You are on page 1of 14

Plant Syst. Evol.

248: 129–142 (2004)


DOI 10.1007/s00606-004-0185-4

Taxonomy of North and Central American diploid wild potato


(Solanum sect. Petota) species: AFLP data
S. I. Lara-Cabrera1 and D. M. Spooner2
1
Facultad de Biologı́a, Universidad Michoacana de San Nicolás de Hidalgo, UMSNH, Michoacán,
Mexico
2
USDA, Agricultural Research Service, Department of Horticulture, University of Wisconsin,
Madison, WI, USA

Received July 8 2003; accepted April 20, 2004


Published online: August 27, 2004
 Springer-Verlag 2004

Abstract. Solanum section Petota, the potato and Key words: AFLP, molecular phylogeny, potato,
its wild relatives, includes about 200 wild species Solanum sect. Petota, systematics, taxonomy.
distributed from the southwestern United States
to central Argentina and adjacent Chile, with There are approximately 30 species of wild
about 30 species in North and Central America. potatoes in the United States, Mexico, and
The North/Central American region and the Central America (here referred to as North
South American region all include diploids, and Central America), out of the total of about
tetraploids, and hexaploids. Chloroplast DNA 200 species in Solanum L. sect. Petota
restriction enzyme data from a prior study
Dumort., that extends to south-central Argen-
showed that 13 of the North/Central American
species formed a clade containing only diploids,
tina and adjacent Chile (Spooner and Hijmans
but there was low resolution within the clade. 2001). The North/Central American region
This Amplified Fragment Length Polymorphism and the South American region include dip-
(AFLP) study is conducted to provide additional loids, tetraploids, and hexaploids. The North/
resolution within the North/Central American Central American diploids are classified in the
diploids and complements the chloroplast results, series Bulbocastana, Morelliformia, Pinnati-
and prior morphological results. Wagner parsi- secta, Polyadenia and Tuberosa (Hawkes
mony and phenetic analyses mostly agreed with 1990). This study considers the species bound-
the morphological data in supporting currently aries and taxonomic relationships of these
recognized species except that they suggest that diploid species.
S. brachistotrichium and S. stenophyllidium are A phylogeny based on chloroplast DNA
conspecific. Our new AFLP data, in combination
(cpDNA) restriction enzyme sites (Spooner
with the cpDNA and morphological data, also
support sister taxon relationships for the follow-
and Sytsma 1992, Spooner and Castillo 1997)
ing diploid species from North and Central divide sect. Petota into four clades (Fig. 1):
America: 1) S. cardiophyllum subsp. ehrenbergii 1) the North/Central American diploid
and S. stenophyllidium, 2) S. tarnii and S. trifidum, species, exclusive of S. bulbocastanum, S. car-
3) S. jamesii and S. pinnatisectum, 4) S. lesteri and diophyllum, and S. verrucosum, 2) S. bulbocas-
S. polyadenium, and 5) S. clarum and S. morel- tanum and S. cardiophyllum, 3) the South
liforme. American series Piurana and some South
130 S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP

American species classified in other series, 4) and clade 2 diploids (Spooner and Sytsma
all remaining South American species and 1992). For ease of discussion, the term ‘‘North/
the North/Central American polyploid spe- Central American diploids’’ is used in this
cies and the diploid species S. verrucosum. paper to include all species from this region
Clade 1 includes series Pinnatisecta, with except S. verrucosum.
internested series Polyadenia, Bulbocastana The separation of S. bulbocastanum and
(S. cardiophyllum) and series Morelliformia, S. cardiophyllum from clade 1 species was
and clade 2 ser. Bulbocastana (S. bulbocas- unexpected based on crossing data (Magoon
tanum) and ser. Pinnatisecta (S. cardiophyl- et al. 1958, Graham et al. 1959, Graham and
lum). Dionne 1961), immunological data (Gell et al.
The only North/Central American diploid 1960), and morphological data (Hawkes 1990)
species not part of clades 1 and 2 is S. verruco- data. Rodrı́guez and Spooner (1997) investi-
sum. This species shares morphology and gated S. bulbocastanum and S. cardiophyllum
crossability relationships with members of by including more accessions and all the
clade 4 and clearly is unrelated to the clade 1 subspecies of each species. The resulting tree

Fig. 1. Abstracted summary results of the chloroplast DNA restriction site analyses of Spooner and Sytsma
(1992), Rodrı́guez and Spooner (1997), and Spooner and Castillo (1997)
S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP 131

maintains all the subspecies in clade 2 except Our present study explores the use of
S. cardiophyllum subsp. ehrenbergii, which is another molecular marker, Amplified Frag-
placed in the North/Central American diploid ment Length Polymorphisms (AFLPs; Vos
clade 1 (Fig. 1). Unpublished data from Spoo- et al. 1995), to study the status of species,
ner and collaborators of the internal tran- and interrelationships, of the North/Central
scribed spacer region of nuclear ribosomal American diploids. The accessions of our
DNA, and the GBSSI gene, however, present AFLP study are a subset of those
show S. bulbocastanum and all subspecies of from the morphological study (Lara-Cabrera
S. cardiophyllum to be sister taxa, showing a and Spooner, in press), with only four excep-
discordance of the cpDNA data and nuclear tions (Table 1).
DNA data. The AFLP technique combines restriction
Lara-Cabrera and Spooner (in press) used enzyme reactions with the Polymerase Chain
morphology and nuclear DNA microsatellites Reaction (PCR), revealing high levels of poly-
developed for S. tuberosum to study the morphism (Vos et al. 1995). The AFLP tech-
validity of the North/Central American dip- nique is becoming the technique of choice
loid species in clade 1 and clade 2. They when there is no previous knowledge of
demonstrated that these microsatellite primers sequence variation or developed molecular
were not useful for this phylogenetic study, probes for a group. AFLPs have high repro-
possibly because of divergence of priming sites ducibility, providing advantages over other
that were developed for S. tuberosum of clade anonymous molecular markers like Random
4. The morphological data, however, provided Amplified Fragment Length Polymorphisms
good support for the North/Central American (RAPDs), and are not as expensive as other
diploid species except that S. brachistotrichium techniques like nuclear Restriction Fragment
and S. stenophyllidium cluster together Length Polymorphisms (RFLPs; Mueller and
(Fig. 2). Wolfenbarger 1999). They have tremendous

Fig. 2. Abstracted summary results of the morphological phenetic analyses of the North/Central American
diploid species by Lara-Cabrera and Spooner (in press). The numbers of accessions analyzed per taxon are
indicated in parentheses)
132 S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP

Table 1. Series affiliation within Solanum sect. Petota, PI (USDA Plant Introduction) number, and
abstracted locality of the accessions examined in this study. Herbarium vouchers are deposited at PTIS. All
of these accessions were studied in Lara-Cabrera and Spooner (in press) except the two outgroups
(S. etuberosum and S. palustre), S. morelliforme, and the S. bulbocastanum · S. cardiophyllum hybrid
Series placement by Taxon PI Country: State
Hawkes (1990) or Departmenta
Bulbocastana (Rydb.) Solanum bulbocastanum 275198 Mexico: México
Hawkes Dunal subsp. bulbocastanum
S. bulbocastanum subsp. bulbocastanum 275184 Mexico: Distrito Federal
S. bulbocastanum subsp. bulbocastanum 275194 Mexico: Oaxaca
S. bulbocastanum subsp. 498224 Mexico: Michoacán
dolichophyllum (Bitter) Hawkes
S. bulbocastanum subsp. dolichophyllum 545752 Mexico: México
S. bulbocastanum subsp. dolichophyllum 255516 Mexico: Jalisco
S. bulbocastanum subsp. partitum 558379 Mexico: Chiapas
(Correll) Hawkes
S. bulbocastanum subsp. partitum 275200 Guatemala:
Huehuetenango
S. clarum Correll 558382 Mexico: Chiapas
Morelliformia Hawkes S. morelliforme 275218 Mexico: México
S. morelliforme 498003 Mexico: México
Pinnatisecta (Rydb.) S. morelliforme 545774 Mexico: Chiapas
Hawkes S. brachistotrichium 498217 Mexico: Chihuahua
S. brachistotrichium 497993 Mexico: Chihuahua
S. brachistotrichium 545832 Mexico: Aguascalientes
S. brachistotrichium 545815 Mexico: Aguascalientes
S. brachistotrichium 255527 Mexico: Aguascalientes
S. cardiophyllum Lindl. subsp. 347759 Mexico: Puebla
cardiophyllum
S. cardiophyllum subsp. cardiophyllum 283062 Mexico: State unknown
S. cardiophyllum subsp. cardiophyllum 283063 Mexico: State unknown
S. cardiophyllum subsp. ehrenbergii Bitter 279272 Mexico: Aguascalientes
S. cardiophyllum subsp. ehrenbergii 184762 Mexico: Quéretaro
S. cardiophyllum subsp. ehrenbergii 341231 Mexico: Jalisco
S. cardiophyllum subsp. ehrenbergii 186548 Mexico: Zacatecas
S. cardiophyllum subsp. ehrenbergii 255520 Mexico: San Luis Potosı́
S. cardiophyllum subsp. ehrenbergii 275213 Mexico: Quéretaro
S. hintonii Correll 607880 Mexico: México
S. jamesii Torr. 458425 USA: Arizona
S. jamesii 564051 USA: Arizona
S. jamesii 564049 USA: New Mexico
S. jamesii 275169 USA: New Mexico
S. jamesii 458423 USA: New Mexico
S. jamesii 275172 USA: Arizona
S. ·michoacanum (Bitter) Rydb. 558497 Mexico: Michoacán
(putative origin is S. bulbocastanum ·
S. pinnatisectum)
S. nayaritense (Bitter) Rydb. 545827 Mexico: Nayarit
S. nayaritense 545820 Mexico: Zacatecas
S. nayaritense 545825 Mexico: Zacatecas
S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP 133

Table 1 (continued)
Series placement by Taxon PI Country: State
Hawkes (1990) or Departmenta
S. pinnatisectum Dunal 275234 Mexico: Jalisco
S. pinnatisectum 184764 Mexico: Guanajuato
S. pinnatisectum 275233 Mexico: Guanajuato
S. pinnatisectum 275236 Mexico: Jalisco
S. pinnatisectum 275230 Mexico: Quéretaro
S. pinnatisectum 190115 Mexico: Michoacán
S. bulbocastanum · S. cardiophyllum Mexico: Jalisco
putative hybrid
S. ·sambucinum Rydb. (putative origin is 595478 Mexico: Quéretaro
S. ehrenbergii · S. pinnatisectum)
S. ·sambucinum 604209 Mexico: Guanajuato
S. stenophyllidium Bitter 558460 Mexico: Jalisco
S. tarnii Hawkes et Hjert. 498048 Mexico: Hidalgo
S. tarnii 545808 Mexico: Hidalgo
S. tarnii 545742 Mexico: Veracruz
S. tarnii 498043 Mexico: Hidalgo
S. tarnii 570641 Mexico: Hidalgo
S. trifidum Correll 255542 Mexico: Michoacán
S. trifidum 558478 Mexico: Michoacán
S. trifidum 558480 Mexico: Michoacán
S. trifidum 283104 Mexico: Jalisco
Polyadenia Bukasov S. lesteri Hawkes et Hjert. 558434 Mexico: Oaxaca
S. lesteri 442694 Mexico: Oaxaca
S. lesteri 558435 Mexico: Oaxaca
S. polyadenium Greenm. 498036 Mexico: Hidalgo
S. polyadenium 347769 Mexico: Puebla
S. polyadenium 347770 Mexico: Veracruz
Sect. Etuberosum S. etuberosum Lindl. 558285 Chile: Santiago
(outgroup) S. palustre Poepp. 558253 Chile: Los Lagos
a
More complete locality data and a map of these accessions are found in Lara-Cabrera and Spooner (in
press).

utility to provide more resolution than cpDNA AFLPs also have been used in Solanum
restriction site data and nuclear ribosomal tuberosum to estimate genetic variation and
internal transcribed spacer data at the genus relationships in cultivars (Kim et al. 1998). In
level (Despres et al. 2003, Beardsley and potato systematics, Kardolus et al. (1998) used
Olmstead 2003). AFLPs in 19 taxa of Solanum sect. Petota and
In the Solanaceae, Mace et al. (1999b) used three taxa of Solanum sect. Lycopersicon, with
AFLP data to study the genetic relationships results concordant with some current taxo-
among cultivated and wild eggplants (Solanum nomic hypotheses. Notably, their one place-
subg. Leptostemonum (Dunal) Bitter) and holder for clade 1 (S. pinnatisectum) was basal
found results that were consistent with ITS-1 in the sect. Petota clade, concordant with the
sequences, isozymes, and morphology. Mace cpDNA data. McGregor et al. (2002) and Van
et al. (1999a) found concordance between den Berg et al. (2002) used AFLPs to distin-
AFLPs and other data sets in Datureae, and guish morphologically similar species. The
proposed a new classification for the tribe. objectives of our AFLP study were to use
134 S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP

AFLPs for phylogenetic inference of the PCR reactions were performed in a 9600 Perkin
North/Central American diploids, to compare Elmer thermocycler. PCR products were sent to the
the results with the cpDNA and morphological Biotechnology Center at the University of Wiscon-
data, with the ultimate goal of using these sin-Madison for fluorescent fragment separation.
insights to produce a taxonomic monograph of Data were collected using a PE Biosystems DNA
Sequencer data collection v. 2.0 and analyzed with
sect. Petota from this region.
GeneScan v. 2.1 (Perkin Elmer Life Sciences).
Trace files were later imported into Genotyper (v.
Materials and methods 2.1 PE, Perkin Elmer Life Sciences) for fragment
sizing. Fragments in the range of 50 to 550 base
Plant material pairs were included in the analysis, because shorter
Each species was sampled from as many geograph- fragments tended to give less sharp peaks, and
ically dispersed sites as possible (Table 1). Only one longer fragments were not always consistently
accession was available for S. ·michoacanum, amplified. Peaks were scored as base-pair sizes
S. clarum, and S. bulbocastanum · S. cardiophyllum and translated into binary presence/absence data.
and our conclusions on relationships of these rare Both primer data sets were combined for analysis.
species are therefore problematical. The accessions Data are available from the authors.
and DNA samples were the same as those used in
Lara-Cabrera and Spooner (in press) with minor
exceptions. In that prior study, morphological data Cladistic analysis
were not available for S. morelliforme because this A cladistic analysis was performed using PAUP
epiphytic species did not flower, nor did we have 4.0b8 (Swofford 2001). The AFLP data were
living material of the putative interspecific hybrid analyzed using Wagner parsimony (Farris 1970),
of S. bulbocastanum · S. cardiophyllum. We used with S. etuberosum and S. palustre as outgroups,
DNA from two outgroup species, S. etuberosum after the elimination of 15 autapomorphous char-
and S. palustre (Table 1). Preparations of total acters, following Spooner et al. (1993). Trees were
DNA were made from fresh leaf tissue from a produced using a four step search following
single individual plant per accession, from plants Olmstead and Palmer (1994): 1) 50,000 random
grown in a greenhouse, following the procedure of replicates were performed with nearest-neighbor
Doyle and Doyle (1987), and purified over CsCl/ interchange (NNI) branch swapping algorithm; 2)
ethidium bromide gradients. the shortest trees from step one were used as
starting trees with tree-bisection and reconnection
(TBR) option; 3) the shortest trees from step 2 were
Amplification and DNA fragment analysis used as starting trees for a search using NNI for
multiple parsimonious trees (MULPARS); 4) trees
The AFLP plant mapping kit (Perkin Elmer Life
from step 3 were used as starting trees for a search
Sciences, Boston, Massachusetts) was used to pro-
using TBR and MULPARS with 1000 as the upper
duce AFLP fragments, following the manufac-
limit of trees saved. The resulting trees were used to
turer’s directions. Three reactions were performed.
compute a strict consensus tree. Bootstrap values
The first reaction ligated specific adaptors to the
(Felsenstein 1985) were obtained from 1000
ends of the restriction fragments generated by
random replicates with NNI, no MULPARS.
digestion of DNA with EcoR1 and Mse1. This was
followed by a preselective procedure that amplified
fragments with adaptors on both ends by using an
Phenetic analysis
EcoR1 complementary primer containing a 3’ C and
an Mse1 complementary primer containing a 3’ A. Because AFLPs are dominant marker data, and
The third reaction was a selective amplification, because they represent anonymous fragments, a case
using one of the two primer combinations: primer 1 has been made that they should be analyzed with
= EcoRI+ ACA/MseI+ CAG, primer 2 = phenetic, as well as cladistic methods (e.g. Koopman
EcoRI+ AAC/MseI+ CAC, based on their suc- et al. 2001). AFLP data were analyzed using
cessful use in the potato study of Kardolus et al. NTSYS-pcR version 2.02k (Applied Biosystematics,
(1998). The primers were fluorescently labeled. The Setauket, New York). The ‘‘FIND’’ option was
S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP 135

enabled to detect all possible trees. Similarity S. trifidum form a terminal clade but not as a
matrices (in SIMQUAL) were generated using the separate clade from S. tarnii. Clade E con-
Dice and Jaccard similarity matrix, which ignores tained all accessions of S. jamesii. Clade F
shared absent bands. Dice and Jaccard algorithms contained all accessions of S. pinnatisectum,
are appropriate algorithms for predominately dom- with the two accessions of the putative hybrid
inant AFLP marker data because they give greater
S. ·sambucinum (S. pinnatisectum · S. cardio-
weight to the greater information content of dom-
inant presence data. Clustering was performed using
phyllum subsp. ehrenbergii) at the base of the
the unweighted pair-group method (UPGMA) and clade. Clade G contained all accessions of
Neighbor Joining in SAHN. Cophenetic correlation S. lesteri and S. polyadenium, and the putative
coefficients were calculated for all combinations of hybrid S. ·michoacanum (S. bulbocastanum ·
similarity and tree building methods using the S. pinnatisectum). Like the S. tarnii and
procedures COPH and MXCOMP. These coeffi- S. trifidum clade, S. lesteri and S. polyadenium
cients indicate the correlation between a similarity did not form species-specific clades. Clade H
matrix and the phenetic tree resulting from it after a contained all accessions of S. clarum and
cluster analysis, indicating goodness of fit of the S. morelliforme. Clade I contained all acces-
cluster analysis to the similarity matrix. Clustering sions of S. cardiophyllum subsp. cardiophyllum.
methods and similarity coefficients are described in Clades A, C, D (containing all accessions of
Rohlf (1993). Bootstrap values were obtained from
S. trifidum and S. tarnii but excluding S. hinto-
the tree with the highest cophenetic coefficient from
5000 random replicates using WinBoot (Yap and
nii), E, G (containing all accessions of S. lesteri
Nelson 1996). and S. polyadenium but excluding S. ·micho-
acanum), and clade I are supported with a
bootstrap values of >51% (as shown on the
Results cladogram) and the others were <50% (not
shown on the cladogram).
Cladistic results
Two primer combinations produced a total of
Phenetic results
234 polymorphic bands (no missing data).
Wagner parsimony analysis of these bands The best cophenetic correlation coefficients
(Fig. 3a) produced 24 most-parsimonious were produced from the combination of
1494-step trees with a consistency index Jaccard/UPGMA (r = 0.904). Successively
(excluding autapomorphies) of 0.157 and a lower r values were produced by the combi-
retention index of 0.516. Some internal clades nation Dice/UPGMA (0.888), Dice/Neighbor
are labeled as A-I for ease of discussion. Clade Joining (0.822), and Jaccard/Neighbor
A contained S. brachistotrichium, S. nayari- Joining (0.789). Rohlf (1993) subjectively
tense, S. stenophyllidium, and one of the six interprets all values above 0.9 as an excellent
accessions of S. cardiophyllum subsp. ehren- fit of the similarity matrix to the resulting
bergii. Clade B contained five of the six tree.
accessions of S. cardiophyllum subsp. ehren- Figure 3b presents the Jaccard/UPGMA
bergii, and the sole accession of a putative phenogram, with dotted lines connecting iden-
hybrid of S. bulbocastanum · S. cardiophyllum tical accessions in the cladogram (Fig. 3a).
subsp. cardiophyllum (Rodrı́guez and Vargas Only a single Jaccard/UPGMA phenogram
1994). Clade C contained all accessions of was found. The cladogram and phenogram
S. bulbocastanum, with the two accessions of outlined almost the same set of species groups,
subsp. partitum forming a clade but imbedded with the major conflict occurring in the
in a clade of the other two species that did not arrangement of species within, not among,
form clades. Clade D contained all accessions clades/groups. Exceptions to this concordance
of S. trifidum and S. tarnii, and the sole of cladistic and phenetic results were: 1) the
accession of S. hintonii. All accessions of exclusion in the phenogram from one
136 S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP

Fig. 3. (a) Cladistic analyses of Amplified Fragment Length Polymorphisms of the North/Central American
diploid wild potato (Solanum sect. Petota) species. The cladogram (left) is one of 24 most-parsimonious 1494-
step Wagner trees, drawn as a phylogram, and is rooted by S. etuberosum and S. palustre, two species in
Solanum sect. Etuberosum. Numbers above the branches represent number of characters supporting each
branch. Numbers following the species are USDA Plant Introduction Numbers as listed in Table 1. The three
arrows indicate branches that break down in a strict consensus tree. Numbers in circles below the branches
represent the bootstrap support values above 50% that were produced in a separate 50% majority rule
consensus tree (not shown here). The letters A-I are clades as discussed in the text. (b) UPGMA phenogram
based on Jaccard coefficient, with overlaid bootstrap values. The dotted lines connect identical accessions in
both analyses

accession of S. cardiophyllum subsp. ehrenber- S. polyadenium, 4) Solanum clarum clustering


gii to the base of the clade including this taxon, at the base of a large cluster in the phenogram,
S. stenophyllidium, and S. bulbocastanum, 2) rather than with S. morelliforme, and outside
Solanum hintonii grouping in the phenogram of any particular group, 5) Solanum pinnati-
with S. bulbocastanum, rather than with sectum, S. ·michoacanum, and S. jamesii not
S. tarnii + S. trifidum, 3) Solanum ·michoaca- clustering in the phenogram as in the clado-
num clustering in the phenogram with gram. Conversely, better species-specific clus-
S. pinnatisectum rather than with S. lesteri + tering is found in the phenogram that placed
S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP 137

S. trifidum and S. tarnii in their own groups, strap support for all North/Central American
rather than the paraphyletic result on the species exclusive of S. cardiophyllum subsp.
cladogram where S. trifidum formed a termi- cardiophyllum, (85% bootstrap support on the
nal clade but not as a separate clade from phenogram, Fig. 3b) none of the external-most
S. tarnii. branches were well supported. Rather, only
more internal branches have bootstrap values
that support sister species, species, or acces-
Discussion sions within species. Such low external branch
support precludes global conclusions of inter-
Comparison of AFLP, chloroplast DNA,
specific relationships for all North/Central
and morphological results
American species. The interpretation of a
Our results match those of Koopman et al. bootstrap value as supporting a ‘‘true’’ phylog-
(2001) and Mace et al. (1999a,b) who also eny is subject to many factors and it is difficult
found that the combination of Jaccard/UP- to choose a value as ‘‘significant.’’ Hillis and
GMA had the similar highest cophenetic Bull (1993) showed, in cladistic methods, that
correlation coefficients in their AFLP studies. under certain conditions thought to be typical
There were many points of agreement in the of most phylogenetic analyses (equal rates of
AFLP, cpDNA (Spooner and Sytsma 1992), change, symmetric phylogenies, and internodal
and morphological (Lara-Cabrera and Spoo- changes of equal to or lesser than 20%),
ner in press) results. In the discussion below, bootstrap values of equal to or greater than
the word ‘‘clade’’ refers to the results of 70% usually corresponded to a probability of
cladistic analysis and ‘‘cluster’’ to the results equal to or greater than 90% that a clade was
of phenetic analysis. Despite the use of the phylogenetically ‘‘true.’’ We are not aware of
same accessions between AFLP and morpho- similar studies of bootstrap values for phenetic
logical studies, we discuss the dominant binary data. The low consistency index (0.157) and
AFLP data and continuously variable mor- general lack of low bootstrap values on the
phological results separately, rather than ana- cladogram indicate much homoplasy that is
lyze them as a common dataset. We do this typical with AFLP data. It also may indicate
because most wild potato species are very recent divergence of some North/Central
similar and are distinguished morphologically American sister taxa. In cases of recent diver-
by a series of largely overlapping character gence, strict monophyly in the AFLP results
states, few if any of which individually are may not yet occur for close sister taxa that are
suitable to exclude or to include a taxon from a definable morphologically (e.g. S. tarnii and S.
group (polythetic support) (Spooner et al. trifidum; S. lesteri and S. polyadenium).
2001, Lara-Cabrera and Spooner in press). The phenetic results of the prior morpho-
The AFLP and morphological data are appro- logical study (Lara-Cabrera and Spooner, in
priately analyzed by different similarity algo- press) and AFLP results given here (Fig. 3b)
rithms (qualitative classes of algorithms for are not designed to show phylogenetic rela-
AFLP, mostly quantitative for morphology). tionships. However, phenetic and cladistic
The cpDNA data included the same species as results of the same data can be similar under
the microsatellite and morphological studies, certain conditions. When similarities due to
but generally only one accession per species. shared ancestral characteristics or homoplasy
However, the few qualitative morphological exceed similarities due to shared derived char-
characters of the North/Central American acteristics, the phenogram will not represent
diploids that could be ‘‘mapped’’ onto the the phylogeny, but if there were no homoplasy,
AFLP cladogram are discussed below. the phenogram constructed by the phenetic
Most clades (Fig. 3a) have low bootstrap analysis would correspond to the true phylo-
support. Except for the 81% cladogram boot- genetic tree (Futuyma 1998).
138 S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP

Solanum stenophyllidium, S. brachistotrich- diophyllum subsp. cardiophyllum is a morpho-


ium, S. nayaritense, S. cardiophyllum subsp. logical synapormorphy of cream-white to
ehrenbergii (Clades A, B). Chloroplast DNA light yellow corollas, a color unique to the
cladistic data (Spooner and Sytsma 1992, North/Central American diploids (the other
Rodrı́guez and Spooner 1997) and morpho- species with corollas pure white or white
logical phenetic data (Lara-Cabrera and Spoo- tinged with blue and purple). The AFLP data,
ner in press) supported S. cardiophyllum subsp. like the cpDNA data, separated S. cardio-
ehrenbergii and S. nayaritense as distinct taxa, phyllum subsp. cardiophyllum from S. cardio-
but suggested that S. stenophyllidium and phyllum subsp. ehrenbergii, and allied the
S. brachistotrichium were synonyms (S. steno- latter with S. brachistotrichium + S. steno-
phyllidium is the earlier name). Hawkes (1990) phyllidium. The cause of the discordance
distinguished S. stenophyllidium by lanceolate between different sister taxon relationships
leaves vs. S. brachistotrichium by linear-lance- of S. bulbocastanum and S. cardiophyllum
olate leaves. Lara-Cabrera and Spooner (in subsp. cardiophyllum with cpDNA and AFLP
press), however, showed no statistically signif- data is unknown, but possibly involves a
icant difference of terminal and lateral leaflet history of ‘‘chloroplast capture,’’ a phenom-
shapes between these species. enon well documented in other groups (Wen-
Clade A also contains one of the six del and Doyle 1998). We await results from
accessions of S. cardiophyllum subsp. ehren- more molecular markers to investigate this
bergii. Solanum cardiophyllum subsp. ehrenber- discordance.
gii is morphologically extremely similar to Solanum tarnii, S. trifidum, S. hintonii
S. stenophyllidium, and they are distinguished (Clade D). AFLP cladistic results place S. tri-
only by minor differences in leaf shape. This fidum and S. tarnii in the same clade, but do
‘‘misplaced accession’’ of S. cardiophyllum not separate them into species-specific clades.
subsp. ehrenbergii suggests misidentification, The AFLP phenetic results, however, place
but it clustered with other accessions of subsp. S. tarnii and S. trifidum into species-specific
ehrenbergii in the phenetic study of Lara- groups. Their relationship likewise was
Cabrera and Spooner (in press). supported with cpDNA (Spooner and Sytsma
Solanum bulbocastanum, S. cardiophyllum 1992) and morphological (Lara-Cabrera and
subsp. cardiophyllum (Clades B, I). Solanum Spooner in press) results. The morphological
bulbocastanum subsp. bulbocastanum and similarity of these two species is not as
subsp. dolichophyllum do not form clades. evident as there is no clear character to unite
The two accessions of S. bulbocastanum subsp. them, and S. tarnii possesses globose fruits
partitum form a clade (76% bootstrap sup- whereas S. trifidum possesses conical fruits.
port), but this is embedded in the above two The concordance of all data sets, however,
subspecies. Previous morphological and nucle- suggests that S. trifidum and S. tarnii are
ar RFLP data (Rodrı́guez and Spooner 2002) sister taxa. Clade D also contains the
failed to support any of these subspecies, and sole accession of S. hintonii, a species
we think that none of the subspecies of clearly distinguishable from S. tarnii and S.
S. bulbocastanum are distinct. trifidum.
Both the AFLP and morphological data Solanum jamesii, S. pinnatisectum,
(Lara-Cabrera and Spooner in press) widely S. ·sambucinum (Clades E, F). The AFLP
separated S. bulbocastanum (Clade C) from results place S. jamesii, S. pinnatisectum, and
S. cardiophyllum subsp. cardiophyllum (Clade the hybrid S. ·sambucinum in the same clade
I), discordant with the cpDNA data that (<50% bootstrap support). However, all three
places them in a single clade (Spooner and species are separate in the morphological
Sytsma 1992). Supporting the sister taxon analysis (Lara-Cabrera and Spooner in press),
relationship of S. bulbocastanum and S. car- and the bootstrap support on the AFLP
S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP 139

cladogram is <50%. Solanum pinnatisectum support for the clade uniting S. clarum and
+ S. ·sambucinum, do not cluster with S. S. morelliforme and their failure to cluster in
jamesii on the AFLP phenogram. Solanum the phenetic results, we consider them as likely
jamesii and S. pinnatisectum are the only two sister taxa because of their morphological
species in sect. Petota with pinnatifid stipules, similarity and AFLP and cpDNA cladistic
in contrast to the lunate stipules of all other results.
species of wild potato. This striking morpho-
logical synapomorphy, and the AFLP cladistic
Hybrid origins
results (despite weak bootstrap support) leads
us to suspect that S. jamesii and S. pinnatisec- This study included three putative hybrids: 1)
tum are sister taxa; the hybrid nature of S. ·michoacanum (S. bulbocastanum · S. pin-
S. ·sambucinum makes statements of cladistic natisectum), 2) S. ·sambucinum (S. cardiophyl-
relationship difficult. lum subsp. ehrenbergii · S. pinnatisectum)
Solanum lesteri, S. polyadenium, S. ·micho- (Hawkes 1990) 3) an unnamed hybrid S. bulbo-
acanum (Clade G). AFLP analyses provide castanum · S. cardiophyllum (Rodrı́guez et al.
support for the sister taxon relationship of 1995). All three of these are distinct from their
S. lesteri and S. polyadenium, also concordant parental species at least with cladistic results.
with cpDNA (Spooner and Sytsma 1992) and Solanum ·michoacanum and S. ·sambucinum
morphological results (Lara-Cabrera and have S. pinnatisectum as one of their putative
Spooner in press). The phenetic similarity of parents, and they always group with this
these two species is clear in the field, as both species, but not the other parent. We cannot
possess a dense indument of sticky short- use additive species-specific AFLP markers to
stalked glands (Type A trichomes; Gibson investigate hybrid origins of these three taxa
1971), and a characteristic ‘‘mousy’’ odor (Rieseberg and Ellstrand 1993) because S. car-
unique within sect. Petota. Both species also diophyllum subsp. ehrenbergii and S. pinnati-
possess pentagonal corollas, unique to the sectum have no species-specific markers useful
North/Central American diploid species (the to address these questions, and because
other species with stellate corollas); S. ·micho- S. bulbocastanum has only two subspecific
acanum had white stellate corollas. markers (only in subsp. partitum).
Solanum clarum, S. morelliforme (Clade
H). AFLP data support a sister taxon rela-
Summary
tionship for S. clarum and S. morelliforme, but
Lara-Cabrera and Spooner (in press) did not Most of the North/Central American diploid
examine the latter morphologically because it species show some degree of support, both
did not grow well for that study. The cpDNA with cladistic and phenetic analyses of AFLP
data (Spooner and Sytsma 1992) also place data and phenetic analyses of morphological
S. clarum and S. morelliforme in the same data. AFLP and morphological data lead us to
clade, but not into species-specific clades. question the validity of separating S. brachis-
Hawkes (1990) classified S. clarum and totrichium and S. stenophyllidium. In combi-
S. bulbocastanum as the sole two representa- nation with cpDNA and morphological data,
tives of ser. Bulbocastana (Rydb.) Hawkes and AFLP data provide some support for sister
placed S. morelliforme as the sole representa- taxon relationships for 1) S. cardiophyllum
tive of ser. Morelliformia Hawkes, also with subsp. ehrenbergii and S. stenophyllidium (to
entire leaves. Solanum morelliforme and S. cla- include S. brachistotrichium), 2) S. tarnii and
rum, however, both grow in moss in epiphytic S. trifidum, 3) S. jamesii and S. pinnatisectum,
like conditions, and share many morphological 4) S. lesteri and S. polyadenium, and 5)
and reproductive traits (discussed in Spooner S. clarum and S. morelliforme. Taxonomic
and Sytsma 1992). Despite low bootstrap decisions on the species boundaries and rela-
140 S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP

tionships of the North/Central American dip- Futuyma D. J. (1998) Evolutionary biology, third
loid species will be finalized with additional edition. Sinauer Associates, Sunderland, Massa-
data to result from a much wider examination chusetts.
of herbarium material for our treatment of Gell, P. G. H., Hawkes J. G., Wright S. T. C.
section Petota from North and Central Amer- (1960) The application of immunological meth-
ods to the taxonomy of species within the genus
ica (Spooner et al. 2004).
Solanum. Proc. Roy. Soc. London, Ser. B, Biol.
Sci. 151: 364–383.
This work represents partial fulfillment for the
Gibson R. W. (1971) Glandular hairs providing
requirements of a Ph.D. degree in Plant Breeding
resistance to aphids in certain wild potato
and Genetics at the University of Wisconsin-
species. Ann. Appl. Biol. 68: 113–119.
Madison. We thank committee members Paul
Graham M. K., Dionne L. A. (1961) Crossability
Berry, Michael Havey, Thomas Osborn, and Ken-
relationships of certain diploid Mexican Solanum
neth Sytsma. We also thank John Bamberg and
species. Canad. J. Genet. Cytol. 3: 121–127.
Staff of the Unites States Potato Genebank for
Graham M. K., Niederhauser J. S., Servin L.
germplasm and locality data; Charles Nicolet and
(1959) Studies on fertility and late blight resis-
staff of the University of Wisconsin Biotechnology
tance in Solanum bulbocastanum Dun. in Mex-
Center for technical help; Lynn Hummel and staff
ico. Canad. J. Bot. 37: 41–49.
at Walnut Street Greenhouse for help in growing
Hawkes J. G. (1990) The potato: evolution, biodi-
plants; and lab partners Brian Karas, Iris Peralta,
versity and genetic resources. Belhaven Press,
Celeste Raker, and Sarah Stephenson for technical
London.
advice. This study was supported by CONACYT
Hillis D. M., Bull J. J. (1993) An empirical test of
(Mexico) scholarship number 116742 granted to
bootstrapping as a method for assessing confi-
Sabina I. Lara-Cabrera, and the United States
dence in phylogenetic analysis. Syst. Biol. 42:
Department of Agriculture. Names are necessary to
182–192.
report data. However, the USDA neither guaran-
Kardolus J. P., Van Eck H. J., Van den Berg R. G.
tees nor warrants the standard of the product, and
(1998) The potential of AFLPs in biosystematics:
the use of the name by the USDA implies no
a first application in Solanum taxonomy (Solan-
approval of the product to the exclusion of others
aceae). Plant Syst. Evol. 210: 87–103.
that may also be suitable.
Kim J. H., Joung H., Kim H. Y., Lim Y. P. (1998)
Estimation of genetic variation and relationship
in potato (Solanum tuberosum L.) cultivars using
References
AFLP markers. Amer. J. Potato Res. 75: 107–
Beardsley P. M., Olmstead R. G. (2003) AFLP 112.
phylogeny of Mimulus section Erythranthe and Koopman W. J. M., Zevenbergen M. J., Van den
evolution of hummingbird pollination. Evolu- Berg R. G. (2001) Species relationships in
tion 57: 1397–1410. Lactuca s.l. (Lactuceae, Asteraceae) inferred
Despres L., Gielly L., Redoutet B., Taberlet P. from AFLP fingerprints. Amer. J. Bot. 88:
(2003) Using AFLP to resolve phylogenetic 1881–1887.
relationships in a morphologically diversified Lara-Cabrera S., Spooner D. S. (2004) Taxonomy
plant species complex when nuclear and chloro- of Mexican diploid wild potato (Solanum sect.
plast sequences fail to reveal variability. Molec. Petota) species: morphological and microsatellite
Phylog. Evol. 27: 185–196. data. Monogr. Syst. Bot., Missouri Bot. Gard.
Doyle J. J., Doyle J. L. (1987) A rapid DNA (in press).
isolation procedure for small quantities of fresh Mace E. S., Gebhardt C. G., Lester R. N. (1999a)
leaf tissue. Phytochem. Bull. 19: 11–15. AFLP analysis of genetic relationships in the
Farris J. S. (1970) Methods for computing Wagner tribe Datureae (Solanaceae). Theor. Appl. Genet.
trees. Syst. Zool. 19: 83–92. 99: 634–641.
Felsenstein J. (1985) Confidence limits on phylog- Mace E. S., Lester R. N., Gebhardt C. G. (1999b)
enies: an approach using bootstrap. Evolution AFLP analysis of genetic relationships among
39: 783–791. the cultivated eggplant, Solanum melongena L.,
S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP 141

and wild relatives (Solanaceae). Theor. Appl. Petota): evidence from chloroplast DNA
Genet. 99: 626–633. restriction site variation. Amer. J. Bot. 84:
Magoon M. L., Cooper D. C., Hougas R. W. 671–685.
(1958) Cytogenetic studies of some diploid Spooner D. M., Hijmans R. J. (2001) Potato
Solanums section Tuberarium. Amer. J. Bot. systematics and germplasm collecting, 1989–
45: 207–221. 2000. Amer. J. Potato Res. 78: 237–268; 395.
McGregor C. E., Van Treuren R., Hoekstra R., Spooner D. M., Sytsma K. J. (1992) Reexamina-
Van Hintum T. J. L. (2002) Analysis of the wild tion of the series relationships of Mexican and
potato germplasm of the series Acaulia with Central American wild potatoes (Solanum sect.
AFLPs: implications for ex situ conservation. Petota): evidence from chloroplast DNA
Theor. Appl. Genet. 104: 146–156. restriction site variation. Syst. Bot. 17: 432–
Mueller U. G., Wolfenbarger L. L. (1999) AFLP 448.
genotyping and fingerprinting. Trends Ecol. Spooner D. M., Van den Berg R. G., Miller J. T.
Evol. 14: 389–394. (2001) Species and series boundaries of Sola-
Olmstead R. G., Palmer J. D. (1994) Chloroplast num series Longipedicellata (Solanaceae) and
DNA systematics: a review of methods and data phenetically similar species in ser. Demissa and
analysis. Amer. J. Bot. 81: 1205–1224. ser. Tuberosa: implications for a practical tax-
Rieseberg L. H., Ellstrand N. C. (1993) What can onomy of sect. Petota. Amer. J. Bot. 88: 113–
molecular and morphological markers tell us 130.
about plant hybridization? Crit. Rev. Plant Sci. Spooner D. M., Van den Berg R. G., Rodrı́guez A.,
12: 213–241. Bamberg J., Hijmans R. J., Lara-Cabrera S. I.
Rodrı́guez A., Spooner D. M. (1997) Chloroplast (2004) Wild potatoes (Solanum section Petota) of
DNA analysis of Solanum bulbocastanum and North and Central America. Syst. Bot. Monogr.
S. cardiophyllum, and evidence for the distinct- 68: 1–209+9pl.
ness of S. cardiophyllum subsp. ehrenbergii (sect. Swofford D. L. (2001) PAUP: phylogenetic analysis
Petota). Syst. Bot. 22: 31–34. using parsimony. Version 4.0b8. Sinauer Asso-
Rodrı́guez A., Spooner D. M. (2002) Subspecies ciates, Sunderland, Massachusetts.
boundaries of the wild potatoes Solanum bulbo- Van den Berg R. G., Bryan G. J., Del Rio A.,
castanum and S. cardiophyllum based on mor- Spooner D. M. (2002) Reduction of species in
phological and nuclear RFLP data. Acta Bot. the wild potato Solanum section Petota series
Mex. 61: 9–25. Longipedicellata: AFLP, RAPD and chloro-
Rodrı́guez A., Vargas O. (1994) Las especies de plast SSR data. Theor. Appl. Genet. 105:
papa silvestre (Solanum L. sección Petota Du- 1109–1114.
mortier) en Jalisco. Bol. Inst. Bot. Guadalajara Vos P., Hogers R., Bleeker M., Reijans M., Van de
Mex. 2: 1–68. Lee T., Hornes M., Frijters A., Pot J., Peleman
Rodrı́guez A., Vargas O., Villegas E., Spooner D. J., Kuiper M., Zabeau M. (1995) AFLP: a new
M. (1995) Wild potato (Solanum sect. Petota) technique for DNA fingerprinting. Nucl. Acids
germplasm collecting expedition to Mexico in Res. 23: 4407–4414.
1993, with special reference to Solanum bulbo- Wendel J. F., Doyle J. J. (1998) Phylogenetic
castanum Dunal and S. cardiophyllum Lindley. incongruence: window into genome history
Potato Res. 38: 47–52. and molecular evolution. In: Soltis D. E.,
Rohlf F. J. (1993) NTSYS-pc: Numerical taxonomy Soltis P. S., Doyle J. J. (eds.) Molecular system-
and multivariate analysis system, version 1.80. atics of plants II: DNA sequencing. Kluwer
Applied Biosystematics, Setauket, New York. Academic Publishers, Boston, MA, pp. 265–
Spooner D. M., Anderson G. J., Jansen R. K. 296.
(1993) Chloroplast DNA evidence for the inter- Yap I., Nelson R. J. (1996) WinBoot: a program
relationships of tomatoes, potatoes, and pepinos for performing bootstrap analysis of binary data
(Solanaceae). Amer. J. Bot. 80: 676–688. to determine the confidence limits of UPGMA-
Spooner D. M., Castillo T. R. (1997) Reexami- based dendrograms. IRRI Discussion Paper
nation of series relationships of South Ameri- Series No. 14, International Rice Research
can wild potatoes (Solanaceae: Solanum sect. Institute, Manila, The Philippines.
142 S.I. Lara-Cabrera and D.M. Spooner: Solanum section Petota AFLP

Address of the authors: Sabina I. Lara-Cabrera Morelia, Michoacán, Cp. 58066, México. David M.
(e-mail: slara@zeus.umich.mx), Facultad de Biolo- Spooner, (e-mail: dspooner@wisc.edu), USDA,
gı́a, Edif. R. Ciudad Universitaria, Universidad Agricultural Research Service, Department of
Michoacana de San Nicolás de Hidalgo-UMSNH, Horticulture, 1575 Linden Drive, University of
Francisco J. Mújica S/N, Col. Felicitas de Rio, Wisconsin, Madison, WI 53706-1590, USA.

You might also like