ArticlePDF Available

The origin and diversification of angiosperms

Authors:

Abstract

The angiosperms, one of five groups of extant seed plants, are the largest group of land plants. Despite their relatively recent origin, this clade is extremely diverse morphologically and ecologically. However, angiosperms are clearly united by several synapomorphies. During the past 10 years, higher-level relationships of the angiosperms have been resolved. For example, most analyses are consistent in identifying Amborella, Nymphaeaceae, and Austrobaileyales as the basalmost branches of the angiosperm tree. Other basal lineages include Chloranthaceae, magnoliids, and monocots. Approximately three quarters of all angiosperm species belong to the eudicot clade, which is strongly supported by molecular data but united morphologically by a single synapomorphy-triaperturate pollen. Major clades of eudicots include Ranunculales, which are sister to all other eudicots, and a clade of core eudicots, the largest members of which are Saxifragales, Caryophyllales, rosids, and asterids. Despite rapid progress in resolving angiosperm relationships, several significant problems remain: (1) relationships among the monocots, Chloranthaceae, magnoliids, and eudicots, (2) branching order among basal eudicots, (3) relationships among the major clades of core eudicots, (4) relationships within rosids, (5) relationships of the many lineages of parasitic plants, and (6) integration of fossils with extant taxa into a comprehensive tree of angiosperm phylogeny.
1614
American Journal of Botany 91(10): 1614–1626. 2004.
T
HE ORIGIN AND DIVERSIFICATION OF ANGIOSPERMS
1
P
AMELA
S. S
OLTIS
2,4
AND
D
OUGLAS
E. S
OLTIS
3
2
Florida Museum of Natural History, University of Florida, Gainesville, Florida 32611 USA; and
3
Department of Botany, University
of Florida, Gainesville, Florida 32611 USA
The angiosperms, one of five groups of extant seed plants, are the largest group of land plants. Despite their relatively recent origin,
this clade is extremely diverse morphologically and ecologically. However, angiosperms are clearly united by several synapomorphies.
During the past 10 years, higher-level relationships of the angiosperms have been resolved. For example, most analyses are consistent
in identifying Amborella, Nymphaeaceae, and Austrobaileyales as the basalmost branches of the angiosperm tree. Other basal lineages
include Chloranthaceae, magnoliids, and monocots. Approximately three quarters of all angiosperm species belong to the eudicot clade,
which is strongly supported by molecular data but united morphologically by a single synapomorphy—triaperturate pollen. Major
clades of eudicots include Ranunculales, which are sister to all other eudicots, and a clade of core eudicots, the largest members of
which are Saxifragales, Caryophyllales, rosids, and asterids. Despite rapid progress in resolving angiosperm relationships, several
significant problems remain: (1) relationships among the monocots, Chloranthaceae, magnoliids, and eudicots, (2) branching order
among basal eudicots, (3) relationships among the major clades of core eudicots, (4) relationships within rosids, (5) relationships of
the many lineages of parasitic plants, and (6) integration of fossils with extant taxa into a comprehensive tree of angiosperm phylogeny.
Key words: Amborella; angiosperms; phylogeny.
The angiosperms, or flowering plants, one of the major
clades of extant seed plants (see Burleigh and Mathews, 2004,
in this issue), are the largest group of embryophytes, with at
least 260000 living species classified in 453 families (APG II,
2003). Angiosperms are amazingly diverse. They occupy ev-
ery habitat on Earth except the highest mountaintops, the re-
gions immediately surrounding the poles, and the deepest
oceans, and they occur as epiphytes, floating and rooted aquat-
ics in both freshwater and marine habitats, and terrestrial
plants that vary tremendously in size, longevity, and overall
form. Furthermore, the diversity in chemistry, reproductive
morphology, and genome size and organization is unparalleled
in the Plant Kingdom.
Despite their diversity, angiosperms are clearly united by a
suite of synapomorphies (i.e., shared, derived features), in-
cluding double fertilization and endosperm formation, the car-
pel, stamens with two pairs of pollen sacs, features of game-
tophyte structure and development, and phloem tissue com-
posed of sieve tubes and companion cells (see Doyle and Don-
oghue, 1986; and P. Soltis et al., 2004, for further discussion).
This evidence strongly negates hypotheses of polyphyletic or-
igins of extant angiosperms.
The fossil record of the angiosperms extends back at least
to the early Cretaceous, conservatively 130 million years ago
(mya) (see Crane et al., 2004). Floral size, structure, and or-
ganization among early angiosperms varied tremendously,
ranging from small (i.e., ,1 cm in diameter) flowers of fossil
Chloranthaceae and many other lineages (reviewed in Friis et
al., 2000), both extant and extinct, to the large, Magnolia-like
flowers of Archaeanthus (Dilcher and Crane, 1984). This floral
diversity in the fossil record is consistent with an early radi-
ation of angiosperms and associated diversification in floral
form (e.g., Friis et al., 2000).
Large-scale collaborations among angiosperm systematists
have greatly improved our understanding of angiosperm phy-
1
Manuscript received 6 February 2004; revision accepted 1 July 2004.
The authors thank M. Chase, J. Palmer, and an anonymous reviewer for
very helpful comments on the manuscript. This research was supported in
part by NSF grants DEB-0090283 and PGR-0115684.
4
E-mail: psoltis@flmnh.ufl.edu.
logeny. Strong support for many clades that correspond to tra-
ditionally recognized families provided early confidence that
the molecular-based trees were producing reasonable recon-
structions of phylogeny. However, some traditional families
and many orders and higher groups have been shown to be
nonmonophyletic, while many groups of previously uncertain
placement have been placed with great confidence. The An-
giosperm Phylogeny Group, an international consortium of
systematists, recognized the need for a new classification that
reflects current views of angiosperm phylogeny (APG, 1998;
APG II, 2003). An abridged version of the classification is
given in Appendix 1 (see Supplemental Data accompanying
online version of this paper) and at the Deep Time website
(http://flmnh.ufl.edu/deeptime).
In this paper, we provide a brief overview of angiosperm
phylogeny as currently understood (Fig. 1) and examine pat-
terns of angiosperm diversification. The monocot and eudicot
clades will be considered in greater detail in the accompanying
papers by Chase (2004) and Judd and Olmstead (2004), re-
spectively.
ANGIOSPERM PHYLOGENY
The root of the tree—A mere decade ago, the possibility
of identifying the basal nodes of the angiosperm clade seemed
remote. However, most analyses of the past five years concur
in placing the monotypic Amborella as the sister to all other
extant angiosperms. Amborella trichopoda, endemic to cloud
forests of New Caledonia, was described in the mid-nineteenth
century (Baillon, 1869) and has since been classified with var-
ious groups of basal angiosperms, most often with Laurales
(e.g., Cronquist, 1981). However, Amborella clearly differs
from most Laurales in having spirally arranged floral organs
(except perhaps the carpels; Buzgo et al., in press), rather than
the whorled phyllotaxis typical of most Laurales (see studies
of floral morphology and development by Endress and Iger-
sheim, 2000b; Posluszny and Tomlinson, 2003; Buzgo et al.,
in press), and lacks those features considered to be synapo-
morphies for Laurales (Doyle and Endress, 2000; see Laurales
later). Amborella has carpels that are closed only by secretion,
October 2004] 1615S
OLTIS AND
S
OLTIS
—D
IVERSIFICATION OF ANGIOSPERMS
Fig. 1. Overview of angiosperm phylogenetic relationships, based on Qiu
et al. (1999), P. Soltis et al. (1999), D. Soltis et al. (2000, 2003), Zanis et al.
(2002), Hilu et al. (2003), Kim et al. (2004b).
rather than by fused tissue as in most angiosperms (Endress
and Igersheim, 2000a)—a feature that may represent a ple-
siomorphy (i.e., ancestral feature) for the angiosperms. Vessels
(Judd et al., 2002; but see Feild et al., 2000; Doyle and En-
dress, 2000) and pollen grains with a reticulate tectum (Doyle
and Endress, 2001) appear to be synapomorphies for all extant
angiosperms except Amborella. Ethereal oil cells—common
throughout basal angiosperms—and columellate pollen grains
with a perforate tectum are synapomorphies for all extant an-
giosperms except Amborella and Nymphaeaceae (Doyle and
Endress, 2001).
The evidence for Amborella—Nearly all multigene analyses
of basal angiosperms have identified Amborella as the sister
to all other extant angiosperms (e.g., Mathews and Donoghue,
1999, 2000; Parkinson et al., 1999; Qiu et al., 1999; P. Soltis
et al., 1999; Graham and Olmstead, 2000; Graham et al., 2000;
D. Soltis et al., 2000; Magallo´n and Sanderson, 2001; Zanis
et al., 2002; see also Nickerson and Drouin, 2004), with vary-
ing levels of support. The genes that support this position
come from all three plant genomes and represent relatively
‘slowly evolving’ protein-coding and ribosomal RNA genes.
Furthermore, analyses of the ‘rapidly evolving’ plastid gene
matK (Hilu et al., 2003) and mostly noncoding trnL-trnF
(Borsch et al., 2003) each showed these same results. In all of
these studies, Nymphaeaceae and Austrobaileyales (both sensu
the Angiosperm Phylogeny Group, APG II, 2003) ‘followed’
Amborella as successive sisters to the remaining extant angio-
sperms. Furthermore, the structural organization of the floral
MADS-box genes Apetala3 and Pistillata also supported the
position of Amborella and Nymphaeaceae as sisters to all other
extant angiosperms, and analyses of nucleotide and amino acid
sequences of these genes also placed Amborella, either alone
or with Nymphaeaceae, in this position (Kim et al., in press).
Alternative views—Despite general support for the place-
ment of Amborella as sister to the rest of the extant angio-
sperms, a few studies have found alternative rootings, using
either different genes or different methods of analysis. For
example, Amborella 1 Nymphaeaceae (e.g., Parkinson et al.,
1999; Barkman et al., 2000; Mathews and Donoghue, 2000;
Qiu et al., 2000; P. Soltis et al., 2000; Kim et al., in press) or
Nymphaeaceae alone (e.g., Parkinson et al., 1999; Graham and
Olmstead, 2000, with partial sampling of Nymphaeaceae; Ma-
thews and Donoghue, 2000) have occasionally been reported
as sister to all other angiosperms. However, statistical analyses
of these alternative rootings using a data set of up to 11 genes
generally favor the tree with Amborella as sister to the rest,
although the Amborella 1 Nymphaeaceae tree could not al-
ways be rejected (Zanis et al., 2002). Nearly all of these stud-
ies are consistent in noting that conflicting topologies are not
strongly supported. Furthermore, the difference among these
three topologies is relatively minor and consists solely of the
relative placement of Amborella and Nymphaeaceae.
A more dramatic alternative, based on a selection of 61
genes from the totally sequenced plastid genomes of 13 plant
species, placed the monocots (represented only by three grass-
es—rice, maize, and wheat) as the sister to all other extant
angiosperms (Goremykin et al., 2003). Whereas all molecular
analyses of angiosperms with dense taxon sampling strongly
supported monophyly of the monocots and most placed this
clade among the basal nodes of the angiosperm tree, none has
indicated that monocots are sister to all other extant angio-
sperms. In the analysis by Goremykin et al. (2003), Amborella
was sister to Calycanthus of Laurales, a position consistent
with the original description of Amborella, but clearly at odds
with other aspects of morphology (see Laurales section). Go-
remykin et al. (2003) attributed their results to the increased
character sampling (30017 nucleotides in their aligned matrix)
in their study relative to other analyses that included fewer
genes but many more taxa. However, further analyses of a data
set of three genes and nearly equivalent taxon sampling indi-
cated that the ‘monocots basal’ topology is an artifact of lim-
ited taxon sampling (Soltis and Soltis, 2004). When either
Nymphaea or Austrobaileya, representing Nymphaeaceae and
Austrobaileyales, respectively, was substituted for Amborella,
each appeared as the sister to Calycanthus, in exactly the same
position that Amborella had occupied, presumably because the
data set, which was limited to a subset of those plant species
for which entire plastid genome sequences are available, con-
tained no other close relatives. Furthermore, representing
monocots by taxa other than grasses, which reside at the end
of a long branch (e.g., Gaut et al., 1992, 1996; Chase et al.,
2000), broke up the long branch to the monocots and resulted
in the Amborella basal’ topology. Likewise, broader sam-
pling of the monocots beyond grasses (the sole monocots in-
cluded by Goremykin et al., 2003) also severed the long mono-
cot branch and yielded the Amborella basal’ tree. Finally,
when plastid sequences of the monocot Acorus were added to
the data set of Goremykin et al., also disrupting the long branch
to the grasses, Amborella resumed its position as sister to the
other angiosperms (S. Stefanovic et al., Indiana University,
1616 [Vol. 91A
MERICAN
J
OURNAL OF
B
OTANY
Fig. 2. Summary of phylogenetic relationships among clades of basal an-
giosperms, based primarily on Zanis et al. (2002).
unpublished data). Although increasing the number of char-
acters will generally lead to greater accuracy (Hillis, 1996) and
support (e.g., Givnish and Sytsma, 1997; Soltis et al., 1998),
the increase in characters cannot come at the expense of ad-
equate taxon sampling (e.g., Chase et al., 1993; Sytsma and
Baum, 1996; Zwickl and Hillis, 2002; Pollock et al., 2002;
Soltis et al., in press-a). Limited taxon sampling, such as that
dictated by the small number of organisms with complete ge-
nome sequences, may lead to artifacts, as apparently occurred
in the analysis by Goremykin et al. (2003).
The fossil record—The fossil record does not clarify basal
groups within the angiosperms. However, it clearly identifies
a number of morphologically diverse lineages early in angio-
sperm evolution (e.g., Crane et al., 1995; Friis et al., 2000).
Although some of these early fossils seem to belong to extant
families, many do not fit easily into extant groups. For ex-
ample, two species of Archaefructus (Sun et al., 1998, 2002)
may be the sister to all other angiosperms (Sun et al., 2002),
although a reanalysis of their data, with the inclusion of ad-
ditional material, indicated alternative placements (Friis et al.,
2003). Notably, with regard to the plastid topology of Gore-
mykin et al. (2003), the monocots are not among the earliest
angiosperm fossils, although both the fossil record (Gandolfo
et al., 2002) and molecular clock estimates (K. Bremer, 2000;
Wikstro¨m et al., 2001; Davies et al., 2004) have indicated that
many lineages of monocots date back at least 80–100 mya.
However, Nymphaeaceae are among the earliest angiosperm
fossils: a water lily from approximately 125 mya (Friis et al.,
2001) is consistent with the basal or near-basal position of the
Nymphaeaceae branch in most molecular-based trees, and the
floral features of Microvictoria (90 mya; Gandolfo et al., 2004)
provide evidence of beetle entrapment pollination in early an-
giosperms. Likewise, the abundance of fossils of Chlorantha-
ceae and Ceratophyllaceae from the early Cretaceous (e.g.,
Couper, 1958; Walker and Walker, 1984; Dilcher, 1989; Friis
et al., 2000; see Endress, 2001, for review) is also consistent
with the placement of these clades among extant angiosperms
in molecular-based trees (Figs. 1, 2).
Basal lineages—The positions of Amborellaceae and Nym-
phaeaceae as successive sisters to the rest of the angiosperms
are followed in turn by Austrobaileyales. Although these first
three nodes are well supported (e.g., Zanis et al., 2002; Hilu
et al., 2003), resolution and support for relationships of the
next few nodes are poor (Fig. 2). Ceratophyllaceae, monocots,
Chloranthaceae, magnoliids, and eudicots are each well sup-
ported, and both the fossil record and molecular-based trees
identify these lineages as ancient. However, their interrelation-
ships remain unclear. It is clear, however, that angiosperms do
not fall into two major groups that correspond to monocots
(Liliopsida) and dicots (Magnoliopsida) of longstanding clas-
sification systems (such as Cronquist, 1981; Takhtajan, 1997,
and their predecessors). Although monocots clearly form a
strongly supported clade, dicots in the traditional sense do not:
most are found in the eudicot clade, but the remaining non-
monocot basal branches (i.e., Amborellaceae, Nymphaeaceae,
Austrobaileyales, Ceratophyllaceae, Chloranthaceae, magno-
liids) were also ‘traditional’’ dicots. The nonmonophyly of the
dicots has long been suspected, and the lack of monophyly
precludes their recognition in current classifications (e.g., APG
II, 2003). The concept of ‘dicot’ should be abandoned in
favor of eudicots, with recognition that considerable diversity
exists outside the monocot and eudicot clades.
Nymphaeaceae—The phylogenetic position of Nymphae-
aceae as one of the two basalmost lineages of extant angio-
sperms is strongly supported by nearly all molecular analyses.
This clade of eight genera has a worldwide distribution, con-
sistent with both the ancient age of this lineage and aquatic
habitats. Although all genera occupy aquatic habitats, these
habitats range from temperate to tropical. Floral diversity
among genera is extensive, ranging from the small, simple,
trimerous, monocot-like flowers of Cabomba to the large,
showy, elaborate flowers of Nymphaea and Victoria. Although
the latter were considered ‘primitive’ by most authors,
Schneider (1979) suggested that the numerous floral organs of
Nymphaea and Victoria resulted from secondary increase.
Phylogenetic analyses (Les et al., 1999) and character recon-
structions (Ronse DeCraene et al., 2003; Soltis et al., in press-
b) supported Schneider’s (1979) hypothesis. Floral diversifi-
cation in Nymphaeacae may be related to changes in pollina-
tion: proliferation of parts in response to beetle pollination in
Nymphaea and Victoria and a reduction in number of parts
associated with a shift to cleistogamy in Euryale (Gottsberger,
1977, 1978; Williams and Schneider, 1993; Lipok et al., 2000).
Austrobaileyales—This small clade comprises Austrobail-
eyaceae (Austrobaileya) and Trimeniaceae (Trimenia) from
Australasia plus Schisandraceae sensu APG II (2003), i.e.,
Schisandraceae (Schisandra and Kadsura) and Illiciaceae (Il-
licium) of most other recent classifications (Qiu et al., 1999;
Renner, 1999; Savolainen et al., 2000a, b; P. Soltis et al., 1999;
D. Soltis et al., 2000). Although the traditional Illiciaceae and
Schisandraceae have typically been united in Illiciales, a re-
lationship between these taxa and Austrobaileya and Trimenia
had not been suspected. No morphological synapomorphies
have been identified for this clade, despite the strong molec-
ular support for its monophyly.
Ceratophyllaceae—Ceratophyllaceae (Ceratophyllum) had
the distinction of appearing as the sister to all other angio-
October 2004] 1617S
OLTIS AND
S
OLTIS
—D
IVERSIFICATION OF ANGIOSPERMS
sperms in the first large molecular phylogenetic analysis based
on rbcL (Chase et al., 1993). The aquatic habit and simple
flowers seemed at odds with most hypotheses about the earliest
angiosperms, although Ceratophyllum has a long fossil record,
going back at least 125 mya (Dilcher, 1989). Subsequent anal-
yses demonstrated that this placement was unique to the rbcL
data set, but the position of Ceratophyllum, based on evidence
from many other genes, is still not clear. It appears as the sister
to monocots in some analyses (e.g., Zanis et al., 2002; Davies
et al., 2004), but further work is needed to identify its proper
position.
Monocots—Among extant angiosperms, monocotyledons
represent the earliest-appearing major clade. Using a molecular
clock, K. Bremer (2000) dated the origin of the monocot clade
to be 134 mya, older than the oldest angiosperm fossils. Al-
though their exact age is unclear from the fossil record (but
see Gandolfo et al., 2002), monocots clearly represent an early
lineage of angiosperms. There are approximately 52000 spe-
cies of monocots (Mabberley, 1993), representing 22% of all
angiosperms. Half of the monocots can be found in the two
largest families, Orchidaceae and Poaceae, which comprise
34% and 17%, respectively, of all monocots.
Phylogenetic studies of nonmolecular data (Donoghue and
Doyle, 1989; Loconte and Stevenson, 1991; Doyle and Don-
oghue, 1992) have identified 13 putative synapomorphies for
the monocots, including, among others, a single cotyledon,
parallel-veined leaves, sieve cell plastids with several cuneate
protein crystals, scattered vascular bundles in the stem, and an
adventitious root system. An often-overlooked synapomorphy
for monocots is their sympodial growth; although there are
other angiosperms with sympodial growth, monocots are near-
ly exclusively so. These synapomorphies are covered in detail
in the paper by Chase in this issue (2004; see also Judd et al.,
2002; Soltis et al., in press-b).
Recognition of the monocots as a distinct group within the
angiosperms dates from Ray (1703) and was largely based on
their possession of a single cotyledon relative to the two cot-
yledons typical of the dicotyledons or ‘dicots.’ As reviewed
earlier, the latter group is now known to be nonmonophyletic,
and the term ‘dicot’’ should be abandoned. There is, however,
a great diversity of form in monocot seedlings (Tillich, 1995)
and not all possess an obvious single cotyledon.
Another major, distinctive trait of the monocots is their vas-
cular system, which is characterized by vascular bundles that
are scattered throughout the medulla and cortex and are closed
(i.e., do not contain an active cambium; reviewed in Tomlin-
son, 1995). In contrast, basal angiosperms formerly considered
dicots (e.g., members of the magnoliid clade) and eudicots
possess open vascular bundles arranged in a ring.
Another widely cited character of the monocots is their par-
ticular form of sieve cell plastids (Behnke, 1969), which are
triangular with cuneate proteinaceous inclusions. Similar sieve
cell plastids are found in Aristolochiaceae (Dahlgren et al.,
1985). This similarity between monocots and Aristolochiaceae
apparently represents convergence, not shared ancestry, be-
cause phylogenetic studies of DNA sequences from all three
genomes (Qiu et al., 1999; Zanis et al., 2002, 2003) have dem-
onstrated a strongly supported relationship of Aristolochiaceae
to other Piperales within the magnoliid clade.
Other traits characteristic of the monocots include parallel
venation without free vein-endings (vs. reticulate venation
with free vein-endings), intercalary meristem, adventitious
roots, and roots without secondary growth. Adventitious roots
are found elsewhere in the angiosperms, in both Piperaceae
and Nymphaeaceae.
Trimerous flowers have long been considered a uniting fea-
ture of the monocots, but it is not an exclusive one because
there are many other basal angiosperms, including Nymphae-
aceae and magnoliids, that also exhibit trimery. In fact, char-
acter-state reconstructions of the angiosperms indicate that tri-
mery arose early in the angiosperms; it may be ancestral for
all angiosperms except Amborella (Ronse De Craene et al.,
2003; Zanis et al., 2003; Soltis et al., in press-b), or perhaps
all angiosperms, if the shift away from trimery in Amborella
occurred along the lineage leading to Amborella. Trimery ap-
pears, therefore, to be a symplesiomorphic feature for mono-
cots and other angiosperms and is not a ‘monocot character.’
Our understanding of monocot phylogenetics has greatly
improved over the past decade, aided greatly by the foci pro-
vided by the international monocot symposia held in 1993,
1998, and 2003. These meetings have focused attention both
on what was known and, more importantly, on which groups
needed additional research. As a result, we now know more
about monocots than any other group of angiosperms of com-
parable size, a situation that is remarkable given the paucity
of information available in 1985 (Dahlgren et al., 1985). This
model should be adopted for the other large groups of angio-
sperms (e.g., rosids, asterids) so that attention is likewise fo-
cused on integration of research programs and gaps in the
database.
There have been several recent analyses of relationships
among the monocots, including the three-gene analyses of
Chase et al. (2000) and D. Soltis et al. (2000) and the seven-
gene analysis of Chase et al. (in press). The first two studies
are based on the same three genes (rbcL, atpB, 18S rDNA);
however, Chase et al. (2000) focused only on the monocots
and employed a larger number of taxa than used in D. Soltis
et al. (2000). The analysis by Chase et al. (2004) included
those three genes, plus partial nuclear 26S rDNA, plastid matK
and ndhF, and mitochondrial atpA. The paper in this issue by
Chase (2004) provides greater detail on monocot phylogeny,
and our coverage will therefore be brief.
All but two molecular phylogenetic analyses of monocots
have placed Acorus alone as sister to all other monocots. The
first exception to this statement was the 18S rDNA analysis
of Bharathan and Zimmer (1995), in which Acorus was placed
outside of the monocots altogether, a result that has to be con-
sidered spurious. Combination of 18S rDNA sequence data
with sequences from rbcL and atpB (Chase et al., 2000; D.
Soltis et al., 2000) resulted in strong support for the mono-
phyly of monocots, as well as strong support for the mono-
phyly of all monocots excluding Acorus. A recent analysis of
two of the seven genes used in Chase et al. (in press), rbcL
and atpA (Davis et al., in press), retrieved an alismatid clade
that included Acorus. This deviating result is perplexing be-
cause neither rbcL (Chase et al., 1993; Duvall et al., 1993)
nor atpA (Davis et al., 1998) analyzed alone produced such a
position for Acorus. In contrast, studies of basal angiosperm
relationships that have employed more genes (six to 11) have
consistently found Acorus sister to the remaining monocots
with strong support (e.g., Qiu et al., 1999, 2000; Zanis et al.,
2002, 2003). A recent angiosperm-wide analysis of matK se-
quence data (Hilu et al., 2003), an analysis of ndhF in mono-
cots (Givnish et al., in press), and a seven-gene analysis of
monocots (Chase et al., in press) found moderate to strong
1618 [Vol. 91A
MERICAN
J
OURNAL OF
B
OTANY
support for the placement of Acorus as sister to other mono-
cots. Hence, most analyses agree on the placement of Acorus
as sister to all other monocots.
Following Acorus, the monophyly of the remaining mono-
cots is strongly supported. Alismatales are sister to the re-
maining monocots, which themselves are strongly supported.
Within this remaining large clade are several component sub-
clades: commelinids, Dioscoreales, Petrosaviaceae, Pandana-
les, Liliales, and Asparagales. Although many of these com-
ponent subclades receive moderate to strong support, relation-
ships among these subclades have been generally poorly re-
solved. In the strict consensus of Chase et al. (2000), the
branching order above Alismatales is Dioscoreales, Pandana-
les, Liliales, and Asparagales 1 commelinids. The seven-gene
analysis of Chase et al. (in press) is consistent with this pat-
tern, except that Dioscoreales and Pandanales are sister taxa,
and most of these relationships received at least moderate
bootstrap support.
Chloranthaceae—Chloranthaceae, with their small, simple
flowers, have an extensive fossil record, dating back 125 my
(e.g., Couper, 1958; Walker and Walker, 1984; Friis et al.,
2000). Chloranthaceae are clearly an isolated lineage separate
from the magnoliid clade (Fig. 2), but their phylogenetic po-
sition remains uncertain. In some analyses (e.g., Zanis et al.,
2002; Davies et al., 2004), they are sister to a clade of mag-
noliids 1 eudicots. Relationships and patterns of evolution
within Chloranthaceae have been addressed by Kong et al.
(2002), Doyle et al. (2003), Zhang and Renner (2003), and
Eklund et al. (2004).
Magnoliids—The magnoliid clade comprises most of those
lineages typically referred to as ‘primitive angiosperms’ in
earlier works (e.g., Stebbins, 1974; Cronquist, 1981, 1988;
Takhtajan, 1997). Although the component families of the
magnoliid clade were loosely associated in previous classifi-
cations, for example, as Cronquist’s (1981) subclass Magno-
liidae, relationships among the families and orders were not
clear. In addition, Magnoliidae contained groups that are not
part of the magnoliid clade as recognized by phylogenetic
analyses. Reconstructing relationships within this clade, and
even recognition of the clade itself, is challenging, given the
age of this clade (some putative members, such as Archaean-
thus, Dilcher and Crane, 1984, date to the early Cretaceous)
and presumably high levels of extinction. Although the major
lineages of the magnoliid clade were identified as well-sup-
ported clades in earlier studies (e.g., Soltis et al., 1999), com-
position and interrelationships of the magnoliid clade did not
become clear until data sets of at least five genes for a broad
sample of taxa were assembled to address these problems (e.g.,
Qiu et al., 1999, 2000; Zanis et al., 2002). Within the mag-
noliids, Magnoliales and Laurales are sisters, and Piperales and
Canellales are sisters (Fig. 2).
Magnoliales. This clade comprises six families (Myristica-
ceae, Degeneriaceae, Himantandraceae, Magnoliaceae, Eu-
pomatiaceae, and Annonaceae), relationships among which are
now clear (e.g., Sauquet et al., 2003; Fig. 2). This same clade
emerged in the nonmolecular analysis of Doyle and Endress
(2000). Apparent synapomorphies for the clade include re-
duced fiber pit borders, stratified phloem, an adaxial plate of
vascular tissue in the petiole, palisade parenchyma, astero-
sclereids in the leaf mesophyll, continuous tectum in the pol-
len, and multiplicative testa in the seed (Doyle and Endress,
2000). Furthermore, all members of this clade examined to
date have a characteristic deletion in their Apetala3 gene (Kim
et al., in press).
Laurales. Laurales, as currently circumscribed (APG II,
2003; see Renner, 1999), comprise seven families: Calycan-
thaceae (including Idiospermaceae), Monimiaceae, Gomorte-
gaceae, Atherospermataceae, Lauraceae, Sipurunaceae, and
Hernandiaceae. Amborellaceae and Trimeniaceae have also
occasionally been placed in Laurales (e.g., Cronquist, 1981,
1988); in fact, both Amborella and Trimenia have even been
considered part of Monimiaceae (Perkins, 1925). Chlorantha-
ceae have also occasionally been placed in Laurales (e.g.,
Thorne, 1974; Takhtajan, 1987, 1997). Laurales are united by
a perigynous flower in which the gynoecium is frequently
deeply embedded in a fleshy receptacle (Endress and Iger-
sheim, 1997; Renner, 1999). Other apparent synapomorphies
include the presence of inner staminodia, ascendant ovules,
and tracheidal endotesta (Doyle and Endress, 2000).
Piperales. Previous circumscriptions of Piperales have var-
ied (e.g., Dahlgren, 1980; Cronquist, 1981, 1988; Takhtajan,
1987, 1997; Thorne, 1992; Heywood, 1993), but molecular
studies clearly united Aristolochiaceae, Lactoridaceae, Piper-
aceae, and Saururaceae (e.g., Qiu et al., 1999; Soltis et al.,
1999; Barkman et al., 2000; D. Soltis et al., 2000; Zanis et
al., 2002). In addition, recent studies have placed Hydnora-
ceae, a family of parasitic plants often placed in Rosidae (e.g.,
Cronquist, 1981; Heywood, 1993), within Piperales, although
the exact position is not certain (Nickrent et al., 2002). Al-
though not recognized as a group prior to molecular analyses,
a number of morphological synapomorphies have been iden-
tified: distichous phyllotaxis, a single prophyll, and oil cells
(Doyle and Endress, 2000).
Canellales. The sister group of Canellaceae and Winteraceae
has been strongly supported in all multigene analyses (e.g.,
Qiu et al., 1999; Soltis et al., 1999; D. Soltis et al., 2000;
Zanis et al., 2002, 2003), and the clade was obtained in Doyle
and Endress’s (2000) nonmolecular analysis as well. However,
these two families have not typically been considered closely
related to each other, and neither was suspected of being re-
lated to any members of Piperales. For example, Winteraceae
have often been considered a close relative of Magnoliaceae
(e.g., Cronquist, 1981, 1988; Heywood, 1993), with Canella-
ceae close to Myristicaceae (e.g., Wilson, 1966; Cronquist,
1981, 1988). Furthermore, Winteraceae have often been re-
garded as perhaps the ‘most primitive’ extant family of an-
giosperms (Cronquist, 1981; Endress, 1986). The phylogenetic
position of Winteraceae clearly indicates that the vesselless
xylem and plicate carpels found in members of the family are
secondarily derived (see also Young, 1981). Possible synapo-
morphies for Canellales are a well-differentiated pollen tube
transmitting tissue, an outer integument with only two to four
cell layers, and seeds with a palisade exotesta (Doyle and En-
dress, 2000). Additional synapomorphies may include an ir-
regular ‘first-rank’ leaf venation (Hickey and Wolf, 1975;
Doyle and Endress, 2000), stelar and nodal structure (Keating,
2000), and vascularization of the seeds (Deroin, 2000).
Eudicots—Eudicots comprise approximately 75% of all an-
giosperm species (Drinnan et al., 1994) and are strongly sup-
October 2004] 1619S
OLTIS AND
S
OLTIS
—D
IVERSIFICATION OF ANGIOSPERMS
Fig. 3. Summary of phylogenetic relationships among clades of eudicots,
based on Hoot et al. (1999), D. Soltis et al. (2000, 2003), and Kim et al.
(2004).
ported by molecular data. However, only a single morpholog-
ical synapomorphy—triaperturate pollen—has been identified.
This pollen type is clearly distinct from the uniaperturate pol-
len of basal angiosperms, monocots, and all other seed plants,
allowing easy assignment of fossil pollen to the eudicots. The
fossil pollen record indicates that the eudicots appeared 125
mya, shortly after the origin of the angiosperms themselves.
The extensive fossil pollen collections worldwide, coupled
with solid dates, make it unlikely that the eudicots arose much
before this time. Although triaperturate pollen is a synapo-
morphy for this clade, not all eudicots have triaperturate pollen
due to subsequent changes in pollen structure. The eudicots
(referred to instead as tricolpates) are covered in greater detail
by Judd and Olmstead (2004).
Basal lineages—A basal grade of five lineages (Ranuncu-
lales, Proteales, Sabiaceae, Trochodendraceae, and Buxaceae)
subtends the large clade of core eudicots (Hoot et al., 1999;
D. Soltis et al., 2000; Kim et al., 2004; Fig. 3). Although
Ranunculales are supported as the sister to all other eudicots,
the relative placements of the remaining four lineages of basal
eudicots are not clear and require additional study.
Core eudicots—The core eudicots comprise the vast major-
ity of eudicot species. Seven major clades (Gunnerales, ‘Ber-
beridopsidales,’ Saxifragales, Santalales, Caryophyllales, ros-
ids, and asterids) have been recognized, but the relationships
among these clades are not clear (Figs. 1, 3; D. Soltis et al.,
2000). The topology indicates a rapid radiation, but additional
data are needed to evaluate this hypothesis. Recent studies
have identified Gunnerales as the sister to all other core eu-
dicots (Hilu et al., 2003; Soltis et al., 2003). Several important
changes in floral genes appear to coincide with the origin of
core eudicots, including duplication of AP3 yielding the
euAP3 lineage (Kramer et al., 1998) and the origin of Apetala1
(Litt and Irish, 2003).
Gunnerales. Gunnerales comprise two small families, Gun-
neraceae (Gunnera with approximately 40 species) and My-
rothamnaceae (Myrothamnus with two species) (or Gunnera-
ceae s.l. sensu APG II, 2003). This relationship had not pre-
viously been suggested on the basis of morphology because
the two genera differ substantially, although molecular support
for their relationship is very strong. Gunneraceae have a dim-
erous perianth (Drinnan et al., 1994), as do many of the basal
eudicot lineages; dimery probably typifies Buxaceae, Trochod-
endraceae, and Proteaceae (but perhaps not the Platanus lin-
eage) and is common and perhaps ancestral in Ranunculales
(van Tieghem, 1897; Drinnan et al., 1994; Douglas and Tucker,
1996). The placement of Gunnerales as sister to the rest of the
core eudicots implies that the pentamerous perianth typical of
most core eudicots was derived from dimerous ancestors
(Ronse De Craene et al., 2003; Soltis et al., 2003).
‘Berberidopsidales’’. Like Gunnerales, ‘Berberidopsida-
les’ comprise two small and morphologically disparate fami-
lies: Berberidopsidaceae (Berberidopsis and Streptothamnus,
which is sometimes included in Berberidopsis) and Aextoxi-
caceae (Aextoxicon, one species). Although this clade has not
been recognized at the ordinal level by APG (hence the quo-
tation marks), it is strongly supported by molecular data and
is isolated from all other clades. Furthermore, both families
have encyclocytic stomata, a rare character and an apparent
synapomorphy for this clade (Soltis et al., in press-b).
Saxifragales. Saxifragales are a morphologically eclectic
clade of annual and perennial herbs, succulents, aquatics,
shrubs, vines, and large trees. Prior to molecular phylogenetics
(Morgan and Soltis, 1993; Fishbein et al., 2001), members of
this clade were classified in three of Cronquist’s (1981) six
subclasses of dicots (see also Takhtajan, 1997). Possible syn-
apomorphies for this clade include a partially fused bicarpel-
late gynoecium, a hypanthium, and glandular leaf teeth (Judd
et al., 2002); aspects of leaf venation and wood anatomy are
similar in the woody members of the clade. The best known
of the 13 families in this clade are Saxifragaceae, Crassula-
ceae, Grossulariaceae, Paeoniaceae, and Hamamelidaceae.
Molecular studies continue to reveal new, unexpected mem-
bers of this clade, such as Peridiscaceae (Davis and Chase,
2004), a family placed in Malpighiales in APG II (2003).
Monophyly of Saxifragales is strongly supported, but the
position of this clade relative to other core eudicots remains
uncertain. Some analyses have placed it as sister to the rosids,
although with weak support (e.g., D. Soltis et al., 2000). The
simple, pentamerous flowers have long been thought to indi-
cate a relationship with Rosaceae and other rosids, but whether
these floral features are synapomorphies for Saxifragales 1
rosids or symplesiomorphies (i.e., shared ancestral features) is
unclear. Despite the fairly constant general floral structure of
Saxifragales, certain aspects of floral evolution within this
clade appear to be quite labile, especially ovary position (e.g.,
Kuzoff et al., 2001). Additional research is needed to resolve
the relationship of Saxifragales within the core eudicots.
Santalales. The seven families of Santalales are united by
molecular characters and aspects of their parasitic habit and
are a strongly supported clade of core eudicots. However, re-
lationships of Santalales to other core eudicots are not clear,
although they occasionally appear near the asterids in at least
some shortest trees. Furthermore, relationships within this
1620 [Vol. 91A
MERICAN
J
OURNAL OF
B
OTANY
clade have not yet been resolved, and the monophyly of some
of the currently recognized families has not been supported by
molecular evidence. The lack of resolution within Santalales
may be explained in part by apparently rapid rates of molec-
ular evolution in all three plant genomes (e.g., Nickrent and
Starr, 1994; Nickrent et al., 1998). Aerial hemiparasites (mis-
tletoes) have evolved multiple times in Santalales.
Caryophyllales. The core of Caryophyllales sensu APG II
(2003) was considered a closely related group of families as
long ago as the mid-nineteenth century (e.g., Braun, 1864;
Eichler, 1876) and was formally recognized as the Centro-
spermae by Harms (1934) based on morphological and em-
bryological characters. Recent molecular studies have identi-
fied a larger clade (Caryophyllales sensu APG II) that includes
the Caryophyllidae of Cronquist (1981; i.e., Caryophyllales,
Polygonales, and Plumbaginales) plus a number of families
previously considered distantly related to Caryophyllales, in-
cluding the carnivorous sundews and Venus’ flytrap (Droser-
aceae) and Old World pitcher plants (Nepenthaceae).
Relationships of Caryophyllales to other core eudicots are
not clear, although Dilleniaceae are sister to Caryophyllales in
some analyses, although with low support (e.g., Chase and
Albert, 1998; D. Soltis et al., 2000; Fig. 3), and some shortest
trees have indicated a possible relationship with the asterids.
Within Caryophyllales, there are two large clades, core and
noncore Caryophyllales (Cue´noud et al., 2002), that corre-
spond to Caryophyllales and Polygonales sensu Judd et al.
(2002). The core Caryophyllales clade generally corresponds
to Caryophyllales of recent classifications (e.g., Cronquist,
1981; Takhtajan, 1997) and comprises 19 families, although
some currently recognized families (e.g., Portulacaceae, Phy-
tolaccaceae) are poly- or paraphyletic and require recircum-
scription (Cue´noud et al., 2002). Synapomorphies for this
clade include unilacunar nodes, stems with concentric rings of
xylem and phloem, phloem sieve tubes with plastids with a
peripheral ring of proteinaceous filaments and a central protein
crystal, betalains (rather than anthocyanins), loss of the intron
in the plastid gene rpl2, a single perianth whorl, free central
to basal placentation, an embryo curved around the seed, and
presence of perisperm with little or no endosperm (Judd et al.,
2002 and references therein). The noncore clade has been
identified on the basis of molecular data and comprises fami-
lies classified in Cronquist’s (1981) Rosidae and Dilleniidae.
Most surprising is the inclusion in this clade of the carnivorous
Droseraceae and Nepenthaceae. Synapomorphies for the non-
core clade are scattered secretory cells containing plumbagin,
an indumentum of stalked, gland-headed hairs, basal placen-
tation, and starchy endosperm (Judd et al., 2002).
Many Caryophyllales are adapted to harsh environments,
such as high-alkaline soils, high-salt conditions, extreme arid-
ity, and nutrient-poor soils (see descriptions of component
families in Heywood, 1993; Judd et al., 2002). Various adap-
tations, such as Crassulacean acid metabolism and C
4
photo-
synthesis, succulence, carnivory, and salt secretion, have
evolved multiple times (e.g., Juniper et al., 1989; Meimberg
et al., 2000; Pyankov et al., 2001; Cameron et al., 2002) and
have allowed Caryophyllales to exploit these habitats.
Rosids. The rosid clade is broader than the traditional sub-
class Rosidae (Cronquist, 1981; Takhtajan, 1980, 1997), also
encompassing many families formerly classified in the poly-
phyletic subclasses Magnoliidae, Dilleniidae, and Hamameli-
dae. The rosids comprise 140 families and close to one-third
of all angiosperm species. Clear synapomorphies for the rosids
have not been identified, although most rosids share several
morphological and anatomical features, such as nuclear en-
dosperm development, reticulate pollen exine, generally sim-
ple perforations of vessel end-walls, alternate intervessel pit-
ting, mucilaginous leaf epidermis, and two or more whorls of
stamens, plus ellagic acid (Hufford, 1992; Nandi et al., 1998).
Relationships within rosids are not clearly resolved. Vita-
ceae may be sister to the rosids, but this relationship is not
strongly supported (Fig. 3; Savolainen et al., 2000a, b; D. Sol-
tis et al., 2000), and Saxifragales may be sister to the Vitaceae
1 rosids clade, but this relationship is not strongly supported
either (D. Soltis et al., 2000). Two large subclades of rosids,
eurosids I (fabids) and II (malvids), have been identified
through molecular analyses (e.g., Chase et al., 1993; Savolai-
nen et al., 2000a, b; D. Soltis et al., 2000; Fig. 1). However,
some orders and families (e.g., Crossosomatales, Geraniales,
Myrtales) do not fit into either eurosid I or eurosid II. The
eurosid I clade comprises Celastrales, Cucurbitales, Fabales,
Fagales, Zygophyllales, Malpighiales, Oxalidales, and Ro-
sales. Of these, Cucurbitales, Fabales, Fagales, and Rosales
form the ‘nitrogen-fixing clade,’ a clade that contains all an-
giosperms known to have symbiotic relationships with nodu-
lating nitrogen-fixing bacteria (see D. Soltis et al., 1995, 1997,
2000). The placements in previous classifications of the spe-
cies that exhibit this symbiosis indicated that symbiotic rela-
tionships with nodulating bacteria must have occurred multiple
times. Current phylogenetic evidence instead indicates a single
origin of the predisposition for symbiosis, with perhaps both
gains and losses of the symbiotic relationship within the nitro-
gen-fixing clade itself (Soltis et al., 1995; Swensen, 1996).
Multiple gains of this association may be more parsimonious
than a single gain followed by multiple losses (Swensen,
1996). The smaller eurosid II clade is composed of Brassi-
cales, Malvales, Sapindales, and Tapisciaceae. Brassicales in-
clude all angiosperms known to produce glucosinolates, a
form of chemical defense, except Drypetes and Putranjiva of
the distantly related rosid family Putranjivaceae of Malpighi-
ales (e.g., Rodman, 1991; Rodman et al., 1993, 1998). Previ-
ous classifications led to the conclusion that glucosinolate pro-
duction had evolved several times in the angiosperms; current
phylogenetic evidence indicates instead only two such origins.
In addition to the large eurosid I and II clades, additional
smaller clades have been recognized (Crossosomatales, Myr-
tales, Geraniales, and Picramniaceae), but their relationships
to each other and to eurosids I and II are not clear. Further-
more, relationships within eurosids I and II are not fully re-
solved, and much additional work is needed to reconstruct
relationships within the rosid clade. In fact, the rosids repre-
sent the largest remaining problematic group of angiosperms.
Several factors may have contributed to the lack of resolu-
tion of relationships within the rosids. The clade is old, dating
at least to the late Santonian to Turonian (approximately 84
89.5 mya; Crepet and Nixon, 1998; Magallo´n et al., 1999),
and possibly to 94 mya, based on an unnamed apparently rosid
flower from the Dakota Formation in Nebraska (Basinger and
Dilcher, 1984). Furthermore, molecular-based age estimates of
Myrtales using penalized likelihood (Sanderson, 2002) placed
the crown radiation of Myrtales at approximately 110 mya
(Sytsma et al., in press), implying an even older age for the
rosids. The age of the rosid clade is therefore sufficient to have
allowed substantial morphological and molecular diversifica-
October 2004] 1621S
OLTIS AND
S
OLTIS
—D
IVERSIFICATION OF ANGIOSPERMS
tion and speciation, although the similar age of the monocots
has not similarly obscured relationships within that clade. The
rosid clade may have diversified via a series of radiations (P.
Soltis et al., 2004), resulting in a pattern of polytomies (see
Remaining Problems and Future Prospects). Furthermore, sub-
tle nonmolecular features that could potentially unite large
groups of families within the rosids have not generally been
identified because, until the results of molecular analyses,
many families of rosids were not suspected of being closely
related, having been placed in four subclasses of dicots (Mag-
noliidae, Hamamelidae, Dilleniidae, and Rosidae), and were
therefore not included together in most previous analyses and
treatments. Gaps in morphological data sets across the rosids
have likewise made it difficult to identify synapomorphies for
groups of families.
Asterids. Like rosids, asterids are a large clade, encompass-
ing nearly one-third of all angiosperm species (80000 species)
and classified in 114 families (Albach et al., 2001b). However,
unlike the rosids, a group of families corresponding closely to
the asterids has been recognized on morphological grounds for
over 200 years (de Jussieu, 1789; Reichenbach, 1828; Warm-
ing, 1879), and several morphological and chemical features
appear to unite all or most asterids. Most notable are iridoid
chemical compounds (e.g., Jensen, 1992), sympetalous corol-
las, unitegmic and tenuinucellate ovules, and cellular endo-
sperm development; however, it is still unclear which of these
features are actually synapomorphies for asterids (cf. Albach
et al., 2001a; Judd et al., 2002). The asterid clade is broader
than Asteridae of recent classifications (e.g., Cronquist, 1981;
Takhtajan, 1980, 1997) and includes also members of the poly-
phyletic subclasses Hamamelidae, Dilleniidae, and Rosidae
(Olmstead et al., 1992, 1993, 2000; Chase et al., 1993; D.
Soltis et al., 1997, 2000; Soltis et al., 1999; Savolainen et al.,
2000a, b).
Many relationships within asterids were resolved by angio-
sperm-wide analyses, but asterids have also been analyzed in
greater detail with extensive taxon sampling and data from
four (Albach et al., 2001b) and six (Bremer et al., 2002) loci.
These studies confirmed earlier results of four major clades
within asterids (Fig. 1): Cornales are sister to all other asterids,
with Ericales sister to a clade of euasterids I 1 euasterids II.
The families of Cornales and Ericales were not considered
closely related to those of Asteridae in previous classifications
and were placed instead mostly in Rosidae and Dilleniidae,
respectively. Euasterids are mostly united by flowers with epi-
petalous stamens that equal the number of corolla lobes and a
gynoecium of two fused carpels.
Within the euasterids, the euasterid I (or lamiid, Bremer et
al., 2002) and euasterid II (or campanulid, Bremer et al., 2002)
clades are sisters and can be distinguished both morphologi-
cally and molecularly (Albach et al., 2001b; Bremer et al.,
2001; Bremer et al., 2002). Most members of euasterids I have
opposite leaves, entire leaf margins, hypogynous flowers,
‘early sympetaly’ with a ring-shaped primordium, fusion of
stamen filaments to the corolla tube, and capsular fruits (Bre-
mer et al., 2001). The euasterid I clade comprises Garryales,
Gentianales, Solanales, and Lamiales, plus Boraginaceae, Vah-
liaceae, and Oncothecaceae 1 Icacinaceae (APG II, 2003).
Most taxa of euasterids II have alternate leaves, serrate-dentate
leaf margins, epigynous flowers, ‘late sympetaly’ with dis-
tinct petal primordia, free stamen filaments, and indehiscent
fruits (Bremer et al., 2001). It is unclear which of the char-
acters that distinguish euasterids I and II are truly synapo-
morphies for these clades and which are symplesiomorphies;
both reversals and parallelisms have contributed to complex
patterns of morphological evolution in the asterids (Albach et
al., 2001a; Bremer et al., 2001). The euasterid II clade is com-
posed of Dipsacales, Aquifoliales, Apiales, and Asterales, plus
Bruniaceae 1 Columelliaceae, a small clade of Tribelaceae,
Polyosmaceae, Escalloniaceae, and Eremosynaceae, and pos-
sibly Paracryphiaceae. The euasterid II clade includes families
previously classified in Asteridae and Rosidae (Cronquist,
1981, 1988).
The supertree approach—The relationships described in
this paper are all based on analyses that use the ‘‘supermatrix’
approach, that is, a taxon-by-character data matrix is assem-
bled and analyzed, directly producing a tree or set of trees. A
problem with this approach is that comprehensive data sets
become extremely large, and analyses become increasingly
computationally complex and time-consuming. In addition,
different gene sets have not always sampled the same taxa,
requiring assumptions of generic or familial monophyly in the
formation of ‘mosaic’ taxa and/or leading to large amounts
of missing data. An alternative to the supermatrix approach is
the supertree approach (e.g., Baum, 1992; Ragan, 1992; Purv-
is, 1995; Ronquist, 1996; Bininda-Emonds and Bryant, 1998;
Sanderson et al., 1998), in which trees that overlap in at least
a single taxon may be joined together algorithmically. Al-
though less satisfying than the supermatrix approach in relat-
ing support or conflict for a topology to specific characters,
the supertree approach is a viable alternative when multiple
data sets overlap in only a small fraction of the taxa or when
the number of taxa to be analyzed is very large. Furthermore,
the two approaches seem to give similar results (e.g., Salamin
et al., 2002).
The supertree approach has not been applied extensively to
angiosperms, but it offers an opportunity for representation of
greater numbers of taxa than the supermatrix analyses con-
ducted to date. A recent supertree analysis combined trees that
included all angiosperm families and produced the first com-
prehensive family-level phylogenetic tree for angiosperms
(Davies et al., 2004). The basic framework of the angiosperm
supertree is largely consistent with the results of large, mul-
tigene analyses of exemplar taxa (e.g., D. Soltis et al., 2000)
on which it was based. Amborella is sister to all other angio-
sperms, followed by Nymphaeaceae, Austrabaileyales, a clade
of monocots 1 Ceratophyllaceae, Chloranthaceae, and a clade
of magnoliids 1 eudicots. Relationships within monocots,
magnoliids, and eudicots are also mostly consistent with the
results of the supermatrix analyses. This congruence indicates
that the placement of those taxa not included in the super-
matrix analyses may be correct, inasmuch as the data set can
convey. Furthermore, some clades that have been difficult to
place appear in resolved locations. For example, Caryophyl-
lales are sister to the asterids, and Saxifragales are sister to the
rosids, positions they occupy in some of the shortest trees ob-
tained in other analyses but not in the strict consensus trees
(e.g., D. Soltis et al., 2000). Although the best methods of
supertree construction remain under debate, supertree ap-
proaches seem a viable alternative to supermatrix analyses as
data sets continue to grow.
1622 [Vol. 91A
MERICAN
J
OURNAL OF
B
OTANY
REMAINING PROBLEMS AND FUTURE PROSPECTS
Despite tremendous progress in angiosperm phylogenetics
during the past 10 years, several difficult problems remain.
Most prominent are (1) relationships among monocots, Chlor-
anthaceae, magnoliids, and eudicots, (2) branching order
among basal eudicots, (3) relationships among the major
clades of core eudicots, (4) relationships within rosids, (5) re-
lationships of the many lineages of parasitic plants (although
this problem has been addressed recently by Barkman et al.,
2004), and (6) integration of fossils with extant taxa into a
comprehensive tree of angiosperm phylogeny. Solving these
problems will require coordinated efforts among angiosperm
systematists and paleobotanists and a large amount of molec-
ular (and other, where appropriate) data.
At least some of those nodes that remain poorly resolved
(e.g., basal eudicots, core eudicots, rosids) may be the results
of rapid radiations (see P. Soltis et al., 2004). If so, increased
sampling of molecular characters coupled with inclusion of
additional taxa (if a clade has not yet been thoroughly sam-
pled) may help to resolve at least some of the remaining po-
lytomies. For example, the addition of 26S rDNA sequences
to the three-gene data set of D. Soltis et al. (2000) for a subset
of eudicots provided evidence for Gunnerales as the sister to
the remaining lineages of core eudicots (D. Soltis et al., 2003),
and increased character sampling for data sets of more than
100 taxa improved relationships among basal angiosperms
(Zanis et al., 2002) and within monocots (e.g., Chase et al., in
press). However, if the lack of resolution is due to a true ra-
diation, it may not be possible to resolve these nodes. Like-
wise, if poor resolution has resulted from other factors, such
as extinction, inadequate sampling of extant lineages, ancient
reticulation, horizontal gene transfer, or unequal evolutionary
rates among lineages, then the prospects for resolution, using
currently available data and methods of analysis, are also poor.
Estimates of the age of the angiosperms and the timing of
important divergences based on molecular data do not gener-
ally agree with each other (ranging from ;125 to .400 mya)
or with dates determined from the fossil record (see e.g., San-
derson and Doyle, 2001; P. Soltis et al., 2002; Sanderson et
al., 2004). Although most molecular-based ages for angio-
sperms, and other groups of organisms (e.g., Heckman et al.,
2001), are much older than the fossil record suggests, many
recent estimates based on methods that do not assume equal
rates of evolutionary change among lineages are similar to, if
slightly older than, dates inferred from the fossil record (San-
derson et al., 2004). Furthermore, estimated ages for specific
angiosperm clades are generally older than inferences from the
fossil record (e.g., Wikstro¨m et al., 2001, compared with Ma-
gallo´n et al., 1999), but these discrepancies are much smaller
than those reported for the age of the angiosperms. However,
room for further reconciliation of age estimates inferred from
fossils and molecular data remains. For example, given the
numerous diverse fossils reported from as early as 115–125
mya, perhaps the earliest angiosperms were older than the con-
servative estimate of ;130 mya. Conversely, molecular meth-
ods tend to overestimate ages (Rodrı´guez-Trelles et al., 2002),
so refinement of dating approaches is needed to compensate
for this bias.
Many of the large clades identified through analysis of mo-
lecular data are not easily recognized morphologically. Al-
though possible synapomorphies for many clades have been
proposed by Doyle and Endress (2000) and Judd et al. (2002),
the identification of nonmolecular synapomorphies is still
needed for many clades. This task will require new morpho-
logical and molecular data for many groups, including both a
search for new characters and filling in data for many families.
Finally, all of this new information—sequences, trees, mor-
phological data—will need to be managed in such a way as
to make it easily accessible to all who are interested via public
databases. The development and maintenance of informatics
tools and resources are therefore also major challenges that lie
ahead for angiosperm systematics. Informatics issues may be-
come particularly important as new methods are needed to
analyze large amounts of sequence data for more taxa than
have yet been analyzed together and to develop algorithms and
methods for constructing supertrees to link new trees with
those that have been archived.
The phylogenetic information currently available for angio-
sperms, and that to come, is fundamentally important for or-
ganizing all that is known about the angiosperm branch of the
tree of life. However, this phylogenetic information is also a
prerequisite for addressing basic questions in a number of oth-
er fields, ranging from genomics to ecology.
LITERATURE CITED
A
LBACH
, D., P. S. S
OLTIS
,
AND
D. E. S
OLTIS
. 2001a. Patterns of embryolog-
ical and biochemical evolution in the asterids. Systematic Biology 26:
242–262.
A
LBACH
, D. C., P. S. S
OLTIS
,D.E.S
OLTIS
,
AND
R. G. O
LMSTEAD
. 2001b.
Phylogenetic analysis of the Asteridae s.l. using sequences of four genes.
Annals of the Missouri Botanical Garden 88: 163–212.
A
NGIOSPERM
P
HYLOGENY
G
ROUP
(APG). 1998. An ordinal classification for
the families of flowering plants. Annals of the Missouri Botanical Garden
85: 531–553.
A
NGIOSPERM
P
HYLOGENY
G
ROUP
(APG II). 2003. An update of the Angio-
sperm Phylogeny Group classification for the orders and families of flow-
ering plants: APG II. Botanical Journal of the Linnean Society 141: 399
436.
B
AILLON
, H. 1869. Histoires des plantes, vol. 1. L. Hachette & Cie, Paris,
France.
B
ARKMAN
, T. J., G. C
HENERY
,J.R.M
CNEAL
,J.L
YONS
-W
EILER
,
AND
C. W.
D
E
P
AMPHILIS
. 2000. Independent and combined analyses of sequences
from all three genomic compartments converge on the root of flowering
plant phylogeny. Proceedings of the National Academy of Sciences, USA
97: 13166–13171.
B
ARKMAN
, T. J., S.-H. L
IM
,K.M.S
ALLEH
,
AND
J. N
AIS
. 2004. Mitochondrial
DNA sequences reveal the photosynthetic relatives of Rafflesia, the
world’s largest flower. Proceedings of the National Academy of Sciences,
USA 101: 787–792.
B
ASINGER
, J.,
AND
D. L. D
ILCHER
. 1984. Ancient bisexual flowers. Science
224: 511–513.
B
AUM
, B. R. 1992. Combining trees as a way of combining data sets for
phylogenetic inference, and the desirability of combining gene trees. Tax-
on 41: 3–10.
B
EHNKE
, H.-D. 1969. Die Siebrohren-Plastiden bei Monocotlen. Nauturwis-
senschaften 55: 140–141.
B
HARATHAN
, G.,
AND
E. A. Z
IMMER
. 1995. Early branching events in mono-
cotyledons—partial 18S ribosomal DNA sequence analysis. In P. J. Ru-
dall, P. J. Cribb, D. F. Cutler, and C. J. Humphries [eds.], Monocotyle-
dons: systematics and evolution, 81–107. Royal Botanic Gardens, Kew,
London, UK.
B
ININDA
-E
MONDS
,O.R.P.,
AND
H. N. B
RYANT
. 1998. The properties of
matrix representation with parsimony analyses. Systematic Biology 47:
497–508.
B
ORSCH
, T., K. W. H
ILU
,D.Q
UANDT
,V.W
ILDE
,C.N
EINHUIS
,
AND
W.
B
ARTHLOTT
. 2003. Non-coding plastid trnT-trnF sequences reveal a
highly supported phylogeny of basal angiosperms. Journal of Evolution-
ary Biology 15: 558–567.
B
RAUN
, A. 1864. Uebersicht der naturlichen Systems nach der Anordnung
derselben. In P. Ascherson [ed.], Flora der Provinz Brandenburg, der
October 2004] 1623S
OLTIS AND
S
OLTIS
—D
IVERSIFICATION OF ANGIOSPERMS
Altmark und des Herzogthums Magdeburg, vol. 1, 22–67. Hirschwald,
Berlin, Germany.
B
REMER
, B., K. B
REMER
,N.H
EIDARI
,P.E
RIXON
,R.G.O
LMSTEAD
,A.A.
A
NDERBERG
,M.K
A
¨
LLERSJO
¨
,
AND
E. B
ARKHORDARIAN
. 2002. Phylo-
genetics of asterids based on 3 coding and 3 non-coding chloroplast DNA
markers and the utility of non-coding DNA at higher taxonomic levels.
Molecular Phylogenetics and Evolution 24: 274–301.
B
REMER
, K. 2000. Early Cretaceous lineages of monocot flowering plants.
Proceedings of the National Acadamy of Sciences, USA 97: 4707–4711.
B
REMER
, K., A. B
ACKLUND
,B.S
ENNBLAD
,U.S
WENSON
,K.A
NDREASEN
,
M. H
JERTSON
,J.L
UNDBERG
,M.B
ACKLUND
,
AND
B. B
REMER
. 2001. A
phylogenetic analysis of 1001 genera and 501 families of euasterids
based on morphological and molecular data with notes on possible higher
level morphological synapomorphies. Plant Systematics and Evolution
229: 137–169.
B
URLEIGH
,J.G.,
AND
S. M
ATHEWS
. 2004. Phylogenetic signal in nucleotide
data from seed plants: Implications for resolving the seed plant tree of
life. American Journal of Botany 91: 1599–1613.
C
AMERON
, K. M., K. J. W
URDACK
,
AND
R. W. J
OBSON
. 2002. Molecular
evidence for the common origin of snap-trees among carnivorous plants.
American Journal of Botany 89: 1503–1509.
C
HASE
, M. W. 2004. Monocot relationships: an overview. American Journal
of Botany 91: 1645–1656.
C
HASE
,M.W.,
AND
V. A. A
LBERT
. 1998. A perspective on the contribution
of plastid rbcL DNA sequences to angiosperm phylogenetics. In D. E.
Soltis, P. S. Soltis, and J. J. Doyle [eds.], Molecular systematics of plants,
vol. 2, DNA sequencing, 488–507. Kluwer, Boston, Massachusetts, USA.
C
HASE
, M. W., D. E. S
OLTIS
,R.G.O
LMSTEAD
,D.M
ORGAN
,D.H.L
ES
,B.
D. M
ISHLER
,M.R.D
UVALL
,R.A.P
RICE
,H.G.H
ILLS
, Y.-L. Q
IU
,K.
A. K
RON
,J.H.R
ETTIG
,E.C
ONTI
,J.D.P
ALMER
,J.R.M
ANHART
,K.J.
S
YTSMA
,H.J.M
ICHAELS
,W.J.K
RESS
,K.G.K
AROL
,W.D.C
LARK
,
M. H
EDREN
,B.S.G
AUT
,R.K.J
ANSEN
, K.-J. K
IM
,C.F.W
IMPEE
,J.F.
S
MITH
,G.R.F
URNIER
,S.H.S
TRAUSS
, Q.-Y. X
IANG
,G.M.P
LUNKETT
,
P. S. S
OLTIS
,S.M.S
WENSEN
,S.E.W
ILLIAMS
,P.A.G
ADEK
,C.J.Q
UINN
,
L. E. E
GUIARTE
,E.G
OLENBERG
,G.H.L
EARN
,J
R
., S. W. G
RAHAM
,S.
C. H. B
ARRETT
,S.D
AYANANDAN
,
AND
V. A. A
LBERT
. 1993. Phyloge-
netics of seed plants: an analysis of nucleotide sequences from the plastid
gene rbcL. Annals of the Missouri Botanical Garden 80: 528–580.
C
HASE
, M. W., D. E. S
OLTIS
,P.S.S
OLTIS
,P.J.R
UDALL
,M.F.F
AY
,W.H.
H
AHN
,S.S
ULLIVAN
,J.J
OSEPH
,T.G
IVNISH
,K.J.S
YTSMA
,
AND
J. C.
P
IRES
. 2000. Higher-level systematics of the monocotyledons: an as-
sessment of current knowledge and a new classification. In K. L. Wilson
and D. A. Morrison [eds.], Monocots systematics and evolution, 3–16.
CSIRO, Melbourne, Australia.
C
HASE
, M. W., M. F. F
AY
,D.S.D
EVEY
,O.M
AURIN
,J.D
AVIES
,Y.P
ILLON
,
G. P
ETERSEN
,O.S
EBERG
,M.N.T
AMURA
,C.B.A
SMUSSEN
,K.H
ILU
,
T. B
ORSCH
,J.I.D
AVIS
,D.W.S
TEVENSON
,J.C.P
IRES
,T.J.G
IVNISH
,
K. J. S
YTSMA
,
AND
S. W. G
RAHAM
. In press. Multi-gene analyses of
monocot relationships: a summary. In J. T. Columbus, E. A. Friar, C. W.
Hamilton, J. M. Porter, L. M. Prince, and M. G. Simpson [eds.], Mono-
cots: comparative biology and evolution. Rancho Santa Ana Botanic Gar-
den, Claremont, California, USA.
C
OUPER
, R. A. 1958. British Mesozoic microspores and pollen grains: a sys-
tematic and stratigraphic study. Palaeontographica, Abteilung B, Pala¨o-
phytologie 103: 75–179.
C
RANE
, P. R., E. M. F
RIIS
,
AND
K. R. P
EDERSEN
. 1995. The origin and early
diversification of angiosperms. Nature 374: 27–33.
C
RANE
, P. R., P. S. H
ERENDEEN
,
AND
E. M. F
RIIS
. 2004. Fossils and plant
phylogeny. American Journal of Botany 91: 1683–1699.
C
REPET
,W.,
AND
K. N
IXON
. 1998. Fossil Clusiaceae from the Late Creta-
ceous (Turonian) of New Jersey and implications regarding the history
of bee pollination. American Journal of Botany 85: 1122–1133.
C
RONQUIST
, A. 1981. An integrated system of classification of flowering
plants. Columbia University Press, New York, New York, USA.
C
RONQUIST
, A. 1988. The evolution and classification of flowering plants,
2nd ed. New York Botanical Garden, Bronx, New York, USA.
C
UE
´
NOUD
, P., V. S
AVOLAINEN
,L.W.C
HATROU
,M.P
OWELL
,R.J.G
RAYER
,
AND
M. W. C
HASE
. 2002. Molecular phylogenetics of Caryophyllales
based on nuclear 18S rDNA and plastid rbcL, atpB, and matK DNA
sequences. American Journal of Botany 89: 132–144.
D
AHLGREN
, R. 1980. A revised system of classification of the angiosperms.
Botanical Journal of the Linnean Society 80: 91–124.
D
AHLGREN
, R., H. C
LIFFORD
,
AND
P. Y
EO
. 1985. The families of the mono-
cotyledons: structure, evolution and taxonomy. Springer-Verlag, Berlin,
Germany.
D
AVIES
, T. J., T. G. B
ARRACLOUGH
,M.W.C
HASE
,P.S.S
OLTIS
,D.E.S
OLTIS
,
AND
V. S
AVOLAINEN
. 2004. Darwin’s abominable mystery: insights from
a supertree of the angiosperms. Proceedings of the National Academy of
Sciences, USA 101: 1904–1909.
D
AVIS
,C.C.,
AND
M. W. C
HASE
. 2004. Elatinaceae are sister to Malpighi-
aceae; Peridiscaceae belong to Saxifragales. American Journal of Botany
91: 262–273.
D
AVIS
, J. I., M. P. S
IMMONS
,D.W.S
TEVENSON
,
AND
J. F. W
ENDEL
. 1998.
Data decisiveness, data quality, and incongruence in phylogenetic anal-
ysis: an example from the monocotyledons using mitochondrial atpA
sequences. Systematic Biology 47: 282–310.
D
AVIS
, J. I., D. W. S
TEVENSON
,G.P
ETERSEN
,O.S
EBERG
,L.M.C
AMPBELL
,
J. V. F
REUDENSTEIN
,D.H.G
OLDMAN
,C.R.H
ARDY
,F.A.M
ICHELAN
-
GELI
,M.P.S
IMMONS
,C.D.S
PECHT
,F.V
ERGARA
-S
ILVA
,
AND
M. A.
G
ANDOLFO
. In press. A phylogeny of the monocots, as inferred from
rbcL and atpA sequence variation. Systematic Botany: in press.
DE
J
USSIEU
, A. L. 1789. Genera plantarum secundum ordines naturals dis-
posita. Heissant and Barrois, Paris, France.
D
EROIN
, T. 2000. Notes on the vascular anatomy of the fruit of Takhtajania
(Winteraceae) and its interpretation. Annals of the Missouri Botanical
Garden 87: 398–406.
D
ILCHER
, D. L. 1989. The occurrence of fruits with affinities to Ceratophyl-
laceae in lower and mid-Cretaceous sediments. American Journal of Bot-
any (Supplement) 76: 162.
D
ILCHER
,D.L.,
AND
P. R. C
RANE
. 1984. Archaeanthus: an early angiosperm
from the Cenomanian of the western interior of North America. Annals
of the Missouri Botanical Garden 71: 351–383.
D
ONOGHUE
,M.J.,
AND
J. A. D
OYLE
. 1989. Phylogenetic studies of seed
plants and angiosperms based on morphological characters. In K. Bremer
and H. Jo¨rnvall [eds.], The hierarchy of life: molecules and morphology
in phylogenetic studies, 181–193. Elsevier Science Publishers, Amster-
dam, Netherlands.
D
OUGLAS
, A.,
AND
S. C. T
UCKER
. 1996. Comparative floral ontogenies
among Persoonioideae including Bellendena (Proteaceae). American
Journal of Botany 83: 1528–1555.
D
OYLE
,J.A.,
AND
M. J. D
ONOGHUE
. 1986. Seed plant phylogeny and the
origin of the angiosperms: an experimental cladistic approach. Botanical
Review 52: 321–431.
D
OYLE
,J.A.,
AND
M. J. D
ONOGHUE
. 1992. Fossils and seed plant phylogeny
reanalzyed. Brittonia 44: 89–106.
D
OYLE
,J.A.,
AND
P. K. E
NDRESS
. 2000. Morphological phylogenetic anal-
yses of basal angiosperms: comparison and combination with molecular
data. International Journal of Plant Sciences 161(Supplement): S121–
S153.
D
OYLE
, J. A., H. E
KLUND
,
AND
P. S. H
ERENDEEN
. 2003. Floral evolution in
Chloranthaceae: implications of a morphological phylogenetic analysis.
International Journal of Plant Sciences 164(Supplement): S365–S382.
D
RINNAN
, A. N., P. R. C
RANE
,
AND
S. B. H
OOT
. 1994. Patterns of floral
evolution in the early diversification of non-magnoliid dicotyledons (eu-
dicots). Plant Systematics and Evolution 8(Supplement): 93–122.
D
UVALL
, M. R., G. H. L
EARN
,L.E.E
GUIARTE
,
AND
M. T. C
LEGG
. 1993.
Phylogenetic analysis of rbcL sequences identifies Acorus calamus as the
primal extant monocotyledon. Proceedings of the National Academy of
Sciences, USA 90: 4611–4644.
E
ICHLER
, A. 1875–1878. Blu¨thendiagramme I/II. Engelmann, Leipzig, Ger-
many.
E
KLUND
, H., J. A. D
OYLE
,
AND
P. S. H
ERENDEEN
. 2004. Morphological phy-
logenetic analysis of living and fossil Chloranthaceae. International
Journal of Plant Sciences 165: 107–151.
E
NDRESS
, P. K. 1986. Reproductive structures and phylogenetic significance
of extant primitive angiosperms. Plant Systematic Evolution 152: 1–28.
E
NDRESS
, P. K. 2001. The flowers in extant basal angiosperms and inferences
on ancestral flowers. International Journal of Plant Sciences 162: 1111–
1140.
E
NDRESS
,P.K.,
AND
A. I
GERSHEIM
. 1997. Gynoecium diversity and system-
atics of the Laurales. Botanical Journal of the Linnean Society 125: 93–
168.
E
NDRESS
,P.K.,
AND
A. I
GERSHEIM
. 2000a. Gynoecium structure and evo-
lution in basal angiosperms. International Journal of Plant Sciences
161(Supplement): S211–S223.
E
NDRESS
,P.K.,
AND
A. I
GERSHEIM
. 2000b. The reproductive structures of
1624 [Vol. 91A
MERICAN
J
OURNAL OF
B
OTANY
the basal angiosperm Amborella trichopoda (Amborellaceae). Interna-
tional Journal of Plant Sciences 161(Supplement): S237–S248.
F
EILD
, T. S., M. A. Z
WEINIECKI
,T.B
RODRIBB
,T.J
AFFRE
´
,M.J.D
ONOGHUE
,
AND
N. M. H
OLBROOK
. 2000. Structure and function of tracheary ele-
ments in Amborella trichopoda. International Journal of Plant Sciences
161: 705–712.
F
ISHBEIN
, M., C. H
IBSCH
-J
ETTER
,D.E.S
OLTIS
,
AND
L. H
UFFORD
. 2001.
Phylogeny of Saxifragales (angiosperms, eudicots): analysis of a rapid,
ancient radiation. Systematic Biology 50: 817–847.
F
RIIS
, E. M., K. R. P
EDERSEN
,
AND
P. R. C
RANE
. 2000. Reproductive structure
and organization of basal angiosperms from the early Cretaceous (Bar-
remian or Aptian) of western Portugal. International Journal of Plant
Sciences 161(Supplement): S169–S182.
F
RIIS
, E. M., K. R. P
EDERSEN
,
AND
P. R. C
RANE
. 2001. Fossil evidence of
water lilies in the Early Cretaceous. Nature 410: 357–360.
F
RIIS
, E. M., J. A. D
OYLE
,P.K.E
NDRESS
,
AND
Q. L
ENG
. 2003. Archaefruc-
tus—angiosperm precursor or specialized early angiosperm? Trends in
Plant Science 8: 369–373.
G
ANDOLFO
, M. A., K. C. N
IXON
,
AND
W. L. C
REPET
. 2002. Triuridaceae
fossil flowers from the Upper Cretaceous of New Jersey. American Jour-
nal of Botany 89: 1940–1957.
G
ANDOLFO
, M. A., K. C. N
IXON
,
AND
W. L. C
REPET
. 2004. Cretaceous flow-
ers of Nymphaeaceae and implications for complex insect entrapment
pollinations mechanisms in early Angiosperms. Proceedings of the Na-
tional Academy of Sciences, USA 101: 8056–8060.
G
AUT
, B. S., B. R. M
ORTON
,B.M.M
C
C
AIG
,
AND
M. T. C
LEGG
. 1996. Sub-
stitution rate comparisons between grasses and palms: synonymous rate
differences at the nuclear gene Adh parallel rate differences at the plastid
gene rbcL. Proceedings of the National Academy of Sciences, USA 93:
10274–10279.
G
AUT
, B. S., S. V. M
USE
,W.D.C
LARK
,
AND
M. T. C
LEGG
. 1992. Relative
rates of nucleotide substitution at the rbcL locus of monocotyledonous
plants. Journal of Molecular Evolution 35: 292–303.
G
IVNISH
,T.J.,
AND
K. J. S
YTSMA
. 1997. Consistency, characters, and the
likelihood of correct phylogenetic inference. Molecular Phylogenetics
and Evolution 7: 320–333.
G
IVNISH
, T. J., J. C. P
IRES
,S.W.G
RAHAM
,M.A.M
C
P
HERSON
,L.M.P
RINCE
,
T. B. P
ATTERSON
,H.S.R
AI
,E.H.R
OALSON
,T.M.E
VANS
,W.J.H
AHN
,
K. C. M
ILLAM
,A.W.M
EEROW
,M.M
OLVRAY
,P.J.K
ORES
,H.E.
O’B
RIEN
,J.C.H
ALL
,W.J.K
RESS
,
AND
K. J. S
YTSMA
. 2004. Phylo-
genetic relationships of monocots based on the highly informative plastid
gene ndhF: evidence for widespread converted convergence. In J. T. Co-
lumbus, E. A. Friar, C. W. Hamilton, J. M. Porter, L. M. Prince, and M.
G. Simpson [eds.], Monocots: comparative biology and evolution. Ran-
cho Santa Ana Botanic Garden, Claremont, California, USA.
G
OREMYKIN
, V., V. K. I. H
IRSCH
-E
RNST
,S.W
O
¨
LFL
,
AND
F. H. H
ELLWIG
.
2003. Analysis of the Amborella trichopoda chloroplast genome se-
quence suggests that Amborella is not a basal angiosperm. Molecular
Biology and Evolution 20: 1499–1505.
G
OTTSBERGER
, G. 1977. Some aspects of beetle pollination in the evolution
of flowering plants. Plant Systematics and Evolution 1: 211–226.
G
OTTSBERGER
, G. 1988. The reproductive biology of primitive angiosperms.
Taxon 37: 630643.
G
RAHAM
,S.W.,
AND
R. G. O
LMSTEAD
. 2000. Utility of 17 chloroplast genes
for inferring the phylogeny of the basal angiosperms. American Journal
of Botany 87: 1712–1730.
G
RAHAM
, S. W., P. A. R
EEVES
,A.C.E.B
URNS
,
AND
R. G. O
LMSTEAD
. 2000.
Microstructural changes in noncoding chloroplast DNA: interpretation,
evolution, and utility of indels and inversions in basal angiosperm phy-
logenetic inference. International Journal of Plant Sciences 161(Supple-
ment): S83–S96.
H
ARMS
, H. 1934. Hamamelidaceae. In A. Engler and K. Prantl [eds.], Die
natu¨rlichen Pflanzenfamilien, vol. 2, 330–343. W. Engelmann, Leipzig,
Germany.
H
ECKMAN
, D. S., D. M. G
EISER
,B.R.E
IDELL
,R.L.S
TAUFFER
,N.L.K
AR
-
DOS
,
AND
S. B. H
EDGES
. 2001. Molecular evidence for the early colo-
nization of land by fungi and plants. Science 293: 1129–1133.
H
EYWOOD
, V. 1993. Flowering plants of the world. B.T. Batsford Ltd., Lon-
don, UK.
H
ICKEY
,L.J.,
AND
A. D. W
OLFE
. 1975. The bases of angiosperm phylogeny:
vegetative morphology. Annals of the Missouri Botanical Garden 62:
538–589.
H
ILLIS
, D. M. 1996. Inferring complex phylogenies. Nature 383: 130.
H
ILU
,K.W.,T.B
ORSCH
,K.M
ULLER
,D.E.S
OLTIS
,P.S.S
OLTIS
,V.S
AVO
-
LAINEN
,M.C
HASE
,M.P
OWELL
,L.A
LICE
,R.E
VANS
,H.S
AUQUET
,C.
N
EINHUIS
,T.S
LOTTA
,J.R
OHWER
,
AND
L. C
HATROU
. 2003. Inference
of angiosperm phylogeny based on matK sequence information. Ameri-
can Journal of Botany 90: 1758–1776.
H
OOT
, S. B., S. M
AGALLO
´
N
,
AND
P. R. C
RANE
. 1999. Phylogeny of basal
eudicots based on three molecular data sets: atpB, rbcL, and 18S nuclear
ribosomal DNA sequences. Annals of the Missouri Botanical Garden 86:
1–32.
H
UFFORD
, L. 1992. Rosidae and their relationships to other nonmagnoliid
dicotyledons: a phylogenetic analysis using morphological and chemical
data. Annals of the Missouri Botanical Garden 79: 218–248.
J
ENSEN
, S. R. 1992. Systematic implications of the distribution of iridoids
and other chemical compounds in the Loganiaceae and other families of
the Asteridae. Annals of the Missouri Botanical Garden 79: 284–302.
J
UDD
,W.S.,
AND
R. G. O
LMSTEAD
. 2004. A survey of tricolpate (eudicot)
phylogeny. American Journal of Botany 91: 1627–1644.
J
UDD
, W. S., C. S. C
AMPBELL
,E.A.K
ELLOGG
,P.F.S
TEVENS
,
AND
M. J.
D
ONOGHUE
. 2002. Plant systematics: a phylogenetic approach. Sinauer
Associates, Inc., Sunderland, Massachusetts, USA.
J
UNIPER
, B. E., R. J. R
OBINS
,
AND
D. M. J
OEL
. 1989. The carnivorous plants.
Academic Press, London, UK.
K
EATING
, R. C. 2000. Anatomy of the young vegetative shoot of Takhtajania
perrieri (Winteraceae). Annals of the Missouri Botanical Garden 87:
335–346.
K
IM
, S., V. A. A
LBERT
, M.-J. Y
OO
,J.S.F
ARRIS
,M.Z
ANIS
,P.S.S
OLTIS
,
AND
D. E. S
OLTIS
. In press. Pre-angiosperm duplication of floral genes
and regulatory tinkering at the base of flowering plants. American Jour-
nal of Botany.
K
IM
, S., D. E. S
OLTIS
,P.S.S
OLTIS
,M.J.Z
ANIS
,
AND
Y. S
UH
. 2004. Phy-
logenetic relationships among early-diverging eudicots based on four
genes: were the eudicots ancestrally woody? Molecular Phylogenetics
and Evolution 31: 16–30.
K
ONG
, H.-Z., Z. C
HEN
,
AND
A.-M. L
U
. 2002. Phylogeny of Chloranthus
(Chloranthaceae) based on nuclear ribosomal ITS and plastid trnL-F se-
quence data. American Journal of Botany 89: 940–946.
K
RAMER
, E. M., R. L. D
ORIT
,
AND
V. F. I
RISH
. 1998. Molecular evolution of
genes controlling petal and stamen development: duplication and diver-
gence within the APETALA3 and PISTILLATA MADS-box gene line-
ages. Genetics 149: 765–783.
K
UZOFF
, R. K., L. H
UFFORD
,
AND
D. E. S
OLTIS
. 2001. Structural homology
and developmental transformations associated with ovary diversification
in Lithophragma (Saxifragaceae). American Journal of Botany 88: 196–
205.
L
ES
, D. H., E. L. S
CHNEIDER
,D.J.P
ADGETT
,M.Z
ANIS
,D.E.S
OLTIS
,
AND
P. S. S
OLTIS
. 1999. Phylogeny, classification, and floral evolution of
water lilies (Nymphaeales): a synthesis of non-molecular, rbcL, matK,
and 18S rDNA data. Systematic Botany 24: 28–46.
L
IPOK
, B., A. A. G
ARDINE
,P.S.W
ILLIAMSON
,
AND
S. S. R
ENNER
. 2000.
Pollination by flies, bees, and beetles of Nuphar ozarkana and N. advena
(Nymphaeaceae). American Journal of Botany 87: 898–902.
L
ITT
, A.,
AND
V. I
RISH
. 2003. Duplication and diversification in the APE-
TALA1/FRUITFULL floral homeotic gene lineage: implications for the
evolution of floral development. Genetics 165: 821–833.
L
OCONTE
, H.,
AND
D. W. S
TEVENSON
. 1991. Cladistics of the Magnoliidae.
Cladistics 7: 267–296.
M
ABBERLEY
, D. J. 1993. The plant book: a portable dictionary of the vascular
plants. Cambridge University Press, Cambridge, UK.
M
AGALLO
´
N
, S., P. R. C
RANE
,
AND
P. S. H
ERENDEEN
. 1999. Phylogenetic
pattern, diversity, and diversification of eudicots. Annals of the Missouri
Botanical Garden 86: 297–372.
M
AGALLO
´
N
, S.,
AND
M. J. S
ANDERSON
. 2001. Absolute diversification rates
in angiosperm clades. Evolution 55: 1762–1780.
M
ATHEWS
, S.,
AND
M. J. D
ONOGHUE
. 1999. The root of angiosperm phylog-
eny inferred from duplicate phytochrome genes. Science 286: 947–949.
M
ATHEWS
, S.,
AND
M. J. D
ONOGHUE
. 2000. Basal angiosperm phylogeny
inferred from duplicate phytochromes A and C. International Journal of
Plant Sciences 161(Supplement): S41–S55.
M
EIMBERG
, H., P. D
ITTRICH
,G.B
RINGMANN
,J.S
CHLAUER
,
AND
G. H
EUBL
.
2000. Molecular phylogeny of Caryophyllidae s.l. based on matK se-
quences with special emphasis on carnivorous taxa. Plant Biology 2:
218–228.
M
ORGAN
,D.R.,
AND
D. E. S
OLTIS
. 1993. Phylogenetic relationships among
October 2004] 1625S
OLTIS AND
S
OLTIS
—D
IVERSIFICATION OF ANGIOSPERMS
Saxifragaceae sensu lato based on rbcL sequence data. Annals of the
Missouri Botanical Garden 80: 631–660.
N
ANDI
, O. I., M. W. C
HASE
,
AND
P. K. E
NDRESS
. 1998. A combined cladistic
analysis of angiosperms using rbcL and nonmolecular data sets. Annals
of the Missouri Botanical Garden 85: 137–212.
N
ICKERSON
, J.,
AND
G. D
ROUIN
. 2004. The sequence of the largest subunit
of RNA polymerase II is a useful marker for inferring seed plant phy-
logeny. Molecular Phylogenetics and Evolution 31: 403–415.
N
ICKRENT
,D.L.,
AND
E. M. S
TARR
. 1994. High rates of nucleotide substi-
tution in nuclear small-subunit (18S) rDNA from holoparasitic flowering
plants. Journal of Molecular Evolution 39: 62–70.
N
ICKRENT
, D. L., R. J. D
UFF
,A.C
OLWELL
,A.D.W
OLFE
,N.D.Y
OUNG
,K.
E. S
TEINER
,
AND
C. W. D
E
P
AMPHILIS
. 1998. Molecular phylogenetic
and evolutionary studies of parasitic plants. In D. E. Soltis, P. S. Soltis,
and J. J. Doyle [eds.], Molecular systematics of plants, vol. 2, 211–241.
Kluwer, Boston, Massachusetts, USA.
N
ICKRENT
, D. L., A. B
LARER
, Y.-L. Q
IU
,D.E.S
OLTIS
,P.S.S
OLTIS
,
AND
M.
Z
ANIS
. 2002. Molecular data place Hydnoraceae with Aristolochiaceae.
American Journal of Botany 89: 1809–1817.
O
LMSTEAD
, R. G., H. M
ICHAELS
,K.S
COTT
,
AND
J. P
ALMER
. 1992. Mono-
phyly of the Asteridae and identification of their major lineages inferred
from DNA sequences of rbcL. Annals of the Missouri Botanical Garden
79: 249–265.
O
LMSTEAD
, R. G., B. B
REMER
,K.M.S
COTT
,
AND
J. D. P
ALMER
. 1993. A
parsimony analysis of the Asteridae sensu lato based on rbcL sequences.
Annals of the Missouri Botanical Garden 80: 700–722.
O
LMSTEAD
, R. G., K.-J. K
IM
,R.K.J
ANSEN
,
AND
S. J. W
AGSTAFF
. 2000. The
phylogeny of the Asteridae sensu lato based on chloroplast ndhF gene
sequence. Molecular Phylogenetics and Evolution 16: 96–112.
P
ARKINSON
, C. L., K. L. A
DAMS
,
AND
J. D. P
ALMER
. 1999. Multigene anal-
yses identify the three earliest lineages of extant flowering plants. Cur-
rent Biology 9: 1485–1488.
P
ERKINS
, J. 1925. Ubersicht uber die Gattungen der Monimiaceae. Engel-
mann, Leipzig, Germany.
P
OLLOCK
, D. D., D. J. Z
WICKL
,J.A.M
CGUIRE
,
AND
D. M. H
ILLIS
. 2002.
Increased taxon sampling is advantageous for phylogenetic inference.
Systematic Biology 51: 664671.
P
OSLUSZNY
, U.,
AND
P. B. T
OMLINSON
. 2003. Aspects of inflorescence and
floral development in the putative basal angiosperm Amborella tricho-
poda (Amborellaceae). Canadian Journal of Botany 81: 28–39.
P
URVIS
, A. 1995. A modification to Baum and Ragan’s method for combining
phylogenetic trees. Systematic Biology 44: 251–255.
P
YANKOV
, V. I., E. G. A
RTYUSHEVA
,G.E.E
DWARDS
,C.C.J.B
LACK
,
AND
P. S. S
OLTIS
. 2001. Phylogenetic analysis of tribe Salsoleae (Chenopo-
diaceae) based on ribosomal ITS sequences: implications for the evolu-
tion of photosynthesis types. American Journal of Botany 88: 1189–
1198.
Q
IU
, Y.-L., J. L
EE
,F.B
ERNASCONI
-Q
UADRONI
,D.E.S
OLTIS
,P.S.S
OLTIS
,M.
Z
ANIS
,Z.C
HEN
,V.S
AVOLAINEN
,
AND
M. W. C
HASE
. 1999. The earliest
angiosperms: evidence from mitochondrial, plastid and nuclear genomes.
Nature 402: 404407.
Q
IU
, Y.-L., J.-Y. L
EE
,F.B
ERNASCONI
-Q
UADRONI
,D.E.S
OLTIS
,P.S.S
OLTIS
,
M. Z
ANIS
,E.Z
IMMER
,Z.C
HEN
,V.S
AVOLAINEN
,
AND
M. C
HASE
. 2000.
Phylogeny of basal angiosperms: analyses of five genes from three ge-
nomes. International Journal of Plant Sciences 161(Supplement): S3–
S27.
R
AGAN
, M. A. 1992. Phylogenetic inference based on matrix representation
of trees. Molecular Phylogenetics and Evolution 1: 53–58.
R
AY
, J. 1703. Methodus plantarum emendata et aucta. Smith and Walford,
London, UK.
R
EICHENBACH
, H. G. L. 1827–1829. Dr. Joh. Christ, Moessler’s Handbuch
der Gewaechskunde, [ed.], vols. 2, 3. Hammerich, Altona, Germany.
R
ENNER
, S. S. 1999. Circumscription and phylogeny of the Laurales: evi-
dence from molecular and morphological data. American Journal of Bot-
any 86: 1301–1315.
R
ODMAN
, J. E. 1991. A taxonomic analysis of glucosinolate-producing
plants.II. Cladistics. Systematic Botany 16: 619629.
R
ODMAN
, J. E., R. A. P
RICE
,K.K
AROL
,E.C
ONTI
,K.J.S
YTSMA
,
AND
J. D.
P
ALMER
. 1993. Nucleotide sequences of the rbcL gene indicate mono-
phyly of mustard oil plants. Annals of the Missouri Botanical Garden
80: 686–699.
R
ODMAN
, J. E., P. S. S
OLTIS
,D.E.S
OLTIS
,K.J.S
YTSMA
,
AND
K. G. K
AROL
.
1998. Parallel evolution of glucosinolate biosynthesis inferred from con-
gruent nuclear and plastid gene phylogenies. American Journal of Botany
85: 997–1006.
R
ODRI
´
GUEZ
-T
RELLES
, F., R. T
ARRI
´
O
,
AND
F. J. A
YALA
. 2002. A methodolog-
ical bias toward overestimation of molecular evolutionary time scales.
Proceedings of the National Academy of Sciences, USA 99: 8112–8115.
R
ONQUIST
, F. 1996. Matrix representation of trees, redundancy, and weight-
ing. Systematic Biology 45: 247–253.
R
ONSE
D
E
C
RAENE
, L. P., P. S. S
OLTIS
,
AND
D. E. S
OLTIS
. 2003. Evolution
of floral structures in basal angiosperms. International Journal of Plant
Sciences 164(Supplement): S329–S363.
S
ALAMIN
, N., T. R. H
ODKINSON
,
AND
V. S
AVOLAINEN
. 2002. Building su-
pertrees: an empirical assessment using the grass family (Poaceae). Sys-
tematic Biology 51: 136–150.
S
ANDERSON
, M. J. 2002. Estimating absolute rates of molecular evolution
and divergence times: a penalized likelihood approach. Molecular Biol-
ogy and Evolution 19: 101–109.
S
ANDERSON
, M. J.,
AND
J. A. D
OYLE
. 2001. Sources of error and confidence
intervals in estimating the age of angiosperms from rbcL and 18S rDNA
data. American Journal of Botany 88: 1499–1516.
S
ANDERSON
, M. J., A. P
URVIS
,
AND
C. H
ENZE
. 1998. Phylogenetic supertrees:
assembling the trees of life. Trends in Ecology and Evolution 13: 105–
109.
S
ANDERSON
, M. J., J. L. T
HORNE
,N.W
ILKSTRO
¨
M
,
AND
K. B
REMER
. 2004.
Molecular evidence of plant divergence times. American Journal of Bot-
any 91: 1656–1665.
S
AUQUET
, H., J. A. D
OYLE
,T.S
CHARASCHKIN
,T.B
ORSCH
,K.W.H
ILU
,L.
W. C
HATROU
,
AND
A. L
E
T
HOMAS
. 2003. Phylogenetic analysis of Mag-
noliales and Myristicaceae based on multiple data sets: implications for
character evolution. Botanical Journal of the Linnean Society 142: 125–
186.
S
AVOLAINEN
, V., M. W. C
HASE
,C.M.M
ORTON
,D.E.S
OLTIS
,C.B
AYER
,
M. F. F
AY
,A.D
E
B
RUIJN
,S.S
ULLIVAN
,
AND
Y.-L. Q
IU
. 2000. Phylo-
genetics of flowering plants based upon a combined analysis of plastid
atpB and rbcL gene sequences. Systematic Biology 49: 306–362.
S
AVOLAINEN
, V., M. F. F
AY
,D.C.A
LBACH
,A.B
ACKLUND
,M.
VAN DER
B
ANK
,K.M.C
AMERON
,S.A.J
OHNSON
,M.D.L
LEDO
´
, J.-C. P
INTAUD
,
M. P
OWELL
,M.C.S
HEAHAN
,D.E.S
OLTIS
,P.S.S
OLTIS
,P.W
ESTON
,
W. M. W
HITTON
,K.J.W
URDACK
,
AND
M. W. C
HASE
. 2000b. Phylogeny
of the eudicots: a nearly complete familial analysis based on rbcL gene
sequences. Kew Bulletin 55: 257–309.
S
CHNEIDER
, E. L. 1979. Pollination biology of the Nymphaeaceae. In D. M.
Caron [ed.], Proceedings of the Fourth International Symposium on Pol-
lination, 419430. Maryland Agricultural Experiment Station Special
Miscellaneous Publication 1. College Park, Maryland, USA.
S
OLTIS
, D. E., V. A. A
LBERT
,V.S
AVOLAINEN
,K.W.H
ILU
, Y.-L. Q
IU
,M.W.
C
HASE
,J.S.F
ARRIS
,S.S
TEFANOVIC
,D.W.R
ICE
,J.D.P
ALMER
,
AND
P. S. S
OLTIS
. In press-a. Genome-scale data, angiosperm relationships,
and ‘ending incongruence’’: a cautionary tale in phylogenetics. Trends
in Plant Science.
S
OLTIS
, D. E., A. E. S
ENTERS
,M.Z
ANIS
,S.K
IM
,J.D.T
HOMPSON
,P.S.
S
OLTIS
,L.P.R
ONSE
D
E
C
RAENE
,P.K.E
NDRESS
,
AND
J. S. F
ARRIS
.
2003. Gunnerales are sister to other core eudicots: implications for the
evolution of pentamery. American Journal of Botany 90: 461–470.
S
OLTIS
,D.E.,
AND
P. S. S
OLTIS
. 2004. Amborella NOT a ‘basal angio-
sperm’’? Not so fast. American Journal of Botany.91: 997–1001.
S
OLTIS
, D. E., P. S. S
OLTIS
,M.W.C
HASE
,
AND
P. K. E
NDRESS
. In press-b.
Angiosperm phylogeny, classification, and evolution. Smithsonian Insti-
tution Press, Washington, D.C., USA.
S
OLTIS
, D. E., P. S. S
OLTIS
,M.W.C
HASE
,M.E.M
ORT
,D.C.A
LBACH
,M.
Z
ANIS
,V.S
AVOLAINEN
,W.J.H
AHN
,S.B.H
OOT
,M.F.F
AY
,M.A
XTELL
,
S. M. S
WENSEN
,L.M.P
RINCE
,W.J.K
RESS
,K.C.N
IXON
,
AND
J. S.
F
ARRIS
. 2000. Angiosperm phylogeny inferred from 18S rDNA, rbcL,
and atpB sequences. Botanical Journal of the Linnean Society 133: 381–
461.
S
OLTIS
, D. E., P. S. S
OLTIS
,D.R.M
ORGAN
,S.M.S
WENSEN
,B.C.M
ULLIN
,
J. M. D
OWD
,
AND
P. G. M
ARTIN
. 1995. Chloroplast gene sequence data
suggest a single origin of the predisposition for symbiotic nitrogen fix-
ation in angiosperms. Proceedings of the National Academy of Sciences,
USA 92: 2647–2651.
S
OLTIS
, D. E., P. S. S
OLTIS
,M.E.M
ORT
,M.W.C
HASE
,V.S
AVOLAINEN
,S.
B. H
OOT
,
AND
C. M. M
ORTON
. 1998. Inferring complex phylogeneis
using parsimony: an empirical approach using three large DNA data sets
for angiosperms. Systematic Biology 47: 32–42.
1626 [Vol. 91A
MERICAN
J
OURNAL OF
B
OTANY
S
OLTIS
, D. E., P. S. S
OLTIS
,D.L.N
ICKRENT
,L.A.J
OHNSON
,W.J.H
AHN
,
S. B. H
OOT
,J.A.S
WEERE
,R.K.K
UZOFF
,K.A.K
RON
,M.W.C
HASE
,
S. M. S
WENSEN
,E.A.Z
IMMER
, S.-M. C
HAW
,L.J.G
ILLESPIE
,W.J.
K
RESS
,
AND
K. J. S
YTSMA
. 1997. Angiosperm phylogeny inferred from
18S ribosomal DNA sequences. Annals of the Missouri Botanical Garden
84: 1–49.
S
OLTIS
, P. S., D. E. S
OLTIS
,
AND
M. W. C
HASE
. 1999. Angiosperm phylogeny
inferred from multiple genes as a tool for comparative biology. Nature
402: 402–404.
S
OLTIS
, P. S., D. E. S
OLTIS
,M.J.Z
ANIS
,
AND
S. K
IM
. 2000. Basal lineages
of angiosperms: relationships and implications for floral evolution. In-
ternational Journal of Plant Science 161(Supplement): S97–S107.
S
OLTIS
, P. S., D. E. S
OLTIS
,V.S
AVOLAINEN
,P.R.C
RANE
,
AND
T. G. B
AR
-
RACLOUGH
. 2002. Rate heterogeneity among lineages of tracheophytes:
integration of molecular and fossil data and evidence for molecular living
fossils. Proceedings of the National Acadamy of Sciences, USA 99:
44304435.
S
OLTIS
, P. S., D. E. S
OLTIS
,M.W.C
HASE
,P.K.E
NDRESS
,
AND
P. R. C
RANE
.
2004. The diversification of flowering plants. In J. Cracraft and M. J.
Donoghue [eds.], Assembling the tree of life, 154–167. Oxford Univer-
sity Press, Oxford, UK.
S
TEBBINS
, G. L. 1974. Flowering plants: Evolution above the species level.
Belknap Press, Cambridge, Massachusetts, USA.
S
UN
, G., D. L. D
ILCHER
,S.Z
HENG
,
AND
Z. Z
HOU
. 1998. In search of the
first flower: a Jurassic angiosperm, Archaefructus, from northeast China.
Science 282: 1692–1695.
S
UN
, G., Q. J
I
,D.L.D
ILCHER
,S.Z
HENG
,K.C.N
IXON
,
AND
X. W
ANG
. 2002.
Archaefructaceae, a new basal angiosperm family. Science 296: 899
904.
S
WENSEN
, S. M. 1996. The evolution of actinorhizal symbioses: evidence for
multiple origins of the symbiotic association. American Journal of Bot-
any 83: 1503–1512.
S
YTSMA
,K.J.,
AND
D. A. B
AUM
. 1996. Molecular phylogenies and the di-
versification of angiosperms. In D. W. Taylor and L. J. Hickey [eds.],
Flowering plant origin, evolution, and phylogeny, 314–340. Chapman
and Hall, New York, New York, USA.
S
YTSMA
, K. J., A. L
ITT
,M.L.Z
JHRA
,J.C.P
IRES
,M.N
EPOKROEFF
,E.C
ONTI
,
AND
P. W
ILSON
. In press. Clades, clocks, and continents: historical and
biogeographical analysis of Myrtaceae, Vochysiaceae, and relatives in
the southern hemisphere. International Journal of Plant Science.
T
AKHTAJAN
, A. 1980. Outline of the classification of flowering plants (Mag-
noliophyta). Botanical Review 46: 225–359.
T
AKHTAJAN
, A. 1987. System of Magnoliophyta. Academy of Sciences, Len-
ingrad, USSR.
T
AKHTAJAN
, A. 1997. Diversity and classification of flowering plants. Co-
lumbia University Press, New York, New York, USA.
T
HORNE
, R. F. 1974. A phylogenetic classification of the Annoniflorae. Aliso
8: 147–209.
T
HORNE
, R. F. 1992. Classification and geography of the flowering plants.
Botanical Review 58: 225–348.
T
ILLICH
, H.-J. 1995. Seedling and systematics in monocotyledons. In P. J.
Rudall, P. J. Cribb, D. F. Cutler, and C. J. Humphries [eds.], Monocot-
yledons: systematics and evolution, 303–352. Royal Botanic Gardens,
Kew, London, UK.
T
OMLINSON
, P. B. 1995. Non-homology of vascular organisation in mono-
cotyledons and dicotyledons. In P. J. Rudall, P. J. Cribb, D. F. Cutler, and
C. J. Humphries [eds.], Monocotyledons: systematics and evolution,
589–622. Royal Botanic Gardens, Kew, London, UK.
VAN
T
IEGHEM
, P. 1897. Sur les Buxace´es. Annales des Sciences Naturelles
Botanique, se´rie 85: 289–338.
W
ALKER
,J.W.,
AND
A. G. W
ALKER
. 1984. Ultrastructure of Lower Creta-
ceous angiosperm pollen and the origin and early evolution of flowering
plants. Annals of the Missouri Botanical Garden 71: 464–521.
W
ARMING
, E. 1879. Haandbog I den systematiske botanik. P. G. Philipsens
Forlag, Copenhagen, Denmark.
W
IKSTRO
¨
M
, N., V. S
AVOLAINEN
,
AND
M. W. C
HASE
. 2001. Evolution of the
angiosperms: calibrating the family tree. Proceedings of the Royal So-
ciety of London, B 268: 2211–2220.
W
ILLIAMS
,P.S.,
AND
E. L. S
CHNEIDER
. 1993. Nelumbonaceae. In K. Ku-
bitzki, J. Rohwer, and V. Bittrich [eds.], The families and genera of vas-
cular plants, 470473. Springer, Berlin, Germany.
W
ILSON
, T. K. 1966. The comparative morphology of the Canellaceae. IV.
Floral morphology and conclusions. American Journal of Botany 53:
336–343.
Y
OUNG
, D. A. 1981. Are the angiosperms primitively vesselless? Systematic
Biology 6: 313–330.
Z
ANIS
, M., D. E. S
OLTIS
,P.S.S
OLTIS
,S.M
ATHEWS
,
AND
M. J. D
ONOGHUE
.
2002. The root of the angiosperms revisited. Proceedings of the National
Academy of Sciences, USA 99: 6848–6853.
Z
ANIS
, M. J., P. S. S
OLTIS
, Y.-L. Q
IU
,E.Z
IMMER
,
AND
D. E. S
OLTIS
. 2003.
Phylogenetic analyses and perianth evolution in basal angiosperms. An-
nals of the Missouri Botanical Garden 90: 129–150.
Z
HANG
, L.-B.,
AND
S. R
ENNER
. 2003. The deepest splits in Chloranthaceae
as resolved by chloroplast sequences. International Journal of Plant Sci-
ences 164(Supplement): S383–S392.
Z
WICKL
,D.J.,
AND
D. M. H
ILLIS
. 2002. Increased taxon sampling greatly
reduces phylogenetic error. Systematic Biology 51: 588–598.
... Among angiosperms, Amborellales (Amborella trichopoda), Nymphaeales (e.g., water lilies), and Austrobaileyales (e.g., star anise), known as the ANA grade (Frohlich, 2006), are placed as successive sisters to all other angiosperms (APG II, 2003;Soltis and Soltis, 2004). The remaining~99.95% of extant angiosperms (mesangiosperms or "core angiosperms"; This tree shows the detailed relationships within asterids. ...
... Monocots (e.g., rice, corn, orchids, pineapple, and banana) and magnoliids (e.g., avocado and black pepper) are the second and third largest clades, respectively, with the smallest clades Chloranthales and Ceratophyllales exhibiting distinct fruit traits (e.g., drupe in Chloranthaceae; achene in Ceratophyllaceae) (APG IV, 2016). The phylogenetic relationships among these clades remain uncertain; for example, the sister lineage of eudicots has been reported to be either Ceratophyllales, Ceratophyllales+Chloranthales, monocots, or monocots+magnoliids (Soltis and Soltis, 2004;Moore et al., 2007;Moore et al., 2010;Qiu et al., 2010;Zeng et al., 2014;APG IV, 2016;Leebens-Mack et al., 2019;Li et al., 2019b). ...
... Only a few taxa showed inconsistencies of placements among different trees (see more details in Figure S11). Amborellales, Nymphaeales and Austrobaileyales successively form sister clades to mesangiosperms with 100% bootstrap supports (Figure 1), consistent with earlier publications (e.g., APG II, 2003;Soltis and Soltis, 2004;Soltis et al., 2011;Leebens-Mack et al., 2019;Li et al., 2019b) ( Figure S12). The relationships in this study among orders and larger groups are further compared with those in other publications (APG IV, 2016;Leebens-Mack et al., 2019;Li et al., 2019b) ( Figure S12). ...
Article
Full-text available
Fruit functions in seed protection and dispersal and belongs to many dry and fleshy types, yet their evolutionary pattern remains unclear in part due to uncertainties in the phylogenetic relationships among several orders and families. Thus we used nuclear genes of 502 angiosperm species representing 231 families to reconstruct a well supported phylogeny, with resolved relationships for orders and families with previously uncertain placements. Using this phylogeny as a framework, molecular dating supports a Triassic origin of the crown angiosperms, followed by the emergence of most orders in the Jurassic and Cretaceous and their rise to ecological dominance during the Cretaceous Terrestrial Revolution. The robust phylogeny allowed an examination of the evolutionary pattern of fruit and ovary types, revealing a trend of parallel carpel fusions during early diversifications in eudicots, monocots, and magnoliids. Moreover, taxa in the same order or family with the same ovary type can develop either dry or fleshy fruits with strong correlations between specific types of dry and fleshy fruits; such associations of ovary, dry and fleshy fruits define several ovary‐fruit “modules” each found in multiple families. One of the frequent modules has an ovary containing multiple ovules, capsules and berries, and another with an ovary having one or two ovules, achenes (or other single‐seeded dry fruits) and drupes. This new perspective of relationships among fruit types highlights the closeness of specific dry and fleshy fruit types, such as capsule and berry, that develop from the same ovary type and belong to the same module relative to dry and fleshy fruits of other modules (such as achenes and drupes). Further analyses of gene families containing known genes for ovary and fruit development identified phylogenetic nodes with multiple gene duplications, supporting a possible role of whole‐genome duplications, in combination with climate changes and animal behaviors, in angiosperm fruit and ovary diversification.
... Angiosperms are more complex in character than fits the original cotyledon-based classification. Angiosperm species are diverse in their genome size and organization, reproductive morphology, and chemistry, among other things, but they share a suite of synapomorphies [48]. Phylogenetic studies using molecular tools in the 1990s and 2000s (e.g., [49]) largely discredited the old cotyledon-based classification of ancestral affinity and revealed the phylogeny of angiosperm plants, showing a significant departure from cotyledon-based classification. ...
... Phylogenetic studies using molecular tools in the 1990s and 2000s (e.g., [49]) largely discredited the old cotyledon-based classification of ancestral affinity and revealed the phylogeny of angiosperm plants, showing a significant departure from cotyledon-based classification. Specifically, molecular-based inferences of angiosperm phylogeny have established a mono-phylogenetic group called monocots (which comprises all the monocot species from the old classification), another group called eudicots (which include most of the species in the original dicots), and other clades that comprise a large diversity of flowering plants [48]. Our current understanding still holds that monocots may be derived from some primitive eudicots [50], as molecular evidence has shown that monophyletic monocots are related to magnoliids, eudicots, Chloranthaceae, and Ceratophyllaceae [51]. ...
Article
Full-text available
Polycotyly, an interesting characteristic of seed-bearing dicotyledonous plants with more than two cotyledons, represents one of the least explored plant characters for utilization, even though cotyledon number was used to classify flowering plants in 1682. Gymnosperm and angiosperm species are generally known to have one or two cotyledons, but scattered reports exist on irregular cotyledon numbers in many plant species, and little is known about the extent of polycotyly in plant taxa. Here, we attempt to update the documentation of reports on polycotyly in plant species and highlight some lines of research for a better understanding of polycotyly. This effort revealed 342 angiosperm species of 237 genera in 80 (out of 416) families and 160 gymnosperm species of 26 genera in 6 (out of 12) families with reported or cited polycotyly. The most advanced research included the molecular-based inference of the phylogeny of flowering plants, showing a significant departure from the cotyledon-based classification of angiosperm plants, and the application of genetic cotyledon mutants as tools to clone and characterize the genes regulating cotyledon development. However, there were no reports on breeding lines with a 100% frequency of polycotyly. Research is needed to discover plant species with polycotyly and to explore the nature, development, genetics, evolution, and potential use of polycotyly.
... Such a consensus among palaeobotany, plant morphology, and molecular studies seems to indicate that we are approaching the truth. 7 There are increasing independent molecular datings suggesting a Jurassic or even Triassic origin of angiosperms [42][43][44][45][46][47]. A recent study using numerous genes of most taxa in angiosperms and 62 fossil calibrations [46,48] suggested that there is a "Jurassic Gap" for angiosperms. ...
Preprint
Full-text available
Carpels are a reproductive feature restricted to angiosperms, therefore they are a focus of many botanical studies. However, there are controversies over the nature of carpels. A reason underly-ing these controversies is mixing implications given by conflicting interpretations on fossil carpels in early fossil angiosperms from the Cretaceous. These controversies hinder a clear understanding of angiosperm evolution and systematics. A key to these questions is older fossil fruits bearing concerned information. Here we report a new fossil fruit, Xenofructus dabuensis gen. et sp. nov, from the Middle Jurassic of Liaoning, China. Unlike previously reported fruits of early angio-sperms that were interpreted as bearing seeds either on adaxial or abaxial margin by various au-thors, our fossil demonstrates clearly that the seeds in Xenofructus are neither borne on the adaxial nor abaxial margin of the fruit, instead the seeds of Xenofructus are borne on an axis positioned between two margins. This new feature implies that a placenta in carpels is an ovule/seed-bearing axis, a carpel is a composite organ comprising an enclosing leaf (fruit wall) and an axis (placenta). The adaxial or abaxial position of ovules/seeds frequently seen in fossil and extant angiosperms is a consequence derived through long time evolution (coalescence of placenta with either margin of fruits). Carpels can be taken as foliar structures enclosing their associated ovulate branches.
... They inhabit every habitat on earth except for the highest mountain peaks, polar regions, and the deepest oceans. They present as epiphytes, aquatic plants that float and have roots in freshwater and marine habitats, and land plants that vary significantly in size, longevity, and overall forms [2]. The main characteristic of angiosperms is the possession of unisexual or bisexual flowers. ...
... Charles Darwin even referred to these gaps in knowledge as an "abominable mystery" (Takhtajan, 1981;Doyle, 1978;Friis and Endress, 1990;Crane et al., 1995;Soltis et al., 2008;Specht and Bartlett, 2009). After decades of arguments, most paleobotanists now agree that angiosperm characteristics developed completely sometime around the Early Cretaceous (Hickey and Doyle, 1977;Crane et al., 1995;Sun et al., 1998;Soltis and Soltis, 2004;Friis et al., 2006). However, a wide range of indirect evidence suggests that angiosperms existed before the Cretaceous. ...
... There is a striking coincidence between the geological evidence of this continent that disappeared in the late Cretaceous and the distribution, around or inside the limits of Zealandia, of many early branching angiosperm families (Fig.1). The age estimates of the divergence of crown group angiosperms using molecular clock data vary considerably, between 140 and 240 Ma or earlier (Martin et al. 1989a,b;Soltis and Soltis 2004;Moore et al. 2007;Bell et al. 2010;Silvestro et al. 2014;Foster et al. 2017). The reconstruction of the genome of the most recent common angiosperm ancestor suggests an age of 214 Ma for its appearance (Murat et al. 2017). ...
Article
Full-text available
The site and time of origin of angiosperms are still debated. The co-occurrence of many of the early branching lineages of flowering plants in a region somewhere between Australia and the SW Pacific islands suggests a possible Gondwanan origin of angiosperms. The recent recognition of Zealandia, a 94% submerged continent in the east of Australia, could explain the discrepancy between molecular clocks and fossil records about the age of angiosperms, supporting the old Darwinian hypothesis of a “lost continent” to explain the “abominable mystery” regarding the origin and rapid radiation of flowering plants.
... C. praecox belongs to the Laurales, a part of Magnoliidae (magnoliids), also including the orders Piperales, Canellales, and Magnoliales [6,7]. Magnoliidae are the third-largest group of Mesangiospermae, containing approximately 9000 species. ...
Article
Full-text available
Mitochondrial genome sequencing is a valuable tool for investigating mitogenome evolution, species phylogeny, and population genetics. Chimonanthus praecox (L.) Link, also known as “La Mei” in Chinese, is a famous ornamental and medical shrub belonging to the order Laurales of the Calycanthaceae family. Although the nuclear genomes and chloroplast genomes of certain Laurales representatives, such as Lindera glauca, Laurus nobilis, and Piper nigrum, have been sequenced, the mitochondrial genome of Laurales members remains unknown. Here, we reported the first complete mitogenome of C. praecox. The mitogenome was 972,347 bp in length and comprised 60 unique coding genes, including 40 protein-coding genes (PCGs), 17 tRNA genes, and three rRNA genes. The skewness of the PCGs showed that the AT skew (−0.0096233) was negative, while the GC skew (0.031656) was positive, indicating higher contents of T’s and G’s in the mitochondrial genome of C. praecox. The Ka/Ks ratio analysis showed that the Ka/Ks values of most genes were less than one, suggesting that these genes were under purifying selection. Furthermore, there is a substantial abundance of dispersed repeats in C. praecox, constituting 16.98% of the total mitochondrial genome. A total of 731 SSR repeats were identified in the mitogenome, the highest number among the eleven available magnoliids mitogenomes. The mitochondrial phylogenetic analysis based on 29 conserved PCGs placed the C. praecox in Lauraceae, and supported the sister relationship of Laurales with Magnoliales, which was congruent with the nuclear genome evidence. The present study enriches the mitogenome data of C. praecox and promotes further studies on phylogeny and plastid evolution.
... Angiosperms dominate the terrestrial biomes and represent one of the largest eukaryotic radiations (Soltis & Soltis, 2004). Angiosperm clades show asymmetrical patterns of diversification, in space and among taxa --some show high species richness, while others show low branching levels and thus, low species richness (see APG IV, 2016). ...
Article
Full-text available
The disparity in species richness among clades of angiosperms is partly explained by differences in evolutionary and biogeographic processes; however, part of this imbalance remains elusive. The relationships between species diversification and key innovations, as well as the impact of clade age, are also predicted to explain such disparities. This relationship has not been examined using phylogeny-based approaches based on holomorphology, i.e., concatenated morphological and molecular datasets. Despite some large-scale evolutionary studies that have contributed to the knowledge of Brazilian flora, the evolutionary history of angiosperms endemic to Brazilian provinces has not yet been elucidated. Therefore, this study had two principal targets. First, we investigated the species richness and distribution patterns of endemic angiosperm genera. Secondly, we perform a phylogenetic analysis based on holomorphology to examine the relationship between species diversification and putative key innovations (by homology assessments) and the effects of clade age on diversification in species-rich genera. We identified the occurrence of 341 exclusive genera (45% monotypic) in 61 families. Our results indicate a positive correlation between diversification and the number of putative key innovations per order but a negative or non-existent one per family. Furthermore, our findings contradict the idea that clade age is associated with species-rich genera, challenging the notion that clade age is a determining factor in species richness. The results showed that 14 traits are closely associated with diversification, and the confluence between biotic and abiotic factors drove the diversification of species-rich genera in the Atlantic Forest, Caatinga, Cerrado, and Parana provinces.
Chapter
Environmental history and Quaternary botany have long been interconnected by palynological methods. In this context, pollen morphology and some aspects of pollen systematics can inform on adaptations to different environmental conditions in short- and long-term approaches. In this chapter, we show that some topics such as pollen aperture evolution, exine ultrastructure, and pollen wall development are critical for past environments. In addition, the inclusion of some niche construction ideas improves palaeoecological reconstructions for cross-scale interactions such as historical legacies, climate change, and microenvironmental conditions. For these reasons, we provide this framework and promising future lines of research in ecology and archaeology. Finally, the environmental history represents a mosaic of methods in pollen morphology and local and regional reconstructions for past societies.
Chapter
This edited volume is provides an authoritative synthesis of knowledge about the history of life. All the major groups of organisms are treated, by the leading workers in their fields. With sections on: The Importance of Knowing the Tree of Life; The Origin and Radiation of Life on Earth; The Relationships of Green Plants; The Relationships of Fungi; and The Relationships of Animals. This book should prove indispensable for evolutionary biologists, taxonomists, ecologists interested in biodiversity, and as a baseline sourcebook for organismic biologists, botanists, and microbiologists. An essential reference in this fundamental area.
Article
Phylogenetic analyses of angiosperm MADS-box genes suggest that this gene family has undergone multiple duplication events followed by sequence divergence. To determine when such events have taken place and to understand the relationships of particular MADS-box gene lineages, we have identified APETALA1/FRUITFULL-like MADS-box genes from a variety of angiosperm species. Our phylogenetic analyses show two gene clades within the core eudicots, euAP1 (including Arabidopsis APETALA1 and Antirrhinum SQUAMOSA) and euFUL (including Arabidopsis FRUITFULL). Non-core eudicot species have only sequences similar to euFUL genes (FUL-like). The predicted protein products of euFUL and FUL-like genes share a conserved C-terminal motif. In contrast, predicted products of members of the euAP1 gene clade contain a different C terminus that includes an acidic transcription activation domain and a farnesylation signal. Sequence analyses indicate that the euAP1 amino acid motifs may have arisen via a translational frameshift from the euFUL/FUL-like motif. The euAP1 gene clade includes key regulators of floral development that have been implicated in the specification of perianth identity. However, the presence of euAP1 genes only in core eudicots suggests that there may have been changes in mechanisms of floral development that are correlated with the fixation of floral structure seen in this clade.
Article
For almost a century, the formation of endosperm from a second and distinctive fertilization event has been viewed as a unique feature of flowering plants. However, until recently, the evolutionary origin of this unique embryo-nourishing entity remained a mystery. Based upon comparative developmental analysis of reproduction among basal angiosperms and their closest extant relatives, the Gnetales (Ephedra, Gnetum, and Welwitschia), it is possible to construct an explicit hypothesis of the events that led to the evolutionary establishment of double fertilization and endosperm. The formulation of this historical record is derived entirely from and dependent upon the determination of reproductive features that are likely to have characterized the common ancestors of angiosperms and Gnetales. Current evidence is most congruent with the concept that a process of double fertilization first evolved in a common ancestor of the Gnetales and angiosperms. Initially, however, the second fertilization product was diploid and yielded a supernumerary embryo. Subsequent to the divergence of the angiosperm lineage from its closest relatives (which include the Gnetales), modification of the development of the supernumerary embryo (derived from the second fertilization event) led to the establishment of an embryo-nourishing endosperm. Comparative analysis of patterns of embryogeny within Gnetales and angiosperms establishes that embryo development in the ancestors of flowering plants (with a rudimentary process of double fertilization) was ab initio cellular, and not free nuclear, as had previously been assumed. Thus, it is likely that the earliest flowering plants displayed an ab initio cellular pattern of endosperm development, whose expression was inherited from that of the supernumerary embryo of the ancestors of flowering plants.
Article
According to morphologically based classification systems, actinorhizal plants, engaged in nitrogen-fixing symbioses with Frankia bacteria, are considered to be only distantly related. However, recent phylogenetic analyses of seed plants based on chloroplast rbcL gene sequences have suggested closer relationships among actinorhizal plants. A more thorough sampling of chloroplast rbcL gene sequences from actinorhizal plants and their nonsymbiotic close relatives was conducted in an effort to better understand the phylogenetic relationships of these plants, and ultimately, to assess the homology of the different actinorhizal symbioses. Sequence data from 70 taxa were analyzed using parsimony analysis. Strict consensus trees based on 24 equally parsimonious trees revealed evolutionary divergence between groups of actinorhizal species suggesting that not all symbioses are homologous. The arrangement of actinorhizal species, interspersed with nonactinorhizal taxa, is suggestive of multiple origins of the actinorhizal symbiosis. Morphological and anatomical characteristics of nodules from different actinorhizal hosts were mapped onto the rbclL-based consensus tree to further assess homology among rbcL-based actinorhizal groups. The morphological and anatomical features provide additional support for the rbcL-based groupings, and thus, together, suggest that actinorhizal symbioses have originated more than once in evolutionary history.
Article
Floral ontogeny is described and compared in five species and four genera of the hypothetically basal proteaceous subfamily Persoonioideae sensu Johnson and Briggs. The hypotheses surrounding the origin of the peculiar proteaceous flower and homologous structures within the flowers are examined using ontogenetic morphological techniques. Ontogenetic evidence reveals that the proteaceous flower is simple, composed of four tepals, each tepal initiated successively with the lateral tepals being initiated first and second followed by the successive initiation of the sagittal tepals. Each of four stamens is initiated opposite a tepal in a similar sequence to tepal initiation. A single carpel develops terminally from the remaining floral meristem. In taxa of Persoonieae, nectaries are initiated from a broadened receptacle in alternistamenous sites after zonal growth beneath and between the tepals and stamens has begun. The nectaries are interpreted as secondary organs, not reduced homologues of a “lost” petal or stamen series. Developmental variation is present among the examined taxa in several forms including the development of a Vorlaüferspitze (spine) on the upper portion of the tepals, adnation between the anthers and tepals, and formation of the carpel. In Placospermum the early formation of the carpel cleft extends to the floral receptacle and in the other taxa, the carpel cleft is distinctly above the receptacle. Different developmental pathways result in similar mature morphologies of the carpel in Persoonia falcata and Placospermum coriaceum. Bellendena montana is unique relative to the other taxa in having free stamens, a punctate stigma, reduced (not lost) floral bracts, and the floral and bract primordia are initiated from a common meristem. This study provides a foundation for future studies of the developmental basis of floral diversity within Proteaceae.
Article
The morphology and anatomy of 105 flowers representing 13 species and 6 genera of the Canellaceae are summarized. The flowers are borne in axillary or terminal racemes, cymes, or small groups, or solitary, in an axillary or terminal position. The flowers are characterized as follows: bisexual, hypogynous; sepals 3, thick and leathery; petals, 5–12, free or united into tube at base, rather thick, in 1 or 2 whorls and/or spirals; androecium of 6–12 stamens united by their filaments forming a tube, anthers with longitudinal extrorse dehiscence; gynoecium of 2–6 carpels fused by their ventral margins; 2–6 placentae. There are 2 vascular bundles (rarely 3) to each sepal, 3 to each petal (some of the inner petals have only 1), 1 to each stamen and 1 trace to each carpel. The petal and stamen bundles have a common origin. All the data accumulated in this series on the Canellaceae indicate that the correct systematic placement of the Canellaceae is in the woody Ranales, perhaps in a complex with the Myristicaceae.