ArticlePDF Available

Pucara (Amaryllidaceae) Reduced to Synonymy with Stenomesson on the Basis of Nuclear and Plastid DNA Spacer Sequences, and a New Related Species of Stenomesson

Authors:
  • Arizona State University (ASU) and Montgomery Botanical Center (MBC)

Abstract and Figures

Pucara leucantha is transferred to Stenomesson as Stenomesson leucanthum based on the phylogenetic position of Pucara resolved by nuclear and plastid DNA sequences. An allied species, Stenomesson chloranthum, is described from the Departments of Amazonas and Cajamarca in Peru, but at lower elevations. Both of these species release their pollen in tetrads, unique within Amaryllidaceae, and have tri-lobed stigmas, unique within Stenomesson. Meerow et al. (1986) described the first known oc-curence of mature pollen in tetrads within the Amar-yllidaceae for a species they determined as Stenomesson elwesii Macbr. from collections by Paul Hutchinson. Meerow later had the opportunity to examine the type of S. elwesii [now Clinanthus elwesii (Macbr.) Meerow] at Kew, and it became obvious that the name was mis-applied by Meerow et al. (1986). The identity of the Hutchinson collections remained unknown. While it clearly bore resemblance to a species of Stenomesson sensu stricto (Meerow et al. 2000), the palynological apomorphy and the low elevation of the collection were unusual. New collections from the Department of Amazonas in Peru in 1999 and 2001 were identified as represent-ing both this unidentified species and the sole species of Pucara Rav., P. leucantha Rav. (Ravenna 1972), allow-ing more detailed study. We observed that P. leucantha has leaf morphology similar to the undescribed spe-cies, and also releases its pollen in tetrads. In this pa-per, we show that the phylogenetic position of these two species based on DNA sequences leaves little doubt that the genus Pucara should not be recognized as distinct from Stenomesson (sensu Meerow et al. 2000). We further describe the misidentified taxon as a new species of Stenomesson, allied to the erstwhile P. leucantha.
Content may be subject to copyright.
511
Systematic Botany (2004), 29(3): pp. 511–517
qCopyright 2004 by the American Society of Plant Taxonomists
Pucara (Amaryllidaceae) Reduced to Synonymy with Stenomesson on the
Basis of Nuclear and Plastid DNA Spacer Sequences, and a New Related
Species of Stenomesson
A
LAN
W. M
EEROW
1
and H
ENK VAN DER
W
ERFF
2
1
USDA-ARS-SHRS, 13601 Old Cutler Road, Miami, Florida, 33158 and Fairchild Tropical Garden, 11935 Old
Cutler Road, Miami, Florida 33156;
2
Missouri Botanical Garden, P. O. Box 299, St. Louis, Missouri 63166-0299
Communicating Editor: James F. Smith
A
BSTRACT
.Pucara leucantha is transferred to Stenomesson as Stenomesson leucanthum based on the phylogenetic position
of Pucara resolved by nuclear and plastid DNA sequences. An allied species, Stenomesson chloranthum, is described from
the Departments of Amazonas and Cajamarca in Peru, but at lower elevations. Both of these species release their pollen in
tetrads, unique within Amaryllidaceae, and have tri-lobed stigmas, unique within Stenomesson.
Meerow et al. (1986) described the first known oc-
curence of mature pollen in tetrads within the Amar-
yllidaceae for a species they determined as Stenomesson
elwesii Macbr. from collections by Paul Hutchinson.
Meerow later had the opportunity to examine the type
of S. elwesii [now Clinanthus elwesii (Macbr.) Meerow] at
Kew, and it became obvious that the name was mis-
applied by Meerow et al. (1986). The identity of the
Hutchinson collections remained unknown. While it
clearly bore resemblance to a species of Stenomesson
sensu stricto (Meerow et al. 2000), the palynological
apomorphy and the low elevation of the collection were
unusual.
New collections from the Department of Amazonas
in Peru in 1999 and 2001 were identified as represent-
ing both this unidentified species and the sole species
of Pucara Rav., P. leucantha Rav. (Ravenna 1972), allow-
ing more detailed study. We observed that P. leucantha
has leaf morphology similar to the undescribed spe-
cies, and also releases its pollen in tetrads. In this pa-
per, we show that the phylogenetic position of these
two species based on DNA sequences leaves little
doubt that the genus Pucara should not be recognized
as distinct from Stenomesson (sensu Meerow et al.
2000). We further describe the misidentified taxon as a
new species of Stenomesson, allied to the erstwhile P.
leucantha.
M
ATERIALS AND
M
ETHODS
DNA Sequencing and Alignment. S
AMPLING
. Genomic DNA
was extracted from silica gel dried leaf tissue of the taxa listed in
Appendix 1 as described by Meerow et al. (2000).
DNA E
XTRACTION
,A
MPLIFICATION AND
S
EQUENCING
P
ROTOCOLS
.
The plastid atpß-rbcL spacer was amplified and sequenced using the
primers and polymerase chain reaction (PCR) protocol of Chiang
et al. (1998). Amplification of the ribosomal DNA ITS1/5.8S/ITS2
region was accomplished using flanking primers (18S, 26S) AB101
and AB102 (Douzery et al. 1999), and the original White et al.
(1990) internal primers ITS2 and 3 to amplify the spacers along
with the intervening 5.8S gene as described by Meerow et al.
(2000). All PCR amplifications were performed on an ABI 9700
(Perkin-Elmer Applied Biosystems, Foster City, California, USA).
Amplified products were purified using QIAquick (Qiagen, Va-
lencia, California, USA) columns, following manufacturer’s pro-
tocols. Cycle sequencing reactions were performed directly on pu-
rified PCR products on the ABI 9700, using standard dideoxycycle
protocols for sequencing with dye terminators on either an ABI
310 or ABI 3100 automated sequencer (according to the manufac-
turer’s protocols; Applied Biosystems, Foster City, California,
USA).
S
EQUENCE
A
LIGNMENT
. Both the ITS and atp-rbcL spacer ma-
trices were readily aligned manually using Sequencher 4.1 (Gene
Codes, Ann Arbor, Michigan, USA). We also used Clustal X (Hig-
gins and Sharp 1988; Thompson et al. 1997) to align the sequences
as a check against our manual alignments to help highlight any
ambiguous regions. The alignments and the parsimony trees used
in this paper can be accessed at TreeBase (study accession number
S1037, matrix accession numbers M1762–M1765).
Phylogenetic Analyses. Both DNA sequence matrices consisted
of 18 taxa (Appendix 1) representing the Andean tetraploid clade
resolved by Meerow et al. (2000) with ITS sequences. Four tribes
are represented: Clinantheae, Eustephieae, Hymenocallideae, and
Stenomesseae. The Eustephieae (Eustephia darwinii and Chlidanthus
fragrans) were used as outgroup, as this tribe is sister to the rest
of the clade (Meerow et al. 2000).
Aligned matrices were analyzed using the parsimony algorithm
of PAUP* for Macintosh (version 4.0b10; Swofford 1998), with the
MULPARS option invoked. Tree branches were retained only if
unambiguous support was available (i.e., branches were collapsed
if the minimum length 50). Gaps were coded as missing char-
acters in the initial analyses, but a gap matrix was also constructed
from the indel-rich plastid spacer region using the program PAUP-
GAP (Anthony Cox, formerly RBG Kew), which applies a strict
interpretation of gaps (i.e., only gaps of equal length are consid-
ered homologous among taxa). As Kelchner (2000) has pointed
out, indels within plastid non-coding regions are frequently as, if
not more, informative than the nucleotide sequences themselves.
This binary matrix was added to the sequence alignment and an-
alyzed in combination. For all matrices, a branch and bound (Hen-
dy and Penny 1982) search was conducted under the Fitch (equal)
weights (Fitch 1971) criterion with simple addition sequence.
We also combined the two data matrices, opting for the ‘‘total
evidence’’ approach (Dubuisson et al. 1998; Seelanan et al. 1997).
However, before combining the ITS and atpß-rbcL spacer data sets,
we performed partition homogeneity tests on the matrices (Farris
et al. 1994, 1995) to assess the degree of congruence between them.
One thousand heuristic searches were conducted for each test,
each with 10 random addition replications, saving no more than
20 trees from each for TBR branch swapping.
Internal support was determined by bootstrapping (Felsenstein
1985; 5000 heuristic replicates with simple addition, TRB branch-
swapping, saving 20 trees per replicate) and calculation of Bremer
(1988) decay indices (DI) using the program TreeRot v. 2.1 (Soren-
512 [Volume 29SYSTEMATIC BOTANY
F
IG
. 1. One of four equally most parsimonious trees found
by a cladistic analysis of the plastid atpß-rbcL spacer region
across 18 species of Amaryllidaceae. Numbers above branches
are branch lengths. Numbers below branches are bootstrap
support percentages/decay indices (italic). Large shadowed
numbers adjacent to nodes are ‘‘clade credibility’’ scoresfrom
500,000 generations of Bayesian analysis. A vertical white bar
indicates a branch that collapses in the strict consensus of all
four trees.
F
IG
. 2. One of three equally most parsimonious trees
found by a cladistic analysis of the plastid atpß-rbcL spacer
region plus a binary coded gap matrix across 18 species of
the Amaryllidaceae. Numbers above branches are branch
lengths. Numbers below branches are bootstrap support per-
centages and decay indices (italic).
son 1996). The cut-off bootstrap value is 50%. A bootstrap value
greater than 75% was considered good support, 65–75% was des-
ignated moderate support, and less than 65% as weak (Meerow
and Snijman 2001; Meerow et al. 2002). A branch and bound
search was implemented for each constraint statement postulated
by TreeRot. A minimum DI 52 was considered to represent good
support for a clade.
We also applied Bayesian analysis (Huelsenbeck et al. 2001) to
each sequence matrix, in order to approximate a bootstrap of max-
imum likelihood estimates of the phylogenetic relationships, and
check for congruence with the results of parsimony analysis. The
program MrBayes v. 3.04 (Huelsenbeck and Rohnquist 2001) was
used for this purpose. We first determined which model of nucle-
otide substitution to impose on our data with ModelTest v. 3.06
(Posada and Crandall 1998). We used the Akaike information cri-
terion (Akaike 1974) to choose the model with the best fit. We then
ran 500,000 generations of four simultaneous Markov chains with
Mr. Bayes, retaining the tree from every 100th generation (i.e., a
total of 5000 trees, less the trees produced before log likelihood
scores stabilized) from which a 50% majority rule consensus tree
was constructed. In each case, the log likelihood scores stabilized
after 10,000 generations of Bayesian analysis, but we dropped the
trees from 20,000 generations (200 of 5000) in constructing the 50%
majority consensus trees. The results of the Bayesian analyses are
reported as ‘‘clade credibility’’ (CC) scores, i.e., the percentage of
trees sampled where a given clade is resolved, which is equal to
the posterior probability of the clade existing (Huelsenbeck and
Rohnquist 2001). Only CC scores in excess of 50% are shown in
our trees (Figs. 1–5).
R
ESULTS
Plastid atpß-rbcL Spacer. The atpß-rbcL spacer ma-
trix yielded 37 parsimony informative characters out
of a total of 1276. The percentage of total matrix cells
coded as missing was 22.5%. A branch and bound
search found four equally parsimonious trees of length
5232, consistency index (CI) 50.93 and retention in-
dex (RI) 50.80). The topology (Fig. 1) is congruent
with the major clades of Andean genera resolved in
the large ITS analysis of American Amaryllidaceae by
Meerow et al. (2000), Pucara leucantha and Stenomesson
chloranthum are sister species with a bootstrap of 63%
and decay index (DI) 51. The nucleotide substitution
model that best fit the atpß-rbcL spacer data was the
Kimura 3-parameter model with unequal base fre-
quencies with gamma distribution (K81uf1G; Kimura
1981), which was approximated in Mr. Bayes with the
settings nst 52. The ‘‘Pucara’’ clade within Stenomes-
seae received a clade credibility score of 99% (Fig. 1).
When a binary gap matrix for the spacer region was
added to the sequence alignment, a total of 86 char-
acters were parsimony informative out of 1406 total.
Three equally parsimonious trees were found, of
length 5420, CI 50.82 and RI 50.69 (Fig. 2). Pucara
leucantha still resolves as sister species to Stenomesson
chloranthum but with weak bootstrap support 560
2004] 513MEEROW AND VAN DER WERFF: PUCARA
F
IG
. 3. Single most parsimonious tree found by a cladistic
analysis of the nrDNA ITS region across 18 species of the
Amaryllidaceae. Numbers above branches are branch lengths.
Numbers below branches are bootstrap support percentages/
decay indices (italic). Large shadowed numbers adjacent to
nodes are ‘‘clade credibility’’ scores from 500,000 generations
of Bayesian analysis.
F
IG
. 4. Single most parsimonious tree found by a cladistic
analysis of combined plastid atpß-rbcL spacer and nrDNA ITS
across 18 species of the Amaryllidaceae. Numbers above
branches are branch lengths. Numbers below branches are
bootstrap support percentages/decay indices (italic). Large
shadowed numbers adjacent to nodes are ‘‘clade credibility’’
scores from 500,000 generations of Bayesian analysis.
(though DI 52). Stenomesson flavum resolves as sister
to the Pucara clade with a bootstrap 566 and DI 52.
Pucara and its sister species are embedded within a
monophyletic Stenomesson clade (bootstrap 553, DI 5
1).
ITS. Of 646 total characters, 129 were parsimony
informative. The percentage of total matrix cells coded
as missing was 6.5%. A single most parsimonious tree
was found by the branch and bound search, of length
5401 steps, CI 50.85 and RI 50.85 (Fig. 3). Pucara
leucantha resolves as sister species to Stenomesson chlor-
anthum with bootstrap 571 and DI 51, nested in a
clade with S. miniatum and S. flavum with bootstrap 5
99 and DI 54. Stenomesson is monophyletic with a
bootstrap 571 and DI 52. For the Bayesian analysis,
a general time reversible model with a proportion of
invariant sites (GTR 1I; Lanave et al. 1984) was the
model that best fit our data. Clade credibility for the
sister relationship of P. leucantha and S. chloranthum was
98%, nested within a clade including S. miniatum and
S. flavum (CC 5100%). Stenomesson was monophyletic
with a CC 598%.
atpß-rbcL Spacer and ITS. The partition homoge-
neity test suggested that the two sequence matrices
were largely incongruent (P 50.01). However, if one
compares the trees supported by each of the two gene
regions (Figs. 1, 3), it is clear that the incongruity is
concentrated in the more terminal branches. The same
main clades are resolved by both sets of data. The
combined sequence matrix yielded 166 parsimony in-
formative characters out of 1922 total. Branch and
bound search found a single tree of length 5637 steps,
CI 50.87 and RI 50.83 (Fig. 4). Pucara leucantha and
Stenomesson chloranthum are again sister species, but
with a bootstrap of 86 (DI 52), within the same clade
resolved by ITS alone (bootstrap 581, DI 52). Boot-
strap support for a monophyletic Stenomesson rises to
85 (DI 52). The Bayesian consensus tree of 4800 sam-
pled trees from the analyses is highly congruent with
the parsimony analysis. Pucara leucantha and Stenomes-
son chloranthum are sister species in a clade with S.
flavum and S. miniatum with a CC5100% in both cases.
While the sister relationship of S. flavum to the Pucara
subclade has a CC of only 54%, Stenomesson is mono-
phyletic with a CC of 100%. Adding the atpß-rbcL spac-
er gap matrix to the combined sequence matrix raised
the number of parsimony informative characters to
215, and resulted in two equally parsimonious trees of
828 steps, CI 50.83, RI 50.78 (not shown), differing
only in the resolution of Stenomesson aurantiacum and
S. miniatum (sister species in one tree; forming a grade
514 [Volume 29SYSTEMATIC BOTANY
F
IG
.5. Stenomesson chloranthum and S. leucanthum. A–B. Flowers of S. chloranthum.A.Meerow 2520 (FTG). B. Meerow 1155
(FLAS). C. Leaf of S. leucanthum (Meerow 2522, FTG). D. Leaf detail of S. chloranthum (Meerow 2520, FTG). E. Flowers of S.
leucanthum (Meerow 2522, FTG). F. Pollen tetrad of S. chloranthum (Meerow 1155, FLAS). Scales A–E 51 cm, F 55mm.
with the Pucara subclade in the other). Other than this,
the tree topologies were the same as the one tree found
without the gap matrix added to the sequences.
D
ISCUSSION
The case for treating Pucara as part of the larger ge-
nus Stenomesson is unambiguously supported by both
plastid and nuclear non-coding sequences, whether an-
alyzed by parsimony or likelihood. Both P. leucantha
and the newly described S. chloranthum have shortly
sub-petiolate leaves with a well-developed midrib,
which is characteristic of the genus Stenomesson,the
limits of which were re-assessed by Meerow et al.
(2000) on the basis of nrDNA ITS sequences. The most
significant synapomorphy of P. leucantha and S. chlor-
anthum is the exclusive presence of pollen tetrads at
anthesis (Fig. 5f), a character state that is not known
to occur in any other species of the genus, or within
2004] 515MEEROW AND VAN DER WERFF: PUCARA
F
IG
.6. Stenomesson chloranthum (Meerow 2520, FTG). A. In-
florescence. B. Leaf. C. Longitudinal cross section of flower. D.
Longitudinal view of flower. E. Oblique view of flower. F. Te-
pals, outer (right), inner (left). G. Staminal cup. All scales 5
1 cm.
the entire family. Ravenna (1972) denoted the alter-
nating position of the free staminal filaments in P. leu-
cantha (those opposite the outer tepals are inserted at
the rim of the staminal cup; those opposite the inner
tepals, below the rim) and its tri-lobed stigma as the
main basis for generic recognition. While S. chloran-
thum has a less complex androecium, polymorphism
of staminal cup morphology within the Andean peti-
olate-leafed clade of Amaryllidaceae has been dem-
onstrated within the limits of a single species (Meerow
1989). Transfer of P. leucantha into Stenomesson is thus
warranted. Stenomesson leucanthum is the only white
flowered species of Stenomesson so far known. Nothing
is known about the pollination biology of the genus.
No other Stenomesson species from the interior of
Peru have ever been found below about 2000 m ele-
vation. Low elevation species have only before been
recorded from the coastal lomas of Peru, where max-
imum temperatures approximate those of Andean
high elevations. The same general region in Peru where
S. chloranthum and S. leucanthum are found is also host
to the three species of the bizarre, green-flowered, suc-
culent-leafed endemic genus Rauhia, which resolved in
a different sub-clade of the tribe Stenomesseae in Mee-
row et al.’s (2000) ITS phylogeny. The geographic con-
centration of such novelties suggests that the area bor-
dered by the lower ´o Utcubamba and middle ´o
Maran˜o´n was a hotspot for diversification in the tribe
Stenomesseae as the Andes rose to their present po-
sition.
T
AXONOMIC
T
REATMENT
Stenomesson leucanthum (Ravenna) Meerow & van
der Werff, comb. nov. (Figs. 5c, e). Pucara leucantha
Ravenna, Ann. Mus. Nat. Valparaiso 5: 85–89
(1972).—Type: Peru, Cajamarca, Jaen, San Anto-
nio (km 81—Pucara), 990 m, 12 Oct 1965, Saga´s-
tegui 5850 (holotype: HUT).
Representative Specimens Examined. PERU. Ama-
zonas: Pedro Ruiz to Chachapoyas, 1500 m, 24 Apr
2002 (pressed from bulbs collected by H. van der Werff
s. n.), Meerow 2523 (FTG); Utcubamba Valley between
Pedro Ruiz and branch to Chachapoyas, 14 Apr 2002
(pressed from bulbs collected by H. van der Werff s.
n.), Meerow 2522 (FTG).
Stenomesson chloranthum Meerow & van der Werff,
sp. nov. (Figs. 5a–b, d, f, 6).—TYPE: PERU. Ama-
zonas: Bagua, near Pongo de Rentema on the ´o
Maran˜on, 1 km east of Olmos on the Mesones-
Muro Highway, 370 m, 26 Jan 1964 (pressed from
living material 23 June 1965), Hutchinson and
Wright 3782 (holotype: UC!).
S. leucantho affine sed foliis olivaceis, floribus pallide
viridibus, staminibus insertis pariter margine coronae,
et praesentia altitudine inferiore differt.
Representative Specimens Examined. PERU. Ama-
zonas: Bagua Grande, 500 m, 25 Mar 2002 (pressed
from bulbs collected by H. van der Werff s. n.), Meerow
2520 (FTG). Cajamarca: Jaen, 2 km north of Chamaya
on rd. to Jaen, 450 m, 7 Feb 1964 (pressed 8 Jul 1965
from living material), Hutchinson & Wright 4123 (UC);
Cajamarca: a few km outside Jaen towards San Ignacio,
600 m 8 Apr 2002 (pressed from bulbs collected by H.
van der Werff s. n.), Meerow 2521 (FTG).
Bulbous perennial herb; bulb globose, with a short
(2–3 cm) neck, 2.5–4 cm diam, producing offset bulbils.
Leaves 2–3, hysteranthous or emerging with the flow-
ers, sub-petiolate, lanceolate, obtuse, 25–29 cm long, 2–
2.5 cm wide, olive-green (Royal Horticultural Society
Color Chart Green 137A), somewhat glaucous, with an
obscure midrib on the adaxial surface (pronounced on
the abaxial). Inflorescence scapose; scape terete, solid,
30–40 cm long, 7–8 mm diam proximally, 5 mm diam
distally, terminated by a pseudoumbel of 5–10 flowers
enclosed by two obvolute, greenish-white, marcescent
bracts before anthesis; bracts ovate-lanceolate, 2.5–3 cm
long, 10–11.5 mm wide at the base, acute. Perianth
campanulate-tubular, pale green, 5.8–6.7 cm long
(from base of tube to apex of limb); tube 3.5–4 cm
long, cylindrical and 2.5–3.1 mm wide at base, abrupt-
ly dilating in the distal half to 8.3–9 mm at the throat,
longitudinally striped white, limb spreading 18.5–19
mm wide. Tepals 6, in two series, outer 17.7–18.5 mm
long, 8–8.5 mm wide, elliptic, acute and minutely apic-
516 [Volume 29SYSTEMATIC BOTANY
F
IG
. 7. Map of Peru showing distributions of Stenomesson
chloranthum (open circles) and S. leucanthum (black squares).
ulate; inner 16.8–18.3 mm long, 8.3–8.9 mm wide, el-
liptic, obtuse, margins of both series hyaline. Stamens
6, white to greenish white, fused for the proximal 4–5
mm into a short staminal cup; free filaments broadly
subulate, edentate, 8.3–8.7 mm long, 2.7–3 mm wide,
abruptly tapering to 1 mm in their distal 2 mm, barely
exserted from the limb; anthers ca. 3 mm long, oblong,
dorsifixed, introrse; pollen white, released in tetrads
(Fig. 5f). Style filiform, ca. 6 cm long, exceeding the
stamens by about 5 mm; stigma trilobed. Ovary ellip-
soid, 7.5–8 mm long, 4.3–5 mm wide, locules 3, ovules
axile, numerous per locule, flattened and superposed.
Fruit a turbinate, loculicidal capsule, turning papery at
dehiscence; seeds numerous per locule, black, flat,
obliquely winged. 2n546.
Stenomesson chloranthum is endemic to the lower
slopes of the seasonally dry interandean valleys of the
Maran˜o´n and Utcubamba drainages in northern Ca-
jamarca and west-central Amazonas departments of
Peru (Fig. 7), between 350–600 m elevation, in dry
scrub, often growing with Opuntia and other Cacta-
ceae.
Stenomesson chloranthum can be distinguished from
S. leucanthum by its more glaucous, olive-green leaves
with a more obscure adaxial midrib (Fig. 5d), its fewer
but larger green and white owers, the more simpli-
fied structure of its androecium, and its lower altitu-
dinal limits. The white-flowered S. leucanthum is not
known from elevations lower than 900 m and occurs
up to 1650 m elevation. The adaxial mibrib is conspic-
uously visible on the less glaucous leaves of this spe-
cies (Fig. 5c).
A
CKNOWLEDGMENTS
. Karen Williams prepared the drawings
of S. chloranthum. This work was partially supported by National
Science Foundation Grants DEB-968787 and 0129179 to AWM.
L
ITERATURE
C
ITED
A
KAIKE
, H. 1974. A new look at the statistical model identification.
IEEE Transactions on Automatic Control 19: 716–723.
B
REMER
, K. 1988. The limits of amino acid sequence data in an-
giosperm phylogenetic reconstruction. Evolution 42: 198–213.
C
HIANG
, T-Y., B. A. S
CHAAL
, and C-I. P
ENG
. 1998. Universal prim-
ers for amplification and sequencing a noncoding spacer be-
tween the atpß and rbcL genes of chloroplast DNA. Botanical
Bulletin of the Academia Sinica 39: 245–250.
D
OUZERY
, J. P., A. M. P
RIDGEON
,P.K
ORES
,H.K
URZWEIL
,P.L
INDER
,
andM.W.C
HASE
. 1999. Molecular phylogenetics of Diseae
(Orchidaceae): a contribution from nuclear ribosomal ITS se-
quences. American Journal of Botany 86: 887–899.
D
UBUISSON
,J.Y.,R.H
EBANT
-M
AURI
,andJ.G
ALTIER
. 1998. Mole-
cules and morphology: conflicts and congruence within the
fern genus Trichomanes (Hymenophyllaceae). Molecular Phy-
logenetics and Evolution 9: 390–397.
F
ARRIS
,J.S,M.K
A
¨LLERSJO
¨
,A.G.K
LUGE
, and C. B
ULT
. 1994. Test-
ing significance of incongruence. Cladistics 10: 315–319.
———, ———, ———, and ———. 1995. Constructing a signifi-
cance test for incongruence. Systematic Biology 44: 570–572.
F
ELSENSTEIN
, J. 1985. Confidence limits on phylogenies: an ap-
proach using the bootstrap. Evolution 39: 783–791.
F
ITCH
, W. M. 1971. Toward defining the course of evolution: min-
imum change for a specific tree topology. Systematic Zoology
20: 406–416.
H
ENDY
,M.D.andD.P
ENNY
. 1982. Branch and bound algorithms
to determine minimal evolutionary trees. Mathematical Biosci-
ence 59: 277–290.
H
IGGINS
,D.G.andP.M.S
HARP
. 1988. CLUSTAL: a package for
performing multiple sequence alignment on a microcomput-
er. Gene 73: 237–244.
H
UELSENBECK
,J.P.,andF.R
ONQUIST
. 2001. MRBAYES: Bayesian
inference of phylogeny. Bioinformatics 17: 754–755.
———, ———, R. N
IELSEN
,andJ.P.B
OLLBACK
. 2001. Bayesian
inference of phylogeny and its impact on evolutionary biol-
ogy. Science 294: 2310–2314.
K
ELCHNER
, S. A. 2000. The evolution of non-coding chloroplast
DNA and its application in plant systematics. Annals of the
Missouri Botanical Garden 87: 482–498.
K
IMURA
, M. 1981. Estimation of evolutionary distances between
homologous nucleotide sequences. Proceedings of the National
Academy of Science, USA 78: 454–458.
L
ANAVE
,C.,G.P
REPARATA
,C.S
ACCONE
,andG.S
ERIO
. 1984. A
new method for calculating evolutionary substitution rates.
Journal of Molecular Evolution 20: 86–93.
M
EEROW
, A. W. 1989. Systematics of the Amazon lilies, Eucharis
and Caliphruria (Amaryllidaceae). Annals of the Missouri Bo-
tanical Garden 76: 136–220.
———, N. B. D
EHGAN
, and B. D
EHGAN
. 1986. Pollen tetrads in
Stenomesson elwesii (Amaryllidaceae). American Journal of Bot-
any 73: 1642–1644.
———, C. L. G
UY
, Q-B. L
I
, and S-L. Y
ANG
. 2000. Phylogeny of the
American Amaryllidaceae based on nrDNA ITS sequences.
Systematic Botany 25: 708–726.
———, ———, ———, and J. R. C
LAYTON
. 2002. Phylogeny of the
tribe Hymenocallideae (Amaryllidaceae) based on morphol-
ogy and molecular characters. Annals of the Missouri Botanical
Garden 89: 400–413.
2004] 517MEEROW AND VAN DER WERFF: PUCARA
——— and D. A. S
NIJMAN
. 2001. Phylogeny of Amaryllidaceae
tribe Amaryllideae based on nrDNA ITS sequences and mor-
phology. American Journal of Botany 88: 2321–2330.
P
OSADA
, D. and K. A. C
RANDALL
. 1998. Modeltest: testing the
model of DNA substitution. Bioinformatics 14: 817–818.
R
AVENNA
, P. 1972. Pucara,ge´nero nuevo de Amaryllidaceae del
norte de Peru. Annales del Museo Nacional de Valparaiso 5: 85–
89.
S
EELANAN
, T., A. S
CHNABEL
,andJ.W
ENDEL
. 1997. Congruence
and consensus in the cotton tribe (Malvaceae). Systematic Bot-
any 22: 259–290.
S
ORENSON
, M. D. 1996. TreeRot. University of Michigan, Ann Ar-
bor.
S
WOFFORD
, D. L. 1998. Phylogenetic analysis using parsimony, v.
4.0 beta. Sutherland: Sinauer Associates,
T
HOMPSON
,J.D.,T.J.G
IBSON
,F.P
LEWNIAK
,F.J
EANMOUGIN
,and
D. G. H
IGGINS
. 1997. The ClustalX windows interface: flexible
strategies for multiple sequence alignment aided by quality
analysis tools. Nucleic Acids Research 24: 4876–4882.
W
HITE
,T.J.,T.B
RUNS
,S.L
EE
,andJ.T
AYLOR
. 1990. Amplification
and direct sequencing of fungal ribosomal RNA genes for
phylogenetics. Pp. 315–322 in PCR protocols: a guide to methods
and applications, eds. M. Innis, D. Gelfand, J. Sninsky, and T.
White. Orlando: Academic Press.
A
PPENDIX
1
Species, vouchers, and Genbank accession numbers (or litera-
ture citations for previously published sequences) of DNA se-
quences used in this paper. All vouchers deposited at FTG unless
otherwise stated.
Clinanthus humilis (Herb.) Meerow—Meerow 2442:atpß-rbcL spac-
er AY460393, ITS Meerow et al. (2000). C. incarnatus (H.B.K.)
Meer ow— Meerow 1120:atpß-rbcL spacer AY460392, ITS Mee-
row et al. (2000). C. mirabilis (Rav.) Meerow—S. Leiva et al.
2000 (HUT): atpß-rbcL spacer AY460391, ITS Meerow et al.
(2000)
Chlidanthus fragrans Herb.—Meerow 2312:atpß-rbcL spacer
AY460390, ITS Meerow et al. (2000)
Eucharis castelnaeana (Baill.) MacBride—Schunke 14156 (FLAS):
atpß-rbcL spacer AY460405, ITS Meerow et al. (2000). E. for-
mosa Meerow—Whitten et al. 95020 (FLAS): atpß-rbcL spacer
AY460406, ITS Meerow et al. (2000)
Eucrosia dodsonii Meerow & Dehgan—Meerow 1115:atpß-rbcL
spacer AY460404, ITS Meerow et al. (2000)
Eustephia darwinii Vargas—Meerow 2436:atpß-rbcL spacer
AY460389, ITS Meerow et al. (2000)
Hymenocallis chiapasiana T. M. Howard—T. M. Howard 1185
(MO): atpß-rbcL spacer AY460396, ITS AY461739. H. glauca
M. Roem.—Meerow 2433:atpß-rbcL spacer AY460395, ITS
Meerow et al. (2000). H. latifolia (Mill.) M. Roem.—Meerow
2438:atpß-rbcL spacer AY460398, ITS Meerow et al. (2000).
H. tubiflora Salisb.—Meerow 2440:atpß-rbcL spacer
AY460399, ITS Meerow et al. (2000)
Ismene vargasii (Velarde) Gereau & Meerow—Meerow 2308:atpß-
rbcL spacer AY460397, ITS Meerow et al. (2000)
Stenomesson aurantiacum Herb.—Meerow 1061 (FLAS): atpß-rbcL
spacer AY460394, ITS Meerow et al. (2000). S. chloranthum
Meerow & van der Werff—Meerow 2520:atpß-rbcL spacer
AY460403, ITS AY461738. S. flavum (R. & P.) Herb.—Meerow
2430:atpß-rbcL spacer AY460402, ITS Meerow et al. (2000). S.
leucanthum (Rav.) Meerow & van der Werff (Pucara leucantha
Rav.)—Meerow 2522:atpß-rbcL spacer AY460401, ITS
AY461737. S. miniatum (Herb.) Ravenna—Meerow 1118:atpß-
rbcL spacer AY460400, ITS AY461736.
... previously reported in Ecuador [7]. These plants usually occur in seasonally dry, grassy vegetation or at the margins of cloud forests above 2000 m elevation, but are also found in Peruvian inter-Andean valleys below 2000 m [8], and in the loma formations along the coast of this country, being all the petiolate-leafed Stenomesseae more closely related to Eucharideae than to the lorate leafed Stenomesseae [5,9]. The genus Clinanthus Herb. ...
... The genus Clinanthus Herb. (Amaryllidaceae), endemic to Peru and Ecuador [10], was separated from Stenomesson, and is primarily known in its same locations [8,11,12]. ...
... Regarding the traditional use of these plants, Urceolina species have been documented as compresses applied to sores and tumors by native people in Ecuador, and the Jíbaro indigenous people in Peru have utilized them for treating facial blemishes and acne [13]. Additionally, archeological findings at Inca ruins in South America suggest that certain Amaryllidaceae genera, such as Stenomesson, were depicted on ceremonial drinking vessels, and may have held significance within popular culture [8,14]. In recent years, the Amaryllidaceae species from Ecuador have garnered significant attention due to their alkaloid profiling and potential therapeutic benefits against Alzheimer's disease. ...
Preprint
Full-text available
Ecuador is an important center of biodiversity of the plant family Amaryllidaceae, specially the subfamily Amaryllidoideae, known as a source of important bioactive molecules. The aim of this study was to assess the chemical and biological potential of four different species of Amaryllidoideae collected in Ecuador: Urceolina formosa, Urceolina ruthiana, Clinanthus incarnatus, and Stenomesson aurantiacum. Sixteen alkaloids were identified in the bulb extract of these species by GC-MS, which the extract of S. aurantiacum exhibiting the greatest structural diversity. The potential of these three genera against Alzheimer’s disease was evaluated by measuring their AChE and BuChE inhibitory activity, revealing that C. incarnatus and U. formosa (from Sucumbíos province) showed the best results with IC50 values of 1.73 ± 0.25 and 30.56 ± 1.56 g·ml-1, respectively. Molecular dynamic assays were conducted to elucidate the possible interactions that occurs among 2-hydroxyanhydrolycorine and AChE enzyme, suggesting some similarity with those observed for galanthamine. This study enhances our understanding of the biodiversity of Amaryllidoideae species from Ecuador, highlighting their potential as a source of chemical compounds with pharmaceutical potential.
... This species was the first branch in Stenomesson (Figures 2, 3), and may represent an isolated eastern relict of early diversification of the subclade (Figure 7), as it is known only from wet forest in the vicinity of Moyobamba, Peru (Meerow, 1989). The two species of Stenomesson that release their pollen in tetrads (Meerow et al., 1986;Meerow and van der Werff, 2004), S. chloranthum Meerow & van der Werff and S. leucanthum, (Ravenna) Meerow & van der Werff, were sister species in every tree generated (Figures 2-4 and Supplementary Figures 2-7). Ravenna (1988) Stenomesson sect. ...
... comm), perhaps an anthropogenically facilitated disjunction. Northern Peru is where the greatest diversity of Stenomesson diversity can be found (Meerow, 1987b;Meerow and van der Werff, 2004). The possible influence of humans on the modern range of species in the Andean clade, though difficult to document, should not be discounted. ...
... Meerow (2010) characterized the Andean tetraploid clade's polyploid genome as grist for the mill of the Andean orogeny. Northern Peru in particular, comprising in part the Neotropical bioprovinces Cauca, Desert and Ecuadorian (Morrone, 2014), with its complex of microhabitats, has been a hotspot of diversification in the Andean clade (Meerow and van der Werff, 2004;Meerow and Cano, 2019;Meerow and Nakamura, 2019;Esquerre and Meerow, 2020). The distributions of Rauhia in Peru and Phaedranassa in Ecuador (Meerow, 1990;Oleas et al., 2012Oleas et al., , 2013Oleas et al., , 2016 fit the allopatric models of dispersal or vicariance (Figure 10 and Supplementary Figure 9) followed by isolation described for Andean complexes of Buddleja L. (Norman, 2000), Calceolaria L. (Molau, 1988), Fuchsia L. (Berry, 1982), and many other taxa (Ayers, 1999;Beck and Richter, 2008;Antonelli et al., 2009;Pennington et al., 2010;Anthelme et al., 2014;Luebert and Weigend, 2014;Bacon et al., 2018). ...
Article
Full-text available
One of the two major clades of the endemic American Amaryllidaceae subfam. Amaryllidoideae constitutes the tetraploid-derived (n = 23) Andean-centered tribes, most of which have 46 chromosomes. Despite progress in resolving phylogenetic relationships of the group with plastid and nrDNA, certain subclades were poorly resolved or weakly supported in those previous studies. Sequence capture using anchored hybrid enrichment was employed across 95 species of the clade along with five outgroups and generated sequences of 524 nuclear genes and a partial plastome. Maximum likelihood phylogenetic analyses were conducted on concatenated supermatrices, and coalescent-based species tree analyses were run on the gene trees, followed by hybridization network, age diversification and biogeographic analyses. The four tribes Clinantheae, Eucharideae, Eustephieae, and Hymenocallideae (sister to Clinantheae) are resolved in all analyses with > 90 and mostly 100% support, as are almost all genera within them. Nuclear gene supermatrix and species tree results were largely in concordance; however, some instances of cytonuclear discordance were evident. Hybridization network analysis identified significant reticulation in Clinanthus, Hymenocallis, Stenomesson and the subclade of Eucharideae comprising Eucharis, Caliphruria, and Urceolina. Our data support a previous treatment of the latter as a single genus, Urceolina, with the addition of Eucrosia dodsonii. Biogeographic analysis and penalized likelihood age estimation suggests an origin in the Cauca, Desert and Puna Neotropical bioprovinces for the complex in the mid-Oligocene, with more dispersals than vicariances in its history, but no extinctions. Hymenocallis represents the only instance of long-distance vicariance from the tropical Andean origin of its tribe Hymenocallideae. The absence of extinctions correlates with the lack of diversification rate shifts within the clade. The Eucharideae experienced a sudden lineage radiation ca. 10 Mya. We tie much of the divergences in the Andean-centered lineages to the rise of the Andes, and suggest that the Amotape—Huancabamba Zone functioned as both a corridor (dispersal) and a barrier to migration (vicariance). Several taxonomic changes are made. This is the largest DNA sequence data set to be applied within Amaryllidaceae to date.
... Of the external selective forces operating on the polyploid genome of the Andean clade, the most obvious is the rise of the Andes itself, an association that has been noted repeatedly (Meerow and Dehgan 1985;Meerow 1987aMeerow ,b, 1989, and has also been implicated in many other plant families (Luteyn 2002;Young et al. 2002). Northern Peru in particular, with its complex of microhabitats has been a hotspot of diversification in the Andean clade (Meerow and van der Werff 2004). The distributions of Rauhia in Peru and Phaedranassa in Ecuador (Meerow 1990) appear similar to the allopatric models of dispersal or vicariance followed by isolation described for Andean complexes of Buddleja L. (Norman 2000), Calceolaria L. (Molau 1988), and Fuchsia L. (Berry 1982). ...
... It is apparent that the Andean tetraploid clade has been able to diversify at a level, and perhaps rate, higher than the hippeastroids. Such novelties as the succulent leaf morphology of Rauhia, the pollen tetrads of two species of Stenomesson endemic to the same valleys as Rauhia (Meerow and van der Werff 2004), and the bulbiform seed of the tribe Hymenocallideae with its thick, chlorenchymous outer integument with a well-developed vascular system and a starch-storing embryo (Whitehead and Brown 1940) bespeak the evolutionary power of genome doubling to overcome canalization during a period of early explosive radiation. Yet canalization is still at play, or has re-asserted itself, mediating convergent shifts in floral morphology implicated in pollination, often in association with altitudinal changes. ...
Chapter
Full-text available
Canalization is defined as the suppression of phenotypic variation, or, in the context of molecular evolution, genetic buffering that has evolved under natural selection in order to stabilize the phenotype. Very little is understood on the processes behind canalization, even in today's genomic era. Canaliza-tion seems to be a powerful force in the Amaryllidaceae from which only a few lineages have escaped, and this has often in the past led to misdiagnosis of phylogenetic relationships. The American Amaryllidaceae resolves into two clades with plastid and nrDNA ITS sequences – the primarily diploid or dysploid hippeastroid clade (x = 11; n = 8, 9, 10, 11, 12), within which there is limited floral morphological diversity, and the Andean tetraploid-derived clade (x = 23), which exhibits greater phenotypic diversity. Moreover, several patterns of morphological diversity in the Andean clade manifest in a homoplastic mosaic at different taxonomic levels in various subclades. Both major clades consist of a small tribe (2-4 genera) that is sister to a larger clade, but only in the hippeastroid group do two genera, Rhodophiala and Zephyranthes, appear polyphyletic based on ITS sequences. Analysis of recombination signal in the ITS alignment supports a hypothesis of reticulation between clades of the hippeastroids, but none in the Andean group. The genus Pyrolirion, historically allied with Zephyranthes on the basis of uniflory, spathe fusion, and upright funnelform, actinomorphic flowers is shown to belong to the tribe Eustephieae in the Andean clade. The formal transfer of Rhodophiala cipoana into Hippeastrum as H. cipoanum is made. The mosaic patterns of phenotypic diversity are discussed in the context of genome duplication, possible reticulation, and biogeographic factors.
... previously reported from Ecuador (Meerow 1990). The genus usually occurs in seasonally dry, grassy vegetation or at the margins of cloud forest above 2000 m elevation, but is also found in Peruvian inter-Andean valleys below 2000 m (Meerow and van der Werff 2004 ...
Article
Full-text available
New species of the genera Stenomesson and Eucharis (Amaryllidaceae) are described from Ecuador. Stenomesson ecuadorense is the second species of the genus reported from that country, and the only endemic one. It is related to Stenomesson miniatum and Stenomesson campanulatum, both from Peru, with which it shares orange flower color and the fusion of the staminal corona to the perianth tube. It differs from Stenomesson miniatum by the non-urceolate perianth, from Stenomesson campanulatum by its shorter stamens and longer perianth, and from both by its lower montane, cloud forest habitat. Eucharis ruthiana, found in the vicinity of Zamora, is related to Eucharis moorei from which it differs by the narrower leaves and tepals; short, deeply cleft staminal corona; the long teeth on either side of the free filaments; the narrowly subulate, incurved free filaments; and the shorter style. The green mature fruit and campanulate floral morphology place it in Eucharis subg. Heterocharis.
... for the lorate-leafed species of Stenomesson , which were placed in the tribe Clinantheae Meerow together with Pamianthe and Paramongaia . A later paper ( Meerow & van der Werff, 2004 ) supported this systematic approach, and showed further that the monotypic genus Pucara belonged with the petiolateleafed species which remained in Stenomesson . The latter paper also suggested that the closest relatives of Stenomesson sensu stricto are Eucrosia Ker Gawl and the rainforest understorey amazon lilies Eucharis Planch. ...
Article
The systematic relationships of Stenomesson Herb. and other Andean genera of Amaryllidaceae are discussed and one of its members, S. pearcei Baker, is compared to similar species, described and illustrated. The cultivation requirements, habitat, distribution and conservation status of S. pearcei are discussed. A lectotype is chosen.
... One hundred heuristic searches with random addition sequence were implemented for each constraint statement postulated by TreeRot, saving no more than 10 trees per search. A minimum DI = 2 was considered to represent good support for a clade [118,119]. Maximum likelihood. ...
Article
Full-text available
As a consequence of there being several ways in which observed patterns of variation in nature can be conveyed in a generic classification, long recognised genera have changed in size over time. The generic rank has its origins in folk taxonomy, where genera were homogenous units of relatively few kinds. In the era of Bentham there was a widespread preference for large genera, many of which were split during the 20th century. In a survey of contemporary (1998-2007) generic delimitation practice we found a significant dichotomy between studies that incorporate molecular data and those that rely exclusively on morphological data. The former lead to delimitation of larger genera whereas the latter in general do not. This finding spurred a broader investigation into what drives changes in overall generic sizes, new data sources or new concepts? Two new data types have been introduced during the course of history: detailed morphology (anatomy, cytology) and chemical data (amino acid and DNA sequence data). Conceptual development has seen several turns: from language and communication, through memory and stability, to evolution and monophyly. We argue that conceptual change has a greater impact than changes in data do, since new data must be interpreted and translated into a classification and since conceptual changes may spur a search for new kinds of data. We conclude that the current trend toward recognising larger genera is a result of a return to study on a broad scale, rather than of incorporation of molecular data.
Article
Full-text available
The Trigoniaceae are represented in Peru by one genus, Trigonia, and six species (Brako & Zarucchi, 1993), between trees, lianas and shrubs. Here we recognize one endemic species, which is found in Humid Lowland Amazonian Forests region, to 100 m elevation. It has not been recorded within Peru's protected areas system. La familia Trigoniaceae es reconocida en el Perú por presentar un género, Trigonia, y seis especies (Brako & Zarucchi, 1993), entre arbustos, lianas y árboles. En este trabajo reconocemos una especie endémica, la cual ha sido encontrada en la región Bosques Húmedos Amazónicos, a 100 m de altitud. No ha sido registrada dentro del Sistema Nacional de Áreas Naturales Protegidas por el Estado.
Article
Full-text available
The Polygalaceae are represented in Peru by six genera and 98 species (Brako & Zarucchi, 1993; Ulloa Ulloa et al., 2004), mostly shrubs. Here we recognize 46 endemic taxa in three genera. Monnina is the genus with the largest number of endemic species. Endemic Polygalaceae are found mainly in Very Humid Montane Forests, Mesoandean and Very Humid Premontane Forests regions, between 600 and 3550 m elevation. We applied IUCN categories and criteria to 19 species. Apparently, only eight species have been recorded within Peru's protected areas system. La familia Polygalaceae es reconocida en el Perú por presentar seis géneros y 98 especies (Brako & Zarucchi, 1993; Ulloa Ulloa et al., 2004), principalmente arbustos. En este trabajo reconocemos 46 endemismos en tres géneros. El género más rico en especies es Monnina. Los taxones endémicos se encuentran principalmente en los Bosques Muy Húmedos Montanos, Mesoandina y Bosques Muy Húmedos Premontanos, entre los 600 y 3500 m de altitud. Se aplicaron las categorías y criterios de la UICN a 19 especies. Aparentemente, sólo ocho taxones endémicos se encuentran representados dentro del Sistema Nacional de Áreas Naturales Protegidas por el Estado.
Article
Full-text available
The Polemoniaceae are represented in Peru by five genera and 24 species (Brako & Zarucchi, 1993; Ulloa Ulloa et al., 2004), between herbs, shrubs and vines. Here we recognize nine endemic taxa in four genera. One genus, Huthia, is endemic to Peru. Cantua is the genus with the largest number of endemic species, and it requires a modern taxonomic evaluation. Endemic species in this family are found mainly in Mesoandean and Desert Shrubland regions, from 1900 m to 3600 m elevation. Apparently, only one species has been recorded in the Peruvian System of Protected Natural Areas. La familia Polemoniaceae es reconocida en el Perú por presentar cinco géneros y 24 especies (Brako & Zarucchi, 1993; Ulloa Ulloa et al., 2004), hierbas, arbustos y bejucos. En este trabajo reconocemos nueve especies endémicas en cuatro géneros. Un género, Huthia, es endémico del Perú. El género más rico en especies es Cantua, el cual requiere de una evaluación taxonómica. Las especies endémicas se encuentran principalmente en las regiones Mesoandina y Matorral Desértico, desde los 1900 m hasta los 3600 m de altitud. Aparentemente, sólo una especie endémica ha sido registrada dentro del Sistema Nacional de Áreas Naturales Protegidas por el Estado.
Article
Full-text available
Universal primers for amplifying and sequencing a noncoding spacer between the atpB and rbcL genes ofthe chloroplast DNA were constructed from the published sequences of Marchantia (a liverwort), tobacco, and rice. Our results indicate an evolutionary trend of increasing spacer size from liverworts, through mosses, to vascular plants. This atpB-rbcL spacer is AT-rich, consistent with other chloroplast noncoding spacers. Due to weak functional constraints, the spacer is evolving rapidly. A sequence identity of 92.2% was observed between spacers of two closely related moss species, Rhytidiadelphus loreus and R. triquestrus. Insertion/deletion events are common in the evolution of this spacer. A 23 bp deletion occurrs in R. loreus. Variation is found between two populations of Amorphophallus henryi (Araceae) and between individuals in a populations of Pasania formosana (Fagaceae). We suggest that this spacer will be useful for molecular systematics at the subspecific, specific, and generic levels and, in some species, for population genetics studies.
Article
Full-text available
Eucharis and Caliphruria are closely related, neotropical genera of petiolate-leaved, white-flowered Amaryllidaceae found in the understory of primary tropical rain forest. Together with the Peruvian endemic Urceolina, Eucharis and Caliphruria form a monophyletic group within "infrafamily Pancratioidinae" on the basis of leaf and seed morphology and ecological specialization. Seventeen species and two natural hybrids within two subgenera are recognized in Eucharis. Eucharis, marked by its crateriform or campanulate flowers, curved perianth tube, pigmented androecium usually developed into a staminal cup, and unicellular stigmatic papillae, is distributed from Guatemala to Bolivia, chiefly in the western Amazon basin and adjacent lower slopes of the eastern Andes. Subgenus Heterocharis represents three relict species and two natural hybrids with many ancestral characters of the genus. Caliphruria (four species, three of which are endemic to Colombia) has funnelform flowers, straight perianth tube, reduced staminal connation, and multicellular stigmatic papillae. Cladistic relationships support the monophyly of Eucharis, Caliphruria, and Urceolina as a unified group. The most parsimonious cladograms suggest that Eucharis is paraphyletic if subg. Heterocharis is included. An alternative hypothesis of generic relationships is tested, and proves less parsimonious. Recognizing the close phenetic relationship of subg. Heterocharis to the rest of Eucharis, the high levels of homoplasy within the cladogram, and ambiguity caused by missing character state data, a less parsimonious phylogeny is accepted as the basis for classification. Keys and descriptions are provided for all species of Eucharis and Caliphruria. Eucharis caucana, a novel hexaploid (2n = 138) species, is described from Colombia.
Article
Full-text available
As a discipline, phylogenetics is becoming transformed by a flood of molecular data. These data allow broad questions to be asked about the history of life, but also present difficult statistical and computational problems. Bayesian inference of phylogeny brings a new perspective to a number of outstanding issues in evolutionary biology, including the analysis of large phylogenetic trees and complex evolutionary models and the detection of the footprint of natural selection in DNA sequences.
Article
Full-text available
ROGERS, DS, IF GREENBAUM, SJ GUNN, AND MD ENGSTROM. 1984. Cytosystematic value of chromosomal inversion data in the genus Peromyscus (Rodentia: Cricetidae). J. Mammal. 65: 457-465. RUVOLO, M. 1992. Molecular evolutionary processes can produce ...
Article
The recently-developed statistical method known as the "bootstrap" can be used to place confidence intervals on phylogenies. It involves resampling points from one's own data, with replacement, to create a series of bootstrap samples of the same size as the original data. Each of these is analyzed, and the variation among the resulting estimates taken to indicate the size of the error involved in making estimates from the original data. In the case of phylogenies, it is argued that the proper method of resampling is to keep all of the original species while sampling characters with replacement, under the assumption that the characters have been independently drawn by the systematist and have evolved independently. Majority-rule consensus trees can be used to construct a phylogeny showing all of the inferred monophyletic groups that occurred in a majority of the bootstrap samples. If a group shows up 95% of the time or more, the evidence for it is taken to be statistically significant. Existing computer programs can be used to analyze different bootstrap samples by using weights on the characters, the weight of a character being how many times it was drawn in bootstrap sampling. When all characters are perfectly compatible, as envisioned by Hennig, bootstrap sampling becomes unnecessary; the bootstrap method would show significant evidence for a group if it is defined by three or more characters.
Article
Amino acid sequence data are available for ribulose biphosphate carboxylase, plastocyanin, cytochrome c, and ferredoxin for a number of angiosperm families. Cladistic analysis of the data, including evaluation of all equally or almost equally parsimonious cladograms, shows that much homoplasy (parallelisms and reversals) is present and that few or no well supported monophyletic groups of families can be demonstrated. In one analysis of nine angiosperm families and 40 variable amino acid positions from three proteins, the most parsimonious cladograms were 151 steps long and contained 63 parallelisms and reversals (consistency index = 0.583). In another analysis of six families and 53 variable amino acid positions from four proteins, the most parsimonious cladogram was 161 steps long and contained 50 parallelisms and reversals (consistency index = 0.689). Single changes in both data matrices could yield most parsimonious cladograms with quite different topologies and without common monophyletic groups. Presently, amino acid sequence data are not comprehensive enough for phylogenetic reconstruction among angiosperms. More informative positions are needed, either from sequencing longer parts of the proteins or from sequencing more proteins from the same taxa.
Article
We explored the evolutionary history of the Gossypieae and Gossypium using phylogenetic analysis of biparentally and maternally inherited characters. Separate and combined data sets were analyzed and incongruence between data sets was quantified and statistically evaluated. At the tribal level, phylogenetic analyses of nuclear ribosomal ITS sequences yielded trees that are highly congruent with those derived from the plastid gene ndhF, except for species that have a reticulate evolutionary history or for clades supported by few characters. Problematic taxa were then pruned from the data sets and the phylogeny was inferred from the combined data. Results indicate that 1) the Gossypieae is monophyletic, with one branch from the first split being represented by modern Cienfuegosia; 2) Thespesia is not monophyletic, and 3) Gossypium is monophyletic and sister to an unexpected clade consisting of the Hawaiian genus Kokia and the east African/Madagascan genus Gossypioides. Based on the magnitude of ndhF sequence divergence, we suggest that Kokia and Gossypioides diverged from each other in the Pliocene, subsequent to their apparent loss of a pair of chromosomes via chromosome fusion. Phylogenetic relationships among species and "genome groups" in Gossypium were assessed using cpDNA restriction site variation and ITS sequence data. Both data sets support the monophyly of each genome group, once taxa known or suspected to have reticulate histories are pruned from the trees. There was little congruence between these two data sets, however, with respect to relationships among genome groups. Statistical tests indicate that most incongruence is not significant and that it probably reflects insufficient information rather than a biological process that has differentially affected the data sets. We propose that the differing cpDNA- and ITS-based resolutions of genome groups in Gossypium reflect temporally closely spaced divergence events early in the diversification of the genus. This "short internode" phenomenon is suggested to be a common cause of phylogenetic incongruence.