp 948 bp G1 G2 G3 36 17 269 bp 185 bp 115 bp bp bp Conserved Domain Section I II III V VI (b) Picea nad1 intron 2 domain IV

Section G1: G1-1 present in all species, G1-2 motifs present one variant per species Section G2: Present in all species, motifs present one variant per species G1-2a CGCCCTCGGAGGGCGAGCGTTCGTTTATTACCCTCTCCCT----- 40 bp G2-1a ------TTTGGTCGAGCGTTCATTTATCA 23 bp * G1-2b CGCCCTCGGAGGGCGAGCGTTCGTTTATTCCCCTCTCCCT----- 40 bp * G2-1b TTTGGTCTCCCTTTTGGTTTGGTCGAGCGTTCATTTATCA 40 bp G1-2c CGCCCTCGGAGGGCGAGCGTTCGTTTATTACCCTCTCCCTTTTCT 45 bp G2-2a CCCTTTATTTCTT-----CCCCTCAAA---GGG 25 bp G2-2b CCCTTGATTTCTT-----CCCCTCAAAGGGGGG 28 bp Section A: Present in clade I and P. morrisonicola, harbors minisatellites G2-2c CCCTTTATTTCATTTCTTCCCCTCAAA---GGG 30 bp A1-a CACCCATATGATGAGTGAGCGTTAACACCCT 31 bp * G2-2d CCCTTATAGGCTT------CCCTCAAA----GG 23 bp * A1-b CACCCATATGATGAGGGAGCGTTAACACCCT 31 bp G2-2e CCCTTATAGGCTT-----CCCCTCAAA---GGG 25 bp A1-c CACCCATATGATGAGKGAGCGTTAACACCCT 31 bp G2-2f CCCTTTATTTCTT-----CCCCTCAAAGGGGGG 28 bp * A2-a CACCCATGAATGGATGAGCGACTTCGTTCCT 31 bp A2-b CACC-ATGAATGGATGAGCGACTTCGTTCCT 30 bp • A3 CTCCCGTTGGTCGAGAGTTCGCTGCCTCACCCT 33 bp Section B: Present in clades II and III, B-a harbors indels and • A4 CGCCCTTTTTGGGTCGAGTCACTTAACGTACCT 33 bp duplications, either B-b or B-c present per species • A5 CGCCCCTTTTGGTCGAGTCACTTAACGTACCT 32 bp B-b CCCTGTATTTATTCCCCTCAAAGGG 25 bp A6-a CGCCTTCCTACCTCAGTCGAGTCACTTAACGTACCT 36 bp • * B-c CCCTGTATTTATTCCCCTAAAAGGG 25 bp A6-b CGCCTTCCTACCTCAGGCGAGTCACTTAACGTACCT 36 bp ** A7 CACCCATCGGATGGATGAGCGACTTCGTACCT 32 bp • • A8 CCCCCTCCGTTGTCAGGGGAGCGACTTCGTACCT 34 bp Section G3: Present in all species

Fig. 2. (a) Schematic representation of nad1 intron 2 in Picea. Variable (darker gray) and conserved (lighter gray) intron domains and sections with lengths based on the ingroup alignment. (b) Motifs identified within both conserved and variable sections of domain IV. Single asterisks () mark motifs present as one variant per species in all species, double asterisks () mark motifs present in single or multiple copies in all species. Single bullets () mark motifs present as one variant per species in some species, a double bullet () mark motifs present in single or multiple copies in some species. performed after removal of autapomorphic insertions, potentially alignments (‘prior-only’ option) and compared the resulting posterior homoplastic microsatellites, and minisatellites. The binary SIC ma- divergence times to assess the influence of the prior settings trix was treated as a separate data partition in BEAST, and a sto- Relative divergence times were translated into absolute time chastic Dollo model was applied to it. This model was using fossil calibrations. The Pinaceae family has an extensive fossil appropriate for the motifs because it is implausible that a motif, record (for a recent review: Klymiuk and Stockey, 2012: their once lost, would be regained in the exact original form. Log files Fig. 5). Fossils used in our study or used in previous Pinaceae dating were analyzed in Tracer v. 1.5 (Rambaut and Drummond, 2007) but that we consider difficult to assign are listed in Table S3 along to assess convergence and to confirm that the effective sample with detailed justifications for why certain fossils were used as cal- sizes for all parameters were larger than 200, indicating that sta- ibrations or instead as outside evidence to provide independent val- tionarity had been reached. TreeAnnotator (part of the BEAST pack- idating arguments in the discussion of results. The first calibration age) was used to discard 10% of the saved trees as burn-in and to used was a Pseudolarix fossil from the Tsagaan Tsab Formation of produce a maximum clade credibility tree from the remaining southern (Krassilov, 1982; Table S3), a locality radiomet- trees. For ML analyses, we first transformed the DNA matrix into rically dated to the Late (156 ± 0.76 Ma; Keller and Hendrix, a five-state matrix in which nucleotides were represented by char- 1997). This provided a minimum age for the stem of Pseudolarix;we acter states 0 to 3, and gaps were treated as a fifth character state. chose a gamma prior distribution offset to 156 Myr, with a shape In a second approach, we removed all sites containing gaps but not parameter of 4.0, allowing 90% of the ages to fall between 163.8 point mutations, and ran the analysis with both a gap-less DNA and 157.4 Myr. Keller and Hendrix’s dating is not for the precise matrix and a binary indel matrix, created using SIC. location of Krassilov’s fossil (Table S3), but a Pinaceae cone dated For Bayesian analyses, posterior probabilities P0.98 were con- to the Lower Kimmeridgian (ca. 155 Ma; Table S3) indicates that sidered good support. For ML analyses, we carried out 100 boot- the Pseudolarix node must be at least that old. strap replicates under the same model as tree searches and Second, permineralized wood of Pinus (form genus Pinuxylon) considered P75% good support. Final trees were viewed and edi- recovered from the Aachen Formation of Belgium (Meijer, 2000), da- ted in FigTree v. 1.3.1 (http://tree.bio.ed.ac.uk). ted to the Santonian (86.3–83.6 Ma; Cohen et al., 2012), provided a minimum age for the two subgenera of Pinus (cf. Willyard et al., 2.4. Molecular clock dating 2007: Pinus; Gernandt et al., 2008: Pinaceae; Lin et al., 2010:Pinaceae; He et al., 2012: Pinus). We chose a gamma prior distribution offset by For the molecular clock dating, we used the nuclear 4CL exon ma- 85 Myr, with a shape parameter of 4.0, which allowed 90% of the ages trix plus the chloroplast genes matK and rbcL. We took out nine Picea to fall between 92.8 and 86.4 Myr. The Pseudolarix and Pinuxylon cal- accessions (P. abies Norway, P. abies , P. abies Austria 1, P. abies ibrations were tested against each other by applying each in isolation, Austria 2, P. abies Serbia 2, P. abies Germany, P. crassifolia, P. koraiensis thereby testing whether they supported each other’s assignment. and P. retroflexa) to reduce the number of very short branches, which Next, we applied a strict clock and the mean Pinaceae substitution cause problems for molecular-clock modeling and do not add infor- rates for matKandrbcL inferred by Gernandt et al. (2008), otherwise mation. We ran strict clocks and uncorrelated lognormal relaxed using the same settings as in the relaxed-clock BEAST runs. Third, we clocks, using the same tree prior, substitution model and burn-in compared the inferred ages with ages inferred in other studies and as used for tree inference. We also ran analyses with empty with the fossil record, using fossils not used as constraints. J.D. Lockwood et al. / Molecular Phylogenetics and Evolution 69 (2013) 717–727 721

2.5. Biogeographic analyses Section A (our Fig. 2a and b: A1–A7) that resemble 32-bp and 34-bp repeats previously described. All but one Picea have either Ancestral area reconstruction (AAR) relied on two approaches: section A or section B in domain IV of nad1. The exception is Trait reconstruction under the Markov k-state 1-parameter model P. morrisonicola in which the presence of both sections accounts (Mk1, a generalization of the Jukes-Cantor model; Lewis (2001))in for its unusually long intron sequence. When aligned with Abies, Mesquite v. 2.75 (Maddison and Maddison, 2011) and the dis- Cedrus, Cathaya, and Pinus, the nad1 intron matrix contained persal--cladogenesis (DEC) model in Lagrange (likeli- 5677 positions. Removal of all autapomorphic insertions and hood analysis of geographic range evolution; Ree and Smith, microsatellites resulted in an alignment of 4020 nucleotide charac- 2008). The AAR analyses were run on the chronogram obtained un- ters with 144 separately coded informative indels. der the fossil-calibrated relaxed clock. The species ranges of the 37 Picea nad5 intron 1 sequences ranged from 1167 to 1198 nucle- Picea species (Farjón, 1990; our Fig. 1) were categorized as follows: otides. Within the highly variable region of this intron (Jaramillo- (1) western North America including Mexico; (2) eastern North Correa et al., 2003; Ran et al., 2006) six mitotypes were identified, America; (3) Europe to the Ural Mountains (including Turkey and one of which is new (Fig. 3 inset, mitotype J). Two indels of five and the Caucasus region), (4) Asia; and (5) ambiguous. These 5 catego- ten bp were detected upstream of this region and an autapomor- ries (areas) provided a balance between having more than one of phic 11-bp duplication unique to P. mariana downstream from it. the 37 species in each area and subdividing the overall range of Alignment of ingroup and outgroup sequences yielded a matrix the genus into biogeographically meaningful units. Two wide- of 1908 bp. A 16-bp region of the Picea intron sequences could spread boreal American species (P. glauca and P. mariana) were not be unambiguously aligned and was excluded from analysis. coded as ‘ambiguous’ in the Mesquite analysis but assigned to re- When autapomorphic insertions, microsatellites, and the ambigu- gions 1 and 2 in Lagrange. The outgroup ancestral areas were coded ous region were removed, the matrix consisted of 1206 nucleotide as follows: Pinus (110–115 species distributed in North America, positions plus 50 informative indels. Europe, Asia and northern Africa): Europe and Asia (Eckert and Alternative treatments (Section 2) of the indel characters from Hall, 2006) in Lagrange but ambiguous in Mesquite; Cathaya (one nad1 and nad5 failed to significantly improve topology or statistical subtropical endemic Chinese species): Once coded as Asian, once support, and completely omitting indel characters yielded topolo- as ambiguous because of North American and Eurasian fossils gies that only differed in the positions of a few species at weakly (Liu and Basinger, 2000); Larix (14 species in North America, Asia, supported branches near the tips of the tree. Excluding the mini- and Europe): Asia (Wei and Wang, 2003); Pseudotsuga (five species satellite region in nad1 intron 2 from the matrix demonstrated that in Asia and North America): Western North America (Wei et al., these minisatellites are informative while not creating spurious 2010); Pseudolarix (one species in subtropical China): Once coded relationships. as Asian, once as ambiguous because of fossils in North America Picea nuclear 4CL gene sequences ranged from 1710 nucleotides and Europe (LePage and Basinger, 1995; our Table S3); Abies (P. glehnii) to 1796 (P. likiangensis var. rubescens), and complete se- (48 species in North America, Europe, and Asia): Ambiguous quences contained four introns, five exons, and a 3 0 untranslated (Xiang et al., 2009); Cedrus (four species in the circum-Mediterra- region. An alignment with outgroup genera had a length of 2610 nean and western ): Asia (Qiao et al., 2007); Keteleeria nucleotides, of which 1245 were coding. (three species in tropical Asia): Asia; Nothotsuga (one species in Plastid matrices consisted of 833 aligned nucleotides of the cod- tropical Asia): Asia; Tsuga (nine species in North America and ing region of matK, 129 bases at the 5 0 end of the trnK intron, and Asia): Ambiguous (Havill et al., 2008) 676 nucleotides of rbcL. Picea trnL-trnF sequences ranged from 698 Python input scripts for Lagrange were generated using an on- to 709 bases, with only short indels of 5 or 6 bases. Ingroup se- line tool (http://www.reelab.net/lagrange/configurator/index), quences of the trnH-psbA intergenic spacer were 601 bp in all sam- with the maximum number of ancestral areas constrained to pled individuals except P. glauca and P. engelmannii, which share a two. In one run, connection probabilities between geographic areas synapomorphic 5-bp insertion. An alignment of the Picea trnS-trnG through time (Table S4) were those developed by Moore and spacer sequences was 626 nucleotides long. Donoghue (2007: Fig. 7) and used also by Havill et al. (2008: There were two statistically supported topological incongruenc- Fig. 4), a model that involves 19 time slices. This model is based es among the data partitions (>87% ML bootstrap support). One in- on Cenozoic paleoclimates and Mesozoic continental positions volved the placement of P. maximowiczii with P. wilsonii, supported and includes such barriers to dispersal as the emergence and re- in the nuclear tree but not the plastid and mitochondrial trees. The treat of the Late Western Interior Seaway separating other involved the Japanese species P. alcoquiana (with only an eastern and western North America, and the Turgai Strait separat- incomplete nuclear 4CL sequence) and the Serbian P. omorika. With ing Europe from Asia from the Jurassic into the Oligocene. The the incomplete nuclear sequence added, these two species were model also accounts for changing connectivity during the Neogene, pulled into an otherwise North American clade, while the remain- which is likely to have affected crown group Picea (Section 3). In ing topology stayed unchanged. other runs, we explored simpler models that either included 11 time slices or no time slice. We also explored the effects of differ- 3.2. Evolutionary relationships in Picea ent outgroup area coding (above). A Bayesian tree shows a monophyletic Picea with three geo- graphically coherent clades (labeled I, II, III in Fig. 3). Maximum 3. Results likelihood yielded the same major clades, except that P. morrison- icola was placed as sister to clade I, and the western North Amer- 3.1. Sequence characterization and utility ican P. breweriana, which in the Bayesian tree is a member of clade III, is sister to all other species with weak support (tree not Picea nad1 intron 2 sequences ranged from 1912 nucleotides in shown). With the exception of P. morrisonicola and P. breweriana, P.jezoensis to 3175 in P. morrisonicola, with the range chiefly due to the three clades fit with the distribution of the nad5 mitotypes, two large indels flanking a 45-bp conserved region in domain IV of which contributed 50 of the total 194 informative indels. Haplo- all spruce species (Fig. 2, sections A and B). Five indels in this do- type D, which lacks insertions in the variable region (Fig. 3 inset), main described by Sperisen et al. (2001) in P. abies were readily occurs in P. breweriana, Taiwanese P. morrisonicola, and the found in other species. We also found minisatellite motifs in Eurasian clade I, while the remaining North American spruces 722 J.D. Lockwood et al. / Molecular Phylogenetics and Evolution 69 (2013) 717–727

Picea retroflexa D Picea D Picea crassifolia D nad5 intron 1 * D mitotypes D ACTTACTT------GAGTGGCA D PiceaPicea abiesabies NorwayNorway D * P. abies ACTTACTTTACTT------GAGTGGCA E PiceaPicea aabiesbies RussiaRussia D ACTTACTT------GACTTGAGTGGCA A * * Baltico-Nordic domain PiceaPicea ababiesies D ACTTACTT------GCTT------GACTTGAGTGGCA C D ACTTACTT-----GACTTGCTT------GACTTGAGTGGCA B Picea koyamaey D ACTTACTTTACTT-----GCTTTACTTGAGACTTGAGTGGCA J 1 PiceaPicea abies AustriaAustria 1 Clade I D * P. abies Asia + 3 European species ACTTACTT------AAGTGGCA Larix PiceaPicea abiesabies SerbiaSerbia 1 D Hercyno-Carpathian ACTTACTT------AG--GCA Pseudotsuga PiceaPicea abiesabies SerbiaSerbia 2 D * and Alpine domains ACTTACTT------GCTT------GAGTGGCA Tsuga * PiceaPicea abiesabies AustriaAustria 2 D ACTTACTT------GAGTGGCA Abies, Cathaya, PiceaPicea abiesabies GermanyGermany D Cedrus, Keteleeria, 1 PiceaPi jezoensisji D D Nothotsuga, Pinus, Picea torano D Pseudolarix * * Picea maximowiczii D Picea omorika D * Picea alcoquiana D Picea orientalis D 1 Picea glauca B 1 Picea engelmannii B 1 A 0.98 Picea mexicana A Picea pungens Clade II A 1 North America 1 Picea mariana C 1 Picea rubens C A A 1 var. rubescens E 0.98 Picea likiangensis var. linzhiensis E 0.98 Picea likiangensis var. hirtella E Picea likiangensis var. likiangensis E 1 Picea farreri E 1 Picea spinulosa E Picea brachytyla Gansu, China E 1 Picea brachytyla , China Clade III E 1 Picea neoveitchii Asia + E 1 North American species J Pinoideae 1 1 E 1 E 1 * E Picea morrisonicola D Picea breweriana D 1 1 Pinus pumila 1 Pinus cembra Pinus sylvestris Cathaya argyrophylla 1 Pseudotsuga menziesii 1 Abietoideae Keteleeria 0.98 Abies 0.98 1 Tsuga canadensis Nothotsuga longibracteata 1 Cedrus atlantica

0.0090

Fig. 3. Bayesian consensus tree based on 9457 plastid, mitochondrial and nuclear nucleotides including coded indels. Numbers at nodes refer to posterior probability (PP) values P0.98. The asterisks mark nodes that have high support (P0.98 PP) if P. koyamae and P. torano, which lack nad1 and 4CL, are removed from the data matrix; support for the remaining nodes remained unaffected. Letters on the right margin represent nad5 intron 1 mitotypes (see inset). Inset shows the six nad5 intron 1 mitotypes detected in Picea and the homologous sequences in outgroup genera.

(P. chihuahuana, P. engelmannii, P. glauca,P. mariana, P. martinezii, closely related to P. spinulosa than it is to P. brachytyla from Yun- P. mexicana, P. pungens, P. rubens, P. sitchensis) have mitotypes A, nan (Fig. 3). B, or C, and the chiefly Asian clade (III) has mitotypes D, E, and J. All members of clade I lack section B of nad1, while members 3.3. Divergence times and substitution rates in Pinaceae and Picea of clades II and III except P. morrisonicola lack section A. Clade I, centered in Asia with three European species, includes A chronogram from a relaxed clock calibrated with Pinuxylon the five Japanese endemics, P. alcoquiana, P. glehnii, P. koyamae, P. and Pseudolarix fossils is shown in Fig. 4. Results from BEAST runs maximowiczii, and P. torano, which did not originate from a single with an empty alignment revealed no contradictions among the radiation. Norway spruce, P. abies, is not monophyletic because its prior constraints and showed that the posterior age distributions Baltico–Nordic populations (Fig. 1) are more closely related to were substantially influenced by the signal in the data (Fig. S1). Asian species of the P. asperata complex (P. asperata, P. crassifolia, With the Pinuxylon and Pseudolarix constraints (gray column in Ta- P. koraiensis, P. koyamae, P. meyeri, P. obovata, and P. retroflexa) ble 1, which also shows the 95% confidence intervals for inferred than to southern Hercyno-Carpathian and Alpine populations of ages), all other Pinaceae divergence events are inferred to times Norway spruce. In clade II, from North America, the western P. predating their earliest reliable fossil record (Table S3). For exam- martinezii is the first to diverge, followed by the nine remaining ple, the oldest conclusive evidence of Larix (Middle to Late , species. The widespread P. mariana and the eastern P. rubens both which would be 44–34 Ma; LePage and Basinger, 1991) post-dates possess mitotype C; the widespread P. glauca and the western P. the inferred time of its divergence from Pseudotsuga, 75 Ma. engelmannii both possess mitotype B. The endemic P. mexicana The Pinaceae crown group emerged in the Upper Triassic, and and the more widespread P. pungens are sister species; the other the divergence between Picea and Pinus occurred in the Lower two Mexican species (P. chihuahuana, P. martinezii) are not closely Jurassic (c. 180 Ma; but inferring these deep splits was not the focus related. Clade III contains another instance of species non-mono- of our study). The Picea crown group emerged in the mid-Oligocene, phyly because P. brachytyla material from Gansu, China is more at 28 Ma (95% CI: 37–21 Ma). Most Picea speciation events date to J.D. Lockwood et al. / Molecular Phylogenetics and Evolution 69 (2013) 717–727 723

the Miocene. The North American clade (II) first diversified 20 (27–14) Ma, the Eurasian clade (I) 23 (32–15) Ma. The Hercyno- Carpathian and Alpine (southern European) lineage of P. abies split from the Baltico–Nordic (northern) lineage 6 Ma. Relaxed clock models with the preferred fossil calibration scheme yielded a matK substitution rate (subs/site/Mya) of 1.92 104 and an rbcL rate of 1.08 104 (Table 1, rates ob- tained with either one of the two calibration fossils are also shown). We also tested the matK and rbcL substitution rates in- ferred by Gernandt et al. (2008; our Table 1).

3.4. Ancestral area reconstruction ). Node ages are median values in millions of

Results of the two AAR approaches, i.e., reconstruction under the Mk1 model in Mesquite and reconstruction under the DEC model in Lagrange, are shown in Table S5. A DEC model with 19 time slices (Section 2) had a global likelihood of 71.11, and

Gernandt et al., 2008 one with 11 time slices 71.27, and without time slices 70.39. We therefore preferred the simplest model. Likelihood values of biogeographic scenarios for ingroup nodes of interest were not significantly affected by the number of time slices or ambiguity in the coding of outgroup ancestral areas (Section 2; Table S5). Using the topology obtained in the Bayesian runs, the most re- results obtained under the preferred calibration scheme. cent common ancestor of today’s species of Picea is inferred to have lived in Asia; using the ML topology in which P. breweriana

inaceae substitution rates ( is sister to all other species yields an ambiguous reconstruction for the root as Asia/North America. Using either tree, the stem lineage of the North American clade (II) apparently entered North America before 20 Ma (Fig. 4, Table 1). Under the Mk1 model, it came from Asia (likelihood of 0.86), while under the 1-time-slice DEC model it lived in Asia and western North America (likelihood of 0.92). The stem lineage of the Asian/North American clade (III), under the Mk1 model, may have occurred in Asia (likelihood of 0.87), while under the 1-time-slice DEC model it occurred in Asia and North America (likelihood of 0.99). The crown node of clade I (the Eurasian clade) under both models existed in Asia (likeli- hoods of 0.97 under Mk1 and 0.90 under DEC). The ancestral area of the clade comprising P. omorika, endemic in Serbia, and P. ori- entalis endemic in the Caucasus region (Figs. 1 and 3), is inferred as Asia (likelihood of 0.83).

4. Discussion

By using signal from large motifs in mitochondrial introns we obtained a well-resolved phylogeny of Picea in which most nodes have strong statistical support. Different from previous molecular phylogenies of the genus, we find three geographically coherent major groups, an American clade with nine species, a predomi- nantly Asian clade, and a Eurasian clade. Surprisingly, our data also suggest the paraphyly of Norway spruce. We now discuss these results in turn.

4.1. Mitochondrial indels under a stochastic Dollo model

Group II introns in the mitochondrial NADH genes, specifically the first intron of nad5 and the second intron of nad1, have long been useful in conifer phylogenetics (Grivet et al., 1999; Gugerli et al., 2001a, 2001b; Sperisen et al., 2001; Ran et al., 2006; Meng et al., 2007; Bouillé et al., 2011). However, aligning length muta- tions in variable domains of introns or spacers is not straightfor- ward (Kelchner, 2000). Previous work by Aleksic´ (2008) had revealed two indels of up to 1500 bp and several complex motifs of about 30 bp in domain IV of nad1 intron 2 that were conserved at the level of species groups, suggesting that introns would be Table 1 Divergence times and substitution rates under strict and relaxed clock models using alternative fossil calibration points or previously inferred P years with 95% confidence intervals shown below. Nucleotide substitution rates are in substitutions/site/million years. Marked in pale gray are the phylogenetically useful. We applied simple indel coding, which 724 J.D. Lockwood et al. / Molecular Phylogenetics and Evolution 69 (2013) 717–727 yielded a binary matrix on which we then used the stochastic Dollo treatment (1990) recognized P. crassifolia as a distinct species, model, appropriate for characters that are unlikely to be regained while Eckenwalder (2009) treats this entity as a variety of P. once lost, which likely is true of the long motifs coded here. The schrenkiana (in clade III in Fig. 3). In our tree it is a member of impact of indel characters was assessed in analyses with and with- the P. asperata complex (clade I in Fig. 3), rejecting the latter place- out coded mitochondrial indels. Results showed that the indel ment. Conversely, Eckenwalder’s (2009) recognition of P. mexicana characters support species groups also seen without coded indels, as a species is supported (Fig. 3). albeit without high statistical support. Population genetic studies of spruces have used some of the 4.3. Biogeography of the major clades of Picea same mt introns (albeit aligned without taking advantage of sec- ondary structure) to infer progenitor–derivative relationships. For Using the preferred topology shown in Fig. 4, in which P. brewe- example, nad1 and nad5 (plus sequence-tagged-site markers of riana is placed in clade III, we inferred that the immediate ancestor expressed genes) were used to infer an ancestor–descendant rela- of the living species of Picea lived in Asia as first proposed by tionship between P. mariana and P. rubens because the latter Wright (1955); with the ML topology in which P. breweriana is sis- species has a subset of the diversity found in the former (Perron ter to all other species the root is reconstructed as Asia/North et al., 2000; Jaramillo-Correa and Bousquet, 2003). In our tree dia- America. Three previous molecular studies (Sigurgeirsson and gram (Fig. 2), they are sister species and uniquely share nad5 mito- Szmidt, 1993; Ran et al., 2006; Bouillé et al., 2011) instead inferred type C. an evolution of Picea in North America, based on the finding that in plastid gene trees western North American species were the first to 4.2. Recent origin and diversification of Picea diverge. Ran et al. (2006) also inferred a back-dispersal from Asia to North America (ancestor of P. chihuahuana) and one from North Spruces provide yet another example of a relatively young America to Europe (ancestor of P. omorika). This last event they diversification atop an unbranched stem lineage. Such stem themselves considered ‘‘nearly impossible’’ (Ran et al., 2006: lineages are thought to reflect major (Harvey et al., 414). Our reconstruction implies two entries from Asia into North 1994). In gymnosperms, extinctions occurred when the global tem- America, likely through Beringia between 25 and 20 Ma, and no peratures declined sharply at the end of the Eocene, followed by re- back-dispersal. radiations during the late Oligocene and early Miocene warming One of the two entries involves the stem lineage of the North (Crisp and Cook, 2011; Nagalingum et al., 2011; Leslie et al., American clade (II), the other the ancestor of clade III, which com- 2012; Mao et al., 2012). The three regional radiations in spruces prises only Asian species and the North American P. breweriana. (clades I, II, III) all fall in this warmer period, between 25 and The latter species is genetically and morphologically highly distinct 20 Ma (Zachos et al., 2001). (Schmidt-Vogt, 1977; Sigurgeirsson and Szmidt, 1993; Weng and Regardless of calibration method and clock model, the most re- Jackson, 2000; Ledig et al., 2004; Ran et al., 2006; Bouillé et al., cent common ancestor of living Picea species was dated to about 2011). With Bayesian inference it formed an early-branching line- 28 Ma (37.2–20.6 Ma, Table 1). Previous studies dated the Picea age in clade III, albeit not as sister to all other spruces as found by crown group to c. 30 Ma (using two species selected to span the Sigurgeirsson and Szmidt (1993), Ran et al. (2006), and Bouillé deepest divergence; Lin et al., 2010) or to 13–20 Ma (with three et al. (2011). With maximum likelihood, however, P. breweriana species, Bouillé and Bousquet, 2005). Leslie et al. (2012: Fig. S4C) was placed as sister to all other spruces, but with low statistical included most species and inferred an age of c. 23 Ma. The age ob- support (tree not shown). That P. breweriana might be part of a tained by Crisp and Cook (2011) was 5.8 Ma, which seems too southern Asian clade would fit with it being ‘‘most similar to some young. Coalescence dating that assumed an average generation south Chinese species [and it likely being] an offshoot of an earlier time in Picea of 50 years and effective population sizes between migration than the one that give rise to the other spruces of north- 100,000 and 200,000 individuals yielded divergence times that west America’’ (Wright, 1955: 328). are younger than those obtained here. Thus, an analysis of 12–16 The regions of highest Picea diversity are the western Cordillera nuclear loci dated the split between P. likiangensis and P. schrenki- of the Rocky Mountains and the Qinghai–Tibetan Plateau (QTP). ana to 10.9 Ma (Li et al. 2010), while we estimated 19.7 Ma for this Western North America (including Mexico) has more distinct node. The same study dated the split between P. purpurea and P. spruces than recognized before this study. Thus, P. chihuahuana schrenkiana to 3.2 Ma, while we dated this split to 11.5 Ma. Picea and P. martinezii are neither close to Asian species (Sigurgeirsson generation times are poorly known and may be 25 or 50 years and Szmidt, 1993; Ran et al., 2006; Bouillé et al., 2011) nor closely (Chen et al., 2010). The different effective population sizes and related to P. glauca [contra Schmidt-Vogt (1977) and Farjón mutation rates of the nuclear and organellar genomes also mean (1990)]. Mexican spruce (P. mexicana, Fig. 3) is not a subspecies that coalescence ages estimated from nuclear data are several of P. engelmannii [contra Schmidt (1988) and Farjón (1990, times higher than those inferred from plastid data (Li et al., 2001)], but instead a sister to the western North American P. pun- 2012), a point that may be insufficiently accommodated in our gens. For the QTP region, studies at the intra- and interspecific lev- molecular clocks. The Pinaceae plastid substitution rates of 1.0 to els have indicated high levels of population differentiation, 1.9 10-4 subs/site/Mya years obtained here are identical to plas- probably attributable to the topography of the QTP, where deep tid rates inferred for other woody angiosperms, for example, le- valleys and high mountain peaks act as barriers to gene flow (Li gumes (Wojciechowski, 2005; Scherson et al., 2008), but are et al., 2010). Clade III (Fig. 3), which comprises nine to ten species slower than the 1.0 to 2.8 subs/site/Mya rates inferred by Gernandt in the QTP region, diversified around 19.8 Ma. Several QTP species et al. in a Pinus-focused study (2008). share mt and cp haplotypes (Du et al., 2011; Zou et al., 2012), per- Wright’s (1955) observation that morphological variation in all haps because it takes 18.75 Mya for 95% of organelle genes to be- of Picea is comparable to that found in species groups of Pinus fits come species-specific, assuming a generation time of 25 years with its young age. Cones of the P.burtonii (Fig. 4 and an effective population size of 150,000 (Chen et al., 2010). inset), the Oligocene P.diettertiana (Miller, 1970), and extant Assuming a generation time of 50 years, it would take 37.5 Mya. species show little change during the last 136 Ma of the genus An unexpected finding is the grouping of the Serbian endemic P. (Klymiuk and Stockey, 2012). While this stasis may facilitate the omorika, the Caucasian P. orientalis, and the two Japanese endemics generic placement of fossils, it reduces the phylogenetic utility of P. alcoquiana and P. maximowiczii, a clade that has 1.0 PP (Fig. 3). morphological characters in living Picea. Thus, Farjón’s taxonomic Since the 4CL sequence of P. alcoquiana is incomplete, this clade J.D. Lockwood et al. / Molecular Phylogenetics and Evolution 69 (2013) 717–727 725

PiPPiceacea asasperataperata PiceaPiP cea memeyeriyeri PiceaP abies NNorwayorway PiceaP obovataobovata Europe Asia PiPPiceacea koyamaekoyamae PiceaP abies SerbiaSerbia Western Eastern PiPPiceacea jezoensisjezoensis North America North America PiPPiceacea gglehniilehnii PiPPiceacea torano PiceaP mmaximowicziiaximowiczii PiceaP oomorikamorika PiPPiceacea aalcoquianalcoquiana PiceaP orientalisorientalis Clade I PiPPiceacea gglaucalauca PiPPiceacea engeengelmanniilmannii PiceaP sisitchensistchensis PiceaPiP cea pungenpungenss PiceaPiP cea mexmexicanaicana PiceaPiP cea rurubensbens Picea burtonii PiceaP mmarianaariana 136 Ma PiceaP chihuahuanachihuahuana PiceaPiP cea martmartineziiinezii Clade IIII PiceaPiP cea likiangensislikiangensis varvar.. lilinzhiensisnzhiensis PiceaPiP cea likiangensislikiangensis var. rrubescensubescens PPiceaicea likiangensilikiangensiss varvar.. hhirtellairtella PiceaPicea likianlikiangensisgensis varvar.. likiangensis PiceaPicea farrerifarreri PiceaPicea sspinulosapinulosa PiceaPicea brachytyla GansuGansu,, China PiPPiceacea bbrachytylarachytyla YYunnan,unnan, CChinahina PiceaP nneoveitchiieoveitchii PiceaP ssmithianamithiana PiPPiceacea ppurpureaurpurea PiPPiceacea wwilsoniiilsonii PiceaPicea schrenkiana PiceaP mmorrisonicolaorrisonicola PiceaPiP cea bbrewerianareweriana CladeClade IIIIII Pinus pumila Pinus cembra Pinus sylvestris Cathaya argyrophylla Larix gmelinii Pseudotsuga menziesii Keteleeria Abies Tsuga canadensis Nothotsuga longibracteata Pseudolarix amabilis Cedrus atlantica

200 150 100 50 0 Ma

Upper Lower Middle Upper Lower Upper Pal. Eoc. Oli. Mio. P Epoch O

Triassic Jurassic Cretaceous Paleogene Neogene Period

Fig. 4. Chronogram of Pinaceae based on two fossil constraints (stars) and ancestral area reconstruction (AAR) without constrained connection probabilities. Gray bars indicate the 95% highest posterior probability of the node age estimates. Pal. = Paleocene, Eoc. = Eocene, Oli. = Oligocene, PO = Pliocene. Inset image shows Picea burtonii from the Early Cretaceous (c. 136 Ma) of Vancouver Island (Klymiuk and Stockey, 2012). needs further investigation. In the chronogram, it is 16.5 Mya old may have several explanations. First, different populations of these (Fig. 4), the time of the mid-Miocene climatic optimum (17– species could have originated from different ancestors but be 15 Ma; Zachos et al., 2001). All four species currently occur at high morphologically so convergent that they look like a single species. elevations (700–2200 m) where annual precipitation ranges from This may the case for P. mexicana, which closely resembles 1000 to 2500 mm and winters are cold, frequently with snow P. engelmannii. Second, because some (but not all) spruce species (Farjón, 1990). Since Picea is generally adapted to boreal and occurring in the same geographic region diversified recently, montane habitats, the increasing seasonality and aridity that arose incomplete lineage sorting and hybridization might cause conflict during the late Miocene (10–5 Ma) may have led to widespread between nuclear and organellar gene trees. The distributional extinction of Picea in the mid latitudes of Eurasia. This greater ranges of spruce species were greatly affected by the last glacial seasonality and aridity also led to the fragmentation of a once con- period from approximately 110,000–10,000 years ago (Tollefsrud tinuous ruminant ungulate fauna ranging from Iberia to central et al., 2008; Parducci et al., 2012), and range retreats and expan- Asia during the early Miocene (23–15 Ma; Fortelius et al., 2002). sions may have caused secondary contact of previously isolated incipient species and led to introgression (Du et al., 2011). This 4.4. Species non-monophyly: Convergence, incomplete lineage sorting, may have occurred in P. abies and P. brachytyla. or interspecific introgressions? Picea abies is a highly polymorphic species, and no consensus on its circumscription has been reached. Thus, P. obovata, which is At least three species, Norway spruce, P. abies, the Chinese part of the P. asperata complex (P. abies, P. asperata, P. crassifolia, P. brachytyla, and the American P. engelmannii are not monophy- P. koraiensis, P. koyamae, P. meyeri, and P. retroflexa; Fig. 1), has letic in their current circumscription (Farjón, 1990, 2001). This been considered a ‘‘minor species’’ near P. abies (Farjón, 1990: 726 J.D. Lockwood et al. / Molecular Phylogenetics and Evolution 69 (2013) 717–727

229), an ‘‘eastern form of P. abies’’ (Schmidt-Vogt, 1974: 195, 1977: Aleksic´, J.M., Geburek, T., in press. Quaternary population dynamics of an endemic 21), or a distinct species (Farjón, 2001; Eckenwalder, 2009). Studies conifer, Picea omorika, in its last refugium in the Balkans, and conservation implications. Conservation Genetics. of P. abies nuclear markers and organelle DNA have consistently re- Bouillé, M., Bousquet, J., 2005. Trans-species shared polymorphisms at orthologous vealed a clear separation between the northern (Baltico–Nordic; nuclear gene loci among distant species in the conifer Picea (Pinaceae): Figs. 1 and 3) and southern (Hercyno-Carpathian, Alpine) popula- implications for the long-term maintenance of genetic diversity in trees. Am. J. Bot. 91, 63–73. tions (e.g., Lagercrantz and Ryman 1990; Sperisen et al., 2001; Col- Bouillé, M., Senneville, S., Bousquet, J., 2011. Discordant mtDNA and cpDNA lignon et al., 2002; Tollefsrud et al., 2008). In our data, the phylogenies indicate geographic speciation and reticulation as driving factors distinctness of the southern populations is seen in seven non- for the diversification of the genus Picea. Tree Genet. Genomes 7, 469–484. Chen, J., Källman, T., Gyllenstrand, N., Lascoux, M., 2010. New insights on the homoplastic point mutations, in indel motifs in nad1 (Sperisen speciation history and nucleotide diversity of three boreal spruce species and a et al., 2001; our Fig. 2). Tertiary relict. Heredity 104, 3–14. So far, studies of P. abies phylogeography (Tollefsrud et al., Christenhusz, M.J.M., Reveal, J.L., Farjon, A., Gardner, M.F., Mill, R.R., Chase, M.W., 2011. A new classification and linear sequence of extant gymnosperms. 2008; Parducci et al., 2012) have either been rooted with P. obov- Phytotaxa 19, 5570. ata, which in cladistic terms is part of the ingroup, or also with Cohen, K.M., Finney, S., Gibbard, P.L., 2012. International chronostratigraphic chart. North American P. glauca, which is too distant to infer the direction . of population expansion. Proper rooting would require inclusion of Collignon, A.-M., Van de Sype, H., Favre, J.-M., 2002. Geographical variation in random amplified polymorphic DNA and quantitative traits in Norway spruce. the Asian species P. glehnii or P. jezoensis (Fig. 2). If one were to di- Can. J. Res. 32, 266–282. vide P. abies into two species, northern P. abies (including all names Crisp, M.D., Cook, L.G., 2011. Cenozoic extinctions account for the low diversity of published in the P. asperata complex) would remain P. abies sensu extant gymnosperms compared with angiosperms. New Phytol. 192, 997–1009. Drummond, A.J., Suchard, M.A., Xie, D., Rambaut, A., 2012. Bayesian phylogenetics stricto because Linnaeus described Scandinavian material. Regard- with BEAUti and the BEAST 1.7. Mol. Biol. and Evol. first published online: less of naming issues, a possible non-monophyly of P. abies means February 25, 2012, doi: 10.1093/molbev/mss075. that the history of this important species, the genome of which has Du, F.K., Petit, R.J., Liu, J.Q., 2009. More introgression with less gene flow: chloroplast vs. mitochondrial DNA in the Picea asperata complex in China, and just become available (Nystedt et al., 2013), has not been fully comparison with other conifers. Mol. Ecol. 18, 1396–1407. understood. Our dating shows that the northern and southern lin- Du, F.K., Peng, X.L., Liu, J.Q., Lascoux, M., Hu, F.S., Petit, R.J., 2011. Direction and eages of P. abies separated 6 Ma, distinctly prior to the ice ages extent of organelle DNA introgression between two spruce species in the Qinghai-Tibetan Plateau. New Phytol. 192, 1024–1033. (2.5 Ma). Pleistocene refugia are thus unlikely to be the explana- Eckenwalder, J.E., 2009. Conifers of the World: The Complete Reference, first ed. tion for the relationships between the European species (P. abies, Timber Press, Portland. P. omorika, P. orientalis) and their East Asian relatives revealed in Eckert, A.J., Hall, B.D., 2006. Phylogeny, historical biogeography, and patterns of diversification for Pinus (Pinaceae): phylogenetic tests of fossil-based this study. However, the interspecific introgressions documented hypotheses. Mol. Phylogenet. Evol. 40, 166–182. in Picea (Section 1), incomplete lineage sorting in large outbreeding eFloras, 2008. Missouri Botanical Garden, St. Louis, MO and Harvard University populations, and differences in the effective sizes and mutation Herbaria, Cambridge, MA. Published on the Internet rates of the sampled DNA regions, mean that our findings concern- [accessed 18 June 2012]. Farjón, A., 1990. Pinaceae: Drawings and Descriptions of the ing species relationships and divergence times need further confir- GeneraAbies,Cedrus,Pseudolarix,Keteleeria,Nothotsuga,Tsuga,Cathaya,Pseudotsuga, mation, especially based on larger genomic data. Larix and Picea. Koeltz Scientific Books, Königstein. Farjón, A., 2001. World Checklist and Bibliography of Conifers, second ed. Royal Botanic Gardens, Kew, London. Acknowledgments Fortelius, M., Eronen, J., Jernvall, J., Liu, L., Pushkina, D., Rinne, J., Tesakov, A., Vislobokova, I., Zhang, Z., Zhou, L., 2002. Fossil mammals resolve regional patterns of Eurasian climate change over 20 million years. Evol. Ecol. Res. 4, We thank U. Pietzarka at the Tharandt Botanical Garden, B. Ja- 1005–1016. quish at the Kalamalka Research Station, and M. Tollefsrud at the Fu, L., Li, N., Mill, R.R., 1999. Picea. In: Wu, Z.-Y., Raven, P.H. (Eds.), Flora of China (4). Norwegian Forest and Landscape Institute for samples; M. Silber Science Press, and Missouri Botanical Garden Press, St. Louis, pp. 25–32. for assistance in the laboratory; B. LePage for discussion of fossils; Gernandt, D.S., Magallón, S., López, G.G., Flores, O.Z., Willyard, A., Liston, A., 2008. Use of simultaneous analyses to guide fossil-based calibrations of Pinaceae B. Moore for a matrix of past continental connection probabilities; phylogeny. Int. J. Plant Sci. 169, 1086–1099. and G. Holman, A. Klein, and C. Campbell for information on Picea Grivet, D., Jeandroz, S., Favre, J.M., 1999. Nad1 b/c intron polymorphism reveals accessions sampled in earlier studies. Financial support came from maternal inheritance of the mitochondrial genome in Picea abies. Theor. Appl. Genet. 99, 346–349. the Ecology, Evolution, and Systematics program of Munich Uni- Gugerli, F., Senn, J., Anzidei, M., Madaghiele, A., Büchler, U., Sperisen, C., Vendramin, versity, the Ministry of Education and Science of the Republic of G.G., 2001a. Chloroplast microsatellites and mitochondrial nad1 intron 2 Serbia (Research Grant 173030 to JA), and the National Natural Sci- sequences indicate congruent phylogenetic relationships among Swiss stone (Pinus cembra), Siberian stone pine (Pinus sibirica), and Siberian dwarf pine ence Foundation of China (Research Grant 30930072 to LJQ) and (Pinus pumila). Mol. Ecol. 10, 1489–1497. the Key Project of International Collaboration Program, the Minis- Gugerli, F., Sperisen, C., Büchler, U., Brunner, I., Brodbeck, S., Palmer, J.D., Qiu, Y.L., try of Science and Technology of China (Research Grant 2001b. The evolutionary split of Pinaceae from other conifers: evidence from an intron loss and a multigene phylogeny. Mol. Phylogenet. Evol. 21, 167–175. 2010DFB63500 to LJQ). Harvey, P.H., May, R.M., Nee, S., 1994. Phylogenies without fossils. Evolution 48, 523–529. Havill, N.P., Campbell, C.S., Vining, T.F., LePage, B., Bayer, R.J., Donoghue, M.J., 2008. Appendix A. Supplementary material Phylogeny and biogeography of Tsuga (Pinaceae) inferred from nuclear ribosomal ITS and chloroplast DNA sequence data. Syst. Bot. 33, 478–489. Supplementary data associated with this article can be found, in He, T., Pausas, J.G., Belcher, C.M., Schwilk, D.W., Lamont, B.B., 2012. Fire-adapted traits of Pinus arose in the fiery Cretaceous. New Phytol. 194, 751–759. the online version, at http://dx.doi.org/10.1016/j.ympev.2013.07. Ickert-Bond, S.M., Rydin, C., Renner, S.S., 2009. A fossil-calibrated relaxed clock for 004. Ephedra indicates an Oligocene age for the divergence of Asian and New World clades, and Miocene dispersal into South America. J. Syst. Evol. 47, 444–456. Jaramillo-Correa, J.P., Bousquet, J., 2003. New evidence from mitochondrial DNA of a References progenitor-derivative species relationship between black spruce and red spruce (Pinaceae). Am. J. Bot. 90, 1801–1806. Aleksic´, J.M., 2008. Genetic structure of natural populations of Serbian spruce [Picea Jaramillo-Correa, J.P., Bousquet, J., Beaulieu, J., Isabel, N., Perron, M., Bouillé, M., omorika (Pancˇ.) Purk.]. PhD thesis, University of Natural Resources and Applied 2003. Cross-species amplification of mitochondrial DNA sequence-tagged-site Life Sciences, Vienna, Austria. . among species in Picea. Theor. Appl. Genet. 106, 1353–1367. Aleksic´, J.M., Geburek, T., 2010. Mitochondrial DNA reveals complex genetic Kelchner, S.A., 2000. The evolution of noncoding chloroplast DNA and its application structuring in a stenoendemic conifer Picea omorika [(Pancˇ.) Purk.] caused by in plant systematics. Ann. MO Bot. Gard. 87, 482–498. its long persistence within the refugial Balkan region. Plant Syst. Evol. 285, 1– Keller, A.M., Hendrix, M.S., 1997. Paleoclimatologic analysis of a Late Jurassic 11. petrified forest, southeastern Mongolia. Palaios 12, 282–291. J.D. Lockwood et al. / Molecular Phylogenetics and Evolution 69 (2013) 717–727 727

Klymiuk, A.A., Stockey, R.A., 2012. A Lower Cretaceous (Valanginian) cone Alsos, Willerslev, E., . Glacial survival of boreal trees in northern Scandinavia. provides the earliest fossil record for Picea (Pinaceae). Am. J. Bot. 99, 1069– Science 335, 1083–1086. 1082. Perron, M., Bousquet, J., 1997. Natural hybridization between black spruce and red Knapp, M., Mudaliar, R., Havell, D., Wagstaff, S.J., Lockhart, P.J., 2007. The drowning spruce. Mol. Ecol. 6, 725–734. of New Zealand and the problem of Agathis. Syst. Biol. 56, 862–870. Perron, M., Perry, D.J., Andalo, C., Bousquet, J., 2000. Evidence from sequence- Krassilov, V., 1982. Early Cretaceous flora of Mongolia. Palaeontogr. Abt. B. 181, 1– tagged-site markers of a recent progenitor-derivative species pair in conifers. 43. Proc. Nat. Acad. Sci. USA 97, 11331–11336. Krutovskii, K.V., Bergmann, F., 1995. Introgressive hybridization and phylogenetic Posada, D., 2008. JModelTest: phylogenetic model averaging. Mol. Biol. Evol. 25, relationships between Norway, Picea abies (L.) Karst., and Siberian, P. obovata 1253–1256. Ledeb., spruce species studied by isozyme loci. Heredity 74, 464–480. Qiao, C.-Y., Ran, J.-H., Li, Y., Wang, X.-Q., 2007. Phylogeny and biogeography of Lagercrantz, U., Ryman, N., 1990. Genetic structure of Norway spruce (Picea abies): Cedrus (Pinaceae) inferred from sequences of seven paternal chloroplast and concordance of morphological and allozymic variation. Evolution 44, 38–53. maternal mitochondrial DNA regions. Ann. Bot. 100, 573–580. Latałowa, M., van der Knaap, W.O., 2006. Late Quaternary expansion of Norway Rajora, O.P., Dancik, B.P., 2000. Population genetic variation, structure, and spruce Picea abies (L.) Karst. in Europe according to pollen data. Quaternary Sci. evolution in Engelmann spruce, white spruce, and their natural Rev. 25, 2780–2805. complex in Alberta. Can. J. Bot. 78, 768–780. Ledig, F.T., Hodgkiss, P.D., Krutovskii, K.V., Neale, D.B., Eguiluz-Piedra, T., 2004. Rambaut, A., Drummond, A.J. 2007. Tracer v1.4. . Relationships among the spruces (Picea, Pinaceae) of southwestern North Ran, J.H., Wei, X.X., Wang, X.Q., 2006. Molecular phylogeny and biogeography of America. Syst. Bot. 29, 275–295. Picea (Pinaceae): implications for phylogeographical studies using cytoplasmic LePage, B.A., Basinger, J.F., 1991. A new species of Larix (Pinaceae) from the early haplotypes. Mol. Phylogenet. Evol. 41, 405–419. Tertiary of Axel Heiberg Island, Arctic Canada. Rev. Palaeobot. Palynol. 70, 89– Ree, R.H., Smith, S.A., 2008. Maximum likelihood inference of geographic range 111. evolution by dispersal, local extinction, and cladogenesis. Syst. Biol. 57, 4–14. LePage, B.A., Basinger, J.F., 1995. History of the genus Pseudolarix Gordon (Pinaceae). Scherson, R.A., Vidal, R., Sanderson, M.J., 2008. Phylogeny, biogeography, and rates Int. J. Plant Sci. 156, 910–950. of diversification of New World Astragalus (Leguminosae) with an emphasis on Leslie, A.B., Beaulieu, J.M., Rai, H.S., Crane, P.R., Donoghue, M.J., Mathews, S., 2012. South American radiations. Am. J. Bot. 95, 1030–1039. Hemisphere-scale differences in conifer evolutionary dynamics. Proc. Natl. Schmidt, P.A., 1988. Taxonomisch-nomenklatorische Notiz zur Gattung Picea A. Acad. Sci. USA 109, 1–5. Dietr.. Haussknechtia 4, 37–38. Lewis, P.O., 2001. A likelihood approach to estimating phylogeny from discrete Schmidt-Vogt, H., 1974. Das natürliche Verbreitungsgebiet der Fichte (Picea abies morphological character data. Syst. Biol. 50, 913–925. [L.] Karst.) in Eurasien. Allg. Forst Jagdztg 145, 185–197. Li, Y., Stocks, M., Hemmilä, S., Källman, T., Zhu, H., Zhou, Y., Chen, J., Liu, J., Lascoux, Schmidt-Vogt, H., 1977. Die Fichte, Band I. Verlag Paul Parey, Hamburg and Berlin. M., 2010. Demographic histories of four spruce (Picea) species of the Qinghai- Sigurgeirsson, A., Szmidt, A., 1993. Phylogenetic and biogeographic implications of Tibetan Plateau and neighboring areas inferred from multiple nuclear loci. Mol. chloroplast DNA variation in Picea. Nord. J. Bot. 13, 233–246. Biol. Evol. 27, 1001–1014. Silvestro, D., Michalak, I., 2011. raxmlGUI: a graphical front-end for RAxML. Org. Li, Z., Zou, J., Mao, K., Lin, K., Li, H., Liu, J., Källman, T., Lascoux, M., 2012. Population Divers. Evol.. genetic evidence for complex evolutionary histories of four high altitude Simmons, M.P., Ochoterena, H., 2000. Gaps as characters in sequence based juniper species in the Qinghai-Tibetan Plateau. Evolution 66, 831–845. phylogenetic analyses. Syst. Biol. 49, 369–381. Lin, C.-P., Huang, J.-P., Wu, C.-S., Hsu, C.-Y., Chaw, S.-M., 2010. Comparative Sperisen, C., Büchler, U., Gugerli, F., Mátyás, G., Geburek, T., Vendramin, G.G., 2001. chloroplast genomics reveals the evolution of Pinaceae genera and subfamilies. Tandem repeats in plant mitochondrial genomes: application to the analysis of Genome Biol. Evol. 2, 504–517. population differentiation in the conifer Norway spruce. Mol. Ecol. 10, 257–263. Liu, Y.S., Basinger, J.F., 2000. Fossil Cathaya (Pinaceae) pollen from the Canadian Stamatakis, A., 2006. RAxML-VI-HPC: maximum likelihood-based phylogenetic High Arctic. Int. J. Plant Sci. 161, 829–847. analyses with thousands of taxa and mixed models. Bioinformatics 22, 2688– Maddison, W.P., Maddison, D.R. 2011. Mesquite: A Modular System for Evolutionary 2690. Analysis. Version 2.75 . Sutton, B.C.S., Flanagan, D.J., Gawley, J.R., Newton, C.H., Lester, D.T., El- Kassaby, Y.A., Mao, K., Hao, G., Liu, J., Adams, R.P., Milne, R.I., 2010. Diversification and 1991. Inheritance of chloroplast and mitochondrial DNA in Picea and biogeography of Juniperus (Cupressaceae): variable diversification rates and composition of hybrids from introgression zones. Theor. Appl. Genet. 82, 242– multiple intercontinental dispersals. New Phytol. 188, 254–272. 248. Mao, K., Milne, R.I., Zhang, L., Peng, Y., Liu, J., Thomas, P., Mill, R.R., Renner, S.S., 2012. Tollefsrud, M.M., Kissling, R., Gugerli, F., Johnsen, Ø., Skrøppa, T., Cheddadi, R., van Distribution of living Cupressaceae reflects the breakup of Pangea. Proc. Natl. der Knaap, W.O., Latałowa, M., Terhürne-Berson, R., Litt, T., Geburek, T., Acad. Sci. USA 109, 7793–7798. Brochmann, C., Sperisen, C., 2008. Genetic consequences of glacial survival Meijer, J., 2000. Fossil woods from the Late Cretaceous Aachen Formation. Rev. and postglacial colonization in Norway spruce: combined analysis of Palaeobot. Palyno. 112, 297–336. mitochondrial DNA and fossil pollen. Mol. Ecol. 17, 4134–4150. Meng, L., Yang, R., Abbott, R.J., Miehe, G., Hu, T., Liu, J., 2007. Mitochondrial and Treutlein, J., Wink, M., 2002. Molecular phylogeny of cycads inferred from rbcL chloroplast phylogeography of Picea crassifolia Kom. (Pinaceae) in the Qinghai- sequences. Naturwissenschaften 89, 221–225. Tibetan Plateau and adjacent highlands. Mol. Ecol. 16, 4128–4137. Wei, X.-X., Wang, X.-Q., 2003. Phylogenetic split of Larix: evidence from paternally Miller, C.N., 1970. Picea diettertiana, a new species of petrified cones from the inherited cpDNA trnT-trnF region. Plant Syst. Evol. 239, 67–77. Oligocene of western Montana. Am. J. Bot. 57, 579–585. Wei, X.-X., Yang, Z.-Y., Li, Y., Wang, X.-Q., 2010. Molecular phylogeny and Mogensen, H.L., 1996. The hows and whys of cytoplasmic inheritance in seed . biogeography of Pseudotsuga (Pinaceae): Insights into the floristic Am. J. Bot. 83, 383–404. relationship between Taiwan and its adjacent areas. Mol. Phylogenet. Evol. Moore, B.R., Donoghue, M.J., 2007. Correlates of diversification in the plant clade 55, 776–785. Dipsacales: geographic movement and evolutionary innovations. Am. Nat. 170 Weng, C., Jackson, S.T., 2000. Species differentiation of North American spruce (Suppl. 2), S28–55. (Picea) based on morphological and anatomical characteristics of needles. Can. J. Müller, K., 2005. SeqState: primer design and sequence statistics for phylogenetic Bot. 78, 1367–1383. DNA datasets. Appl. Bioinform. 4, 65–69. Willyard, A., Syring, J., Gernandt, D.S., Liston, A., Cronn, R., 2007. Fossil calibration of Nagalingum, N.S., Marshall, C.R., Quental, T.B., Rai, H.S., Little, D.P., Mathews, S., molecular divergence infers a moderate mutation rate and recent radiations for 2011. Recent synchronous radiation of a living fossil. Science 334, 796–799. Pinus. Mol. Biol. Evol. 24, 90–101. Nystedt, B., Street, N.R., Wetterbom, A., Zuccolo, A., Lin, Y.-C., Scofield, D.G., Vezzi, F., Wojciechowski, M.F., 2005. Astragalus (Fabaceae): a molecular phylogenetic Delhomme, N., Giacomello, S., Alexeyenko, A., Vicedomini, R., Sahlin, K., perspective. Brittonia 57, 382–396. Sherwood, E., Elfstrand, M., Gramzow, L., Holmberg, K., Hällman, J., Keech, O., Won, H., Renner, S.S., 2006. Dating dispersal and radiation in the gymnosperm Klasson, L., Koriabine, M., Kucukoglu, M., Käller, M., Luthman, J., Lysholm, F., Gnetum (Gnetales) – clock calibration when outgroup relationships are Niittylä, T., Olson, A., Rilakovic, N., Ritland, C., Rossello, J.A., Sena, J., Svensson, T., uncertain. Syst. Biol. 55, 610–622. Talavera-López, C., Theißen, G., Tuominen, H., Vanneste, K., Z-Q, Wu., Zhang, B., Wright, J.W., 1955. Species crossability in spruce in relation to distribution and Zerbe, P., Arvestad, L., Bhalerao, R., Bohlmann, J., Bousquet, J., Garcia Gil, R., . Forest Sci. 1, 319–349. Hvidsten, T.R., de Jong, P., MacKay, J., Morgante, M., Ritland, K., Sundberg, B., Xiang, Q.-P., Xiang, Q.-Y., Guo, Y.-Y., Zhang, X.-C., 2009. Phylogeny of Abies Thompson, S.L., Van de Peer, Y., Andersson, B., Nilsson, O., Ingvarsson, P.K., (Pinaceae) inferred from nrITS sequence data. Taxon 58, 141–152. Lundeberg, J., Jansson, S., 2013. The Norway spruce genome sequence and Zachos, J., Pagani, M., Sloan, L., Thomas, E., Billups, K., 2001. Trends, rhythms, and conifer genome evolution. Nature. http://dx.doi.org/10.1038/nature12211. aberrations in global climate 65 Ma to present. Science 292, 686–693. Parducci, L., Jørgensen, T., Tollefsrud, M.M., Elverland, E., Alm, T., Fontana, S.L., Zou, J.-B., Peng, X.-L., Li, L., Liu, J.-Q., Miehe, G., Opgenoorth, L., 2012. Molecular Bennett, K.D., Haile, J., Matetovici, I., Suyama, Y., Edwards, M.E., Andersen, K., phylogeography and evolutionary history of Picea likiangensis in the Qinghai– Rasmussen, M., Boessenkool, S., Coissac, E., Brochmann, C., Taberlet, P., Tibetan Plateau inferred from mitochondrial and chloroplast DNA sequence Houmark-Nielsen, M., Larsen, N.K., Orlando, L., Gilbert, M.T.P., Kjær, K.H., variation. J. Syst. Evol. 50, 341–350.