ArticlePDF Available

Nuclear and Plastid DNA Sequences Confirm the Placement of the Enigmatic Canacomyrica monticola in Myricaceae

Authors:

Abstract and Figures

Phylogenetic analyses of DNA sequences of the plastid rbcL gene were used to obtain a phylogenetic framework for the New Caledonian endemic genus Canacomyrica (monotypic). A further analysis of selected genera within Fagales combining DNA sequences of the rbcL gene, plastid trnL-F region and nuclear ITS region was also performed. In all analyses Canacomyrica fell into a well-supported clade in which it occupied a position as sister to the remaining genera of Myricaceae. A chromosome number of 2n = 16, consistent with Myricaceae, is reported for Canacomyrica for the first time. On the basis of the phylogenetic data and numerous shared morphological features, the original placement of Canacomyrica in Myricaceae is accepted. Canacomyrica is distinguished from other members of the family by the presence of staminodes in the female flower, a six-lobed perianth and a lamellular, laciniate style. The affinities of Myricaceae within Fagales are re-evaluated in the light of the unusual morphological features of Canacomyrica. The significance of staminodes in Canacomyrica is discussed with reference to pollination syndrome and incomplete suppression of male function in Myricaceae. A lectotype is designated for the name Canacomyrica monticola Guillaumin.
Content may be subject to copyright.
INTRODUCTION
The monotypic genus Canacomyrica is an endan-
gered shrub or small tree, endemic to the ultramafic (ser-
pentine) soils of New Caledonia (Herbert, in press).
Since its description by Guillaumin (1941; see also
Guillaumin, 1939), who placed it in Myricaceae, there
has been confusion over its morphology and doubt about
its affinity. Prior to this study, little research had been
carried out into the biology of this species and basic
information, such as chromosome number, was un-
known.
Most modern taxonomists consider Myricaceae to
comprise three genera: Myrica L. (ditypic); Comptonia
L’Hérit. ex Aiton (monotypic); and Morella Lour. (c. 50
species) (Wilbur, 1994; Polhill & Verdcourt, 2000;
Herbert, 2005a). All three genera share features such as a
simple bifid style, no perianth and an orthotropous ovule.
In his original description, Guillaumin (1941) listed a
number of anomalous features that distinguish
Canacomyrica monticola from all other Myricaceae
species: androdioecious breeding system (male flowers
and hermaphrodite flowers on separate plants); lamellu-
lar, laciniate style; fleshy perianth; inferior ovary; and
ovule hanging from a long funicle with an inferior
micropyle. He later modified his description, stating that
the inflorescences were co-sexual and the ovule anat-
ropous (Guillaumin, 1948).
In a study of the floral morphology of Canaco-
myrica, Leroy (1949) showed that Guillaumin had incor-
rectly described several characters, stating that the flow-
ers were unisexual with the female flowers bearing ster-
ile anthers and the ovule was erect, sessile, and
orthotropous with a superior micropyle. On the basis of
this re-evaluation of morphology, Leroy (1949) accepted
the placement of Canacomyrica in Myricaceae. How-
ever, in recognition of the remaining unusual features he
placed it in the new subfamily Canacomyricoideae. The
work of Leroy (1949) appears to have been largely over-
looked, resulting in continued misinterpretation of the
morphology, particularly regarding the true nature of the
flowers (e.g., Macdonald, 1977, 1989; Cronquist, 1981;
but see Kubitzki, 1993b).
Several authors (e.g., Elias, 1971; Raven & Axelrod,
1974) have questioned the inclusion of Canacomyrica in
Myricaceae based on its distinctive morphology and dis-
tribution (Myricaceae as previously circumscribed does
not extend into Australasia). Thorne (1973) placed it in
349
Herbert & al. • Systematic position of Canacomyrica55 (2) • May 2006: 349–357
Nuclear and plastid DNA sequences confirm the placement of the enigmatic
Canacomyrica monticola in Myricaceae
Jane Herbert1, Mark W. Chase2, Michael Möller3& Richard J. Abbott4
1School of Biology, Sir Harold Mitchell Building, University of St. Andrews, KY16 9TH, U.K.
janeherbert2@yahoo.co.uk (author for correspondence).
2Molecular Systematics Section, Jodrell Laboratory, Royal Botanic Gardens, Kew, Richmond, Surrey, TW9
3DS, U.K.
3Royal Botanic Garden, Edinburgh, 20A Inverleith Row, Edinburgh, EH3 5LR, U.K.
4School of Biology, Sir Harold Mitchell Building, University of St. Andrews, KY16 9TH, U.K.
Phylogenetic analyses of DNA sequences of the plastid rbcL gene were used to obtain a phylogenetic frame-
work for the New Caledonian endemic genus Canacomyrica (monotypic). A further analysis of selected gen-
era within Fagales combining DNA sequences of the rbcL gene, plastid trnL-F region and nuclear ITS region
was also performed. In all analyses Canacomyrica fell into a well-supported clade in which it occupied a posi-
tion as sister to the remaining genera of Myricaceae. A chromosome number of 2n = 16, consistent with
Myricaceae, is reported for Canacomyrica for the first time. On the basis of the phylogenetic data and numer-
ous shared morphological features, the original placement of Canacomyrica in Myricaceae is accepted.
Canacomyrica is distinguished from other members of the family by the presence of staminodes in the female
flower, a six-lobed perianth and a lamellular, laciniate style. The affinities of Myricaceae within Fagales are
re-evaluated in the light of the unusual morphological features of Canacomyrica. The significance of stamin-
odes in Canacomyrica is discussed with reference to pollination syndrome and incomplete suppression of male
function in Myricaceae. A lectotype is designated for the name Canacomyrica monticola Guillaumin.
KEYWORDS:
Canacomyrica, dioecy, Fagales, Myricaceae, New Caledonia, rbcL sequences, staminodes.
his list of Taxa Incertae Sedis, later including it in Myr-
icaceae without comment (Thorne, 1992a, b, 2000).
Others have accepted the original placement of Canaco-
myrica (Cronquist, 1981; Kubitzki, 1993b; Takhtajan,
1997). The recent placement of Canacomyrica in a sepa-
rate family (Canacomyricaceae; Doweld, 2000) added
little insight into the problem of the systematic position
of this taxon.
To date Canacomyrica has not been included in a
phylogenetic study, but it has been suggested that its
anomalous features may be primitive in Myricaceae
(Macdonald, 1989; Carlquist, 2002). If this is the case,
then a better understanding of the relationship of
Canacomyrica to Myricaceae may shed light on the
affinities of the family. Based on molecular data,
Myricaceae are placed in Fagales (sensu APG, 1998,
2003) in the nitrogen-fixing clade, which are nested
within the eurosid I clade (fabids) of the eudicots (Chase
& al., 1993; Soltis & al., 1995, 2000; Swensen, 1996 ;
APG, 1998, 2003). The expanded concept of Fagales
comprises Nothofagaceae (sister to the rest of the order),
Fagaceae, Ticodendraceae, Betulaceae, Casuarinaceae,
Juglandaceae (including Rhoipteleaceae) and Myrica-
ceae. With one exception (Maggia & Bousquet, 1994),
all recent molecular studies of Fagales have reported
Fagaceae (s.s.) as sister to the remaining five families
(the core “higher” hamamelids of Manos & Steele,
1997). Manos & Steele (1997) showed Myricaceae to be
sister to Betulaceae, Ticodendraceae and Casuarinaceae.
Li & al. (2002) placed it sister to all other core “higher”
hamamelids. In an analysis that sampled widely from
throughout Fagales, Li & al. (2004) placed Myricaceae
as sister to only Juglandaceae.
Many studies have employed phylogenetic analysis
of DNA sequences of the plastid rbcL gene (ribulose bis-
phosphate carboxylase/oxygenase, large subunit) to as-
sess the affinities of taxonomically enigmatic angio-
sperm taxa (e.g., Fay & al., 1997; Chase & al., 2002;
Längström & Chase, 2002; Sosa & Chase, 2003; Whit-
lock & al., 2003). A wide range of rbcL sequences are
now available (Savolainen & al., 2000), permitting
analyses across a diverse array of flowering plants.
In this study we used rbcL sequences to investigate
the phylogenetic relationships of Canacomyrica. We
chose to combine rbcL sequences with two more variable
regions: the plastid trnL-F region (consisting of the trnL
intron and the trnL-trnF intergenic spacer) and the nucle-
ar ITS region (consisting of the internal transcribed spac-
er region of the 18S-5.8S-26S nuclear ribosomal cistron).
Many studies have shown the trnL-F region to be useful
in resolving relationships above family level (e.g.,
Richardson & al., 2000; Li & al., 2002). Although ITS
has been widely used for resolving sub-familial level
relationships (Alvarez & al., 2003), it is seldom used at
higher taxonomic levels due to alignment difficulties.
The availability of ITS sequence data for Fagales species
motivated its use in this study as an additional source of
potentially informative characters from a different
genome.
MATERIALS AND METHODS
Plant material. —
Silica gel-dried leaf material
(Chase & Hills, 1991) was obtained for C. monticola,
Myrica hartwegii Watson, Morella cordifolia (L.)
Killick, and Comptonia peregrina (L.) L’Hérit. Voucher
information and GenBank accession numbers for plant
material used in this study are listed in the Appendix.
DNA was extracted using a 2X CTAB method adapted
from (Doyle & Doyle, 1990), modified to include a wash
with ammonium acetate (7.5 M NH
4
AC; Weising & al.,
1995) to remove impurities co-precipitated with the
DNA.
PCR amplification and sequencing.
Ampli-
fication of the rbcL gene was carried out in two overlap-
ping fragments using the forward primers 1F and 636F
and the reverse primers 724R and 1460R (Fay & al.,
1997). Amplification of the ITS1-5.8S-ITS2 region was
carried out using the forward primer ITS5 and the
reverse primer ITS4 (White & al., 1990). Amplification
of the trnL-F region was carried out using the forward
primer c and the reverse primer f (Taberlet & al., 1991).
PCR reactions of 50 !l contained: 2 !l DNA tem-
plate, 0.2 mM of each dNTP, 0.3 !M of each primer, 2
units Taq polymerase (Bioline, London, U.K.), 2 mM
MgCl
2
, and 5 !l reaction buffer (160 mM (NH
4
)
2
SO
4
,
670 mM Tris HCl, 0.1% Tween 20, pH 8.8). The follow-
ing PCR profile was used for both rbcL and trnL-F: 1
cycle at 94ºC for 4 min; 30 cycles at 94ºC for 45 s, 55ºC
for 45 s and 72ºC for 3 min; 1 cycle at 72ºC for 10 min.
The PCR profile used for the ITS region was as follows:
1 cycle at 94ºC for 3 min; 30 cycles at 94ºC for 1 min,
55ºC for 1 min and 72ºC for 1 min 30 s; 1 cycle at 72ºC
for 5 min. The resulting PCR products were purified
using QIAquick purification kits (QIAGEN Ltd., Craw-
ley, U.K.) according to the manufacturers instructions.
Sequencing reactions of 20 !l contained the CEQ
DTCS Quick Start Kit (Beckman Coulter Ltd, High
Wycombe, U.K.) and the following primers: the same
primers as in the PCR for the rbcL gene; primers ITS2,
ITS3, ITS4 and ITS5 (White & al., 1990) for the ITS
region; and primers c, d, e and f (Taberlet & al., 1991) for
the trnL-F region. Reactions were performed using the
following PCR profile: 25 cycles at 96ºC for 10 s, 50ºC
for 5 s, 60ºC for 4 min. The sequence reaction products
were cleaned according to manufacturer’s instructions
before being run on a CEQ
8000 Genetic Analysis
Herbert & al. • Systematic position of Canacomyrica 55 (2) • May 2006: 349–357
350
System (Beckman Coulter Ltd, High Wycombe, U.K.).
Forward and reverse sequences were manually assem-
bled in Chromas version 2.12 (Technelysium Pty. Ltd.,
Helensvale, Australia).
Alignment and analyses. —
A preliminary study
of over 500 eudicot taxa using the Savolainen & al.
(2000) matrix was first performed to establish the broad-
scale relationships of Canacomyrica (results not shown).
Based on these results an rbcL dataset of 35 taxa (the
“narrow rbcL dataset) was assembled from the fabid
(eurosid I) clade. Fabales taxa were chosen as outgroup.
New sequences included in the narrow rbcL dataset are
listed in Appendix; the remainder were previously pub-
lished (Savolainen & al., 2000).
A combined dataset of rbcL, ITS and trnL-F se-
quences focusing on 11 Fagales taxa (the “combined
Fagales” dataset) was assembled using new sequences
produced in this study and previously published
sequences (Appendix). Following the approach of
authors such as Wiens (1998) and Reeves & al. (2001) it
was considered appropriate to combine these datasets
after separate analyses showed there to be no strongly
supported (>85%) incongruent clades among the individ-
ual datasets (results not shown). Fagaceae taxa were used
as outgroup in this dataset. Alignment of all sequences
was carried out by eye following the guidelines in
Kelchner (2000). Alignment of rbcL sequences required
no gaps. Alignment of the trnL-F sequences required the
insertion of gaps. Alignment of the ITS sequences
required the insertion of gaps and the exclusion of some
regions where alignment was ambiguous. An aligned
matrix may be obtained on request from MWC (m.chase-
@kew.org.)
Phylogenetic analyses were performed using PAUP*
Herbert & al. • Systematic position of Canacomyrica55 (2) • May 2006: 349–357
351
Table 1. Comparison of morphological characters of Fagales (excluding Nothofagus) (Goldberg, 1986; Wilson &
Johnson, 1989; Kubitzki, 1993a, b; Zheng-yi & al., 1999).
Character Fagaceae Myricaceae Canacomyrica Juglandaceae Betulaceae Casuarinaceae
N-fixing root Absent Present Unknown Absent Present Present
nodules
Stipules Present Absent or Absent Absent (present Present Absent
foliaceous in Rhoiptelea)
Style Linear, short Linear, Lamellular, Plumose or Linear, Linear,
elongate laciniate lamellular elongate elongate
Inflorescence Usually lax Erect, spicate Erect, spicate Lax Lax Erect,
structure spicate
Ovule Anatropous Orthotropous Orthotropous Orthotropous Anatropous Orthotropous
(hemitropous
in Rhoiptelea)
Fruit size and Nut (>1 cm Drupe (to 3 cm Drupe (to 5 mm Winged nutlet (to 9 Winged nutlet (to Samara (2–12
structure long) borne diam.) usually diam.) with a mm diam.) or large 5 mm diam.) or mm long) en-
in cupule covered with smooth fleshy nut (to 6 cm diam.) nut (to 2 cm closed in woody
fleshy papillae pericarp with much lobed diam.) enclosed bracteoles
cotyledons in enlarged bracts
Leaf form Simple Simple or Simple Pinnately or impari- Simple Simple, reduced
pinnatifid pinnately compound
Tepals/perianth 6 tepals Absent 6 lobed perianth 0–4 tepals (0–)1–4(–6) tepals Absent
Integuments 2 (1) 1 2 1 (2) 1, 2 2
Chromosome nr. x = 12, 11, 22 x = 8 x = 8 x = 16 x = 8, 14 x = 8, 14
Fig. 1. One of the 12 most parsimonious trees produced
from analysis of the narrow rbcL dataset. Branch lengths
are shown above the branches (DELTRAN optimization),
bootstrap percentages "50 are shown in bold below the
branches. Asterisks indicate groups not present in the
strict consensus tree.
*
*
11
96
89
69
97
74
53
65
73
69
81
94
64
57
99
50
58
64
97
96
95
92
73
68
24
72
14
3
7
7
8
7
721
28
21
28
35
48
27
11
11 20
34
721
44
19
8
42
7
7
8
10
11 3
5
2
7
1
81
7
5
5
7
27
4
11
11
5
7
6
8
10
13
15
20
9
20 310
9
7
3
Morella
Myrica
Comptonia
Canacomyrica
Juglans
Carya
Rhoiptelea
Alnus
Carpinus
Corylus
Betula
Casuarina
Allocasuarina
Ceuthostoma
Gymnostoma
Ticodendron
Fagus
Quercus
Trigonobalanus
Chrysolepis
Nothofagus
Celtis
Humulus
Cecropia
Ficus
Ulmus
Elaeagnus
Rhamnus
Dryas
Rosa
Spiraea
Cucurbita
Datisca
Pisum
Albizia
Myricaceae
Juglandaceae
Betulaceae
Casuarinaceae
Ticodendraceae
Fagaceae
Nothofagaceae
Celtidaceae
Cannabaceae
Urticaceae
Moraceae
Ulmaceae
Elaeagnaceae
Rhamnaceae
Rosaceae
Cucurbitaceae
Datiscaceae
Fabaceae (outgroup)
Fagales
Rosales
Cucurbitales
100
100
100
Version 4.0b10 (Swofford, 1998). Maximum parsimony
analysis was carried out with tree-bisection-reconnection
(TBR) and saving multiple trees (MulTrees). Maximum
parsimony trees were obtained using the heuristic search
option for the narrow rbcL dataset (Fitch parsimony).
The small size of the combined Fagales dataset made it
possible to carry out a branch and bound search. Branch
lengths were calculated using the delayed transformation
(DELTRAN) optimization.
Support for the clades obtained was assessed using
bootstrap analysis (Felsenstein, 1985) performed using
the heuristic search option and 1000 replicates for the
narrow rbcL dataset and using the branch and bound
search option and 100 replicates for the combined
Fagales dataset. The following categories were used to
describe bootstrap percentage (BP) results: 50–74, weak
support; 75–84, moderate support; 85–100, strong sup-
port.
Chromosome count. —
Root tip preparations
were made from living seedlings of Canacomyrica
housed in the research collection at the Royal Botanic
Garden, Edinburgh (RBGE). Root tips were treated, fol-
lowing the protocol of Jong & Möller (2000), in
colchicine (0.05%) for 4 hours at room temperature, or in
saturated aqueous 1-bromonapthalene for 2 to 8 hours at
room temperature, or in 0.002 M 8-hydroxyquinoline for
4 to 8 hours at 12ºC. They were variously stained in
Feulgen’s reagent (Fox, 1969), lacto-propionic orcein or
aceto carmine, but no method gave satisfactorily intense
staining. After the root tips had been softened in an en-
zyme solution of 5% pectinase and 5% cellulase at 35ºC
for 20 min and mounted on slides, they were viewed
using phase-contrast (Zeiss, Axiophot) at 40#1.6 or 100#
magnification. Images were captured digitally.
RESULTS
Sequencing. —
The narrow rbcL matrix contained
35 taxa and 1343 characters, of which 206 were poten-
tially parsimony informative. Maximum parsimony
analysis of the narrow rbcL dataset produced 12 most
parsimonious trees (tree length 869 steps, CI = 0.51, RI
= 0.58). One of the 12 trees is shown in Fig. 1 with
branch lengths above the branches (DELTRAN opti-
mization) and bootstrap percentages (BP) equal to or
greater than 50 shown below the branches.
In all trees Canacomyrica is sister to the other
Myricaceae taxa in a strongly supported clade (92 BP).
In Fig. 1 (asterisks indicate nodes not present in all trees)
and in the strict consensus tree (not shown), Myricaceae
form a polytomy with two other clades, the first com-
prising Juglandaceae (including Rhoiptelea) and the sec-
ond Betulaceae, Casuarinaceae and Ticodendraceae.
The combined Fagales matrix contained 11 taxa and
2728 characters, of which 210 were potentially parsimo-
ny informative. The rbcL data contributed 51 potentially
parsimony informative characters (total 1286 charac-
ters), the trnL-F data contributed 75 potentially parsimo-
ny informative characters (total 1014 characters) and the
ITS data contributed 84 potentially parsimony informa-
tive characters (total 428 characters). In our combined
analysis 40% of informative characters were from the
ITS region. Of a total length of 664 aligned characters
(including gaps) sequenced for the region, we were
unable to use 236 characters (36%; 183 at the start of
ITS1, 53 at the start of ITS2) due to alignment difficul-
ties. Although the ITS region is seldom used above the
genus level (but see Loockerman & al., 2003; Muellner
& al., 2003) our results, in accordance with the findings
of Coleman (2003), suggest that this region has utility at
higher taxonomic levels within Fagales. Maximum parsi-
mony analysis of the combined Fagales dataset produced
a single most parsimonious tree (tree length 754, CI =
0.82, RI = 0.67). Figure 2 shows the single most parsi-
monious tree with branch lengths (DELTRAN optimiza-
tion) above the branches, and bootstrap percentages
equal to or greater than 50 shown below the branches.
Herbert & al. • Systematic position of Canacomyrica 55 (2) • May 2006: 349–357
352
Fig. 2. The single most parsimonious tree produced from
analysis of the combined Fagales dataset. Branch
lengths are shown above the branches (DELTRAN opti-
mization), bootstrap percentages "50 are shown in bold
below the branches. Asterisks indicate groups with boot-
strap percentage of less than 50.
Myricaceae
Juglandaceae
Betulaceae
Casuarinaceae
Fagaceae (outgroup)
Fagus
Quercus
Casuarina
Carpinus
Alnus
Rhoiptelea
Juglans
Canacomyrica
Comptonia
Myrica
Morella
2
91
15
12
8
29
100
41
48
100
13
33
47
31
100
41
56
33
18
*157
36
97
71
71
10
99
*
In the single most-parsimonious tree, Canacomyrica
is sister to a strongly supported clade (100 BP) compris-
ing the rest of Myricaceae. Monophyly of Myricaceae s.l.
(including Canacomyrica) is also strongly supported
(100 BP). Myricaceae s.l. are sister to Juglandaceae, but
this relationship received bootstrap support of less than
50%.
Chromosome count.
The root material avail-
able for this study was of inferior quality due to the dif-
ficulty of cultivation for this rare material, which subse-
quently died. Under phase-contrast, chromosomes were
clearly discernible in a small number of metaphase or
pro-metaphase cells. Figure 3 shows a total of 16 chro-
mosomes (2n= 16) counted for Canacomyrica montico-
la.
DISCUSSION
Systematic implications.
In all our analyses
Canacomyrica formed a strongly supported monophylet-
ic group with Myricaceae s.s. The combined use of DNA
sequences from three different regions gave greater boot-
strap support for this group than the analysis of rbcL
alone (100 BP versus 92 BP). There was insufficient
variability within the DNA markers used to resolve the
sister group of Myricaceae. In the single most parsimo-
nious tree produced by the combined analysis,
Juglandaceae were sister to Myricaceae s.l. (including
Canacomyrica), although this relationship received boot-
strap support of less than 50%. Li & al. (2004) recovered
a sister group relationship between Myricaceae (exclud-
ing Canacomyrica) and Juglandaceae with weak boot-
strap support (66 BP) and high Bayesian posterior prob-
ability (0.95 PP). A morphological phylogenetic study by
Hufford (1992) also supported a sister group relationship
between Myricaceae and Juglandaceae (but not
Rhoiptelea), and a close relationship between these fam-
ilies has been recognized previously by several authors
(e.g., Hjelmqvist, 1948; Leroy, 1949; Melchior, 1964;
Cronquist, 1981; Macdonald, 1989; Thorne, 1992a, b,
2000; Kubitzki, 1993b).
The principal differences between Myricaceae s.l.
and Juglandaceae are leaf form, inflorescence structure,
fruit morphology and size (Table 1). The two families
share the synapomorphies of chains of cuboidal crystal-
containing cells in the wood (Carlquist, 2002) and aro-
matic resin glands. The lamellular, laciniate style and
fleshy perianth that differentiate Canacomyrica from the
rest of Myricaceae are characters that can be found in
Juglandaceae (Table 1). The ovule in Canacomyrica is
bitegmic (Herbert, 2005b; A. Doweld, pers. comm.), a
feature that is shared with Rhoiptelea and may be sym-
plesiomorphic in Fagales.
The sister relationship of Canacomyrica to the other
Myricaceae genera, indicated in this study, is consistent
with the original proposal to include it in the family
(Guillaumin, 1939, 1941). Myricaceae are generally de-
fined by simple, entire (pinnatifid in Comptonia) leaves
with resinous (usually aromatic) “balloon” glands
(Chourey, 1974), short, erect inflorescences, solitary
ovules and small drupaceous fruits with small (typically
<10 mm diameter) seeds. Canacomyrica shares all these
features with other genera of Myricaceae. Furthermore,
studies have shown that Canacomyrica has pollen with a
Myrica-type aperture (Sundberg, 1985) and myricaceous
wood anatomy (Carlquist, 2002). The chromosome count
reported here for Canacomyrica (2n = 16) is consistent
with counts a basic chromosome number of x= 8 report-
ed for other members of Myricaceae, e.g., Myrica 2n=
16, 48, 96 (Löve, 1980, 1982; Morawetz & Samuel,
1989; Al-Bermani & al., 1993), Comptonia 2n = 32
(Löve, 1982), and Morella 2n= 16 (Oginuma & Tanaka,
1987).
The sister relationship between Canacomyrica and
Myricaceae s.s. is also consistent with the opinion of
Leroy (1949), who argued for subfamily status for
Canacomyrica. The presence of a long terminal branch
for Canacomyrica (Fig. 2) indicates significant diver-
gence between the two lineages as might be expected
Herbert & al. • Systematic position of Canacomyrica55 (2) • May 2006: 349–357
353
Fig. 3. Chromosomes of
Canacomyrica
monticola. Photo-
graphs A, B, and C are of the same cell taken at different
depths of focus to capture all the chromosomes present:
(A) clearly shows two chromosomes, these can be seen
less clearly in (B) overlaid by a third chromosome; (B)
clearly shows three chromosomes; (C) clearly shows ele-
ven chromosomes and less clearly the position of the
cluster of three shown in (B). Total = 16 chromosomes.
(D) Drawing of the metaphase spread illustrated in A to C,
showing the relative position of the 16 chromosomes.
given its unique morphological features. On the basis of
the strongly supported monophyly of Myricaceae s.l. in
our analyses and the numerous morphological features
shared between Canacomyrica and Myricaceae,
described above, we accept the original placement of
Canacomyrica in Myricaceae.
Dioecy in Canacomyrica. —
Many authors (e.g.,
Macdonald, 1977; 1989; Cronquist, 1981; Zomlefer,
1994) have continued to follow Guillaumin’s (1941)
description of Canacomyrica in which he stated that
there are both hermaphrodite and male flowers, despite
later studies showing that the flowers are unisexual with
female flowers bearing six sterile anthers or staminodes
(Kubitzki, 1993b; Leroy, 1949). Based on observations
in the field and examination of a total of 50 flowers from
16 individual plants (Herbert, 2005b), we confirm that
male flowers and functionally female flowers are found
on separate plants in Canacomyrica. The confirmation
that Canacomyrica is dioecious is consistent with its
placement in Myricaceae (at least 75% of species are
dioecious), but staminodes are unknown in the rest of the
family.
In Canacomyrica, the catkin-like inflorescence, the
lax attachment of the anthers and the laciniate stigma
indicate that the plant is wind-pollinated. It could be
argued, however, that the staminodes in Canacomyrica
function as attractants for generalist pollinators.
Although Myricaceae are generally considered to be an
anemophilous family, in Hawai’i it has been observed
that introduced honey bees (Apis mellifera) visit the
flowers of Morella faya (Aiton) Wilbur (Vitousek &
Walker, 1989). Elsewhere, among the predominantly
wind-pollinated Fagales, staminodes have been recorded
in Lithocarpus Blume, Triganobalanus Forman, Quercus
subg. Cyclobalanopsis (Oersted) C. K. Schneider and all
Castanoideae (Fagaceae; Kaul, 1985; Kubitzki, 1993a).
In Lithocarpus, the presence of staminodes is acknowl-
edged to be associated with entomophily (Kaul, 1985)
and in Juglandaceae, insect pollination has been
observed in Platycarya (Endress, 1986). This raises the
possibility that Canacomyrica may be insect pollinated
and thus indicates an entomophilous ancestry for
Myricaceae. Such a suggestion is consistent with the
notion that wind-pollination in Fagales may have multi-
ple origins from insect-pollinated ancestral lineages
(Manos & al., 2001).
Alternatively, in the case of species in which monoe-
cy or dioecy are common, the occurrence of staminodes
may be due to incomplete suppression of male function
(Walker-Larsen & Harder, 2000). Incomplete suppres-
sion of male function is known in other Myricaceae, for
example: within a single population of Myrica gale L.
both dioecy and monoecy can occur, and sex expression
within a single stem has been reported to be unstable
from year to year (Lloyd, 1981); androgynous or mixed
inflorescences have been observed in several species of
Myrica and Morella; and, perhaps most significantly of
all, occasional functional stamens are often observed on
the fruit wall of Morella species (Macdonald, 1989; J.
Herbert, pers. obs.). In Canacomyrica the presence of
staminodes could be viewed as an extreme example of
the incomplete suppression of male function found
throughout Myricaceae.
Lectotypification. —
In the original description
of this species, Guillaumin (1941) listed several speci-
mens but the type was not indicated. Therefore, a lecto-
type is designated below. In addition, a detailed descrip-
tion of the species is given.
Canacomyrica monticola Guillaumin, Bull. Soc. Bot.
France 87: 300. 1941. – Lectotype (designated here):
NEW CALEDONIA, Noumea, Balansa 564 (P!).
Dioecious shrub or small tree to 7 m in height.
Unscented resinous glands present on leaves and inflo-
rescences. Leaves evergreen, typically to 14 cm long,
coriaceous, oblanceolate, serrate at apex (margin deeply
serrate in leaves of young plants), dark green above,
whitish below, trichomes absent; stomata anomocytic
(confirmed for this species on the basis of unpublished
data R.S. Hill and R. Paull, University of Adelaide,
pers. comm.); stipules absent. Inflorescence an erect
spike; flowers spirally arranged and widely spaced on
floral stem, trichomes present on floral stem. Male flow-
ers surrounded by 3 bracts; 6 stamens, filaments free,
anthers dorsifixed; vestigial style and vestigial perianth
present. Female flowers surrounded by 3 bracts; 6 sta-
minodes present; style pink-red, bifid, lamellular with
laciniate margin; 6-lobed fleshy perianth present; ovary
sessile, 1-locular; ovule 1, erect, orthotropous, bitegmic.
Fruit a drupe to 1 cm diam., pericarp fleshy, white or
pink becoming black at maturity, endocarp hard.
Conclusion. —
In conclusion, we accept the place-
ment of Canacomyrica in Myricaceae on the basis of the
molecular data presented here. This decision is support-
ed by the confirmation of a dioecious breeding system
and a chromosome count of 2n= 16. The data are
ambiguous with regards to the sister group of
Myricaceae, but the clarification of the placement of
Canacomyrica enhances our understanding of pollina-
tion syndrome within the family, and further highlights
the morphological characters shared by Myricaceae and
Juglandaceae.
ACKNOWLEDGEMENTS
The authors thank L. Ronse Decraene (RBG, Edinburgh) for
Herbert & al. • Systematic position of Canacomyrica 55 (2) • May 2006: 349–357
354
comments on an earlier draft of this paper; S. Schuster (RBG,
Kew) for additional sequencing; P. Hollingsworth, M. Gardner
and A. Ponge (RBG, Edinburgh) for their help during fieldwork;
Administration, Province Sud, New Caledonia for permission to
collect; J. Dahl Regional Parks Botanic Garden, Berkeley (U.S.
A.), for supplying plant material; Administration of the Western
Cape Nature Conservation Board and the Cape Peninsula National
Park, South Africa, for permission to collect; and T. Meagher
(University of St. Andrews, U.K.) for providing plant material.
The first author was funded by a Natural Environment Research
Council studentship (ref.: NER/S/A/2000/03638) with additional
funding for field work from the Merlin Trust (Kettering, U.K.), the
Russell Trust (University of St. Andrews, U.K.) and the Davis
Expedition Fund (University of Edinburgh, U.K.).
LITERATURE CITED
Al-Bermani, K.-A. K. A., Al-Shammary, K. I. A., Bailey, J.
P. & Gornall, R. J. 1993. Contributions to a cytological
catalogue of the British and Irish flora, 3. Watsonia 19:
269–271.
Alvarez, I. & Wendel, J. F. 2003. Ribosomal ITS sequences
and plant phylogenetic inference. Molec. Phylog. Evol. 29:
417–434.
Angiosperm Phylogeny Group. 1998. An ordinal classifica-
tion for the families of flowering plants. Ann. Missouri
Bot. Gard. 85: 531–553.
Angiosperm Phylogeny Group. 2003. An update of the
Angiosperm Phylogeny Group classification for the orders
and families of flowering plants: APG II. Bot. J. Linn. Soc.
141: 399–436.
Bousquet, J., Strauss, S. H. & Li, P. 1992. Complete congru-
ence between morphological and rbcL-based molecular
phylogenies in birches and related species (Betulaceae).
Molec. Biol. Evol. 9: 1076–1088.
Carlquist, S. 2002. Wood and bark anatomy of Myricaceae:
relationships, generic definitions, and ecological interpre-
tations. Aliso 21: 7–29.
Chase, M. W. & Hills, H. H. 1991. Silica-gel — an ideal mate-
rial for field preservation of leaf samples for DNA studies.
Taxon 40: 215–220.
Chase, M. W., Soltis, D. E., Olmstead, R. G., Morgan, D.,
Les, D. H., Mishler, B. D., Duvall, M. R., Price, R. A.,
Hills, H. G., Qiu, Y. L., Kron, K. A., Rettig, J. H., Conti,
E., Palmer, J. D., Manhart, J. R., Sytsma, K. J.,
Michaels, H. J., Kress, W. J., Karol, K. G., Clark, W. D.,
Hedren, M., Gaut, B. S., Jansen, R. K., Kim, K. J.,
Wimpee, C. F., Smith, J. F., Furnier, G. R., Strauss, S.
H., Xiang, Q. Y., Plunkett, G. M., Soltis, P. S., Swensen,
S. M., Williams, S. E., Gadek, P. A., Quinn, C. J.,
Eguiarte, L. E., Golenberg, E., Learn, G. H., Graham,
S. W., Barrett, S. C. H., Dayanandan, S. & Albert, V. A.
1993. Phylogenetics of seed plants — an analysis of
nucleotide sequences from the plastid gene rbcL. Ann.
Missouri Bot. Gard. 80: 528–580.
Chase, M. W., Zmarzty, S., Lledó, M. D., Wurdack, K. J.,
Swensen, S. M. & Fay, M. F. 2002. When in doubt, put it
in Flacourtiaceae: a molecular phylogenetic analysis based
on plastic rbcL DNA sequences. Kew Bull. 57: 141–181.
Chen, Z.-D., Wang, X.-Q., Sun, H.-Y., Han, Y., Zhang, Z.-
X., Zou, Y.-P. & Lu, A.-M. 1998. Systematic position of
the Rhoipteleaceae: evidence from nucleotide sequences
of rbcL gene. Acta Phytotax. Sin. 36: 1–7.
Chourey, M. S. 1974. A study of the Myricaceae from Eocene
sediments of southeastern North America. Palaeonto-
graphica B 146: 88–153.
Coleman, A. W. 2003. ITS2 is a double-edged tool for eukary-
ote evolutionary comparisons. Trends Genet. 19: 370–375.
Cronquist, A. 1981. An Integrated System of Classification of
Flowering Plants. Columbia Univ. Press, New York.
Denk, T., Grimm, G., Stogerer, K., Langer, M. & Hemleben,
V. 2002. The evolutionary history of Fagus in western
Eurasia: evidence from genes, morphology and the fossil
record. Pl. Syst. Evol. 232: 213–236.
Doweld, A. B. 2000. Validation of some suprageneric taxa in
dicotyledons (Rosopsida, Seu Magnoliopsida). Bull.
Mosc. Soc. Nat. 105: 59–60.
Doyle, J. J. & Doyle, J. L. 1990. Isolation of plant DNA from
fresh tissue. Focus 12: 13–15.
Elias, T. S. 1971. The genera of Myricaceae in the southeast-
ern United States. J. Arnold Arbor. 52: 305–318.
Endress, P. K. 1986. An entomophily syndrome in
Juglandaceae: Platycarya strobilacea. Veröff. Geobot.
Inst. Rübel, Zürich 87: 100–111.
Fay, M. F., Swensen, S. M. & Chase, M. W. 1997. Taxonomic
affinities of Medusagyne oppositifolia (Medusagynaceae).
Kew Bull. 52: 111–120.
Felsenstein, J. 1985. Confidence limits on phylogenies — an
approach using the bootstrap. Evolution 39: 783–791.
Fox, D. P. 1969. Some characteristics of the cold hydrolysis
technique for staining plant tissues by the Feulgen reac-
tion. J. Hist. Cytochem. 17: 226.
Fujii, N., Tomaru, N., Okuyama, K., Koike, T., Mikami, T.
& Ueda, K. 2002. Chloroplast DNA phylogeography of
Fagus crenata (Fagaceae) in Japan. Pl. Syst. Evol. 232:
21–33.
Gielly, L. & Taberlet, P. 1994. The use of chloroplast DNA to
resolve plant phylogenies noncoding versus rbcL
sequences. Molec. Biol. Evol. 11: 769–777.
Goldberg, A. 1986. Classification, evolution, and phylogeny
of the families of dicotyledons. Smithsonian Contr. Bot.
58: 1–314.
Guillaumin, A. 1939. La présence inattendue d’une Myricacée
en Nouvelle-Calédonie. Compt. Rend. Hebd. Séances
Acad. Sci. 209: 233–234.
Guillaumin, A. 1941. Matériaux pour la flore de la Nouvelle-
Calédonie. LVII. La présence d’une Myricacée. Bull. Soc.
Bot. France 87: 299–300.
Guillaumin, A. 1948. Flore Analytique et Synoptique de la
Nouvelle-Calédonie: Phanerogames. Office de la
Recherche Scientifique Coloniale, Paris.
Herbert, J. 2005a. New combinations and a new species in
Morella (Myricaceae). Novon 15: 293–295.
Herbert, J. 2005b. Systematics and Biogeography of Myrica-
ceae. Ph.D. Thesis, University of St. Andrews [Unpubl.].
Herbert, J. In press. Distribution, habitat and Red List status
of the New Caledonian endemic tree Canacomyrica mon-
ticola (Myricaceae). Biodiv. Cons.
Hjelmqvist, H. 1948. Studies on the floral morphology and
phylogeny of the Amentiferae. Bot. Not. Suppl. 2, 1:
Herbert & al. • Systematic position of Canacomyrica55 (2) • May 2006: 349–357
355
1–171.
Hufford, L. 1992. Rosidae and their relationships to other non-
magnoliid dicotyledons: a phylogenetic analysis using
morphological and chemical data. Ann. Missouri Bot.
Gard. 79: 218–248.
Jong, K. & Möller, M. 2000. New chromosome counts in
Streptocarpus (Gesneriaceae) from Madagascar and the
Comoro Islands and their taxonomic significance. Pl. Syst.
Evol. 224: 173–182.
Kaul, R. B. 1985. Reproductive morphology of Quercus
(Fagaceae). Amer. J. Bot. 72: 1962–1977.
Kelchner, C. A. 2000. The evolution of non-coding chloroplast
DNA and its application in plant systematics. Ann.
Missouri Bot. Gard. 87: 482–498.
Kubitzki, K. 1993a. Fagaceae. Pp. 301–309 in: Kubitzki, K.,
Rohwer, J. G. & Bittrich, V. (eds.), The Families and
Genera of Vascular Plants, vol. 2. Springer-Verlag, Berlin.
Kubitzki, K. 1993b. Myricaceae. Pp. 453–457 in: Kubitzki,
K., Rohwer, J. G. & Bittrich, V. (eds.), The Families and
Genera of Vascular Plants, vol. 2. Springer-Verlag, Berlin.
Längström, E. & Chase, M. W. 2002. Tribes of Boragino-
ideae (Boraginaceae) and placement of Antiphytum,
Echiochilon, Ogastemma and Sericostoma: a phylogenet-
ic analysis based on atpB plastid DNA sequence data. Pl.
Syst. Evol. 234: 137–153.
Leroy, J.-F. 1949. De la morphologie florale et de la classifi-
cation des Myricaceae. Compt. Rend. Hebd. Séances
Acad. Sci. 229: 1162–1163.
Li, R. Q., Chen, Z. D., Hong, Y. P. & Lu, A. M. 2002.
Phylogenetic relationships of the “higher” hamamelids
based on chloroplast trnL-F sequences. Acta Bot. Sin. 44:
1462–1468.
Li, R. Q., Chen, Z. D., Lu, A. M., Soltis, D. E. & Soltis, P. S.
2004. Phylogenetic analysis of Fagales based on multiple
DNA sequences from three genomes. Int. J. Plant Sci. 165:
311–324.
Lloyd, D. G. 1981. The distribution of sex in Myrica gale. Pl.
Syst. Evol. 138: 29–45.
Loockerman, D. J., Turner, B. L. & Jansen, R. K. 2003.
Phylogenetic relationships within the Tageteae
(Asteraceae) based on nuclear ribosomal ITS and chloro-
plast ndhF gene sequences. Syst. Bot. 28: 191–207.
Löve, A. 1980. Chromosome number reports LXIX. Taxon 29:
703–730.
Löve, A. 1982. IOPB Chromosome number reports LXXIV.
Taxon 31: 119–128.
Macdonald, A. D. 1977. Myricaceae: floral hypothesis for
Gale and Comptonia. Canad. J. Bot. 55: 2636–2651.
Macdonald, A. D. 1989. The morphology and relationships of
the Myricaceae. Pp. 147–165 in: Crane, P. R. & Black-
more, S. (eds.), Evolution, Systematics, and Fossil History
of the Hamamelidae, vol. 2. Clarendon Press, Oxford.
Maggia, L. & Bousquet, J. 1994. Molecular phylogeny of the
actinorhizal hamamelidae and relationships with host
promiscuity towards Frankia. Molec. Ecol. 3: 459–467.
Manos, P. S., Doyle, J. J. & Nixon, K. C. 1999. Phylogeny,
biogeography, and processes of molecular differentiation
in Quercus subgenus Quercus (Fagaceae). Molec. Phylog.
Evol. 12: 333–349.
Manos, P. S. & Steele, K. P. 1997. Phylogenetic analyses of
“higher” Hamamelididae based on plastid sequence data.
Amer. J. Bot. 84: 1407–1419.
Manos, P. S. & Stone, D. E. 2001. Evolution, phylogeny, and
systematics of the Juglandaceae. Ann. Missouri Bot. Gard.
88: 231–269.
Manos, P. S., Zhe-Kun Zhou, I. & Cannon, C. H. 2001.
Systematics of Fagaceae: phylogenetic tests of reproduc-
tive trait evolution. Int. J. Plant Sci. 16: 1361–1379.
Martin, P. G. & Dowd, J. M. 1993. Using sequences of rbcL
to study phylogeny and biogeography of Nothofagus
species. Australian Syst. Bot. 6: 441–447.
Melchior, H. 1964. Myricaceae. 40, Engler’s Syllabus der
Pflanzenfamilien. 12 ed. Gebrüder Borntraeger, Berlin-
Nikolassee.
Morawetz, W. & Samuel, M. R. A. 1989. Karyological pat-
terns in the Hamamelidae. Pp. 131–135 in: Crane, P. R. &
Blackmore, S. (eds.), Evolution, Systematics and Fossil
History of the Hamamelidae, vol. 2. Clarendon Press,
Oxford.
Muellner, A. N., Samuel, R., Johnson, S. A., Cheek, M.,
Pennington, T. D. & Chase, M. W. 2003. Molecular phy-
logenetics of Meliaceae (Sapindales) based on nuclear and
plastid DNA sequences. Amer. J. Bot. 90: 471–480.
Navarro, E., Bousquet, J., Moiroud, A., Munive, A., Piou,
D. & Normand, P. 2003. Molecular phylogeny of Alnus
(Betulaceae), inferred from nuclear ribosomal DNA ITS
sequences. Pl. & Soil 254: 207–217.
Oginuma, K. & Tanaka, R. 1987. Karyomorphological stud-
ies on three species of Myrica. J. Jap. Bot. 62: 183–188.
Polhill, R. M. & Verdcourt, B. 2000. Myricaceae. Pp. 1–12 in:
Beentje, H. J. & Smith, S. A. L. (eds.), Flora of Tropical
East Africa. Balkema, Rotterdam.
Potter, D., Gao, F. Y., Baggett, S., McKenna, J. R. &
McGranahan, G. H. 2002. Defining the sources of para-
dox: DNA sequence markers for North American walnut
(Juglans L.) species and hybrids. Sc. Hort. 94: 157–170.
Raven, P. H. & Axelrod, D. I. 1974. Angiosperm biogeogra-
phy and past continental movements. Ann. Missouri Bot.
Gard. 61: 539–673.
Reeves, G., Chase, M. W., Goldblatt, P., Rudall, P., Fay, M.
F., Cox, A. V., Lejeune, B. & Souza-Chies, T. 2001.
Molecular systematics of Iridaceae: evidence from four
plastid DNA regions. Amer. J. Bot. 88: 2074–2087.
Richardson, J. E., Fay, M. F., Cronk, Q. C. B., Bowman, D.
& Chase, M. W. 2000. A phylogenetic analysis of
Rhamnaceae using rbcL and trnL-F plastid DNA
sequences. Amer. J. Bot. 87: 1309–1324.
Savolainen, V., Fay, M. F., Albach, D. C., Backlund, A., van
der Bank, M., Cameron, K. M., Johnson, S. A., Lledó,
M. D., Pintaud, J.-C., Powell, M., Sheahan, M. C.,
Soltis, D. E., Soltis, P. S., Weston, P., Whitten, W. M.,
Wurdack, K. J. & Chase, M. W. 2000. Phylogeny of the
eudicots: a nearly complete familial analysis based on
rbcL gene sequences. Kew Bull. 55: 257–309.
Sogo, A., Setoguchi, H., Noguchi, J., Jaffré, T. & Tobe, H.
2001. Molecular phylogeny of Casuarinaceae based on
rbcL and matK gene sequences. J. Pl. Res. 114: 459–464.
Soltis, D. E., Soltis, P. S., Chase, M. W., Mort, M. E.,
Albach, D. C., Zanis, M., Savolainen, V., Hahn, W. H.,
Hoot, S. B., Fay, M. F., Axtell, M., Swensen, S. M.,
Prince, L. M., Kress, W. J., Nixon, K. C. & Farris, J. S.
2000. Angiosperm phylogeny inferred from 18S rDNA,
rbcL, and atpB sequences. Bot. J. Linn. Soc. 133:
381–461.
Herbert & al. • Systematic position of Canacomyrica 55 (2) • May 2006: 349–357
356
Soltis, D. E., Soltis, P. S., Morgan, D. R., Swensen, S. M.,
Mullin, B. C., Dowd, J. M. & Martin, P. G. 1995.
Chloroplast gene sequence data suggest a single origin of
the predisposition for symbiotic nitrogen-fixation in
angiosperms. Proc. Natl. Acad. Sci. U.S.A. 92: 2647–2651.
Soltis, P. S., Soltis, D. E. & Chase, M. W. 1999. Angiosperm
phylogeny inferred from multiple genes as a tool for com-
parative biology. Nature 402: 402–404.
Sosa, V. & Chase, M. W. 2003. Phylogenetics of Crossosoma-
taceae based on rbcL sequence data. Syst. Bot. 28: 96–105.
Sundberg, M. D. 1985. Pollen of the Myricaceae. Pollen et
Spores 27: 15–28.
Swensen, S. M. 1996. The evolution of actinorhizal symbios-
es: evidence for multiple origins of the symbiotic associa-
tion. Amer. J. Bot. 83: 1503–1512.
Swofford, D. L. 1998. PAUP*. Phylogenetic Analysis Using
Parsimony (*and Other Methods). Version 4. Sinauer
Associates, Sunderland, Massachusetts.
Taberlet, P., Gielly, L., Pautou, G. & Bouvet, J. 1991.
Universal primers for amplification of three non-coding
regions of chloroplast DNA. Pl. Molec. Biol. 17:
1105–1109.
Takhtajan, A. 1997. Myricaceae. 153–154 in: Takhtajan, A.
(ed.), Diversity and Classification of Flowering Plants.
Columbia Univ. Press, New York.
Thorne, R. F. 1973. The “Amentiferae” or Hamamelidae as an
artificial group: a summary statement. Brittonia 25:
395–405.
Thorne, R. F. 1992a. Classification and geography of the flow-
ering plants. Bot. Rev. 58: 225–348.
Thorne, R. F. 1992b. An updated phylogenetic classification
of the flowering plants. Aliso 13: 365–389.
Thorne, R. F. 2000. The classification and geography of the
flowering plants: dicotyledons of the class Angiospermae.
Bot. Rev. 66: 442–647.
Vitousek, P. E. & Walker, L. R. 1989. Biological invasion by
Myrica faya in Hawai’i: plant demography, nitrogen fixa-
tion, ecosystem effects. Ecol. Monogr. 59: 247–265.
Walker-Larsen, J. & Harder, L. D. 2000. The evolution of
staminodes in angiosperms: patterns of stamen reduction,
loss, and functional re-invention. Amer. J. Bot. 87:
1367–1384.
Weising, K., Nybom, H., Wolff, K. & Meyer, W. 1995. DNA
Fingerprinting in Plants and Fungi. CRC, London.
White, T. J., Bruns, T., Lee, S. & Taylor, J. 1990. Ampli-
fication and direct sequencing of fungal ribosomal RNA
genes for phylogenetics. 315–322 in: Innis, M. D.,
Gelfand, D., Sninsky, J. & White, T. (eds.), PCR
Protocols: A Guide to Methods and Applications.
Academic Press, San Diego.
Whitlock, B. A., Karol, K. G. & Alverson, W. S. 2003.
Chloroplast DNA sequences confirm the placement of the
enigmatic Oceanopapaver within Corchorus (Grewioi-
deae: Malvaceae s.l., formerly Tiliaceae). Int. J. Plant Sci.
164: 35–41.
Wiens, J. J. 1998. Combining datasets with different phyloge-
netic histories. Syst. Biol. 47: 568–581.
Wilbur, R. L. 1994. The Myricaceae of the United States and
Canada: genera, subgenera, and series. Sida 16: 93–107.
Wilson, K. L. & Johnson, L. A. S. 1989. Casuarinaceae. Pp.
100–174 in: George, A. S. (ed.), Flora of Australia, vol. 3.
Australian Government Publishing Service, Canberra.
Yoo, K. O. & Wen, J. 2002. Phylogeny and biogeography of
Carpinus and subfamily Coryloideae (Betulaceae). Int. J.
Plant Sci. 163: 641–650.
Zheng-yi, W. & Raven, P. H. 1999. Flora of China, vol. 4.
Missouri Botanical Garden Press, St. Louis.
Zomlefer, W. B. 1994. Myricaceae. Pp. 174–176 in: (eds.),
Guide to Flowering Plant Families. The Univ. of North
Carolina Press, Chapel Hill & London.
Herbert & al. • Systematic position of Canacomyrica55 (2) • May 2006: 349–357
357
Appendix. Accessions sampled. Voucher information and GenBank accession number are given for sequences report-
ed here for the first time. Taxa for which sequences were previously published are listed with the original reference and
GenBank accession number.
Taxon, Voucher/Source, Genbank Acc. No.: rbcL, trnL-F, ITS.
Alnus firma Siebold & Zucc. (Betulaceae), Kamiya, unpubl., AB060562; Kamiya & Harada, unpubl., AB063524 and AB 063548;
Navarro & al., 2003, AJ251684. Canacomyrica monticola Guillaumin (Myricaceae), Herbert 934 (E), DQ310504; Herbert 934 (E),
DQ310508; Herbert 934 (E), DQ310500. Carpinus laxiflora (Siebold & Zucc.) Blume (Betulaceae), Kamiya & Harada, unpubl.,
AB060585; Kamiya & Harada, unpubl., AB063571 and AB063541; Yoo & al., 2002, AF432038. Casuarina equisetifolia L.
(Casuarinaceae), Sogo & al., 2001, AY033859; Li & al., 2002, AY147090; Steane, unpubl., AY864057. Comptonia peregrina (L.) L’Hérit.
(Myricaceae), Meagher s.n. (E), DQ310505; Meagher s.n. (E), DQ310509; Meagher s.n. (E), DQ310501. Fagus crenata Blume
(Fagaceae), Martin & al., 1993, L13339; Fujii & al., 2002, AB046508; Denk & al., 2002, AF456969. Juglans nigra L. (Juglandaceae),
Soltis & al., 1999, AF206785; Manos & al., 2001, AF303783; Potter & al., 2002, AF338491. Morella cordifolia (L.) Killick (Myricaceae),
Herbert 1007 (E), DQ310502; Herbert 1007 (E), DQ310506; Herbert 1007 (E), DQ310498. Myrica hartwegii Watson (Myricaceae),
Edwards 93 (RPBG), DQ310503; Edwards 93 (RPBG), DQ310507; Edwards 93 (RPBG), DQ310499. Quercus rubra L. (Fagaceae),
Bousquet & al., 1992, M58391; Gielly & al., 1994, X75707; Manos & al., 1999, AF098418. Rhoiptelea chiliantha Diels & Hand.-Mazz.
(Juglandaceae), Chen & al., 1998, AF017687; Manos & al., 2001, AF303773; Manos & al., 2001, AF303800.
... Fagales are strongly supported (BS = 99%), and include Nothofagaceae (BS = Fagaceae and then Juglandaceae are subsequent sisters to a clade of Myricaceae + (Casuarinaceae + (Betulaceae + Ticodendraceae) with strong support (Fig. 2A). These relationships differ from published topologies concerning the position of Myricaceae (Li et al., 2004;Herbert et al., 2006;Zhu et al., 2007;Soltis et al., 2011). For example, in Soltis et al. (2011), Myricaceae + Juglandaceae are sister to Betulaceae + Casuarinaceae. ...
... Within Myricaceae, Canacomyrica and Comptona are successive sisters to Myrica + Morella with strong support, in agreement with previous studies (Herbert et al., 2006;Xiang et al., 2014;Fig. S3 BS support, respectively. ...
Article
Full-text available
Rosidae, a clade of ∼ 90 000 species of angiosperms, exhibits remarkable morphological diversity and extraordinary heterogeneity in habitats and life forms. Resolving phylogenetic relationships within Rosidae has been difficult, in large part due to nested radiations and the enormous size of the clade. Current estimates of phylogeny contain areas of poor resolution and/or support, and there have been few attempts to synthesize the available data into a comprehensive view of Rosidae phylogeny. We aim to improve understanding of the phylogeny of Rosidae with a dense sampling scheme using both newly generated sequences and data from GenBank of the chloroplast rbcL, atpB, and matK genes and the mitochondrial matR gene. We combined sequences from 9300 species, representing 2775 genera, 138 families, and 17 orders into a supermatrix. Although 59.26% of the cells in the supermatrix have no data, our results generally agree with previous estimates of Rosidae phylogeny and provide greater resolution and support in several areas of the topology. Several noteworthy phylogenetic relationships are recovered, including some novel relationships. Two families (Euphorbiaceae and Salvadoraceae) and 467 genera are recovered as non-monophyletic in our sampling, suggesting the need for future systematic studies of these groups. Our study demonstrates the value of a botanically informed bioinformatics approach and dense taxonomic sampling for resolving rosid relationships. The resulting tree provides a starting point for large-scale analyses of the evolutionary patterns within Rosidae.
... Within Myricaceae, previous analyses-based on plastid, nuclear, or combined plastid and nuclear data (Herbert et al., 2006;Xing et al., 2014;Liu et al., 2015;Sun et al., 2016)-have mostly supported the topology of ((Myrica, Morella), Comptonia). Although several of our analyses supported this relationship (i.e., the concatenated analysis of PCN_outlier_removed_MT12 and all ASTRAL analyses), our concatenated coding datasets supported the topology (Myrica, (Morella, Comptonia)), whereas our concatenated noncoding datasets supported (albeit weakly) the topology (Morella, (Myrica, Comptonia)). ...
Article
Plastid phylogenomic analyses have shed light on many recalcitrant relationships across the angiosperm Tree of Life and continue to play an important role in plant phylogenetics alongside nuclear data sets given the utility of plastomes for revealing ancient and recent introgression. Here we conduct a plastid phylogenomic study of Fagales, aimed at exploring contentious relationships (e.g., the placement of Myricaceae and some intergeneric relationships in Betulaceae, Juglandaceae, and Fagaceae) and dissecting conflicting phylogenetic signals across the plastome. Combining 102 newly sequenced samples with publically available plastomes, we analyzed a dataset including 256 species and 32 of the 34 total genera of Fagales, representing the largest plastome-based study of the order to date. We find strong support for a sister relationship between Myricaceae and Juglandaceae, as well as strongly supported conflicting signal for alternative generic relationships in Betulaceae and Juglandaceae. These conflicts highlight the sensitivity of plastid phylogenomic analyses to genic composition, perhaps due to the prevalence of uninformative loci and heterogeneity in signal across different regions of the plastome. Phylogenetic relationships were geographically structure in subfamily Quercoideae, with Quercus being non-monophyletic and its sections forming clades with co-distributed Old World or New World genera of Quercoideae. Compared against studies based on nuclear genes, these results suggest extensive introgression and chloroplast capture in the early diversification of Quercus and Quercoideae. This study provides a critical plastome perspective on Fagales phylogeny, setting the stage for future studies employing more extensive data from the nuclear genome.
... Resolving the affinities of phylogenetically enigmatic lineages of flowering plants is necessary for building robust taxonomic classifications (Herbert et al., 2006;Moore et al., 2007;Saarela et al., 2007;Soltis et al., 2007;Li et al., 2011;Shipunov and Shipunova, 2011;Su et al., 2012;Xiang et al., 2012;Cardoso et al., 2012aCardoso et al., , 2012bCardoso et al., , 2015Chacón et al., 2016;Chen et al., 2016;Ramos et al., 2016), for understanding the evolution of florally disparate architectures Saarela et al., 2007;Espíndola et al., 2010;Cardoso et al., 2012a) and the biogeographic histories of widely disjunct distributions (Ellison et al., 2012;Deng et al., 2015). ...
... -Myricaceae comprise four genera, the monotypic Comptonia and Canacomyrica, the small Myrica (two species) and Morella with c. 47 species (e.g. Wilbur 1994;Huguet et al. 2005;Herbert 2005aHerbert , 2005bHerbert et al. 2006 ...
Article
Full-text available
An ongoing investigation of the middle Miocene (Sarmatian) palynoflora from the Lavanttal Basin continues to show that it contains an extremely rich assemblage of angiosperm taxa. The Fagales to Rosales pollen record documented here contains 34 different taxa belonging to the Betulaceae (Alnus, Betula, Carpinus, Corylus, Ostrya), Fagaceae (Castanea, Fagus, Quercus Groups Cerris, Ilex, Cyclobalanopsis, Quercus/Lobatae), Juglandaceae (Engelhardioideae, Carya, Juglans, Pterocarya), Myricaceae (Morrella vel Myrica), Cannabaceae (Celtis), Elaeagnaceae (Elaeagnus), Rhamnaceae, Rosaceae (Prunus) and Ulmaceae (Cedrelospermum, Ulmus, Zelkova). Two of the pollen types represent extinct genera, Trigonobalanopsis and Cedrelospermum, and are also reported for the first time from the Lavanttal Basin along with pollen of Rhamnaceae and Prunus. The different types of Quercus pollen are now affiliated with Groups Cerris, Cyclobalanopsis, Ilex and Quercus/Lobatae based on sculpturing elements observed using scanning electron microscopy (SEM). Köppen signatures of potential modern analogues of the fossil Fagales and Rosales suggest a subtropical (Cfa, Cwa) climate at lower elevation and subsequent subtropical to temperate climate with altitudinal succession (Cfa → Cfb/Dfa→ Dfb; Cwa → Cwb → Dwb) in the Lavanttal area during accumulation of the palynoflora. Most of the fossil taxa have potential modern analogues that can be grouped as nemoral and/or merido-nemoral vegetation elements, and the diversity of Fagales indicates a varying landscape with a high variety of niches.
... Close relatives to the oak are chestnut (Castanea), hazelnut (Corylus), pecan (Carya), and walnut (Juglans). Evolutionary relationships among these species have been difficult to estimate (Manos and Steele 1997;Soltis et al. 2000;Ruiqi et al. 2002;Cook and Crisp 2005;Sauquet et al. 2012), but analyses of DNA sequence data are converging on the topology shown in Fig. 1 (Soltis et al. 2000;Stevens 2001 onwards;Li et al. 2004;Herbert et al. 2006). ...
Article
Full-text available
Tree nut allergies are some of the most common and serious allergies in the United States. Patients who are sensitive to nuts or to seeds commonly called nuts are advised to avoid consuming seeds from a variety of different species, even though these may be distantly related in terms of their evolutionary history. Clinical studies have uncovered preliminary patterns of cross-reactivity to different nut species, but have been limited in patient sample size and the number of nut species included in each test. Here, we investigate the evolutionary patterns of three allergenic seed-storage proteins (vicilin, legumin, and 2S albumin) from species of vascular plants, by generating estimates of their phylogenetic relationships to test the hypothesis that primary sequence data is a good predictor of allergic cross-sensitivity. In general, evolutionary relationships of the three proteins are congruent with the current understanding of plant species relationships. We find little evidence for convergent evolution between distantly related species with vicilin, legumin, or 2S albumin amino acid sequences. Thus, features of the proteins other than their amino acid sequences may be driving the cross-reactivity observed during in vitro tests and skin tests. Our results support current treatment guidelines to limit nut and seed consumption if allergies are present in a patient, but we propose that more studies are necessary to accurately test the extent of cross-reactivity between nut species and to better understand the protein characteristics that cause allergies.
... Hufford, 1992) and, more significantly, molecular phylogenetic analyses (Manos et al., 1993;Stevens, 2001 onwards;Soltis et al., 2005, and references therein) have shown the polyphyletic relationships of these taxa and have broken the group into a number of clades. Based on molecular data, the Fagales currently includes seven families, with Nothofagaceae sister to the remaining families, followed by Fagaceae, which is sister to the 'core' fagalean families (Manos et al., 1993;Manos & Steele, 1997;Soltis et al., 2000Soltis et al., , 2005Stevens, 2001 onwards;Judd et al., 2002;Li et al., 2002Li et al., , 2004Herbert et al., 2006;Zhu et al., 2007). 'Core' Fagales include a clade with Myricaceae and Juglandaceae (including recently submerged Rhoipteleaceae; APG III, 2009) and a sister clade including Casuarinaceae, Betulaceae and Ticodendraceae, although the relationship of Myricaceae remains somewhat problematic. ...
Article
Full-text available
Flowers of many living Fagales exhibit unusual developmental characteristics. At anthesis, ovulate flowers have carpels bearing immature orthotropous ovules. After pollination, the ovules increase in size and become anatropous and the ovary enlarges. Simultaneously, the pollen tubes extend from the stigma to the ovules with several phases of growth and quiescence. Finally, after the first fertilization, the remaining ovules abort, resulting in a single‐seeded fruit. Three‐dimensionally preserved potentially fagaceous mesofossil flowers from the Campanian of Massachusetts, USA, provide evidence on the evolution of these characters. The fossils share putative synapomorphies with the Fagales (six tepals, mostly inferior, three‐carpellate ovary with each locule initially containing two pendant ovules, punctate‐rugulate, tricolporate pollen and fruit with a single seed). However, the fossil is bisexual and has nectaries, characters shared with the sister order Cucurbitales, and both lack the fagalean immature orthotropous developmental stage. The fossil shares synapomorphies of an inferior ovary and a single‐seeded indehiscent fruit with both living orders and appears to be transitional. Comparison of ontogenetic changes between the fossil and related fagalean taxa suggests independent stepwise changes in development in which some characters of the modern clades were in place at ∼ 75 Myr and others evolved more recently. © 2012 The Linnean Society of London, Botanical Journal of the Linnean Society, 2012, 168, 353–376.
Article
New Caledonia is a biodiversity hotspot located in the south-western Pacific, well known for its rich, unique and endangered flora. The island flora has a high level of endemism not only at the species level (75%), but also at the generic and family (three endemic) levels. We review here the taxonomic validity of the c. 100 endemic New Caledonian genera of vascular plants (13%) by using the monophyly criterion based on the available phylogenetic data. As observed in other island floras, some of these genera were recovered nested in larger genera and are consequently likely to lose their rank. After a critical review, we concluded that the New Caledonian plant vascular flora contains between 62 and 91 endemic genera. This large variation in the number of endemic genera is mainly caused by a lack of DNA sequences (eight genera) and limited phylogenetic evidence. This work highlights gaps of knowledge that will have to be addressed to stabilize the taxonomy of the New Caledonian flora. Although this study shows that several genera are not monophyletic, New Caledonia still harbours more endemic genera than any other islands in the Pacific Ocean. Preliminary results indicate that the high level of endemism at higher taxonomic levels could be explained by an accumulation of relictual lineages, rather than adaptive radiations. Hypotheses explaining this phenomenon are provided in this study
Article
Full-text available
There has been increasing interest in integrating a regional tree of life with community assembly rules in the ecological research. This raises questions regarding the impacts of taxon sampling strategies at the regional versus global scales on the topology. To address this concern, we constructed two trees for the nitrogen-fixing clade: (i) a genus-level global tree including 1023 genera; and (ii) a regional tree comprising 303 genera, with taxon sampling limited to China. We used the supermatrix approach and performed maximum likelihood analyses on combined matK, rbcL, and trnL-F plastid sequences. We found that the topology of the global and the regional tree of the N-fixing clade were generally congruent. However, whereas relationships among the four orders obtained with the global tree agreed with the accepted topology obtained in focused analyses with more genes, the regional topology obtained different relationships, albeit weakly supported. At a finer scale, the phylogenetic position of the family Myricaceae was found to be sensitive to sampling density. We expect that internal support throughout the phylogeny could be improved with denser taxon sampling. The taxon sampling approach (global vs. regional) did not have a major impact on fine-level branching patterns of the N-fixing clade. Thus, a well-resolved phylogeny with relatively dense taxon sampling strategy at the regional scale appears, in this case, to be a good representation of the overall phylogenetic pattern and could be used in ecological research. Otherwise, the regional tree should be adjusted according to the correspondingly reliable global tree.
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
Details of inflorescence, floral, and fruit morphology have been studied in more than 120 species of Asiatic and American Quercus. Of the two subgenera, subgenus Cyclobalanopsis has fewer species but greater diversity of reproductive morphology than subgenus Quercus. Some character states of subgenus Cyclobalanopsis, such as more numerous stamens, male flowers sometimes grouped in dichasia, abortive ovules well developed, prominent intrusive septae in the nut, and lamellate cupules, are shared with Lithocarpus. The diversity of cupular sizes, coverage, and ornamentation raises questions about the adaptive nature, ecological function, and phylogeny of the cupule, which clearly has by now evolved as a structure with its own qualities.
Article
The recently-developed statistical method known as the "bootstrap" can be used to place confidence intervals on phylogenies. It involves resampling points from one's own data, with replacement, to create a series of bootstrap samples of the same size as the original data. Each of these is analyzed, and the variation among the resulting estimates taken to indicate the size of the error involved in making estimates from the original data. In the case of phylogenies, it is argued that the proper method of resampling is to keep all of the original species while sampling characters with replacement, under the assumption that the characters have been independently drawn by the systematist and have evolved independently. Majority-rule consensus trees can be used to construct a phylogeny showing all of the inferred monophyletic groups that occurred in a majority of the bootstrap samples. If a group shows up 95% of the time or more, the evidence for it is taken to be statistically significant. Existing computer programs can be used to analyze different bootstrap samples by using weights on the characters, the weight of a character being how many times it was drawn in bootstrap sampling. When all characters are perfectly compatible, as envisioned by Hennig, bootstrap sampling becomes unnecessary; the bootstrap method would show significant evidence for a group if it is defined by three or more characters.
Article
Wood anatomy of the single species of Canacomyrica (hitherto not studied) shows that it belongs in Myricaceae, although it differs from other genera in several respects (axial parenchyma grouped in bands or columns as well as diffuse; Heterogeneous Type I rays; more numerous bars per perforation plate). The latter two features are primitive for the family. The four genera (Canacomyrica. Comptonia, Morella, and Myrica s.s.) differ from each other not only by qualitative features but by quantitative features (feature means in genera mostly non- overl apping). Wood of Comptonia and Myri ca s.s. lacks chambered crystals in axial parenchyma and ray crystals. Wood of Myrica s.s. has tracheids in latewood but fiber-tracheids in earl ywood. Diagnostic generic summaries are presented. Features of Myricaceae such as scalarifom perforation plates, presence of (true) tracheids , ray types , chambered encapsulated crystals in axial parenchyma, and bark anatomy correspond with character states and expressions in Betulaceae, Casuarinaceae, Corylaceae, Juglandaceae (including Rhoipteleaceae), Ticodendraceae and, to a lesser extent, Fagaceae and Nothofagaceae. This grouping of families can be found as Fagales in recent DNA trees. The predominance of tracheids in basal Fagales such as Myricaceae and Ticodendraceae suggests that origin of vasicentric tracheids which occur in combination with libriform fibers in Fagaceae is the product of tracheid dimorphism. Low imperforate tracheid length to vessel element length ratios (FN ratios) in Myricaceae are a probable indication of wood primitiveness. Quantitative vessel features of Myricaceae, as combined in Mesomorphy Ratio values, characterize wood of Myricaceae as a whole, but at the species level such values correspond to respective habitats; notably high vessel density in Comptonia may represent greater conductive safety appropriate to relatively dry habitats.
Article
The molecular support for current generic concepts in Myricaceae has made it necessary to make new combinations in Morella. Morella comprises approximately 47 species found in the Old and New World tropics. Three new combinations are made here: Morella adenophora, M. nana, and M. punctata. Morella nana is lectotypified, and a new species, M. rivas-martinezii, is described.