Sei sulla pagina 1di 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/230864722

Apomixis and reticulate evolution in the Asplenium monanthes fern complex

Article  in  Annals of Botany · September 2012


DOI: 10.1093/aob/mcs202 · Source: PubMed

CITATIONS READS

45 198

3 authors:

Robert J Dyer Vincent Savolainen


Natural History Museum, London Imperial College London
6 PUBLICATIONS   90 CITATIONS    257 PUBLICATIONS   20,644 CITATIONS   

SEE PROFILE SEE PROFILE

Harald Schneider
Xishuangbanna Tropical Botanical Garden
352 PUBLICATIONS   11,235 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Evolution of New Zealand ferns View project

Evolutionary history of the Karst Fern Diversity in Southeast Asia View project

All content following this page was uploaded by Harald Schneider on 31 May 2014.

The user has requested enhancement of the downloaded file.


Annals of Botany Page 1 of 15
doi:10.1093/aob/mcs202, available online at www.aob.oxfordjournals.org

Apomixis and reticulate evolution in the Asplenium monanthes fern complex


Robert J. Dyer1,2,*, Vincent Savolainen2,3 and Harald Schneider1,4,*
1
Department of Botany, Natural History Museum, London SW7 5BD, UK, 2Imperial College London, Silwood Park Campus,
Ascot, Berkshire SL5 7PY, UK, 3Royal Botanic Gardens, Kew TW9 3DS, UK and 4State Key Laboratory of Systematic and
Evolutionary Botany, Institute of Botany, The Chinese Academy of Sciences, Beijing, China
* For correspondence. Email r.dyer@nhm.ac.uk, robert.dyer07@imperial.ac.uk or h.schneider@nhm.ac.uk

Received: 10 May 2012 Returned for revision: 16 July 2012 Accepted: 7 August 2012

† Background and Aims Asexual reproduction is a prominent evolutionary process within land plant lineages and

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


especially in ferns. Up to 10 % of the approx. 10 000 fern species are assumed to be obligate asexuals. In the
Asplenium monanthes species complex, previous studies identified two triploid, apomictic species. The
purpose of this study was to elucidate the phylogenetic relationships in the A. monanthes complex and to inves-
tigate the occurrence and evolution of apomixis within this group.
† Methods DNA sequences of three plastid markers and one nuclear single copy gene were used for phylogenetic
analyses. Reproductive modes were assessed by examining gametophytic and sporophyte development, while
polyploidy was inferred from spore measurements.
† Key Results Asplenium monanthes and A. resiliens are confirmed to be apomictic. Asplenium palmeri,
A. hallbergii and specimens that are morphologically similar to A. heterochroum are also found to be apomictic.
Apomixis is confined to two main clades of taxa related to A. monanthes and A. resiliens, respectively, and is
associated with reticulate evolution. Two apomictic A. monanthes lineages, and two putative diploid sexual pro-
genitor species are identified in the A. monanthes clade.
† Conclusions Multiple origins of apomixis are inferred, in both alloploid and autoploid forms, within the
A. resiliens and A. monanthes clades.

Key words: Apomixis, Asplenium monanthes, asexual reproduction, apogamy, diplospory, reticulate evolution,
hybridization, polyploidy.

IN T RO DU C T IO N taxa and because apomictic taxa are usually found in complex


reticulation networks with other apomictic and sexually repro-
Asexually reproducing organisms are at the forefront of evolu-
ducing lineages (Manton, 1950; Lovis, 1977; Walker, 1979;
tionary theory and continue to be the subject of new inquiries
Grusz et al., 2009). A recent study by Beck et al. (2011a) sup-
(Judson and Normark, 1996; Fontaneto et al., 2007; ports the role of hybridization in apomictic origins, showing
Schwander et al., 2011). Although the evolutionary transition that it drives the transition to apomixis in diploid Boechera
to obligate asexual reproduction is relatively rare across eukar- species. The interpretation of genetic patterns in other apomic-
yotes, it appears to be an important evolutionary process in tic lineages often reveals a cryptic range of genotypes and
plants (Mogie, 1990, 1992; Asker and Jerling, 1992). morphotypes (Grusz et al., 2009; Schneller and Krattinger,
However, the taxonomic distribution of apomictic (asexual) 2010). These patterns can be caused by a number of factors,
taxa across plant lineages is uneven, with estimations of including: ongoing apomict formation caused by coexisting
0.1 % in angiosperms, up to 10 % in ferns, and with little or sexual progenitor species (clonal turnover); extinction or
no evidence in gymnosperms, mosses, liverworts or hornworts rarity of sexual progenitors; hybridization between apomicts
(Walker, 1966a; Lovis, 1977; Asker and Jerling, 1992; Mogie, and closely related sexual taxa through functional antheridia;
1992; Pichot et al., 2001; Park and Kato, 2003). ongoing somatic mutation; and homoeologous chromosome
Sexual reproduction is ancestral in homosporous ferns, and pairing (Klekowski, 1970; Klekowski and Hickok, 1974;
the transition to apomixis requires the evolution and alterna- Ishikawa et al., 2003). Crucial to the understanding of these
tion of two distinct phenomena: diplospory and apogamy. patterns is the level of gene flow to and from apomictic
They represent the avoidance of meiotic reduction during sporo- lineages. In elucidating origins of apomixis, it is important
genesis (diplospory), and the spontaneous development of a therefore to consider two hypotheses: a single origin and the
sporophyte without fertilization (apogamy) (Lovis, 1977). All subsequent diversification of apomictic lineages, or multiple
except one of the documented apomict fern taxa are obligate origins by hybridization between different sexual relatives.
(Lloyd, 1973; Lovis, 1977), and the reverse transition from apo- In this study, we aim to investigate the occurrence and
mixis to sexual reproduction has not been observed. origins of apomixis in the Asplenium monanthes complex,
The transition to apomixis has long been associated with for which a modern taxonomic treatment is overdue. The
hybridization, polyploidy and female sterility (Mogie, 1992). complex consists of up to 22 terrestrial and saxicolous
This is due to the meiotic chromosome behaviour observed species that occur mainly in Mesoamerica, with its centre of
in apomictic taxa, the high proportion of triploid apomictic diversity in southern Mexico (based on Mickel and Smith,
# The Author 2012. Published by Oxford University Press on behalf of the Annals of Botany Company. All rights reserved.
For Permissions, please email: journals.permissions@oup.com
Page 2 of 15 Dyer et al. — Apomixis and reticulate evolution in a fern complex

2004). The complex is part of the ‘black-stemmed’ spleenwort The approximated taxonomic coverage may be inflated as
group that also includes the A. normale complex, the result of the limited knowledge on the taxonomy of this
A. trichomanes complex, A. viride and the Hawaiian Diellia group. First, some unsampled taxa were erroneously consid-
complex (Schneider et al., 2004, 2005). The A. monanthes ered as members of the lineages based on morphology alone.
complex roughly corresponds to ‘Grupo 5’ of Adams (1995), Secondly, several of the unsampled taxa were known from a
and contains two documented triploid apomictic species, single or a few samples in herbaria. Studies on the putative
A. monanthes (n ¼ 2n ¼ 108 and 2n ¼ 144) (Manton, 1950; sister lineage, the A. trichomanes complex, showed the limita-
Manton and Vida, 1968; Manton et al., 1986; Wagner et al., tions of morphological-based studies in assessing species di-
1970; Tryon et al., 1973; Lovis et al., 1977; Smith and versity of Asplenium species complexes (Lovis, 1977;
Mickel, 1977) and A. resiliens (n ¼ 2n ¼ 108). It also contains H. Schneider, unpublished). Most specimens were collected
a rare pentaploid apomict species, A. heteroresiliens. The latter during fieldwork in Mexico, El Salvador and Costa Rica.
is thought to be a hybrid between the apomictic triploid DNA samples were preserved in silica gel, and voucher speci-
A. resiliens and the sexual tetraploid/hexaploid A. heterochroum mens were deposited at the herbarium of the Natural History
(2n ¼ 144 or 216) (Morzenti and Wagner, 1962; Wagner, Museum in London (BM), as well as at MEXU, LAGU and
1963, 1966; Morzenti, 1966; Wagner and Wagner, 1966; INBIO. Additional silica-dried material was provided by col-

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


Walker, 1966b; Wagner et al., 1970; Windham, 1983; leagues (collectors, localities and additional voucher informa-
Haufler and Soltis, 1986). Besides these species, the complex tion are given in Supplementary Data, Table S1). Some
includes the diploid sexual A. formosum (2n ¼ 72) (Manton, samples were obtained from herbarium specimens stored at
1959; Walker, 1966b; Ghatak, 1977; Ammal and Bahavanandan, BM. Unfortunately we were unable to obtain material for
1991; Guillén and Daviña, 2005), and several less well-studied A. gentryi, A. heteroresiliens, A. nesioticum, A. oligosorum,
taxa such as A. castaneum, A. blepharodes, A. fibrillosum, A. olivaceum, A. pringlei, A. sanchezii, A. stolonipes, A. tryonii,
A. hallbergii, A. palmeri, A. polyphyllum and A. soleirolioides A. underwoodii and A. vespertinum. These species are rare
(Mickel and Smith, 2004). Little or no evidence exists concern- and/or have restricted geographic ranges that were not visited
ing the reproductive biology and ploidy level of these taxa. (Mickel and Smith, 2004). All included samples were carefully
Asplenium monanthes is a widespread and presumed long- identified using the keys provided in Mickel and Smith (2004),
lived apomictic species that forms 32 unreduced spores via and by critical comparison with herbarium collections held at
the Döpp– Manton scheme of sporogenesis (Manton, 1950). the BM. If needed, specimen identifications were restudied in
Currently the complex lacks comprehensive revision, and so the context of the obtained DNA sequences, to address incon-
here we follow the treatment of the fern flora of Mexico by sistencies in results and identifications. Outgroup taxa were
Mickel and Smith (2004). Although Mexico is the centre of chosen based on Schneider et al. (2004, 2005).
species diversity for this complex, species are distributed
from southern North America, thorough Central and South
DNA extraction, PCR amplification and sequencing
America (Stolze, 1981; Tryon et al., 1993; Wagner et al.,
1993; Stolze et al., 1994; Adams, 1995; Smith et al., 1999; Total genomic DNA was extracted from leaf tissue using
Mickel and Smith, 2004; Zuloaga et al., 2008; Jørgensen cetyltrimethylammonium bromide (CTAB) (Doyle, 1987).
et al., 2010; Kessler and Smith, 2011). Apomictic triploid Three plastid markers commonly used in ferns were success-
A. monanthes and the sexual diploid A. formosum show a fully amplified using polymerase chain reactions (PCRs): (1)
wider distribution than the other species within this range, the psbA – trnH intergenic spacer (IGS) region (approx. 600
and are distinct as they also occur in other localities throughout nucleotides) (Aldrich et al., 1988); (2) the rps4 plus rps4 –
the sub-tropics, including Africa and various islands in the trnS IGS region (approx. 1000 nucleotides) (Smith and
Atlantic and Indian oceans. Current knowledge makes this Cranfill, 2002); and (3) the trnL – trnF region including the
group a prime candidate to explore the origins and long-term trnL intron and the trnL – trnF IGS region (approx. 900 nucleo-
fate of apomictic ferns, along with aspects of their biogeog- tides) (Taberlet et al., 1991; Trewick et al., 2002). Samples
raphy and ecology. from A. formosum proved difficult to amplify for trnL – trnF,
Here, we use plastid and nuclear sequence data to recon- so only the trnL IGS was amplified using E and F primers
struct species relationships within the A. monanthes complex. (Taberlet et al., 1991; Trewick et al., 2002).
This phylogenetic framework is used in combination with PCR was carried out using Promega Gotaqw Green Master
some experiments to determine the reproductive modes and Mix, and amplification was performed using an ABI Veriti
ploidy levels in the taxa. Then, we evaluate the following hy- Thermo Cycler (Applied Biosystems, Warrington, Cheshire,
potheses: (1) apomixis has multiple origins within this complex; UK) using protocols from Shaw et al. (2007). We also
(2) apomixis is associated with reticulate evolution; and (3) apo- sequenced a fragment of the low copy nuclear gene pgiC
mictic lineages have sexual progenitors present in the complex. (approx. 600 nucleotides) for a sub-set of specimens.
Specimens were chosen in order to attain sufficient phylogen-
etic coverage, based on the relationships observed in the
M AT E R IA L S A ND M E T HO DS plastid data set. The generated fragment included the introns
spanning exons 14 and 16. The region was amplified using
Taxonomic sampling
primers pgiC14FN– 16RN (Ishikawa et al., 2002), and PCR
In total, 140 specimens were studied, representing approx. conditions as above were used. Successfully amplified pro-
50 % of the species in the Asplenium monanthes complex ducts were isolated by gel extraction using a QIAquick gel ex-
(Mickel and Smith, 2004; Supplementary Data, Table S1). traction kit (www.qiagen.com). Isolated bands were then
Dyer et al. — Apomixis and reticulate evolution in a fern complex Page 3 of 15

cloned using the Promega pGEMw-T Vector system to isolate had to consider the possibility of polymerase error, which
multiple copies present within the nuclear genome. Only one can result in two clones of the same allele being amplified
PCR amplicon per individual was cloned, which unfortunately with some base pair differences. In order not to confuse
increases the potential effect of PCR copy preference (Brysting these for different alleles, clones that had very similar
et al., 2011). Where possible, eight colonies were randomly sequences were only treated as unique if they were separated
chosen from the plated transformants and amplified by diluting by significant branch support (Bayesian posterior probability
colonies in 30 mL of sterile water and using 1.5 mL of this ≥0.6). If sequence differences between these clones had no
colony suspension as DNA template. The plasmid-specific significant branch support then they were treated as being
primers, puc/M13, were used to amplify the inserted fragments the same allele, and any observed base differences were in-
using the thermocycling conditions: 3 min initial denaturation ferred to be due to polymerase error. The disadvantage of
at 94 8C, followed by 35 cycles with 15 s denaturation at this method is that real alleles that are recovered only once
94 8C, 30 s annealing at 55 8C, 1 min elongation at 72 8C (due to PCR bias) may be disregarded if supported within a
and no final extension. Products were eluted in 40 mL of deio- clade of other alleles. No chimeric sequences were observed.
nized water and then purified using Montage PCR micro 96 According to these criteria, the full data set was reduced
Plates. Purified products were quantified using NanoDrop from 385 clones to 141, representing only one sequence per

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


ND-8000 (Thermo Scientific). Sequencing reactions were set unique clone per specimen (Supplementary Data, Table S2).
up using 2 ng of purified PCR product per 100 bp, 3.5 ml of Maximum likelihood (ML) analyses on the combined
buffer (2×), 1 mL of 1 mM stock solution of primer, 0.5 mL plastid and nuclear data sets were carried out in PhyML 3.0
of Big Dye version 1.1, and the final volume made up to with standard search options and parameters estimated simul-
10 mL with deionized water. Cycle sequencing reactions taneously from the data (Guindon et al., 2005, 2010). The nucleo-
were run on an Applied Biosystemsw 9700 96-well Dual tide substitution model with the least number of parameters that
Block Thermal Cycler (Life Technologies Ltd, Paisley, UK) best fit the data was determined using a likelihood ratio test
as follows: 28 cycles of 10 s denaturation at 96 8C, 5 s anneal- and the Akaike information criterion (AIC) as implemented in
ing at 50 8C, and 4 min elongation at 60 8C. Both forward and jModeltest version 3.7 (Posada and Crandall, 1998; Posada,
reverse strands were sequenced. The fragments were 2008), and branch support was estimated using 300 bootstrap
assembled and edited using Sequencher v4.8 (Gene Codes replicates (Supplementary Data, Table S4).
Corporation) and aligned manually using MacClade 4.08 Bayesian analysis (BY) was performed on the same two data
(Maddison and Maddison, 1989, 2005). All newly generated sets using substitution models according to the Bayesian infor-
sequences were confirmed using BLAST searches in mation criterion (BIC) determined in jModeltest (Supplementary
GenBank (Benson et al., 2011). The three plastid regions Data, Table S4). Analyses were run in MrBayes 3.1
were analysed both separately and in combination. All align- (Huelsenbeck and Ronquist, 2001; Ronquist and Huelsenbeck,
ments were checked visually for ambiguous regions, which 2003), with Markov Chain Monte Carlo (MCMC) run for 5
were excluded from all analyses. Gaps in the alignment were million generations and sampled every 500 generations to ap-
treated as missing data. All sequences are available from proximate the posterior probabilities of trees. The two analyses
GenBank (see Supplementary Data, Tables S1 and S2 for ac- were run simultaneously and a conservative burn-in phase of
cession numbers). 25 % was implemented in both data sets to disregard trees
prior to convergence on the ML. The remaining trees were
then compiled to give 7500 trees for each run, from which a
Phylogenetic analyses 50 % majority rule consensus was calculated.
Topological congruence among the three plastid regions
was evaluated using bootstrap resampling with maximum
Reproductive modes
parsimony (MP) (Felsenstein, 1985) using PAUP* 4.0b8
(Swofford, 1993, 2002). The bootstrap heuristic searches Diplospory involves a departure from the normal pathway of
were as follows: 1000 replicates, ten random sequence addi- sporogenesis in ferns, resulting in the production of unreduced
tions per replicate, TBR branch swapping and MULTrees (2n) spores (Döpp, 1932; Manton, 1950; Braithwaite, 1964). In
option on, collapse zero-length branches off, and saving all derived ferns (see Pryer et al., 2004), the number of spores per
trees. The strict bootstrap consensus trees of the individual sporangium often can be used as a reliable indicator for the oc-
regions were compared by eye to detect topological con- currence of diplospory because the number of spore mother
flicts among the three regions using the criteria outlined cells and the number of cell divisions are highly conserved
by Mason-Gamer and Kellogg (1996). We did not find (Schneider et al., 2009). Thus, the vast majority of sexual
any major topological incongruence between regions ferns produce 64 haploid spores per sporangium. In contrast,
(Supplementary Data, Figs S1– S3); therefore, the three apomicts exhibiting diplospory produce 32 unreduced spores
plastid regions were compiled into a single matrix for analysis. per sporangium (Döpp, 1932; Manton, 1950; Tryon, 1956;
Specimens that did not have sequence data for all three regions Knobloch, 1966; Vida, 1970).
were not included in the combined analysis. This data set was We determined the number of spores per sporangia for most
then reduced to represent unique haplotypes only of the collected specimens. Where possible, whole sporangia
(Supplementary Data, Table S3); this resulted in a data set were isolated, but in some instances no unopened sporangia
of 54 specimens. The pgiC data set was reduced to represent were found due to the age of some of the specimens. The spor-
unique clones per specimen. To designate unique clones we angia/spore samples were mounted using glycerine jelly, and
Page 4 of 15 Dyer et al. — Apomixis and reticulate evolution in a fern complex

spore number per sporangia was counted for up to ten sporan- Reticulate evolution
gia per specimen using a translucent light microscope.
In order to illustrate reticulate evolution (see Hörandl et al.,
Apogamy involves the initiation of sporophytic growth from
2005; Pirie et al., 2009; Huson and Scornavacca, 2011), we
a gametophyte without fertilization and indeed without change
reconstructed a phylogenetic network using SplitsTree
in chromosome numbers (Manton, 1950; Regalado Gabancho
version 4.12.6 (Huson and Bryant, 2006). Plastid and nuclear
et al., 2010). To confirm the presence/absence of apogamous
data sets were reduced to a sub-set of compatible samples
growth, we carried out experiments in which prothalli were
(Supplementary Data, Tables S5 and S6). One sample for
grown from the spore, allowing direct observations on the gam-
each unique combination of plastid vs. nuclear loci was
etophyte stage of the fern life cycle. Spores were filtered and
included. However, unique combinations did not include
sterilized using diluted HCl and then applied to the surface
those that varied in nuclear copy number, only those that
of sterilized potted compost and sealed in a re-sealable
varied in nuclear copy distribution, i.e. nuclear copies that oc-
plastic bag. The pots were kept in dappled light to shaded con-
curred in different clades. The nuclear data set was further
ditions at 10– 20 8C. The pots were monitored regularly and
divided into four sub-sets, so that each specimen was repre-
growth typically began after 4 weeks and took between 8
sented only once in a data set, i.e. multiple copies were
and 24 weeks until prothalli were mature. For determination

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


present in separate data sets. The optimal trees resulting
of reproductive mode, select mature prothalli were mounted
from ML bootstrap analyses of the plastid and nuclear data
on slides and observed in water. Euparal mounting medium
sets were used as input files. A reticulation network was com-
(ANSCO Laboratories, Manchester, UK) was used to obtain
puted using the ‘RECOMB2007’ method as implemented in
images. The prothalli were then examined for the presence
SplitsTree (Klöpper and Huson, 2008). The resulting
of antheridia and archegonia and evidence of sporophyte
network was then used to inform the manual construction of
emergence (see examples in Fig. 1). Where sporophytes oc-
a reticulation cladogram, which summarizes the reticulate rela-
curred, their emergent position, character and shape were
tionships between and within species of this complex.
recorded. Observation of the outgrowth of the sporophyte
from vegetative cells of the gametophyte is considered as the
best approach to identify apogamous reproduction in ferns R E S U LT S
(Huang et al., 2011), although apogamy can be induced in
Plastid phylogenetic analyses
sexual fern gametophytes under some conditions (Somer
et al., 2009; Kawakami et al., 2011). Analysis of the combined plastid data set using BY (Fig. 2)
and ML (Supplementary Data, Fig. S4) resulted in trees with
similar topology: see Fig. 2 for posterior support values, and
Supplementary Data, Fig. S4 for bootstrap support values.
Ploidy level
Asplenium formosum ( pFO) was found to be the sister to the
In ferns, the size of spores and stomatal guard cells can often remainder of the A. monanthes complex, which formed a polyt-
be employed to infer differences in ploidy levels among closely omy consisting of three well-supported clades: the A. castaneum
related species (Barrington et al., 1986; Huang et al., 2006). clade, the A. resiliens clade and the A. monanthes clade (Fig. 2).
Spore measurements were successfully employed in several The A. castaneum clade consisted of distinct lineages of
recent studies on the evolution of apomixis in ferns (Beck A. polyphyllum (pPO) and A. soleirolioides (pSO), two
et al., 2010; Sigel et al., 2011) despite the occurrence of A. castaneum sub-clades (pCA1 and pCA2), and two lineages
aborted spores, which may add some error to the estimate containing a paraphyletic A. fibrillosum (pFI) (Fig. 2). In ana-
range (Huang et al., 2011). In the A. trichomanes complex, lysis of the psbA–trnH data set, A. blepharodes is recovered
the sister group to the A. monanthes complex, it was possible within the A. castaneum clade (Supplementary Data, Fig. S1).
to distinguish ploidy level according to spore size classes, The A. resiliens clade consisted of three lineages: two distinct
with a mean spore length of 29.4 mm in diploids and 41.6 mm A. resiliens sub-clades (pRE1 and pRE2), and one mixed sub-
in tetraploids (Moran, 1982; Tutin et al., 1993). Here, spores clade consisting of A. palmeri and A. aff. heterochroum speci-
were measured for a representative sampling of 49 specimens mens (pPA) (Fig. 2). Three exceptions were observed, but not
(i.e. maximizing phylogenetic coverage). Spores from individ- present in the combined analysis: two A. palmeri specimens
ual specimens were mounted using glycerine jelly and measured (CJR2494 and ES486) were recovered in the pRE2 sub-clade
for length and width using AxioVision on a calibrated light (Supplementary Data, Figs S1–S3), and one A. resiliens
microscope (v4.8.2, www.zeiss.com). An average of 25 spores specimen (CJR2504) was recovered in sub-clade pPA
per specimen were measured, although with some specimens (Supplementary Data, Fig. S3). The A. monanthes clade con-
very few spores were recovered. Box plots of width and sisted of two isolated accessions designated spec.nov.1 (pSP1)
length were plotted separately to determine the variance and spec.nov.2 (pSP2), and three well-supported monophyletic
within each. Ploidy level was inferred based on statistical ana- groups. These included two sub-clades of A. monanthes (pMO1
lysis of spore length variation. Analyses included a one-way and pMO2), and one sub-clade of A. hallbergii (pHA) (Fig. 2).
analysis of variance (ANOVA) and Tukey–Kramer analyses at
95 % confidence levels using R (R Development Core Team,
Nuclear phylogenetic analyses
2011). Ploidy level was also estimated based upon calibrations
of the mean values and variation in spore measurements to The BY (Fig. 3) and ML (Supplementary Data, Fig. S5)
known spore measurements and ploidy levels in the analysis of the pgiC 14FN – 16RN data set (approx. 600 bp)
A. trichomanes complex (Moran, 1982; Tutin et al., 1993). resulted in trees of similar topology. However, the sub-clades
Dyer et al. — Apomixis and reticulate evolution in a fern complex Page 5 of 15

A B
500 µm

500 µm

C D

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


500 µm 1000 µm

E F

500 µm 200 µm

G 500 µm H

500 µm

F I G . 1. Sexual vs. apogamous prothalli. Pictures of cultivated prothalli from different specimens (Supplementary Data, Table S7): (A) sexual A. formosum,
RD28 – the dashed arrow show the archegonia; (B) apogamous A. monanthes, RD110; (C) apogamous A. monanthes, RD101b; (D) apogamous A. resiliens,
RD127; (E) apogamous A. resiliens, RD128; (F) apogamous A. monanthes, RD132; (G) apogamous A. resiliens RD107; (H) apogamous A. monanthes,
RD17. Filled circles indicate apical meristems, and sporophytes are indicated by filled squares. Solid arrows indicate apogamous growth.

recovered were not analogous to the sub-clades present in the one nuclear copy with the A. palmeri/A. aff. heterochroum
plastid-based phylogenetic trees. There was evidence of reticu- sub-clade (nPA).
late evolution as individual specimens often had multiple Sequences obtained from A. castaneum specimens were
copies of nuclear sequences that were distributed in different recovered in three different sub-clades, two of which (nPO
sub-clades. Reticulate evolution generally was restricted to and nSO) were sister to the remainder of the complex.
species within the three major clades (A. castaneum, Asplenium polyphyllum sequences were also found in these
A. resiliens and A. monanthes) identified in the plastid ana- two sub-clades, and A. soleirolioides sequences were present
lysis, although one A. fibrillosum accession (RD10b) shared in sub-clade nSO. The remainder of the complex formed an
Page 6 of 15 Dyer et al. — Apomixis and reticulate evolution in a fern complex

Diellia erecta
1

1 A. normale 2
1 A. hobdyi
0·81 A. normale 1 Outgroup
1 A. trich subsp trich
1 A. trich subsp quad

A. trich subsp inex

A. formosum IJ2436
1
A. formosum RD27

A. formosum RD28
pFO
A. formosum RD33

A. fibrillosum RD10b
1 pFI
A. fibrillosum RD22
1 A. castaneum RD49 *
1 pCA1

A. castaneum clade
A. castaneum RD54

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


1 A. castaneum RD46.b *
1
0·7 A. castaneum RD47
pCA2
A. castaneum RD52
1 A. soleirolioides RD71

A. soleirolioides RD114 pSO


1 A. polyphyllum RD95 *
0·64
1 A. polyphyllum RD115

A. polyphyllum ALG08-146
pPO
0·97 1 A. polyphyllum Mehltreter

A. palmeri RD130
1

A. resiliens clade
A. aff. heterochroum RD9a * pPA
1 A. aff. heterochroum RD75
1

1 A. resiliens RD63 *

1 A. resiliens RD72
pRE2
A. resiliens RD128
1
A. resiliens RD4a
pRE1
A. resiliens RD121
Spec. nov. 1. JM1339
pSP1
A. hallbergii RD23 *

A. hallbergii RD81

A. aff. hallbergii RD85


1 1 pHA
A. aff. hallbergii RD90

A. aff. hallbergii RD111

A. aff. hallbergii RD112

A. aff. hallbergii RD136


1 Spec. nov. 2. SK10151 pSP2
0·68
A. monanthes RD16
1
A. monanthes RD24 *
A. monanthes clade

A. monanthes RD99 pMO2


1
A. monanthes ALG08-145 *
0·99 A. monanthes RD45
0·91

0·95 A. monanthes RD28b *

A. monanthes RD117a *
1 A. monanthes RD80 *

A. monanthes RD2a *

A. monanthes RD74
1
A. monanthes RD76
A. monanthes RD1a
0·98
A. monanthes RD17 pMO1
A. monanthes RD26

0·65 A. monanthes RJ11 *

A. monanthes RD19 *
1
A. monanthes RD20

A. monanthes RD8a

0·98 A. monanthes RD10a

A. monanthes RD110
Dyer et al. — Apomixis and reticulate evolution in a fern complex Page 7 of 15

unresolved polytomy. The third sub-clade of A. castaneum Extensive reticulation was observed between the
(nCA) was recovered within this polytomy as the weakly sup- A. monanthes and A. hallbergii clades designated nMO1,
ported ( posterior probability ¼ 0.75) putative sister to a clade nMO2 and nHA. Asplenium monanthes specimens with the
comprised of A. hallbergii clone sequences, and one plastid pMO1 haplotype had up to two nuclear copies in
A. monanthes sequence (nHA). The polytomy also included: clade nMO1 and a third in nMO2. Asplenium hallbergii speci-
one clade of A. fibrillosum clone sequences (nFI); one clade mens showed reticulation, having up to two nuclear copies in
comprising sequences of A. monanthes and spec.nov.1 the nHA clade, and a single copy in clade nMO2. Specimen
(nMO1); one clade comprising sequences of A. monanthes, RD23 was the exception, having three copies only in the
A. hallbergii and one sequence of spec.nov.2 (nMO2); and a nHA clade. One of the A. aff. hallbergii specimens (RD85)
clade comprising sequences of A. formosum, A. resiliens, had four nuclear copies, two of which were derived from the
A. palmeri, A. aff. heterochroum and one sequence of nHA clade and two from the nMO2 clade. In addition to
A. fibrillosum. The latter clade was a polytomy consisting of these clear cases of reticulate evolution, there were a few spe-
four sub-clades: nFO, formed of A. formosum sequences cimens that showed no evidence of reticulation. Asplenium
only; nRE1, formed of A. resiliens sequences and one monanthes specimens with the plastid haplotype pMO2 were
A. palmeri sequence; nRE2, formed of A. resiliens sequences monophyletic (2 – 3 copies) in clade nMO2, and spec.nov.2

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


only; and nPA, formed of A. palmeri and A. aff. heterochroum (of sub-clade nMO2) and spec.nov.1 (of sub-clade nMO1)
sequences, one sequence of A. resiliens and one sequence of had one and two copies of pgiC, respectively. The exception
A. fibrillosum (see Fig. 3 for details). in pMO2 was specimen RD99, which had a nuclear copy in
each of the clades nMO1, nMO2 and nHA.

Evidence for reticulate evolution


Evidence for diplospory and apogamy
The unique clones (i.e. nuclear copies) within individual
specimens were interpreted as different alleles. The distribu- Diplospory was inferred for taxa producing 32 spores per
tion of these unique clones showed evidence of reticulation sporangium; these included A. monanthes, A. hallbergii, A.
between clades (Fig. 3; Supplementary Data, Table S5). aff. hallbergii, A. resiliens, A. palmeri and A. aff. hetero-
These reticulate relationships are summarized in a hybridization chroum (Table 1; Supplementary Data, Table S5). No diplosp-
network (Fig. 4; Supplementary Data, Fig. S6 and Table S6). ory was evident in A. formosum, A. fibrillosum, A. castaneum,
The nuclear copy distribution of A. formosum specimens A. polyphyllum, A. soleirolioides, spec.nov.1 or spec.nov.2, all
and A. fibrillosum specimen RD22 showed no evidence of re- of which had counts of 64 spores per sporangium. One speci-
ticulation, and these specimens only had one or two copies. men of A. aff. hallbergii (RD85) yielded counts of both 32 and
The A. fibrillosum accession designated RD10b had four 64 spores per sporangium from the same individual; all of
copies, one of which was recovered in the nPA clade. these spores appeared viable.
Asplenium castaneum specimens appeared in three forms Our spore germination experiments showed that
according to copy number and distribution. One showed evi- A. formosum produced normal sexual organs, and showed no
dence of reticulation, sharing one copy in sub-clade nCA evidence of apogamous growth (Fig. 1A). All studied speci-
and the other in sub-clade nPO. Of the remaining inferred mens of A. hallbergii, A. aff. heterochroum, A. monanthes
A. castaneum forms, one had only a single clone sequenced and A. resiliens produced gametophytes without sexual
( present in sub-clade nCA), and the other showed no evidence organs (see Fig. 1B –H). Sporophytes in these taxa emerged
of reticulation, having a single copy in sub-clade nSO. either through a callus-like sporophytic growth from the
Asplenium polyphyllum specimens were present in two central or lower region of the gametophyte (Fig. 1B– E), or
forms: one with a single copy in sub-clade nPO, whereas the by extension of the apical meristem (Fig. 1E – G). In some
other showed reticulation with one copy in sub-clade nPO cases, prothalli were deformed where the apical cell appears
and a second copy in sub-clade nSO. Asplenium soleirolioides to have become detached (Fig. 1H). Apomictic reproduction
showed no reticulation. Asplenium resiliens specimens showed appeared to be obligate in all cases, with apogamous prothalli
clear evidence of reticulation, with all except one specimen exclusive to specimens exhibiting evidence of diplospory
having one, two or three copies distributed between sub-clades (Table 1; Supplementary Data, Table S7).
nRE1 and nRE2. The exception was specimen CJR2504,
which had one sequence in sub-clade nPA, and two in sub-
Inferring ploidy level
clade nRE1. Most specimens of A. palmeri and A. aff. hetero-
chroum showed no evidence of reticulation. There was one ex- Analysis of spore length variation between the different
ception, whereby one A. palmeri specimen (CJR2494) had species showed that the spores of A. formosum are significantly
three copies, two distributed in sub-clade nPA and one in sub- smaller than those of all other species (P , 0.001), with
clade nRE1. Details are shown in Fig. 3 and a summary is pre- the exception of spec.nov.1 and spec.nov.2 (Table 1, Fig. 5;
sented in Fig. 4. Supplementary Data, Table S8 and Fig. S7). Spec.nov.1 and

F I G . 2. Bayesian phylogenetic tree based on the combined plastid data. Posterior support values of P . 0.6 are shown. Tips are labelled with species names
followed by a voucher accession, and colour-coded according to species. Sub-clades are abbreviated and summarized to the right of the tree. Abbreviations
for sub-clades correspond to those used in the text. Asterisks indicate tips representing multiple haplotypes (Supplementary Data, Table S3). Filled black
circles indicate apomictic specimens, and open circles indicate specimens inferred to have a sexual mode of reproduction.
Page 8 of 15 Dyer et al. — Apomixis and reticulate evolution in a fern complex
1 A. trichomanes NA5. CL1
1 A. trichomanes NA5. CL4
1 A. viride NA8. CL1 Outgroup
D. erecta NA3. CL3
A. soleirolioides RD71. C2/2
1 0·95 A. castaneum RD49. C1/1
0·99
A. soleirolioides RD71. C1/2
A. polyphyllum RD95. C2/2 nSO
1 A. polyphyllum RD115. C2/2
A. castaneum RD46b. C1/2
1 A. castaneum RD47. C1/2
A. castaneum RD52. C1/2
A. polyphyllum RD98. C1/1
0·98 A. polyphyllum RD95. C1/2 nPO
A. polyphyllum RD115. C1/2
1 A. polyphyllum ALG08-146. C1/1
A. polyphyllum Mehltreter. C1/1
1 A. fibrillosum RD10b. C4/4
0·75 A. fibrillosum RD10b. C2/4
0·92
A. fibrillosum RD10b. C3/4
A. fibrillosum RD22. C1/2
nFO
A. fibrillosum RD22. C2/2
1 A. formosum ES1398. C1/2
1 A. formosum IJ2436. C1/1
A. formosum RD33. C1/1
1 A. formosum RD28. C2/2
0·76 1 A. formosum RD27. C1/1 nFO
A. formosum RD28. C1/2
1 A. formosum MK12699. C1/2
A. formosum MK12699. C2/2
1 A. resiliens RD127. C1/2
0·86 A. resiliens RD3b. C1/2
0·94 1 A. resiliens RD63. C1/3
1 A. resiliens RD4a. C1/2
A. resiliens RD121. C1/2
nRE2
1 0·77 A. resiliens RD121. C2/2
A. resiliens RD128. C1/3
A. resiliens RD128. C2/3

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


A. resiliens CJR2504. C1/3
1 A. palmeri CJR2494. C1/3
1 1 A. aff. heterochroum RD9a. C2/2
A. fibrillosum RD10b. C1/4
1 A. palmeri RD130. C1/1 nPA
0·94 A. palmeri CJR2494. C2/3
A. aff. heterochroum RD9a. C1/2
A. aff. heterochroum RD75. C1/1
A. resiliens CJR2504. C3/3
1 A. palmeri CJR2494. C3/3
A. resiliens RD4a. C2/2
0·97 0·97 A. resiliens CJR2504. C2/2
A. resiliens RD63. C3/3
A. resiliens RD127. C2/2 nRE1
0·96 A. resiliens RD72. C2/2
A. resiliens RD3b. C2/2
0·98 A. resiliens RD63. C2/3
A. resiliens RD72. C1/2
A. resiliens RD128. C3/3
0·99 A. castaneum RD54. C1/1
1 A. castaneum RD46b. C2/2
A. castaneum RD47. C2/2
nCA
0·75 A. castaneum RD52. C2/2
A. hallbergii RD18. C2/3
1 A. hallbergii RD23. C3/3
1 A. hallbergii RD81. C2/3
A. monanthes RD99. C2/3
A. hallbergii RD111. C1/3
A. hallbergii RD23. C2/3
0·88 A. hallbergii RD18. C1/3
1 1 A. hallbergii RD23. C1/3
A. hallbergii RD81. C1/3 nHA
A. hallbergii RD112. C1/3
A. hallbergii RD136. C1/2
0·97 A. aff. hallbergii RD85. C1/4
1 A. aff. hallbergii RD90. C1/3
A. hallbergii RD112. C2/3
A. hallbergii RD111. C2/3
1 A. aff. hallbergii RD85. C2/4
A. aff. hallbergii RD90. C2/3
Spec. nov. 2. SK10151 C1/1
A. monanthes ALG08-145. C1/3
0·99 A. monanthes RD16. C2/3
A. hallbergii RD81. C3/3
A. aff. hallbergii RD85. C3/4
1 1 A. monanthes RD16. C1/3
A. monanthes RD24. C1/2
A. monanthes RD45. C1/2
A. monanthes RD53. C1/2
A. monanthes ALG08-145. C2/3
A. monanthes RD8a. C2/2
1 A. monanthes RD10a. C3/3
A. monanthes RD16. C3/3
A. monanthes RD17. C2/2
0·62 A. monanthes RD20. C2/2
A. monanthes RD25. C3/3
A. monanthes RD26. C3/3
0·97 A. monanthes RD28b. C3/3 nMO2
A. monanthes RD74. C2/2
A. monanthes RD76. C3/3
A. monanthes RD110. C3/3
0·96 A. monanthes RD19. C3/3
A. hallbergii RD112. C3/3
A. monanthes ALG08-145. C3/3
A. hallbergii RD18. C3/3
A. monanthes RD24. C2/2
A. monanthes RD45. C2/2
A. monanthes RD53. C2/2
0·97 A. monanthes RD80. C3/3
A. aff. hallbergii RD85. C4/4
A. aff. hallbergii RD90. C3/3
A. monanthes RD97. C2/2
A. hallbergii RD111. C3/3
A. monanthes RD117a. C2/3
A. hallbergii RD136. C2/2
0·97 A. monanthes RD1a. C2/2
A. monanthes RD2a. C3/3
A. monanthes RD99. C3/3
1 Spec. nov. 1 JM1339. C1/2
Spec. nov. 1 JM1339. C2/2
A. monanthes RD74. C1/2
A. monanthes RD99. C1/3
0·83 1 A. monanthes RD2a. C2/3
A. monanthes RD80 C2/3
1 A. monanthes RD117a. C1/3
0·91 A. monanthes RD117a. C3/3
1 A. monanthes RD25. C1/3
1 A. monanthes RD28b. C1/3
A. monanthes RD97. C1/2
1 A. monanthes RD76. C1/3
0·98 A. monanthes RD1a. C1/2
A. monanthes RD2a. C1/3
A. monanthes RD26. C2/3
A. monanthes RD80. C1/3 nMO1
0·97 A. monanthes RD25. C2/3
A. monanthes RD28b. C2/3
0·98 A. monanthes RD10a. C2/3
1 A. monanthes RD17. C1/2
A. monanthes RD26. C1/3
0·73 A. monanthes RD110. C2/3
1 A. monanthes RD20. C1/2
A. monanthes RD19. C1/3
0·71 A. monanthes RD19. C2/3
A. monanthes RD76. C2/3
A. monanthes RD110. C1/3
A. monanthes RD8a. C1/2
A. monanthes RD10a. C1/3

F I G . 3. Bayesian phylogenetic tree based on the nuclear data. Posterior support values of P . 0.6 are shown. Tips are labelled with species names, followed by a
voucher accession (e.g. RD112), and indication of the clone identifier out of the total number of clones (e.g. C2/3 ¼ clone no. 2 out of three unique clones; see
text for details). Tips are colour-coded according to species. Sub-clades are abbreviated and summarized to the right of the tree.
Dyer et al. — Apomixis and reticulate evolution in a fern complex Page 9 of 15

A. formosum (1–2) pFO, nFO, nFO

A. resiliens clade
A. resiliens R1. (2–3) pRE1, nRE1, nRE2, nRE2

A. resiliens R2. (2–3) pRE2, nRE1, nRE1, nRE2

A. palmeri CJR2494. (3) pRE2, nPA, nPA, nRE1

A. resiliens CJR2504. (3) pPA, nRE1, nRE1, nPA


A. palmeri (1) pPA, nPA, nPA
A. aff. heterochroum (1–2)

A. fibrillosum RD10b. (4) pFI, nFI, nFI, nFI, nPA

A. fibrillosum RD22. (2) pFI, nFI, nFI

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


A. soleirolioides (2) pSO, nSO, nSO

A. castaneum clade
A. castaneum RD49. (1) pCA1, nSO
A. polyphyllum (2) pPO, nPO, nSO

A. polyphyllum (1) pPO, nPO

A. castaneum (2) pCA2, nCA, nPO

A. castaneum RD54. (1) pCA1, nCA

A. aff. hallbergii (3–4) pHA, nHA, nHA, nMO2, nMO2

A. hallbergii (2–3) pHA, nHA, nHA, nMO2

A. monanthes clade
Spec. nov. 2. (1) pSP2, nMO2

MO2. A. monanthes (2–3) pMO2, nMO2, nMO2, nMO2


A. monanthes RD99. (3) pMO2, nMO1, nMO2, nHA

MO1. A. monanthes (2–3) pMO1, nMO1, nMO1, nMO2


Spec. nov. 1. (2) pSP1, nMO1, nMO1

F I G . 4. A reticulation network illustrating the evolutionary history of the A. monanthes complex. This summarizes a hybridization network (Supplementary
Data, Fig. S6) performed in SplitsTree (Huson and Bryant, 2006) and illustrates hypothetical hybrid relationships observable by comparison between nuclear
and plastid trees (Figs 1 and 2). The terminals illustrated are inferred as different species forms. Where copy numbers for one-species forms are variable, we
have included the range of values in parentheses after the species label. Branches are colour-coded according to species. Filled black circles indicate apomictic
species, and open circles indicate species inferred to have a sexual mode of reproduction. The plastid group and nuclear copy distribution are summarized adjacent
to terminal labels. The dashed red line indicates inferred association of this spec.nov.1 with the MO1 A. monanthes lineage.

spec.nov.2 had significantly smaller spore sizes (P , 0.001) species (,0.05, with the exception of A. aff. heterochroum,
than all other specimens within the A. monanthes MO1 A. soleirolioides and A. fibrillosum). However, this variation
clade and the A. monanthes MO2 clade, respectively. Mean was difficult to classify into distinguishable groups and so
spore lengths in A. formosum, spec.nov.1 and spec.nov.2 inference of ploidy level was not possible. All remaining
were ,32 mm, comparable with those of sexual diploids in species did have large mean spore sizes (35 – 52 mm), and
the related A. trichomanes complex (see Fig. 5). We infer were significantly larger (P , 0.001) than inferred diploid
from these results that these three taxa represent sexually re- species (A. formosum, spec.nov.1 and spec.nov.2)
producing diploids. (Table 1, Fig. 5; Supplementary Data, Table S8 and Fig.
The remainder of the A. monanthes complex showed sig- S7). We therefore infer that the remaining species were all
nificant spore size variation between (,0.001) and within polyploids.
Page 10 of 15 Dyer et al. — Apomixis and reticulate evolution in a fern complex

TA B L E 1. Spore characteristics and apogamy as recovered in that the power of these inferences is potentially limited by
this study by investigating spore characters, and cultivation and PCR copy preference (Brysting et al., 2011), or processes
observations of gametophytes such as diploidization, which could produce misleading
results.
Mean spore Mean spore According to spore number and cultivation experiments,
Spore length range width range
Species number Apogamy (mm) (mm)
apomixis has evolved in the A. monanthes and A. resiliens
clades but appears to be absent in the A. castaneum clade.
A. castaneum 64 – 36.54–46.53 28.39– 37.69 Several previously published chromosome counts indicate
A. fibrillosum 64 – 44.34–47.90 34.85– 37.97 that A. formosum is a sexual diploid (2n ¼ 72) (Manton, 1959;
A. formosum* 64 No 27.85–32.20 20.64– 25.02 Walker, 1966b; Ghatak, 1977; Ammal and Bahavanandan,
A. hallbergii 32 Yes 39.15–46.83 27.07– 32.84 1991; Guillén and Daviña, 2005), which is consistent with the
A. aff. hallbergii 32/64 Yes 43.32–44.97 30.45– 31.86
A. aff. 32 Yes 37.43–38.62 25.89– 27.83
spore measurements and gametophyte experiments reported
heterochroum here. We did not find any evidence to support suggestions of
A. monanthes 32 Yes 38.02–47.38 26.68– 32.08 A. formosum as one parent of A. monanthes.
(MO1)* Fern spore size has been successfully used to determine

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


A. monanthes 32 Yes 38.45–48.54 26.63– 34.72 ploidy level in several recent studies on apomixis in xeric
(MO2)*
A. palmeri † 32 – – – ferns (Beck et al., 2010; Sigel et al., 2011). Also, studies cor-
A. polyphyllum 64 – 35.87–39.87 25.06– 30.06 relating cytofluorometry and spore measurements in the
A. resiliens † 32 Yes 39.78–44.70 28.17– 31.10 A. trichomanes complex have indicated that spore size pro-
A. soleirolioides 64 – 36.88 27.55 vides a useful estimator of ploidy level among the black-
Spec.nov.1 64 – 31.26 21.42
Spec.nov.2 64 – 29.22 20.77 stemmed rock spleenworts (Ekrt and Stech, 2008). However,
it is not always clear if spore size is reliable for inferring
Mean spore length and width are reported as ranges where multiple ploidy level (Barrington et al., 1986), and in this study we
specimens were measured per species, and as single values where only one were only able to use spore size measurements to distinguish
specimen was measured per species. between diploid and polyploid specimens, and were not able
* Taxa for which chromosome counts are available: A. formosum 2n ¼ 72 to infer polyploid levels.
(Manton, 1959; Walker, 1966b; Ghatak, 1977; Ammal and Bahavanandan,
1991; Guillén and Daviña, 2005); A. monanthes n ¼ 2n ¼ 108 and 2n ¼ 144
(Manton, 1950; Manton and Vida, 1968; Wagner et al., 1970; Tryon et al.,
1973; Smith and Mickel, 1977; Manton et al., 1986) and A. resiliens n ¼
The A. resiliens clade
2n ¼ 108 (Morzenti and Wagner, 1962; Wagner, 1963, 1966; Morzenti, We report that A. palmeri and A. heterochroum-like mor-
1966; Wagner and Wagner, 1966; Walker, 1966b; Wagner et al., 1970;
Windham, 1983).
photypes produce 32 spores per sporangium and show an apo-

Wagner et al. (1993) reported 64 spores for A. palmeri whereas Mickel mictic mode of reproduction. An unpublished chromosome
and Smith (2004) show no spore count. count of n ¼ 2n ¼ 108 for A. palmeri (M. D. Windham,
pers. comm.) confirms that at least some specimens assigned
to this taxon are apomictic triploids. Wagner et al. (1993)
DISCUSSION reported 64 spore counts per sporangium in the Flora of
North America treatment of this species, whereas Mickel and
Evolutionary relationships and reticulate evolution
Smith (2004) did not record the spore number for this
Plastid sequence data identified four main lineages within this species. Asplenium heterochroum was previously known as a
complex, and some of these lineages have also been recovered sexual polyploid (4x and 6x) with 64 spores (Morzenti and
in the pgiC data set. The backbone of the pgiC-based tree is a Wagner, 1962; Wagner, 1963, 1966; Morzenti, 1966;
polytomy, and this nuclear marker only provides limited evi- Wagner and Wagner, 1966; Walker, 1966b; Wagner et al.,
dence to reconstruct relationships among these species. This 1970). Our conflicting results may reflect reported taxonomic
region has not commonly been used in ferns, and other difficulties (Stolze, 1981; Tryon et al., 1993; Stolze et al.,
studies have encountered similar problems resolving species 1994) and suggest that these species require further cytological
evolutionary relationships (Juslén et al., 2011). However, study. This complex was not exhaustively sampled, and infer-
pgiC is better suited to trace reticulate evolution when com- ence of reticulate relationships and apomictic origins calls for
pared with the plastid marker data (see also Juslén et al., caution.
2011).
In the plastid phylogenetic tree, A. formosum was recovered
The A. castaneum clade
as the putative sister to three clades that broadly correspond to
taxonomic species assemblages of A. castaneum, A. resiliens The distinction between A. polyphyllum and A. castaneum
and A. monanthes. Relationships among these three clades has previously posed taxonomic difficulties (Mickel and
are unresolved, though the lineages themselves are strongly Smith, 2004), and our reconstruction of phylogenetic relation-
supported. The nuclear phylogenetic tree shows evidence of ships was unable to resolve these. Our spore size data provide the
extensive reticulate evolution within these clades but not first evidence for polyploidy in this species complex. Spore data
between them, except for one hybrid specimen originally iden- indicate that A. fibrillosum, A. polyphyllum, A. soleirolioides and
tified as A. fibrillosum (Figs 3 and 4). We are able to infer A. castaneum are polyploids.
parentage of putative hybrids based on the phylogenetic distri- The plastid phylogenetic tree identifies species lineages
bution of nuclear clones. However, it is important to consider broadly corresponding to proposed morphological groups.
Dyer et al. — Apomixis and reticulate evolution in a fern complex Page 11 of 15

60

50
Length (µm)

40 3x

2x

30 x

*
A. formosum RD28
A. formosum RD33
Spec. nov. 1. JM1339
A. monanthes MO1. RD73
A. monanthes MO1. RD70
A. monanthes MO1. RD17
A. monanthes MO1. RD117
A. monanthes MO1. RD80
A. monanthes MO1. RD74
A. monanthes MO1. RD110
A. monanthes MO1. RD20
A. monanthes MO1. RD103
A. monanthes MO1. RD2a
Spec. nov. 2. SK10151
A. monanthes MO2. RD24
A. monanthes MO2. RD16
A. monanthes MO2. RD45
A. monanthes MO2. RD41
A. monanthes MO2. ALG08-145
A. monanthes MO2. RD53
A. monanthes MO2. RD99
A. hallbergii RD81
A. hallbergii RD137
A. aff. hallbergii RD90
A. aff. hallbergii RD85
A. hallbergii RD112
A. hallbergii RD118
A. hallbergii RD18
A. hallbergii RD136
A. hallbergii RD125b
A. aff. heterochroum RD75
A. aff. heterochroum RD9a
A. heterochroum ML601
A. resiliens RD121
A. resiliens RD63
A. resiliens RD72
A. resiliens RD4a
A. soleirolioides RD71
A. polyphyllum RD95
A. polyphyllum RD116
A. polyphyllum RD115
A. castaneum RD50b
A. castaneum MC5271
A. castaneum RD54
A. castaneum RD52
A. castaneum RD49
A. castaneum RD47
A. fibrillosum RD22
A. fibrillosum RD10b

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


Specimen

A. formosum A. monanthes A. resiliens A. castaneum

F I G . 5. Boxplots illustrating variation in spore length of species/specimens of the A. monanthes complex. Each boxplot is labelled below with its corresponding
species name and specimen voucher, and coloured according to species as in Fig. 2. Boxplots are grouped together according to clades and ordered from lowest to
highest mean lengths. Each boxplot represents the variance of measurements of spores within each specimen, the thick horizontal line is the median, the box
indicates the variation observed between the 25th and 75th percentiles, the whiskers show the variance range, and small circles identify extreme outliers.
Dashed lines running across the graph illustrate average spore measurements in comparison with the ploidy levels of the spores, based on average counts of
triploid A. resiliens (spore ploidy ¼ 3x) specimens, hexaploid A. heterochroum (spore ploidy ¼ 3x) specimens (Morzenti, 1966) and diploids and tetraploids
of the A. trichomanes complex (spore ploidy ¼ x and 2x, respectively) (Tutin et al., 1993). Filled black circles indicate apomictic specimens that produce 32
spores per sporangium. Open circles indicate specimens that produce 64 spores per sporangium and are inferred to have a sexual mode of reproduction. The
asterisk indicates a specimen that produces both 32 and 64 spores per sporangium.

Asplenium castaneum comprises two sub-clades, with members Both the MO1 and A. hallbergii lineages consist of allopoly-
of the pCA1 sub-clade more similar to A. fibrillosum in appear- ploid hybrids formed with the MO2 lineage, which itself
ance, and representatives of the pCA2 sub-clade more likely to be appears to be autoploid (Figs 4 and 6). Specimen RD99 is
misidentified as A. polyphyllum. However, the nuclear phylogen- an exception to the inference of autoploidy in lineage MO2,
etic tree indicates extensive reticulation and multiple hybrid for- as it contains clones from three lineages. This could indicate
mation (Fig. 4). The observed pattern is similar in complexity PCR bias in the cloning method, but we would expect that
to the pervasive reticulation observed in the A. trichomanes the high sample number within this lineage would overcome
polyploid complex (Lovis, 1977). Denser sampling of speci- this bias. RD99 could therefore represent a case of gene flow
mens and additional cytological information (e.g. chromosome from an apomictic lineage, through hybridization with a
counts) are required to reconstruct the relationships within this sexual relative, via functional antheridia.
widespread complex. The A. aff. hallbergii specimens (RD85 and RD90) are
nested within the A. hallbergii lineage, but appear physically
larger in size and show up to four nuclear copies, implying
The A. monanthes clade
that they could be tetraploid variants (Figs 4 and 5). One of
This clade comprises two A. monanthes lineages (MO1 and these specimens (RD90) produced both 32- and 64-spored
MO2) and one A. hallbergii (HA) lineage, which are supported sporangia. Both types of spores appeared normal and viable.
in both plastid and nuclear genomes. It also contains two This would suggest that RD90 is producing both reduced
accessions (spec.nov.1 and spec.nov.2), which are isolated in and unreduced spores, and thus might be a facultative
the plastid genome but associated with MO1 and MO2 in apomict, though separate cultivation experiments are neces-
the nuclear genome. Specimens of A. monanthes and sary to test this hypothesis.
A. hallbergii are apomictic and appear to be triploid and pos- Spec.nov.1 (JM1339) and spec.nov.2 (SK10151) are inferred
sibly tetraploid, as indicated by previous chromosome counts to be sexual diploids, and have affinities for the nuclear
for A. monanthes (Manton, 1950; Smith and Mickel, 1977). A. monanthes lineages nMO1 and nMO2, respectively
Page 12 of 15 Dyer et al. — Apomixis and reticulate evolution in a fern complex

(1) The weak resolution among clades and putative sister


HHMM relationships of the A. monanthes clade and A. resiliens
clade may be seen as arguments for a single origin of apo-
A. aff. hallbergii mixis in these ferns. This would then imply that the
various apomictic lineages were the result of diversifica-
tion within an asexual lineage (Fontaneto et al., 2007).
However, this would require the sexual taxa (spec.nov.1
and spec.nov.2) within the A. monanthes clades to have
HHM MMM PPM undergone evolutionary reversals from asexual to sexual
A. hallbergii A. monanthes (MO2) A. monanthes (MO1) reproduction (Figs 1 and 4). To our knowledge, this
process has not been documented for apomictic ferns yet.
(2) The alternative hypothesis of multiple origins is supported
by evidence for the association of apomixis with reticulate
HH MM PP evolution in both the A. resiliens and A. monanthes clades.
Here, apomictic taxa are present in multiple polyploid and

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


Spec. nov. 2 Spec. nov. 1 hybrid forms, including putative autoploid and alloploid
lineages. Although reticulation is common throughout
F I G . 6. A reticulogram illustrating the apomictic polyploid species relation-
ships and hypothetical origins of apomixis within the A. monanthes clade. the complex (including sexual taxa), and there are some
Extant cytotypes are used to illustrate hypothesized lineage origins and paren- non-hybrid apomictic lineages present, the pattern of hy-
tal relationships. The inferred ploidy levels within the three lineages are repre- bridization observed does not correspond to the single
sented using circles for diploids, triangles for triploids and squares for origin hypothesis.
tetraploids. The diploids are sexually reproducing species, whereas the tri-
ploids and tetraploid forms are apomictic. The dashed circle represents a
hypothesized (either unsampled or extinct) parent cytotype. Dashed arrows in-
dicate paternal parent relationships. Solid arrows indicate maternal parents (i.e. We therefore favour this second scenario of multiple origins
unreduced gametophytes). within the whole complex. Within the A. monanthes clade in
particular, there is evidence for three lineages. The pattern
of nuclear copy distribution between these lineages indicates
(Figs 4 and 6). Although no morphological analysis was per- three origins of apomixis from up to three different sexual pro-
formed, some patterns were observed during visual identifica- genitor species (Fig. 6; Supplementary Data, Tables S5 and
tion of these specimens. Spec.nov.2 is morphologically similar S6). Under the assumption of rarity of regaining sexual repro-
to some A. monanthes specimens, but spec.nov.1 appears dis- duction in obligate apomictic lineages, the observed pattern
tinct, having previously been identified as A. polyphyllum does not support a single origin, and gene flow between
(Monterrosa et al., 2009). Further study confirmed the morpho- asexual and sexual lineages by hybridization (via functional
logical differences of the later specimen from A. monanthes antheridia) with sexual relatives. This inference is strengthened
and A. polyphyllum (as defined in Mickel and Smith, 2004). by the lack of antheridia observations in gametophyte studies.
The A. hallbergii group is morphologically distinct from the However, one specimen (RD99) has nuclear copies from all
two A. monanthes groups (Mickel and Smith, 2004). three apomictic lineages. This indicates that apomixis may
Morphological distinctions between the MO1 and MO2 lineages have spread via functional hybridization in some instances.
are unclear, but further investigation may uncover correlations Nuclear copies from the MO2 lineage are present in both the
between morphological and genetically identified taxa. A. monanthes MO1 and the A. hallbergii apomictic lineages
In general, our approach using chloroplast DNA (cpDNA) (Figs 4 and 6). This indicates that the A. monanthes MO2
and nuclear ribosomal DNA (nrDNA) markers improved our lineage is crucial in alloploid apomictic origin. The
understanding of the relationships within the group, especially A. monanthes MO2 lineage itself is a putative autoploid
when polyploidy is considered. Further improvements will lineage. Although the precise pathways to autoploid apomict
probably be achieved by expanding the taxonomic sampling, origin are difficult to infer, this process almost certainly
using a more sensitive method to determine polyploidy (e.g. involved the cross-fertilization of normal haploid gameto-
flow cytometry) and by using several nuclear markers. phytes and unreduced gametes, the latter produced by either
Single nuclear markers can result in misleading interpretations sexual diploids or autotetraploids (Fig. 6).
as a result of factors such as incomplete lineage sorting which The affinity of the sexual diploids, spec.nov.1 and
resemble patterns of ongoing hybridization (Nitta et al., 2011). spec.nov.2, for the A. monanthes MO1 and MO2 lineages, re-
spectively, indicates that these accessions may represent sexual
progenitor species for each group. No putative diploid sexual
progenitor specimen was observed for the A. hallbergii
Origins of apomixis
lineage but, due to the clear difference in morphology and dis-
The occurrence of apomixis in this complex can be tinction in both the plastid and nuclear genomes, we suggest
explained in two ways: (1) a single origin with a complex the participation of at least one additional undetected sexual
history of reversals and post-origin differentiation/divergence progenitor. The A. castaneum sub-clade nCA (weakly sup-
of apomictic lineages; or (2) multiple independent origins ported as sister to A. hallbergii by the nuclear gene analysis)
from more than one sexual progenitor species (e.g. Beck and the putative facultative tetraploid A. aff. hallbergii are
et al., 2011a, b; Sigel et al., 2011). potential progenitors (Fig. 4).
Dyer et al. — Apomixis and reticulate evolution in a fern complex Page 13 of 15

Similar patterns are observed within the A. resiliens clade, Table S7: analysis of prothalli and evidence of reproductive
although we make no inferences here, as more exhaustive sam- mode. Table S8: summary of spore size measurements. Fig.
pling of species within this clade is necessary. S1: phylogenetic analysis of the plastid psbA– trnH intergenic
These hypotheses of origin need to consider the limits of spacer (IGS) region using maximum parsimony. Fig. S2:
single gene markers; additional studies would benefit from phylogenetic analysis of the plastid of the rps4 plus rps4 –
the use of several nuclear genes or a better marker system trnS IGS region using maximum parsimony. Fig. S3: phylo-
(such as co-dominant microsatellites) that sample a larger genetic analysis of the plastid trnL – trnF region, including
portion of the genomes. We also need to be aware that the dy- the trnL intron and the trnL – trnF IGS region, using
namics of genomic re-arrangement in apomictic polyploids maximum parsimony. Fig. S4: phylogenetic analysis of the
may create patterns similar to those observed by entirely dif- combined plastid data sets using maximum likelihood infer-
ferent processes such as homoeologous chromosome pairing ence. Fig. S5: phylogenetic analysis of the nuclear
(Ishikawa et al., 2003). The large ranges of spore size in pgiC14FN – 16RN region using maximum likelihood infer-
some specimens such as RD24 and RD99 (Fig. 5) suggest ence. Fig. S6: hybridization network computed in SplitsTree
that such processes may be present in this group and may version 4.12.6 from ML consensus trees of plastid and
also affect spore development. nuclear data sets. Fig. S7: boxplots illustrating spore width

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


These caveats aside, our results are congruent to those of variation of species and specimens.
several recent studies on cheilanthoid ferns (Grusz et al.,
2009; Beck et al., 2011b; Sigel et al., 2011) and filmy ferns
(Nitta et al., 2011) that reported multiple recent origins of apo- ACK NOW LED GE MENTS
mixis (both alloploidy and autoploid) from a variety of ances- We would like to thank A. Aitken, A. Hall and S. Russell for
tral diploid progenitor species. help in the laboratory; the DNA sequencing services at the
Natural History Museum; P. Acock, M. Christenhusz,
Conclusions M. Dickinson, A. Grusz, S. Hennequin, I. Jimenez, R. Jonas,
P. Kamau, M. Kessler, S. Knapp, T. Kroemer, A. Larsson,
Within this complex there appear to be a number of relative- J. Launert, M. Lehnert, T. Lemieux, C. Marsh, K. Mehlttreter,
ly well-established apomictic lineages of independent origins. A. Monro, J. Monterosa, H. Muth, E. Otto, C. Rothfels,
However, we find no strong evidence for long-lived apomictic P. Rothfels, E. Schuettpelz, D. Tejero-Diez, M. Windham and
lineages that outlast or replace their diploid progenitors com- G. Yatskievych for providing samples and helping in the field;
pletely. Sexual progenitors are rare, and the lack of evid- S. Ansell, T. Barraclough, M. Carine, A. Dyer, A. de Groot,
ence for functional antheridia in apomictic prothalli provides C. Gueiden, S. Hennequin, H. Hipperson, M. Lewis,
little evidence for ongoing clonal turnover. Apomictic lineages J. Marquardt, C. Múrria, A. Papadopulos, N. Robson,
are associated with reticulate evolution. However, we are as yet A. Smith, J. Vogel, L. Wang and Y. Yu for helpful discussions;
unable to determine the precise mechanisms of origin or, for and M. D. Windham, an anonymous referee and the editor for
that matter, whether this association of apomixis with reticula- helpful comments. This work was supported by the Natural
tion is the cause or the effect of apomictic origins within this History Museum in London, the Chinese Academy of Science,
complex. the Royal Society and the European Research Council.
Our study is limited by the fact that nuclear copy distribu-
tion is not the best indicator of parentage/reticulation patterns,
and our conclusions should be viewed primarily as a founda- L I T E R AT U R E CI T E D
tion for further investigation. Our findings also call for some Adams CD. 1995. Asplenium. In: Davidse G, Sousa M, Knapp S. eds. Flora
taxonomic studies in the complex. For example, following Mesoamericana. Mexico: Universidad Nacional Autónoma de México,
the collection of two previously unknown sexual diploid taxa 290–324.
for the flora of El Salvador (Monterrosa et al., 2009), more ex- Aldrich J, Cherney BW, Merlin E, Christopherson L. 1988. The role of
tensive sampling is needed and may lead to the discovery of insertions/deletions in the evolution of the intergenic region between
psbA and trnH in the chloroplast genome. Current Genetics 14: 137–146.
several new taxa that are either rare and/or have small distribu- Ammal LS, Bahavanandan KV. 1991. Cytological studies on the genus
tion ranges. We believe that future work would benefit from Asplenium Linn. Indian Fern Journal 8: 69– 73.
using population genetic approaches and more in-depth cyto- Asker SE, Jerling L. 1992. Apomixis in plants. London: CRC Press.
logical analysis to disentangle fully the mechanisms by Barrington DS, Paris CA, Ranker TA. 1986. Systematic inferences from
spore and stomate size in the ferns. American Fern Journal 76: 149– 159.
which asexuality evolves and is maintained. Beck JB, Windham MD, Yatskievych G, Pryer KM. 2010. A diploids-first
approach to species delimitation and interpreting polyploid evolution in
the fern genus Astrolepis (Pteridaceae). Systematic Botany 35: 223– 234.
S U P P L E M E N TARY D ATA Beck JB, Alexander PJ, Allphin L, Al-Shehbaz IA, Rushworth C, Bailey
CD, Windham MD. 2011a. Does hybridization drive the transition to
Supplementary data are available online at www.aob.oxford- asexuality in diploid Boechera. Evolution 66: 985–995.
journals.org and consist of the following. Table S1: sampling Beck JB, Windham MD, Pryer KM. 2011b. Do asexual polyploid lineages
information. Table S2: nuclear clone information and lead short evolutionary lives? A case study from the fern genus
GenBank accession numbers. Table S3: plastid haplotype in- Astrolepis. Evolution 65: 3217–3229.
formation. Table S4: information on phylogenetic analyses. Benson DA, Karsch-Mizrachi I, Lipman DJ, Ostell J, Sayers EW. 2011.
GenBank. Nucleic Acids Research 39(suppl): D32–D37. http://
Table S5: information on phylogenetic relationships and dx.doi.org/10.1093/nar/gkq1079.
spore counts per sporangium for each specimen. Table S6: Braithwaite AF. 1964. A new type of apogamy in ferns. New Phytologist 63:
nuclear sub-clade and terminal label information (Fig. 4). 293–305.
Page 14 of 15 Dyer et al. — Apomixis and reticulate evolution in a fern complex

Brysting AK, Mathiesen C, Marcussen T. 2011. Challenges in polyploid Klöpper TH, Huson DH. 2008. Drawing explicit phylogenetic networks
phylogenetic reconstruction: a case story from the arctic-alpine and their integration into SplitsTree. BMC Evolutionary Biology 8: 22.
Cerastium alpinum complex. Taxon 60: 333– 347. http://dx.doi.org/10.1186/1471-2148-8-22.
Döpp W. 1932. Die Apogamie bei Aspidium remotum Al.Br. Planta 17: Knobloch IW. 1966. A preliminary review of spore number and apogamy
86– 152. within the genus Cheilanthes. American Fern Journal 56: 163–167.
Doyle JJ. 1987. A rapid DNA isolation procedure for small quantities of fresh Lloyd RM. 1973. Facultative apomixis and polyploidy in Matteuccia orienta-
leaf tissue. Phytochemical Bulletin 19: 11– 15. lis. American Fern Journal 63: 43– 48.
Ekrt E, Stech M. 2008. A morphometric study and revision of the Asplenium Lovis JD. 1977. Evolutionary patterns and processes in ferns. Advances in
trichomanes group in the Czech Republic. Preslia 80: 325–347. Botanical Research 4: 230 –415.
Felsenstein J. 1985. Confidence-limits on phylogenies: an approach using the Lovis JD, Rasbach H, Rasbach K, Reichstein T. 1977. Asplenium azoricum
bootstrap. Evolution 39: 783– 791. and other ferns of the A. trichomanes group from the Azores. American
Fontaneto D, Herniou EA, Boschetti C, Caprioli M, Melone G, Ricci C, Fern Journal 67: 81–93.
Barraclough TG. 2007. Independently evolving species in asexual bdel- Maddison DR, Maddison WP. 2005. MacClade 4: analysis of phylogeny and
loid rotifers. PLoS Biology 5: e87. http://dx.doi.org/10.1371/journal. character evolution. Version 4.08a. http://macclade.org.
pbio.0050087. Maddison WP, Maddison DR. 1989. Interactive analysis of phylogeny and
Ghatak J. 1977. Biosystematic survey of Pteridophytes from Shevaroy Hills, character evolution using the computer program MacClade. Folia
South India. The Nucleus 20: 105–108. Primatologica 53: 190– 202.
Grusz AL, Windham MD, Pryer KM. 2009. Deciphering the origins of apo- Manton I. 1950. Problems of cytology and evolution in the Pteridophyta.

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


mictic polyploids in the Cheilanthes yavapensis complex (Pteridaceae). Cambridge: Cambridge University Press.
American Journal of Botany 96: 1636– 1645. Manton I. 1959. Cytological information of the ferns of West Tropical Africa.
Guillén RH, Daviña JR. 2005. Estudios cromosómicos en especies de In: Alson AHG. ed. The ferns and fern allies of West Tropical Africa.
Asplenium (Aspleniaceae) de la Argentina. Darwiniana 43: 44–51. Cambridge: Cambridge University Press, 75– 81.
Guindon S, Lethiec F, Duroux P, Gascuel O. 2005. PHYML Online: a web Manton I, Vida G. 1968. Cytology of the ferns flora of Tristan da Cunha.
server for fast maximum likelihood-based phylogenetic inference. Proceedings of the Royal Society of London B: Biological Sciences
Nucleic Acids Research 33(suppl.): W557–W559. http://dx.doi.org/ 170: 361–371.
10.1093/nar/gki352. Manton I, Vida G, Gibby M. 1986. Cytology of the fern flora of Madeira.
Guindon S, Dufayard JF, Lefort V, Anisimova M, Hordijk W, Gascuel O. Bulletin of the British Museum (Natural History). Botany. 15: 123– 161.
2010. New algorithms and methods to estimate maximum-likelihood phy- Mason-Gamer RJ, Kellogg EA. 1996. Testing for phylogenetic conflict
logenies: assessing the performance of PhyML 3.0. Systematic Biology among molecular data sets in the tribe Triticeae (Gramineae).
59: 307–321. Systematic Biology 45: 524– 545.
Haufler CH, Soltis DE. 1986. Genetic evidence suggests that homosporous Mickel JT, Smith AR. 2004. The Pteridophytes of Mexico. New York:
ferns with high chromosome numbers are diploid. Proceedings of the New York Botanical Gardens Press.
National Academy of Sciences, USA 88: 4389–4393. Mogie M. 1990. Homospory and the cost of asexual reproduction. Evolution
Hörandl E, Paun O, Johansson JT, Lehnebach C, Armstrong T, Chen L, 44: 1707–1710.
Lockhart P. 2005. Phylogenetic relationships and evolutionary traits in Mogie M. 1992. The evolution of asexual reproduction in plants. London:
Ranunculus s.l. (Ranunculaceae) inferred from ITS sequence analysis. Chapman and Hall.
Molecular Phylogenetics and Evolution 36: 305–327. Monterrosa JS, Pena-chocarro MC, Knapp S, Lechuga RE. 2009. Guia de
Huang Y-M, Chou H-M, Hsieh T-H, Wang J-C, Chiou W-L. 2006. Cryptic indentificacion de helechos de El Salvador. Antiguo Cuscatlan.
characteristics distinguish diploid and triploid varieties of Pteris fauriei Moran RC. 1982. The Asplenium trichomanes complex in the United States
(Pteridaceae). Canadian Journal of Botany 84: 261 –268. and adjacent Canada. American Fern Journal 72: 5– 11.
Huang YM, Hsu SM, Hsieh TS, Chou HU, Chiou WL. 2011. Three Pteris Morzenti VM. 1966. Morphological and cytological data on southeastern
species (Pteridacae, Pteridophyta) reprodcue by apogamy. Botanical United States species of the Asplenium heterochroum-resiliens
Studies 52: 79–87. complex. American Fern Journal 56: 167–177.
Huelsenbeck JP, Ronquist F. 2001. MRBAYES: Bayesian inference of Morzenti VM, Wagner WH. 1962. Southeastern American blackstem spleen-
phylogenetic trees. Bioinformatics 17: 754–755. worts of the Asplenium heteochroum-resilines complex. Association of
Huson DH, Bryant D. 2006. Application of phylogenetic networks in evolu- Southeastern Biologists Bulletin 9: 40– 41.
tionary studies. Molecular Biology and Evolution 23: 254– 267. Nitta JH, Ebihara A, Ito M. 2011. Reticulate evolution in the Crepidomanes
Huson DH, Scornavacca C. 2011. A survey of combination methods for minutum species complex (Hymenophyllaceae). American Journal of
phylogenetic networks. Genome Biology Evolution 3: 23– 35. Botany 98: 1782–1800.
Ishikawa H, Watano Y, Kano K, Ito M, Kurita S. 2002. Development of Park CH, Kato M. 2003. Apomixis in the interspecific triploid hybrid fern
primer sets for PCR amplification of the pgiC gene in ferns. Journal of Cornopteris christenseniana (Woodsiaceae). Journal of Plant Research
Plant Research 115: 65–70. 116: 93–103.
Ishikawa H, Ito M, Watano Y, Kurita S. 2003. Electrophoretic evidence for Pichot C, Maataoui M, Raddi S, Raddi P. 2001. Surrogate mother for endan-
homoeologous chromosome pairing in the apogamous fern species gered Cupressus. Nature 412: 39–39.
Dryopteris nipponensis (Dryopteridaceae). Journal of Plant Research Pirie MD, Humphreys AM, Barker NP, Linder HP. 2009. Reticulation, data
116: 165– 167. combination, and inferring evolutionary history: an example from
Jørgensen PM, Nee M, Beck SG. 2010. Catálogo de las plantas vasculares Danthonioideae (Poaceae). Systematic Biology 58: 612– 628.
de Bolivia. http://www.tropicos.org/Project/BC. Posada D. 2008. jModelTest: phylogenetic model averaging. Molecular
Judson OP, Normark BB. 1996. Ancient asexual scandals. Trends in Ecology Biology and Evolution 25: 1253– 1256.
and Evolution 11: 41– 46. Posada D, Crandall KA. 1998. MODELTEST: testing the model of DNA
Juslén A, Väre H, Wikström N. 2011. Relationships and evolutionary origins substitution. Bioinformatics 14: 817–818.
of polyploid Dryopteris (Dryopteridaceae) from Europe inferred using Pryer KM, Schuettpelz E, Wolf PG, Schneider H, Smith AR, Cranfill R.
nuclear pgiC and plastid trnL-F sequence data. Taxon 60: 1284– 1294. 2004. Phylogeny and evolution of ferns (monilophytes) with a focus on
Kawakami SM, Kawakami S, Katsuhiko K, Shmakov A. 2011. Cytological the early leptosporangiate divergences. American Journal of Botany 91:
studies of Russian Altai ferns and the haploid sporophyte formation. 1582–1598.
Chromosome Botany 6: 21– 23. R Development Core Team. 2011. R: a language and environment for stat-
Kessler M, Smith AR. 2011. Ferns of Bolivia. http://www.systbot.uzh.ch/ istical computing. http://www.R-project.org/.
static/fernsofbolivia/. Regalado Gabancho L, Prada C, Gabriel y Galán JM. 2010. Sexuality
Klekowski EJ Jr. 1970. Populational and genetic studies of a homosporous and apogamy in the Cuban Asplenium auritum– monodon complex
fern – Osmunda regalis. American Journal of Botany 57: 1122– 1138. (Aspleniaceae). Plant Systematics and Evolution 289: 137– 146.
Klekowski EJ Jr, Hickok LG. 1974. Nonhomologous chromosome pairing in Ronquist F, Huelsenbeck JP. 2003. MrBayes 3: Bayesian phylogenetic infer-
the fern Ceratopteris. American Journal of Botany 61: 422–432. ence under mixed models. Bioinformatics 19: 1572– 1574.
Dyer et al. — Apomixis and reticulate evolution in a fern complex Page 15 of 15

Schneider H, Russell SJ, Cox CJ, et al. 2004. Chloroplast phylogeny of asple- Taberlet P, Gielly L, Pautou G, Bouvet J. 1991. Universal primers for amp-
nioid ferns based on rbcL and trnL-F spacer sequences (Polypodiidae, lification of 3 non-coding regions of chloroplast DNA. Plant Molecular
Aspleniaceae) and its implications for biogeography. Systematic Botany Biology 17: 1105– 1109.
29: 260–274. Trewick SA, Morgan-Richards M, Russell SJ, et al. 2002. Polyploidy, phy-
Schneider H, Ranker TA, Russell SJ, et al. 2005. Origin of the endemic fern logeography and Pleistocene refugia of the rockfern Asplenium ceterach:
genus Diellia coincides with the renewal of Hawaiian terrestrial life in the evidence from chloroplast DNA. Molecular Ecology 11: 2003–2012.
Miocene. Proceedings of the Royal Society B: Biological Sciences 272: Tryon RM. 1956. A revision of the American species of Notholaena.
455– 460. Contributions from the Gray Herbarium 179: 1 –106.
Schneider H, Smith AR, Pryer KM. 2009. Is morphology really at odds with Tryon R, Voeller B, Tryon A, Riba R. 1973. Fern biology in Mexico (a class
molecules in estimating fern phylogeny. Systematic Botany 34: 455– 475. field program). BioScience 23: 28–33.
Schneller J, Krattinger K. 2010. Genetic composition of Swiss and Austrian Tryon RM, Stolz RG, Leon B. 1993. Pteridophyta of Peru. Part V. 18:
members of the apogamous Dryopteris affinis complex (Dryopteridaceae, Aspleniaceae – 21. Polpodiaceae. Fieldiana, Botany 32: 1– 190.
Polypodiopsida) based on ISSR markers. Plant Systematics and Evolution Tutin TG, Burges NA, Chater AO, et al. 1993. Flora Europea. Cambridge:
286: 1– 6. Cambridge University Press.
Schwander T, Henry L, Crespi Bernard J. 2011. Molecular evidence Vida G. 1970. The nature of polyploidy in Asplenium ruta-muraria L. and
for ancient asexuality in Timema stick insects. Current Biology 21: A. lepidium C. Presl. Caryologia 23: 525– 547.
1129–1134. Wagner WH. 1963. A biosystematic study of United States ferns – prelimin-
Shaw J, Lickey EB, Schilling EE, Small RL. 2007. Comparison of whole ary abstract. American Fern Journal 53: 1– 16.

Downloaded from http://aob.oxfordjournals.org/ at Institute of Botany, CAS on September 15, 2012


chloroplast genome sequences to choose noncoding regions for phylogen- Wagner WH. 1966. Two new species of ferns from the United States.
etic studies in angiosperms: the tortoise and the hare III. American American Fern Journal 56: 3 –17.
Journal of Botany 94: 275–288. Wagner WH, Wagner FS. 1966. Pteridophytes of the mountain lake aras,
Sigel EM, Windham MD, Huiet L, Yatskievych G, Pryer KM. 2011. Gilles Co., Virginia: Biosystematic Studies 1964–65. Castanea 31:
Species relationships and farina evolution in the Cheilanthoid fern 121–140.
genus Argyrochosma (Pteridaceae). Systematic Botany 36: 554–564. Wagner WH, Farrar DR, McAlpin BW. 1970. Pteridology of the Highlands
Smith AR, Cranfill RB. 2002. Intrafamilial relationships of the thelypteroid Biological Station Area, Southern Appalachians. Journal of the Elisha
ferns (Thelypteridaceae). American Fern Journal 92: 131–149. Mitchell Scientific Society 86: 1–24.
Smith AR, Mickel JT. 1977. Chromosome counts for Mexican ferns. Wagner WH, Moran RC, Werth CR. 1993. Aspleniaceae. In: Flora of North
Brittonia 29: 391 –398. America. New York: Oxford University Press, 237–238.
Smith AR, Kessler M, Gonzales J. 1999. New records of pteridophytes from Walker TG. 1966a. Apomixis and vegetative reproduction in ferns. In:
Bolivia. American Fern Journal 89: 244– 266. Hawkes JG. ed. Reproductive biology and taxonomy of vascular plants.
Somer M, Arbesú R, Menéndez V, Revilla MA, Fernández H. 2009. The report of the Conference held by the Botanical Society of the
Sporophyte induction studies in ferns in vitro. Euphytica 171: 203– 210. British Isles, at Birmingham University in 1965. Oxford: Pergamon
Stolze RG. 1981. Ferns and fern allies of Guatemala. Part II. Polypodiaceae. Press, 152– 161.
Fieldiana, Botany 6: 1– 522. Walker TG. 1966b. A cytotaxonomic survey of the pteridophytes of Jamaica.
Stolze RG, Pacheco L, Øllgaard B. 1994. 14(5B). Polypodiaceae – Transactions of the Royal Society of Edinburgh 66: 169– 237.
Dryopteridoideae – Physematieae. In: Harling G, Andersson L. eds. Walker TG. 1979. The cytogenetics of ferns. In: Dyer AF. ed. The experimen-
Flora of Ecuador. Goteborg: University of Goteborg, 108. tal biology of ferns. New York: Academic Press, 87–132.
Swofford DL. 1993. PAUP: a computer-program for phylogenetic inference Windham MD. 1983. The ferns of Elden Mountain, Arizona. American Fern
using maximum parsimony. Journal of General Physiology 102: A9 –A9. Journal 73: 85– 93.
Swofford DL. 2002. PAUP* phylogenetic analysis using parsimony (* and Zuloaga FO, Morrone O, Belgrana MJ. 2008. Catalogo de las plantas vas-
other methods). Version 4.0b10. Sunderland, MA: Sinauer Associates. culares del Cono Sur. St. Louis: MBG Press.

View publication stats

Potrebbero piacerti anche