07.04.2013 Views

A NEW BROWN ALGAL ORDER, ISHIGEALES (PHAEOPHYCEAE ...

A NEW BROWN ALGAL ORDER, ISHIGEALES (PHAEOPHYCEAE ...

A NEW BROWN ALGAL ORDER, ISHIGEALES (PHAEOPHYCEAE ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

J. Phycol. 40, 921–936 (2004)<br />

r 2004 Phycological Society of America<br />

DOI: 10.1111/j.1529-8817.2004.03160.x<br />

A <strong>NEW</strong> <strong>BROWN</strong> <strong>ALGAL</strong> <strong>ORDER</strong>, <strong>ISHIGEALES</strong> (<strong>PHAEOPHYCEAE</strong>), ESTABLISHED ON<br />

THE BASIS OF PLASTID PROTEIN-CODING rbcL, psaA, AND psbA REGION<br />

COMPARISONS 1<br />

Ga Youn Cho, Sang Hee Lee, and Sung Min Boo 2<br />

Department of Biology, Chungnam National University, Daejon 305-764, Korea<br />

The brown algal family Ishigeaceae currently<br />

includes a single genus, Ishige Yendo, with two species.<br />

The relationship of the family to other brown<br />

algal lineages is less studied in terms of their plastid<br />

ultrastructure and molecular phylogeny. We determined<br />

the sequences of rbcL from four samples of<br />

the two Ishige species and nine putative relatives<br />

and the psaA andpsbA sequences from 37 representatives<br />

of the brown algae. Analyses of individual<br />

and combined data sets resulted in similar trees;<br />

however, the concatenated data gave greater resolution<br />

and clade support than each individual gene.<br />

In all the phylogenies, the Phaeophyceae was well<br />

resolved, the Ectocarpales being placed in a terminal<br />

position and the Ishigeaceae ending up in a<br />

basal position. From our ultrastructural study,<br />

we concluded that the pyrenoid is absent in the<br />

Ishigeaceae, despite the presence of a rudimentary<br />

pyrenoid in I. okamurae. These results suggest that<br />

the Ishigeaceae is an early diverging brown lineage.<br />

Our molecular and morphological data, therefore,<br />

lead us to exclude the Ishigeaceae from the Ectocarpales<br />

s.l., which have an elaborate pyrenoid, and<br />

to propose its own order Ishigeales ord. nov. The<br />

Ishigeales is distinguished by oligostichous structure<br />

of thalli, phaeophycean hairs formed within<br />

cryptostomata, unilocular sporangia transformed<br />

from terminal cortical cells, and plurilocular sporangia<br />

lacking sterile terminal cells. This study is the<br />

first to document the utility of the psaA and psbA<br />

sequences for brown algae and also the first report<br />

on the multigene phylogeny of the Phaeophyceae<br />

based on three protein-coding plastid genes.<br />

Key index words: Ishige; Ishigeaceae; Ishigeales ord.<br />

nov.; Phaeophyceae; phylogeny; psaA; psbA; pyrenoid;<br />

rbcL; taxonomy; ultrastructure<br />

Abbreviations: BS, bootstrap value; LSU, large subunit<br />

of rDNA; ML, maximum likelihood; MP, maximum<br />

parsimony; psaA, PSI P700 chl a apoprotein<br />

A1 gene; psbA, PSII thylakoid protein D1 gene;<br />

rbcL, large subunit of RUBISCO gene; SH,<br />

Shimodaira-Hasegawa; SSU, small subunit of rDNA<br />

1 Received 4 September 2003. Accepted 9 June 2004.<br />

2 Author for correspondence: e-mail smboo@cnu.ac.kr.<br />

921<br />

Ishige is a genus of brown algae that contains two<br />

species. Both are summer annuals and occur exclusively<br />

in the warm waters of the Pacific Ocean (Yendo<br />

1907, Setchell and Gardner 1924, Tseng 1983, Lee<br />

et al. 2003). The genus is characterized by cylindrical<br />

to foliose thalli, hairs in cryptostomata, uniseriate<br />

plurilocular sporangia lacking sterile terminal cells,<br />

and an isomorphic life history (Yendo 1907, Ajisaka<br />

1989, Tanaka in Hori 1993, Lee et al. 2003). The<br />

genus was based on I. okamurae, which was described<br />

from two different forms (filiform and foliose forms)<br />

collected in Shimoda on the Pacific coast of Japan. The<br />

foliose thalli are considered abnormal branches of the<br />

filiform type, because the former is commonly epiphytic<br />

on the latter. Ishige okamurae occurs in the upper<br />

intertidal zone along the coasts of Korea (Lee et al.<br />

2003), Japan (Yendo 1907, Yoshida 1998), and China<br />

(Tseng 1983). Okamura (in Segawa 1935) described<br />

I. foliacea, the second member of the genus, based on<br />

the foliose type, which has a cortical layer much thinner<br />

than that of I. okamurae. However, Chihara (1969)<br />

showed that I. foliacea is a later homonym of Polyopes<br />

sinicola Setchell et Gardner (1924), revising the name to<br />

I. sinicola (Setchell et Gardner) Chihara. Ishige sinicola<br />

occurs as an epiphyte on I. okamurae in Korea,<br />

Japan, and China (Tseng 1983, Yoshida 1998, Lee<br />

et al. 2003), whereas the former occurs alone along<br />

the northern coast of the Gulf of California (Setchell<br />

and Gardner 1924).<br />

On establishing the genus Ishige, Yendo (1907)<br />

putatively put it within the family Fucaceae on the<br />

grounds of cryptostomata in the foliose form ( 5 I. sinicola).<br />

However, Okamura (in Segawa 1935) established<br />

the monotypic family Ishigeaceae based on the zoospores<br />

in the unilocular sporangia of I. sinicola, and<br />

then Okamura (1936) placed the Ishigeaceae within<br />

the Punctariales, because the apical cells of the assimilatory<br />

filaments change into unilocular sporangia<br />

where zoospores with two lateral flagella are produced.<br />

Finally, Arasaki (1943) concluded that the Ishigeaceae<br />

belonged to the order Chordariales on the basis of its<br />

heteromorphic life history, motile gametes in microscopic<br />

gametophytes, and plurilocular sporangia similar<br />

to those of some chordarialean species. In contrast,<br />

Ajisaka (1989) proposed that the family Ishigeaceae<br />

may not belong to the Chordariales, because I. okamurae<br />

has uniseriate plurilocular sporangia lacking<br />

sterile terminal cells, whereas the Chordariales have


922<br />

unilocular sporangia formed on the basal or middle<br />

part of the assimilatory filaments. Nevertheless, the<br />

family Ishigeaceae is still classified in the Chordariales<br />

in textbooks of algae (Yoshida 1998).<br />

The presence of a pyrenoid is a diagnostic character<br />

for the Phaeophyceae at the ordinal level (Evans 1966,<br />

Hori and Ueda 1975, Kawai 1992). This is supported<br />

by previous studies using small subunit (SSU), large<br />

subunit (LSU), and rbcL DNAs (Tan and Druehl 1994,<br />

de Reviers and Rousseau 1999, Peters and Ramírez<br />

2001). The Ectocarpales have a large pedunculate<br />

pyrenoid with a cap layer in the discoid plastids of<br />

vegetative cells (Evans 1966, Hori and Ueda 1975).<br />

The Scytothamnales have a pyrenoid in the center of<br />

stellate plastids (Peters and Clayton 1998), whereas<br />

Asterocladon, Asteronema, and Bachelotia, which are insertae<br />

sedis, have several terminal pyrenoids in the ribbon-shaped<br />

plastids (Müller et al. 1998). In contrast,<br />

the pyrenoid is absent in other brown algae (Evans<br />

1966, 1968, Chi 1971, Hori 1971, Hori and Ueda<br />

1975, Henry 1984). The absence/presence of the pyrenoid<br />

in the Ishigeaceae has been contentious. Hori<br />

(1971) observed a small pyrenoid in the plastids of the<br />

vegetative cells of I. okamurae from Japan but found no<br />

pyrenoids in the plastids of I. sinicola. He therefore<br />

questioned the absence of the pyrenoid in the latter<br />

species because of the congeneric relationship of the<br />

two species (Hori 1971, Hori and Ueda 1975).<br />

Rousseau and de Reviers (1999) proposed the<br />

broad concept of the order Ectocarpales, including<br />

the Chordariales, Dictyosiphonales, Punctariales, and<br />

Scytosiphonales and excluding the Ralfsiales and taxa<br />

with stellate plastids, based on combined SSU þ LSU<br />

rDNA sequences. This is corroborated by other DNA<br />

phylogenies (de Reviers and Rousseau 1999, Rousseau<br />

et al. 2000, 2001, Draisma et al. 2001, 2003, Peters and<br />

Ramírez 2001, Cho et al. 2003). de Reviers and Rousseau<br />

(1999) placed the Ishigeaceae and 22 other families,<br />

formerly assigned to the aforementioned different<br />

orders, within the Ectocarpales s.l. Then, Peters<br />

and Ramírez (2001) reclassified all 23 families into<br />

just 5 families, viz. Acinetosporaceae, Adenocystaceae,<br />

Chordariaceae, Ectocarpaceae, and Scytosiphonaceae,<br />

using DNA data, life history, and plastid structure data.<br />

In this classification, the previous chordarialean and<br />

dictyosiphonalean families are synonymized with the<br />

Chordariaceae (Peters and Ramírez 2001). This forces<br />

Ishige, the only member of the family Ishigeaceae, to be<br />

included in the Chordariaceae s.l. Based on partial sequences<br />

of SSU rDNA, Lee et al. (2003) suggested that<br />

the Ishigeaceae form an independent lineage with<br />

some affinity with the early lineage of the Ectocarpales<br />

s.l. However, the relationship of the family Ishigeaceae<br />

to the other brown algal lineages has not been addressed<br />

because taxon sampling was limited.<br />

The present study investigated the phylogenetic relationships<br />

of the family Ishigeaceae in the class<br />

Phaeophyceae by analyses of sequence data derived<br />

from three independent protein-coding plastid genes.<br />

First, we analyzed the rbcL gene from Ishige and puta-<br />

GA YOUN CHO ET AL.<br />

tive relatives and subsequently compiled an rbcL data<br />

set that included previously published sequences from<br />

GenBank. The rbcL gene has frequently been used to<br />

study various brown algal groups at a variety of taxonomic<br />

levels (Siemer et al. 1998, Draisma et al. 2001,<br />

Peters and Ramírez 2001, Cho et al. 2003). The second<br />

coding gene we chose was the psaA gene, which encodes<br />

the PSI P700 chl a apoprotein A1 (Morden and<br />

Sherwood 2002). To date, psaA has been analyzed for<br />

only two species, Fucus vesiculosus (Pearson et al. 2001)<br />

and Pylaiella littoralis (Yoon et al. 2002a), in the Phaeophyceae.<br />

The third protein-coding gene we used is the<br />

psbA gene, which encodes the PSII thylakoid protein<br />

D1 (Morden and Sherwood 2002). The psbA has been<br />

analyzed for two species, Ectocarpus siliculosus (Winhauer<br />

et al. 1991) and Pylaiella littoralis (Yoon et al.<br />

2002a), in the Phaeophyceae. Samples of 37 algae including<br />

two outgroup taxa, selected from the taxa used<br />

in forming the rbcL data set described above, were<br />

available for the psaA and psbA analyses. Because psaA<br />

and psbA gene sequences provide better resolution for<br />

the deep branches of the red and dinophycean algal<br />

phylogenies (Morden and Sherwood 2002, Yoon et al.<br />

2002a, b, Yang and Boo 2004), both genes should be<br />

useful for reconstructing the phaeophycean phylogeny<br />

at higher ranks.<br />

When multiple independent data sets are available<br />

for a phylogenetic study, the investigator uses combined<br />

data sets as well as individual data sets for phylogenetic<br />

reconstructions. For combing individual data<br />

sets, Bull et al. (1993) suggested that congruent data<br />

sets only can be combined for phylogenetic analyses,<br />

whereas Gatesy et al. (1999a), who advocated a ‘‘total<br />

evidence’’ methodology, showed that the combination<br />

of incongruent data can increase the resolution and the<br />

support within phylogenetic trees, revealing if a ‘‘hidden<br />

signal’’ is present in the different data sets. However,<br />

advocates of these two approaches agree that use<br />

of multiple data sets improve phylogeny estimation<br />

(Lavoué et al. 2003, Shimabukuro-Dias et al. 2004). In<br />

this study, we follow an empirical strategy, conducting<br />

both separate and simultaneous analyses. Together<br />

with molecular analyses of the Ishigeaceae, we studied<br />

the ultrastructure of plastids in the vegetative cells of<br />

I. okamurae, the type species of the genus.<br />

MATERIALS AND METHODS<br />

Samples. The starting point for this work was our collection<br />

of previously published rbcL sequences of the brown<br />

algaefromGenBank.WecompiledrbcL sequences from 57<br />

taxa, which consisted of the type genus or representatives of<br />

13 brown algal orders, 4 ‘‘ordinal-level taxa’’ (Choristocarpaceae,<br />

Onslowiaceae, Phyllariaceae, and Asterocladon/Asteronema)<br />

(see Table 2 in Draisma et al. 2003), and 3 outgroup<br />

taxa (Schizocladia ischiensis, Tribonema aequale, andPhaeothamnion<br />

confervicola). Thirteen rbcL sequences we determined<br />

here were added to this data set: two samples from each of<br />

Ishige okamurae and I. sinicola and nine other brown algae.<br />

Then, psaA andpsbA regions from 37 taxa, including Tribonema<br />

aequale and Schizocladia ischiensis as outgroup taxa, were<br />

analyzed. Both psaA and psbA sequences from Pylaiella littoralis


and the psbA sequence from Ectocarpus siliculosus were downloaded<br />

from GenBank (Winhauer et al. 1991, Yoon et al.<br />

2002a). The psaA sequence of Fucus vesiculosus was not used<br />

because of its insufficient length (330 nt, Pearson et al. 2001).<br />

The specimens and their corresponding GenBank accession<br />

numbers are listed in Table 1.<br />

Analyses of rbcL, psaA, and psbA regions. Total DNA was<br />

extracted from approximately 0.01 g of dried thalli ground<br />

in liquid nitrogen using a DNeasy Plant Mini Kit (Qiagen<br />

Gmbh, Hilden, Germany), according to the manufacturers’<br />

instructions, and then dissolved in 150 mL of distilled water.<br />

Extracted DNA was stored at 201 Candusedtoamplifythe<br />

rbcL, psaA, and psbA regions.<br />

The rbcL region was amplified and sequenced using the<br />

method of Kogame et al. (1999) and Yoon and Boo (1999).<br />

Primers PRB-FO, F2, F3, R1A, R2, R3A, RS1, and RS2 were<br />

used for most brown algae. For Cutleria, Ishige, Sargassum, and<br />

Sphacelaria, primer RbcL68F (Draisma et al. 2001), instead of<br />

PRB-FO, was used. The same DNA aliquot was used for amplifying<br />

the psaA andpsbA regions, and the amplification and<br />

sequencing reactions for these regions were the same as those<br />

used for rbcL. The psaA region was amplified and sequenced<br />

using primers psaA130F, psaA870F, psaA970R, and<br />

psaA1760R (Yoon et al. 2002a). The psbA region was amplified<br />

using primers psbA-F and psbA-R2 and sequenced using<br />

primers psbA-F, psbA-600R, psbA-500F, and psbA-R2 (Yoon<br />

et al. 2002a). The PCR products were purified using a High<br />

Pure TM PCR Product Purification Kit (Roche Diagnostics<br />

GmbH, Mannheim, Germany) according to the manufacturers’<br />

instructions. The sequences of the forward and reverse<br />

strands were determined for all taxa using an ABI PRISM TM<br />

377 DNA Sequencer (Applied Biosystems, Foster City, CA,<br />

USA). The electropherogram output for each sample was edited<br />

using the program Sequence Navigator v. 1.0.1 (Applied<br />

Biosystems).<br />

All sequences of the rbcL gene from 70 taxa (67 brown algae<br />

and 3 outgroups) were collated using the multisequence editing<br />

program, SeqPup (Gilbert 1995), and aligned by eye to<br />

compare our sequences with those published previously (Draisma<br />

et al. 2001, Cho et al. 2003). The undefined sequences at<br />

the 5 and 3 0 ends of the rbcL data set were coded as missing<br />

data. The psaA sequences from 38 taxa (36 brown algae, including<br />

previously published sequence of Pylaiella littoralis, and<br />

2 outgroups) were also aligned by eye (Table 1). The psbA sequences<br />

from 39 taxa (37 brown algae, including previously<br />

published sequences of Ectocarpus siliculosus and Pylaiella littoralis,<br />

and 2 outgroups) were aligned by eye, as for rbcL gene<br />

(Table 1). All alignments posed no problems because there<br />

were no gaps in each data set of these three protein-coding<br />

genes. The alignments are available at Treebase (study accession<br />

S1095) under accession numbers M1870 (psbA), M1871<br />

(rbcL), M1872 (psaA), and M1873 (rbcL þ psaA þ psbA).<br />

Phylogenetic analyses. Four data sets were used for the<br />

phylogenetic analyses: 70 taxa for rbcL, 39 taxa for psbA, and<br />

38 taxa for both the psaA andcombinedrbcL þ psaA þ psbA<br />

data sets. To infer the level of nucleotide saturation of the<br />

rbcL, psaA, and psbA sequences, uncorrected p-distances were<br />

plotted against corrected pairwise distances using the Hasegawa-Kishino-Yano<br />

85 model (Hasegawa et al. 1985) for the<br />

first, second, and third codon positions and all positions. We<br />

also conducted the partition homogeneity test (incongruence<br />

length difference [ILD] test of Farris et al. 1994), implemented<br />

in PAUP* 4.0b8 (Swofford 2002). The partition homogeneity<br />

test used 1000 replicates, each with 100 random<br />

sequence-addition replicates using tree bisection-reconnection<br />

(TBR) branch swapping.<br />

Maximum parsimony (MP) trees were constructed for each<br />

data set with PAUP* using a heuristic search algorithm with the<br />

following settings: 100 random sequence-addition replicates,<br />

<strong>ISHIGEALES</strong> ORD. NOV. 923<br />

TBR branch swapping, MulTrees, all characters unordered<br />

and unweighted, and branches with a maximum length of zero<br />

collapsed to yield polytomies. The bootstrap values (BS) for the<br />

resulting nodes were assessed using bootstrapping with 1000<br />

replicates.<br />

For maximum likelihood (ML) and Bayesian analyses, we<br />

performed a likelihood ratio test using Modeltest 3.06 (Posada<br />

and Crandall 1998) to determine the best available model for<br />

individual and combined data sets. For all data sets, the best<br />

model was a general time reversible (GTR) model with a gamma<br />

correction for among-site variation (G) and invariant sites<br />

(I). ML analyses (heuristic search with 10 random sequenceaddition<br />

replicates, TBR branch swapping, and MulTrees on)<br />

were performed using the GTR þ G þ I model. BS analysis was<br />

conducted by performing replicate ML searches, with two random<br />

sequence-addition replicates, using the same search conditions<br />

as described above. We succeeded in performing 100<br />

bootstrap replicates for each of individual and combined data<br />

sets using two processors. To run simultaneous bootstrap replicates<br />

in different processors, we prepared two batch files that<br />

differed in their starting random seed number (seed number,<br />

0), each specified to run 50 bootstrap replicates and to save the<br />

resulting bootstrap trees into files. A 50% majority rule consensus<br />

bootstrap tree was estimated by aggregating and<br />

weighting trees accordingly to the number of trees found in<br />

each bootstrap replicate, so that the bootstrap replicates had<br />

equal weight.<br />

Bayesian phylogenetic analyses were performed using<br />

MrBayes 3.0 (Huelsenbeck and Ronquist 2001). Each analysis<br />

was initiated from a random starting tree, and the program<br />

was set to run four chains of Markov chain Monte Carlo iterations<br />

simultaneously for 2,000,000 generations with trees<br />

sampled every 100th generation. The likelihood scores stabilized<br />

at approximately 50,000 generations, and thus the first<br />

500 trees were burned. For the purpose of comparison with<br />

bootstrapping, we chose to consider nodes with Bayesian posterior<br />

probabilities (PP) greater than 0.9 (e.g. the node appears<br />

in greater than 90% of sampled trees) as being well supported.<br />

The SH test (Shimodaira and Hasegawa 1999) was used to<br />

compare statistically alternative phylogenetic hypotheses, focusing<br />

on the inclusion of the Ishigeaceae within the Ectocarpales<br />

or other putative relatives. The SH test was conducted<br />

using PAUP*, with resampling estimated log-likelihood optimization<br />

and 10,000 bootstrap replicates.<br />

Morphology of plastids in Ishige okamurae. Apical parts of<br />

field-collected thalli were prepared for EM according to the<br />

following protocol. Specimens were fixed with 3% glutaraldehyde<br />

in 0.1 M cacodylate buffer (pH 7.2) containing 2%<br />

NaCl and 0.1% CaCl2 for 3 h at 41 C and then postfixed in 2%<br />

osmium tetroxide in 0.1 M sodium cacodylate buffer for 2 h.<br />

The apical parts were rinsed with cold distilled water, cut,<br />

and transferred into test tubes at 41 C. They were dehydrated<br />

in a graded series of ethanol and propylene oxide at 41 Cand<br />

infiltrated gradually with Spurr’s epoxy resin (Spurr 1969).<br />

After polymerization of the resin at 701 C for 24 h, serial sections<br />

(120–200 sections per sample) were cut with a diamond<br />

knife on an Ultracut E ultramicrotome (Reichert-Jung, Germany)<br />

and mounted on Formvar-coated slot grids. Thin sections<br />

were stained with Reynold’s lead citrate (Reynolds<br />

1963) and uranyl acetate and examined with a JEM-1010<br />

transmission electron microscope (JEOL, Ltd., Tokyo, Japan)<br />

at the Center for Research Facilities, Chungnam National<br />

University.<br />

RESULTS<br />

The rbcL alignment. The sequences determined in<br />

the present study were 1467 nt long, except for<br />

the two Ishige species (1325 and 1368 nt), Cutleria


924<br />

TABLE 1. Taxa, collection site or data source, and GenBank accession number of the rbcL, psaA, and psbA sequences.<br />

GenBank accession no.<br />

rbcL/psaA/psbA<br />

Taxa Collection site, date, voucher, or reference of rbcL/psaA/psbA<br />

GA YOUN CHO ET AL.<br />

Phaeophyceae<br />

Ectocarpales<br />

Acinetosporaceae<br />

Geminocarpus austro-georgiae Skottsberg Peters and Ramírez (2001) AJ295830/ — / —<br />

Pylaiella littoralis (Linnaeus) Kjellman Assali et al. (1990)/Yoon et al. (2002a)/ibid X55372/AY119724/AY119760<br />

Adenocystaceae<br />

Adenocystis utricularis (Bory) Skottsberg Peters and Ramírez (2001)/Barton, Maxwell Bay, Antarctica, 25 January 2000, PE001/ibid AJ295823/AY372939/AY528824<br />

Caepidium antarcticum J. Agardh Peters and Ramírez (2001) AJ295826/ — / —<br />

Utriculidium durvillei Skottsberg Peters and Ramírez (2001) AJ295835/ — / —<br />

Chordariaceae<br />

Asperococcus fistulosus (Hudson) Hooker Cho et al. (2003)/Port Erin Bay, Isle of Man, UK, 9 July 2002, PE002/ibid AY095321/AY372940/AY528825<br />

Chordaria flagelliformis (O.F.Müller) J. Agardh Cho et al. (2003)/Avacha Bay, Kamchatka, Russia, 24 July 1998, PE003/ibid AY095324/AY372941/AY528826<br />

Coelocladia arctica Rosenvinge Siemer et al. (1998) AF055395/ — / —<br />

Delamarea attenuata (Kjellman) Rosenvinge Siemer et al. (1998)/Nakhodka, Russia, 22 May 2002, PE004/ibid AF055396/AY372942/AY528827<br />

Dictyosiphon foeniculaceus (Hudson) Greville Avacha Bay, Kamchatka, Russia, 28 July 1998, PE005 AY372973/AY372943/AY528828<br />

Elachista fucicola (Velley) Areschoug Siemer et al. (1998) AF055398/ — / —<br />

Giraudia sphacelarioides Derbès et Solier Siemer et al. (1998) AF055399/ — / —<br />

Hummia onusta (Kützing) Fiore Siemer et al. (1998) AF055402/ — / —<br />

Isthmoplea sphaerophora (Harvery) Kjellman Siemer et al. (1998) AF055403/ — / —<br />

Myriotrichia clavaeformis Harvey Siemer et al. (1998) AF055408/ — / —<br />

Punctaria latifolia Greville Cho et al. (2003)/Hoedong, Jindo, Korea, 9 March 2001, PE010/ibid AY095322/AY372948/AY528833<br />

Sphaerotrichia divaricata (C. Agardh) Kylin Siemer et al. (1998) AF055412/ — / —<br />

Streblonema maculans (Hamel) South et Tittley Rousseau and de Reviers (1999) AY157694/ — / —<br />

Striaria attenuata (Greville) Greville Siemer et al. (1998) AF055415/ — / —<br />

Ectocarpaceae<br />

Ectocarpus sp. Hoedong, Jindo, Korea, 9 March 2001, PE011 AY372978/AY372949/AY528834<br />

E. siliculosus (Dillwyn) Lyngbye Winhauer et al. (1991) —/ —/ X56695<br />

Scytosiphonaceae<br />

Chnoospora implexa J. Agardh Kogame et al. (1999) AB022231/ — / —<br />

Colpomenia sinuosa (Mertens ex Roth) Derbès etSolier Kogame et al. (1999)/Guryongpo, Pohang, Korea, 6 November 2000, PE012/ibid AB022234/AY372950/AY528835<br />

in Castagne<br />

Hydroclathrus clathratus (C. Agardh) Howe Kogame et al. (1999)/Tsuyazaki, Fukuoka, Japan, 7 March 1999, PE013/ibid AB022233/AY372951/AY528836<br />

Myelophycus simplex (Harvey) Papenfuss Cho et al. (2003)/Daesado, Wando, Korea, 13 June 1999, PE014/ibid AY095320/AY372952/AY528837<br />

Petalonia fascia (O.F.Müller) Kuntze Kogame et al. (1999)/Ile de Batz, Roscoff, France, 5 April 2000, PE015/ibid AB022243/AY372953/AY528838<br />

Rosenvingea intricate (J. Agardh) Boergesen Kogame et al. (1999) AB022232/ — / —<br />

Scytosiphon lomentaria (Lyngbye) Link Kogame et al. (1999)/Seongsan, Jeju, Korea, 22 March 2000, PE016/ ibid AB022238/AY372954/AY528839<br />

Choristocarpaceae<br />

Choristocarpus tenellus (Kützing) Zanardini Draisma et al. (2001) AJ287862/ — / —<br />

Cutleriales<br />

Cutleria cylindrica Okamura Hoedong, Jindo, Korea, 9 March 2001, PC001 AY372979/AY372955/AY528840<br />

C. multifida (J. E. Smith) Greville Burrowes et al. (2003) AY157692/ — / —<br />

Zanardinia prototypes (Nardo) Nardo Burrowes et al. (2003) AY157693/ — / —<br />

Desmarestiales<br />

Desmarestia ligulata (Stackhouse) Lamouroux Draisma et al. (2001) AJ287848/ — / —<br />

D. viridis (Müller) Lamouroux Siemer and Pedersen (unpublished data) AF207799/ — / —<br />

Desmarestia sp. Penguin Rookery, Maxwell Bay, Antarctica, 11 January 2000, PD001 AY372980/AY372956/AY528841<br />

Himantothallus grandifolius (A. et E. Gepp) Zinova Draisma et al. (2001) AJ287850/ — / —<br />

Dictyotales<br />

Dictyota dichotoma (Hudson) Lamouroux Draisma et al. (2001) AJ287852/ — / —<br />

Dictyota sp. Guryongpo, Pohang, Korea, 15 March 2001, PDI001 AY422654/AY372957/AY528842<br />

Zonaria diesingiana J. Agardh Ishigaki Island, Okinawa, Japan, 19 January 1998, PDI002 AJ295823/AY372958/AY528843<br />

Fucales<br />

Ascophyllum nodosum (Linnaeus) Le Jolis Draisma et al. (2001)/Neeltjejans, Netherlands, 13 August 1997, PF001/ibid AJ287853/AY372959/AY528844


Table 1. (Contd.)<br />

GenBank accession no.<br />

rbcL/psaA/psbA<br />

Taxa Collection site, date, voucher, or reference of rbcL/psaA/psbA<br />

Fucus vesiculosus Linnaeus Burrowes et al. (2003)/Coos Bay, Oregon, USA, 16 May 2001, PF002/ibid AY157695/AY372960/AY528845<br />

Sargassum horneri (Turner) C. Agardh Seosang, Namhaedo, Korea, 3 November 2002, PF003 AY372981/AY372961/AY528846<br />

Ishigaceae<br />

Ishige okamurae Yendo Hanrim, Jeju, Korea, 4 December 2002, PE006 AY372974/AY372944/AY528829<br />

I. okamurae Kominato, Chiba, Japan, 28 July 2002, PE007 AY372975/AY372945/AY528830<br />

I. sinicola (Setchell et Gardner) Chihara Hanrim, Jeju, Korea, 4 December 2002, PE008 AY372976/AY372946/AY528831<br />

I. sinicola Kominato, Chiba, Japan, 28 July 2002, PE009 AY372977/AY372947/AY528832<br />

Laminariales<br />

Alaria crassifolia Kjellman in Kjellman et Petersen Hakodate, Hokkaido, Japan, 21 August 1999, PL001 AY372982/AY372962/AY528847<br />

Chorda filum (Linnaeus) Stackhouse Oshoro, Hokkaido, Japan, 25 April 2002, PL002 AY372983/AY372963/AY528848<br />

Laminaria digitata (Linnaeus) Lamouroux Port Erin Bay, Isle of Man, UK, 9 July 2000, PL003 AY372984/AY372964/AY528849<br />

Onslowiaceae<br />

Onslowia endophytica Searles Draisma et al. (2001) AJ287864/ — / —<br />

Verosphacela ebrachia Henry Draisma et al. (2001) AJ287867/ — / —<br />

<strong>ISHIGEALES</strong> ORD. NOV. 925<br />

Phyllariaceae<br />

Phyllariopsis brevipes ssp. brevipes (C. Agardh)<br />

Sasaki et al. (2001) AB045244/ — / —<br />

Henry et South<br />

Saccorhiza polyschides (Lightfoot) Batters Sasaki et al. (2001)/Ile Callot, Roscoff, France, September 1999, PP001/ibid AB045256/AY372965/AY528850<br />

Ralfsiales<br />

Analipus japonicus (Harvey) Wynne Cho et al. (2003)/Boiler Bay, Oregon, USA, 16 May 2001, PR001/ibid AY095323/AY372966/AY528851<br />

Scytothamnales<br />

Scytothamnus australis (J. Agardh) Hooker et Harvey Peters and Ramírez (2001)/Scorching Bay, Wellington, New Zealand, 3 August 2001, AJ295833/AY372967/AY528852<br />

PS001/ibid<br />

Splachnidium rugosum (Linnaeus) Greville Peters and Ramírez (2001)/Scorching Bay, Wellington, New Zealand, 3 August 2001, AJ295834/AY372968/AY528853<br />

PS002/ibid<br />

Sphacelariales<br />

Alethocladus corymbosus (Dickie) Sauvageau Draisma et al. (2001) AJ287860/ — / —<br />

Halopteris filicina (Grateloup) Kützing Draisma et al. (2001)/Seongsan, Jeju, Korea, DNA from Y. S. Keum/ibid AJ287894/AY372969/AY528854<br />

Sphacelaria divaricata Montagne Draisma et al. (2001) AJ287889/ — / —<br />

S. divaricata Seongsan, Jeju, Korea, DNA from Y. S. Keum AY372985/AY372970/AY528855<br />

Sporochnales<br />

Carpomitra costata (Stackhouse) Batters Sasaki et al. (2001)/Munseom, Jeju, Korea, 8 April 2003, PSP001/ibid AB045257/AY372971/AY528856<br />

Sporochnus radiciformis (B. Brown ex Turner) C. Agardh Kawai and Sasaki (2000) as S. scoparuis Harvey /Cheonbu, Ulreungdo, Korea,<br />

AB037142/AY528861/AY528857<br />

26 August 2003, PSP002/ibid<br />

Syringodermatales<br />

Microzonia velutina (Harvey) J. Agardh Burrowes et al. (2003) AY157697/ — / —<br />

Syringoderma phinneyi Henry et Müller Draisma et al. (2001)/Moe 2, Culture of Prof. D. G. Müller (Konstanz, Germany)/ibid AJ287868/AY528862/AY528858<br />

Tilopteridales<br />

Haplospora globosa Kjellman Kawai and Sasaki (2000) AB037138/ — / —<br />

Tilopteris mertensii (Turner in Smith) Kützing Sasaki et al. (2001) AB045260/ — / —<br />

Incertae sedis<br />

Asterocladon lobatum Müller et al. Peters and Ramírez (2001) AJ295824/ — / —<br />

Asteronema rhodochortonoides (Boergesen) Müller et Parodi Peters and Ramírez (2001) AJ295825/ — / —<br />

Phaeothamniophyceae<br />

Phaeothamnion confervicola G. Lagerheim Bailey et al. (1998) AF064746/ — / —<br />

Schizocladiophyceae<br />

Schizocladia ischiensis Henry, Okuda et Kawai Kawai et al. (2003)/CCMP2287/ibid AB085615/AY528863/AY528859<br />

Tribophyceae<br />

Tribonema aequale Pascher Bailey and Andersen (1998)/UTEX 50/ibid AF084611/AY372972/AY528860<br />

CCMP, Provasoli-Guillard National Center for Culture of Marine Phytoplankton; UTEX, Culture Collection of Algae at the University of Texas at Austin. Bold numbers<br />

indicate new sequences published in the present study.


926<br />

TABLE 2. Nucleotide composition of the rbcL, psaA, and psbA and statistics from MP analyses of the individual and combined<br />

data sets including outgroups.<br />

cylindrica (1372 nt), Sargassum horneri (1292 nt), and<br />

Sphacelaria divaricata (1383 nt), which were amplified<br />

with different primers. The 70 aligned rbcL sequences<br />

had 685 (46.7%) variable bases and 558 (38.0%)<br />

parsimoniously informative sites. There were excesses<br />

of adenine (29.49%) and thymine (32.21%) at all<br />

codon positions. Transversions were more common<br />

than transitions for all codon positions (Ti/Tv 5 0.95)<br />

(Table 2).<br />

The uncorrected sequence divergence (p-distance)<br />

values for the rbcL region within Ishige ranged from<br />

6.77% between I. okamurae and I. sinicola from Korea to<br />

7.10% between I. okamurae and I. sinicola from Japan.<br />

The samples of I. okamurae from Korea and Japan differed<br />

by seven nucleotides (0.51% sequence divergence),<br />

and there was a difference of two nucleotides<br />

(0.15% sequence divergence) between I. sinicola from<br />

Korea and Japan. Ishige differed from other ectocarpalean<br />

algae between 14.01% (between I. okamurae from<br />

Japan and Dictyosiphon foeniculaceus) and 16.61%<br />

(between I. sinicola from Japan and Pylaiella littoralis)<br />

sequence divergence.<br />

The psaAalignment. The sequences determined for<br />

the psaA region totaled 1488 nt. For the 38 aligned<br />

sequences of the psaA gene, 661 (44.4%) bases were<br />

variable and 578 (38.8%) were parsimoniously informative.<br />

There were excesses of adenine and thymine<br />

at all codon positions (29.52% and 35.62%,<br />

respectively). Transversions and transitions occurred<br />

at approximately similar frequencies for all codon<br />

positions (Ti/Tv 5 1.02) (Table 2).<br />

The sequence divergence for the psaA gene within<br />

Ishige ranged from 8.27% (between I. okamurae from<br />

Japan and I. sinicola from Korea) to 8.74% (between<br />

I. okamurae Korea and I. sinicola from Japan). Ishige<br />

okamurae from Korea and Japan differed by 11 nt<br />

(0.74% sequence divergence), and I. sinicola from Korea<br />

and Japan differed by 5 nt (0.34% sequence divergence).<br />

Ishige diverged from other ectocarpalean algae<br />

between 16.4% (between I. sinicola from Korea and<br />

Dictyosiphon foeniculaceus) and 18.82% (between<br />

I. okamurae from Japan and Adenocystis utricularis).<br />

The psbA alignment. The sequences determined<br />

for the psbA region totaled 885 bp. For the 39 aligned<br />

rbcL psaA psbA Combined<br />

Number of taxa 70 38 39 38<br />

Nucleotides (bp) 1467 1488 885 3840<br />

Base frequency (A/C/G/T) 0.2949/0.1627/0.2203/0.3221 0.2952/0.1564/0.1922/0.3562 0.2467/0.1825/0.2077/0.3632<br />

Number of transitions/ 191379/200661 (0.95) 74550/73082 (1.02) 29095/23450 (1.24)<br />

transversions (Ti/Tv ratio)<br />

GA YOUN CHO ET AL.<br />

Variable sites (%) 685 (46.7) 661 (44.4) 306 (34.6) 1583 (41.2)<br />

Informative sites (%) 558 (38.0) 578 (38.8) 226 (25.5) 1296 (33.8)<br />

Number of MP trees 2 2 4 1<br />

MP tree length 4062 3387 1021 7243<br />

Consistency index 0.278 0.333 0.378 0.344<br />

Retention index 0.561 0.457 0.554 0.477<br />

sequences of the psbA gene, 306 (34.6%) bases were<br />

variable and 226 (25.5%) were parsimoniously informative.<br />

There were excesses of adenine and thymine<br />

at all codon positions (24.67% and 36.32%,<br />

respectively). Transitions were more common than<br />

transversion for all codon positions (Ti/Tv 5 1.24)<br />

(Table 2).<br />

The sequence divergence for the psbA gene within<br />

Ishige ranged from 6.44% (between I. okamurae and I.<br />

sinicola from Japan) to 6.89% (between I. okamurae and<br />

I. sinicola from Korea). Ishige okamurae from Korea and<br />

Japan differed by 6 nt (0.68% sequence divergence),<br />

and I. sinicola from Korea and Japan differed by 2 nt<br />

(0.23% sequence divergence). Ishige diverged from<br />

other ectocarpalean algae between 9.49% (between<br />

I. okamurae from Korea and Petalonia fascia) and<br />

11.86% (between I. sinicola from Korea and Adenocystis<br />

utricularis).<br />

Saturation tests. Saturation tests showed that the<br />

first and second codon positions for all three genes<br />

were conserved and the third codon positions were<br />

the most variable. The scatter plots (not shown) for<br />

the rbcL, psaA, and psbA sequences were linear and<br />

showed no evidence of multiple hit problems for all<br />

three codon positions. Therefore, we included all<br />

events for the three codon positions in the phylogenetic<br />

analysis.<br />

Partition homogeneity tests. Partition homogeneity<br />

tests between rbcL and psaA were significant<br />

(P 5 0.002, mean of three tests); however, with<br />

Desmarestia or the kelp genera (Chorda, Alaria, and<br />

Laminaria) excluded, the results were not significant<br />

(P 5 0.102 and 0.150, respectively), indicating congruence.<br />

The psbA data was incongruent with the<br />

psaA andrbcL data sets (at the P 0.01 level) even<br />

with Desmarestia and/or the kelp genera excluded.<br />

The P value was not changed when outgroup taxa or<br />

uninformative sites were excluded. So, we provided<br />

all the phylogenetic trees reconstructed from individual<br />

and combined data sets.<br />

Phylogenetic relationships. The independent analyses<br />

of rbcL, psaA, psbA, and rbcL þ psaA þ psbA data<br />

sets resulted in congruent, though not identical,<br />

phylogenetic reconstructions. The statistics for the


MP analyses are compared among individual and<br />

combined data sets (Table 2), and the ML trees for<br />

all four data sets are shown in Figures 1 through 4.<br />

The rbcL tree (Fig. 1) showed that all the phaeophycean<br />

algae investigated here were strongly monophyletic<br />

(100% BS for ML, 92% BS for MP, and<br />

PP 5 1), having the Schizocladiophyceae as the sister<br />

taxon. The basal-most taxon of the Phaeophyceae was<br />

the Choristocarpaceae. Ishige okamurae and I. sinicola<br />

from Korea and Japan were strongly monophyletic<br />

(100% BS for ML and MP and PP 5 1), and the Ishigeaceae<br />

was the sister taxon to all other remaining<br />

taxa (100% BS for ML, 79% BS for MP, and PP 5 1).<br />

The Dictyotales, Onslowiaceae, Sphacelariales, and<br />

Syringodermatales formed a clade, although the clade<br />

was only supported by Bayesian PP. The remaining 11<br />

higher taxa, including the Phyllariaceae and Asterocladon/Asteronema<br />

group, formed a monophyletic clade<br />

(56% BS for ML, 51% BS for MP, and PP 5 1). The<br />

Ectocarpales, comprising the Acinetosporaceae, Adenocystaceae,<br />

Chordariaceae, Ectocarpaceae, and Scytosiphonaceae,<br />

was strongly monophyletic (94% BS for<br />

ML, 90% BS for MP, and PP 5 1) and was placed in a<br />

terminal position. The Laminariales and Asterocladon/<br />

Asteronema group clustered with the Ectocarpales (88%<br />

BS for ML, 85% BS for MP, and PP 5 1). However,<br />

interrelationships of most of other orders and families<br />

were not well resolved.<br />

The psaA tree (Fig. 2) is similar to the rbcL tree in<br />

having the monophyletic Phaeophyceae with the Ishigeaceae<br />

in a basal position and the Ectocarpales in a<br />

terminal position. The clade of the Sphacelariales, Dictyotales,<br />

and Syringodermatales was not supported<br />

with bootstrap values or Bayesian probabilities. The<br />

remaining taxa formed a monophyletic clade (61% BS<br />

for ML, 59% BS for MP, and PP 5 1). However, the<br />

psaA tree is different from the rbcL tree in having the<br />

Desmarestiales as the sister taxon to the Laminariales<br />

(66% BS for ML, 75% BS for MP, and PP 5 1), and the<br />

Laminariales is not monophyletic.<br />

The psbA tree (Fig. 3) had a similar topology to the<br />

rbcL and psaA trees in the positions of the Ishigeaceae<br />

and the Ectocarpales within the Phaeophyceae. However,<br />

the psbA tree is different from the rbcL and psaA<br />

trees in having the Syringodermatales separated from<br />

the clade of the Dictyotales and Sphacelariales. The<br />

latter two orders were not supported with ML and MP<br />

bootstrap values but were supported with 0.91 and<br />

0.97 PP, respectively.<br />

Although the tree of the concatenated data set was<br />

similar to the rbcL tree than the psaA andpsbA trees, all<br />

these trees were congruent in having the monophyletic<br />

Phaeophyceae with the Ishigeaceae in a basal position<br />

and the Ectocarpales in a terminal position (Fig. 4).<br />

Most of orders and families included in the present<br />

study received higher bootstrap values and Bayesian<br />

probabilities than each individual data set. The Dictyotales,<br />

Sphacelariales, and Syringodermatales formed a<br />

clade supported with 65% BS for ML and 1 PP. The<br />

other remaining taxa produced a monophyletic clade<br />

<strong>ISHIGEALES</strong> ORD. NOV. 927<br />

(93% BS for ML, 94% BS for MP, and PP 5 1). The<br />

Laminariales was monophyletic (97% BS for ML, 90%<br />

BS for MP, and PP 5 1) and was shown to be the sister<br />

taxon to the Ectocarpales, but the sister relationship<br />

was not supported with bootstrap values. The interrelationships<br />

between most of orders and families were<br />

not well resolved.<br />

The SH tests using the combined data set showed<br />

that the best position for the Ishigeaceae was on<br />

the basal branch of the Phaeophyceae. The alternative<br />

topologies that force the genus to be placed into a clade<br />

within the Chordariaceae, Ectocarpales, Fucales,<br />

Sphacelariales, Dictyotales, or Syringodermatales<br />

were significantly worse than the best tree (Po0.05,<br />

Table 3).<br />

Morphology of plastids and ultrastructure of Ishige okamurae.<br />

Thalli of I. okamurae (Fig. 5a) contain plastids<br />

in the vegetative cells of the assimilatory<br />

filaments. They are discoid, and the number is three<br />

to four per cell (Fig. 5b). The ultrastructure is typical<br />

of other brown algal plastids. No pyrenoid was found<br />

in most of the plastids in our samples, but a small<br />

exserted pyrenoid was very rarely observed (Fig. 5, c<br />

and d).<br />

DISCUSSION<br />

This study presents the first psaA and psbA phylogenies<br />

of the Phaeophyceae based on 38–39 representatives<br />

from 13 orders or ‘‘ordinal-level families,’’<br />

including two outgroup taxa. The psaA and psbA phylogenies<br />

are compared with the rbcL tree from 70 taxa<br />

from 17 orders, including 4 ordinal-level taxa and 3<br />

outgroup species. To our knowledge, such a large data<br />

set containing three protein-coding plastid genes has<br />

not been used for the Phaeophyceae. Although the<br />

rbcL region is more variable (46.7%) than the psaA<br />

(44.4%) and psbA (34.6%) regions, the psaA region contains<br />

more informative sites (38.8%) than the psbA<br />

(25.5%) and rbcL (38%) regions. Lower consistency index<br />

of the rbcL region (0.278) among three genes is<br />

probably due to the number of taxa investigated. The<br />

result that the consistency index values of the psaA<br />

(0.333) and psbA (0.378) in the present study are lower<br />

than those (0.493 in psaA and 0.383 in psbA) in previous<br />

study of red algae (Yang and Boo 2004) may be<br />

assigned to taxon sampling of brown algae at the ordinal<br />

level than that of red algae at the genus level. A<br />

comparison of other phylogenetic signals in individual<br />

and combined data sets is given in Table 2.<br />

Recently, several studies pointed out that the results<br />

of congruence tests should not be used as criteria for<br />

combining individual data sets. The reliability of the<br />

congruence tests such as the ILD test has been questioned<br />

(Lavoué et al. 2003, Shimabukuro-Dias et al.<br />

2004), and some different data sets, being in conflict<br />

over some parts of a tree, may be congruent over other<br />

parts (Gatesy et al. 1999b). Although there are no signs<br />

of saturation of each gene sequence, the ILD test detected<br />

a significant incongruence (at the Po0.05 level)


928<br />

GA YOUN CHO ET AL.<br />

FIG. 1. ML tree for the Ishigeaceae and other phaeophyceaen algae estimated from the rbcL sequence data (GTR þ G þ I model,<br />

–ln likelihood 5 20839.33098; G 5 0.798767; I 5 0.463896; A–C 5 1.223060, A–G 5 4.104599, A–T 5 1.183624, C–G 5 2.299422,<br />

C–T 5 7.838568, G–T 5 1). The bootstrap values shown above the branches are from ML/MP methods and dashes indicate o50%<br />

support of bootstrap. Thick branches indicate Bayesian posterior probabilities 0.9.


among the rbcL, psaA, and psbA data sets in the present<br />

study. This is a contrast to congruence among the same<br />

three genes in the ceramiaceous red algal tribe Griffithsieae<br />

(Yang and Boo 2004). However, the combined<br />

sequences of the three genes have more resolving<br />

power and clade support for all ordinal taxa including<br />

the three orders, the Laminariales not resolved in the<br />

psaA and psbA trees and the Dictyotales and Sphacelariales<br />

not supported in the psbA tree, than the sequences<br />

of individual gene. The individual and combined data<br />

sets are congruent in phylogenetic positions of the Ishigeaceae<br />

and the Ectocarpales within the Phaeophyceae<br />

investigated. It appears that incongruence among<br />

the three genes is probably attributed to unstable relationships<br />

among orders, which are a feature well<br />

known in phylogenetic reconstructions of the Phaeophyceae<br />

(Draisma et al. 2001, 2003, Rousseau et al.<br />

2001). Our studies show that statistically incongruent<br />

data can be combined for a better understanding of<br />

brown algal phylogeny, although the rbcL, psaA, and<br />

<strong>ISHIGEALES</strong> ORD. NOV. 929<br />

FIG. 2. ML tree for the Ishigeaceae<br />

and other phaeophyceaen algae<br />

estimated from the psaA sequence<br />

data (GTR þ G þ I model, –ln likelihood<br />

5 16551.60179; G 5 0.568987;<br />

I 5 0.492528; A–C 5 2.626896, A–G 5<br />

4.898248, A–T 5 0.072912, C–G 5<br />

3.757704, C–T 5 12.922112, G–T 5 1).<br />

The bootstrap values shown above the<br />

branches are from ML/MP methods<br />

and dashes indicate o50% support of<br />

bootstrap. Thick branches indicate<br />

Bayesian posterior probabilities 0.9.<br />

psbA genes possess different phylogenetic information.<br />

Our results indicate that the psaA and psbA genes, both<br />

being independent to each other (Morden and Sherwood<br />

2002), are useful for other brown algal phylogeny<br />

and have more resolution when combining with<br />

other molecular data such as the rbcL gene.<br />

In the present study, the Phaeophyceae are well resolved,<br />

having the Schizocladiophyceae as sister taxon<br />

in the rbcL and combined data sets, although the<br />

Tribophyceae was the sister to the Phaeophyceae in<br />

the psaA and psbA data sets. Based on trees of individual<br />

and combined data sets, the Phaeophyceae<br />

consist of two large groups. The ‘‘basal group’’ included<br />

the Choristocarpaceae, Ishigeaceae, Dictyotales,<br />

Sphacelariales, and Syringodermatales and the ‘‘crown’’<br />

contained the Ectocarpales, Fucales, Laminariales, and<br />

other many lineages. These two groups are also reflected<br />

in the trees of previous rbcL (Draisma et al. 2001,<br />

2003) and LSU þ partial SSU data sets (Rousseau et al.<br />

2001). However, because the interrelationships of the


930<br />

orders or families within each of the two groups were<br />

not resolved, we focus on the phylogenetic and taxonomic<br />

implications of the Ishigeaceae compared with<br />

other relative orders.<br />

In all phylogenies reconstructed from three protein-coding<br />

genes, the Ectocarpales was placed in a<br />

terminal position, whereas the Ishigeaceae ended up<br />

in a basal position. The monophyly of the Ectocarpales,<br />

which includes the largest number of taxa in our study,<br />

is strongly supported and, within the order, the five<br />

families established by Peters and Ramírez (2001) are<br />

well resolved, although the Chordariaceae appear paraphyletic.<br />

The long-held concept of an ancestral Ectocarpales<br />

and derived Fucales (van den Hoek et al.<br />

1995) is not supported by our psaA andpsbA gene<br />

phylogenies as well as rbcL gene, as is seen in previous<br />

rbcL (Draisma et al. 2001, 2003) and LSU þ partial<br />

SSU trees (Rousseau et al. 2001).<br />

The main finding of our study is that the family Ishigeaceae<br />

does not cluster with members of the Chord-<br />

GA YOUN CHO ET AL.<br />

FIG. 3. ML tree for the Ishigeaceae<br />

and other phaeophyceaen algae estimated<br />

from the psbA sequence data<br />

(GTR þ G þ I model, –ln likelihood 5<br />

5953.44945; G 5 0.472723; I 5 0.402634;<br />

A–C 5 1.245033, A–G 5 9.15766, A–T 5<br />

8.589725, C–G 5 1.088511, C–T 5<br />

23.338746, G–T 5 1]. The bootstrap<br />

values shown above the branches are<br />

from ML/MP methods and dashes indicate<br />

o50% support of bootstrap.<br />

Thick branches indicate Bayesian posterior<br />

probabilities 0.9.<br />

ariaceae such as Chordaria flagelliformis or other taxa of<br />

the Ectocarpales. These results contradict the current<br />

classifications of Yoshida (1998) that assign the family<br />

to the Chordariales (or the Chordariaceae s.s.). Our<br />

trees confirm that the Ectocarpales s.l. (Peters and<br />

Ramírez 2001), which is circumscribed by a pedunculated<br />

pyrenoid in discoid plastids and mostly a heteromorphic<br />

life history (Peters and Ramírez 2001, Draisma<br />

et al. 2003), cannot contain the genus Ishige. Instead,<br />

all analyses of our rbcL, psaA, psbA, and combined data<br />

sets indicate that the family Ishigeaceae consistently<br />

forms a basal group in the Phaeophyceae. The Ishigeaceae<br />

was the basal most in the psaA, psbA, combined<br />

data sets, and SH tests because Choristocarpus is<br />

not included; however, the Ishigeaceae was the penultimate<br />

basal taxon, after the Choristocarpaceae, in the<br />

rbcL tree. These results suggest that the Ishigeaceae<br />

might have diverged early in brown algal evolution.<br />

This is the first time that such a hypothesis has<br />

been proposed on either morphological or molecular


evidence. A closer look on morphological features of<br />

the Ishigeaceae that support their probable early divergence<br />

is described below.<br />

From the ultrastructural details of the vegetative<br />

cells of I. okamurae from Korea, it is clear that most cells<br />

do not have a pyrenoid or vary rarely have a pyrenoid.<br />

The pyrenoid is very small, exserted (Fig. 5, c and d),<br />

and similar to that of I. okamurae from Japan (Fig. 12 in<br />

Hori 1971). The small exserted pyrenoid of the species<br />

is considered a relatively rudimentary type when compared<br />

with the elaborate pedunculate pyrenoid of the<br />

Ectocarpales (Evans 1968). Ishige sinicola also lacks a<br />

pyrenoid (Hori 1971, Hori and Ueda 1975). Therefore,<br />

we conclude that the Ishigeaceae lack a pyrenoid<br />

in plastid. We suspect that Hori (1971) might wonder<br />

at the absence of a pyrenoid in Ishige, believing it to<br />

belong to the Chordariales, which have pyrenoids. The<br />

vegetative cells of the Dictyotales, Fucales, Laminariales,<br />

and Sphacelariales also lack pyrenoids, although<br />

a rudimentary form occurs in their gametes or zoo-<br />

<strong>ISHIGEALES</strong> ORD. NOV. 931<br />

FIG. 4. ML tree for the Ishigeaceae<br />

and other phaeophyceaen algae<br />

estimated from combined rbcL þ<br />

psaA þ psbA sequence data (GTR þ G þ<br />

I model, –ln likelihood 5 37313.27150;<br />

G 5 0.756446; I 5 0.516176; A–C 5<br />

2.052989, A–G 5 5.063074, A–T 5<br />

1.442179, C–G 5 2.684154, C–T 5<br />

12.351116, G–T 5 1). The bootstrap<br />

values shown above the branches are<br />

from ML/MP methods and dashes indicate<br />

o50% support of bootstrap.<br />

Thick branches indicate Bayesian posterior<br />

probabilities 0.9.<br />

spores (Evans 1966, Chi 1971). Therefore, the presence<br />

of the ‘‘true’’ pyrenoid (not a ‘‘rudimentary’’ one<br />

in which the exact nature is not demonstrated and<br />

perhaps not homologous) is apomorphic in the Phaeophyceae.<br />

Our results do not accord with the view of<br />

Evans (1966, 1968) that the presence of the pyrenoid is<br />

plesiomorphic. However, because the occurrence and<br />

development of the pyrenoid varies with the life history<br />

stage and physiological conditions (Bourne and<br />

Cole 1968), further ultrastructural studies of the<br />

plastids of various brown algae are required before<br />

the phylogenetic significance of the pyrenoid can be<br />

substantiated.<br />

The two species of Ishigeaceae grow via small apical<br />

cells (Fig. 1 in Lee et al. 2003). Apical growth occurs in<br />

all other members of the basal group, such as the<br />

Choristocarpaceae, Dictyotales, Sphacelariales, Onslowiaceae,<br />

and Syringodermatales, which have an apical<br />

cell, a group of apical cells, or marginal cells cutting<br />

off segments proximally (Prud’homme van Reine


932<br />

TABLE 3. Results of the Shimodaira-Hasegawa tests used to<br />

evaluate alternative hypotheses of the Ishigeaceae position<br />

in the phylogeny of the psaA þ psbA þ rbcL data set (see<br />

Fig. 4).<br />

Hypotheses tested –ln L Difference –ln L P<br />

Basal position 37313.27150 best<br />

Chordariaceae 37415.16034 101.88883 0.0000 a<br />

Ectocarpales 37366.85999 53.58849 0.0020 a<br />

Fucales 37395.40131 82.12981 0.0000 a<br />

Sphacelariales 37368.18768 54.91617 0.0118 a<br />

Dictyotales 37366.71234 53.44084 0.0143 a<br />

Syringodermatales 37359.43507 46.16357 0.0193 a<br />

a Significant at 0.05 level.<br />

1982, Henry 1984, Bold and Wynne 1985). We agree<br />

with Draisma et al. (2001, 2003) and Rousseau et al.<br />

(2001) that apical growth is plesiomorphic. In this context,<br />

it is interesting to speculate whether apical growth<br />

in spermatochnacean ectocarpoids, such as Chordariopsis,<br />

Spermatochnus, and Stilophora (Fritsch 1945), is<br />

plesiomorphic or convergent. However, the apical<br />

growth in the Fucales may be derived in that the apical<br />

meristem is composed of one or more cells that can<br />

divide in several directions, generating three-dimensional<br />

thalli (Graham and Wilcox 2000). In other<br />

crown group algae, thalli grow for trichothallic, intercalary,<br />

or marginal meristems (Bold and Wynne 1985,<br />

Graham and Wilcox 2000).<br />

The internal tissues of the thallus of the Ishige<br />

species consist of cortex and medulla; the cortex<br />

is composed of assimilatory filaments and pseudoparenchymatous<br />

tissue (Ajisaka 1989, Figs. 5 and 11<br />

in Lee et al. 2003), and the medulla contains colorless<br />

hypheal cells connected to adjacent hypheal cells or<br />

directly to cortical cells (Ajisaka 1989, Tanaka in Hori<br />

1993, Lee et al. 2003). This oligostichous organization<br />

is similar to the haplostichous structure of uniseriate<br />

filament of the Choristocarpaceae (Fritsch 1945).<br />

Both haplostichous and polystichous structures are<br />

found in the Sphacelariales (Fritsch 1945, Prud’homme<br />

van Reine 1982). However, the Dictyotales and<br />

Syringodermatales have polystichous organization<br />

alone (Fritsch 1945, Henry 1984, Draisma et al.<br />

2003). Therefore, a series of evolutionary changes<br />

may have led from haplostichous structures to polystichous<br />

structures in the internal organization of the<br />

thalli of the basal brown algae. However, it appears that<br />

the internal organization of the thallus might have<br />

evolved independently several times among the crown<br />

group of the brown algae (Draisma et al. 2003).<br />

One of the most distinctive characters of the Ishigeaceae<br />

is the presence of the cryptostomata on the<br />

surface of thalli, which are well illustrated in previous<br />

studies (Fig. 8 in Yendo 1907, Figs. 2 and 10 in Lee<br />

et al. 2003). Phaeophycean hairs originate from cells<br />

in the medulla. In the presence of cryptostomata,<br />

the Ishigeaceae is similar to crown groups, such as<br />

the Fucales, Scytothamnales, and Adenocystis of the<br />

Ectocarpales (Clayton 1984, 1985). However, the<br />

GA YOUN CHO ET AL.<br />

scytothamnalean cryptostomata mature into conceptacles<br />

containing unilocular sporangia, whereas the<br />

fucalean cryptostomata do not form gametangial<br />

branches. There are a variety of other genera of the<br />

Ectocarpales, such as Colpomenia, Chnoospora, and Leathesia,<br />

in which there are groups of phaeophycean<br />

hairs growing from shallow epithermal pits (Fritsch<br />

1945, Clayton 1984). Our molecular phylogenies<br />

based on the rbcL, psaA, and psbA genes place taxa<br />

with cryptostomata, viz. Ishigeaceae, Scytothamnales,<br />

and Fucales, in different clades. Adenocystis formed a<br />

clade independent from the scytosiphonacean genera,<br />

despite their belonging to the Ectocarpales. These results<br />

demonstrate that the cryptostomata in the family<br />

Ishigeaceae is a homoplastic character that misled<br />

Yendo (1907) into classifying the genus in the family<br />

Fucaceae. At the present state of knowledge, the<br />

occurrence of cryptostomata in the Phaeophyceae is<br />

evolutionarily enigmatic. On the other hand, the<br />

phaeophycean hairs are present on branches in the<br />

Sphacelariales (Prud’homme van Reine 1982) or are<br />

associated with sporangial sori in the Dictyotales (Bold<br />

and Wynne 1985), whereas they are absent in the<br />

Choristocarpaceae, Onslowiaceae, and Syringodermatales<br />

(Henry 1984, 1987, Womersley 1987, Draisma<br />

and Prud’homme van Reine 2001). Our multigene<br />

trees do not provide phylogenetic correlation between<br />

the cryptostomata and the phaeophycean hairs.<br />

The unilocular sporangia of I. okamurae are produced<br />

from the outermost cortical cells (Tanaka in<br />

Hori 1993, Lee et al. 2003). The plurilocular sporangia<br />

of the species originate from the assimilatory filaments,<br />

are uniseriate, and lack sterile terminal cells<br />

(Ajisaka 1989, Tanaka in Hori 1993, Lee et al. 2003).<br />

Both unilocular and plurilocular sporangia occur in<br />

the thalli of I. sinicola from the south coast of Korea<br />

(Lee et al. 2003). The observations of Ajisaka (1989)<br />

and Lee et al. (2003) contradict those of Arasaki (1943)<br />

that the plurilocular sporangia of I. sinicola in culture<br />

were gametangia and similar to those of certain species<br />

of the Chordariales, in which the unilocular sporangia<br />

are always formed in the basal or middle part of the<br />

assimilatory filaments, and both types of sporangia are<br />

present on the sporophytes of some species (Inagaki<br />

1958, Ajisaka 1989, Yoshida 1998). However, the<br />

Dictyotales have meiosporic tetrasporangia (Bold and<br />

Wynne 1985), and the Choristocarpaceae, Onslowiaceae,<br />

and Sphacelariales have unilocular zoidangia,<br />

plurilocular zoidangia, or multicellular propagules<br />

(Prud’homme van Reine 1982).<br />

Ajisaka (1989) reported that plurispores of I. okamurae,<br />

released from the plurilocular sporangia, develop<br />

directly into typical filamentous thalli through a<br />

pseudoparenchymatous prostrate disc, which is sterile<br />

and functions as a perennial holdfast system. A similar<br />

perennial basal system is found in the Sphacelariales,<br />

including the Choristocarpaceae (Fritsch 1945). Perennial<br />

systems are found in Ralfsia verrucosa (Areschoug)<br />

Areschoug, Desmarestia aculeata (Linnaeus)<br />

Lamouroux, and the Aglaozonia-stage of Cutleria monoica


Ollivier (Fritsch 1945). Because the latter three taxa<br />

were not included in the present study, they are beyond<br />

the scope of our discussion. Tanaka (in Hori<br />

1993) demonstrated that I. okamurae has an isomorphic<br />

life cycle, consisting of sporophytes that form plurilocular<br />

sporangia and gametophytes bearing unilocular<br />

sporangia. In the same chapter, Tanaka doubted the<br />

heteromorphic life history of I. sinicola observed by<br />

Arasaki (1943). Both plurilocular and unilocular sporangia<br />

occurred in erect thalli of I. okamurae (Figs. 7 and<br />

8 in Lee et al. 2003) and I. sinicola (Figs. 15 and 16 in<br />

Lee et al. 2003) during a year-round collection from<br />

Namhaedo on the south coast of Korea. All these studies<br />

led us to conclude that the family Ishigeaceae has<br />

an isomorphic life history, with macroscopic sporophytes<br />

and gametophytes. Isomorphic life cycles have<br />

been reported for the Dictyotales, Sphacelariales, and<br />

Onslowiaceae (Prud’homme van Reine 1982, Henry<br />

<strong>ISHIGEALES</strong> ORD. NOV. 933<br />

FIG. 5. Thallus and plastid of Ishige okamurae. (a) A herbarium specimen collected on 28 June 2001 in Geojedo, Korea. Scale bar,<br />

2 cm. (b) Discoid plastids (arrows) in assimilatory cell. Scale bar, 10 mm. (c) Ultrastructure of a plastid, showing thylakoids, without<br />

pyrenoid. Scale bar, 1 mm. (d) A small exserted pyrenoid (arrow) in a plastid. Scale bar, 0.4 mm.<br />

1987, van den Hoek et al. 1995, Draisma and Prud’homme<br />

van Reine 2001). In the Choristocarpaceae, thalli<br />

with either uni- or plurilocular sporangia look similar<br />

and could also indicate an isomorphic life history<br />

(Fritsch 1945, Burrowes et al. 2003). However, the<br />

Syringodermatales have a heteromorphic life history<br />

(Henry 1984). Although some members of the Scytosiphonaceae<br />

have an isomorphic type (Cho et al.<br />

2003), the Ectocarpales and other brown algae have<br />

heteromorphic life histories or single diploid generations<br />

(Graham and Wilcox 2000).<br />

Interestingly, both species of Ishige are distributed in<br />

the warmer waters of the Pacific Ocean (Yendo 1907,<br />

Setchell and Gardner 1924, Tseng 1983, Lee et al.<br />

2003). Choristocarpus tenellus of the Choristocarpaceae<br />

inhabits the Mediterranean Sea (Fritsch 1945). The<br />

Dicytotales are very diverse in tropical and subtropical<br />

waters (Bold and Wynne 1985), and the Sphacelariales


934<br />

TABLE 4. A taxonomic comparison of the Ishigeaceae with other basal brown algae that have symplesiomorphic characters such as apical growth and discoid pyrenoidless<br />

plastids.<br />

Ishigeaceae Choristocarpaceae Onslowiaceae Sphacelariales Dictyotales Syringodermatales<br />

Thallus construction Oligostichous Haplostichous Oligostichous Haplostichous or Polystichous Oligostichous<br />

of sporophytes<br />

polystichous<br />

Perennial disc Present Present Present? Present Rarely present Absent<br />

Cryptostomata Present Absent Absent Absent Absent Absent<br />

Phaeophycean hairs Within cryptostomata Absent Absent On branches Associated with<br />

Absent<br />

reproductive organs<br />

Propagules Absent With a large apical cell Without a large Without a large<br />

Absent Absent<br />

apical cell<br />

apical cell<br />

Plurilocular From cortical layer of On branches of On branches of On branches of From the surface From the surface cells of<br />

sporangia<br />

gametophytes gametophytes<br />

gametophytes<br />

gametophytes cells of gametophytes gametophytes<br />

Unilocular Many spores, from the Many spores, on Many spores, on Many spores, on Tetraspores, from the Bi-, tetraspores, from the<br />

sporangia<br />

cortical layer of<br />

branches of<br />

branches of branches of sporophytes surface cells of surface cells of<br />

sporophytes<br />

sporophytes<br />

sporophytes<br />

sporophytes<br />

sporophytes<br />

Gametes Isogametes Isogametes Anisogametes Mostly iso- and Egg and spermatozoids Isogametes<br />

anisogametes<br />

Life cycle Isomorphic Probably isomorphic Isomorphic Mostly isomorphic Isomorphic Heteromorphic<br />

Pheromone Not detected Not detected Not detected Ectocarpene/ Dictyotene, multifidene Viridene<br />

hormosirene,<br />

desmarestene<br />

GA YOUN CHO ET AL.<br />

Data from Fritsch (1945), Prud’homme van Reine (1982), Henry (1984, 1987), Womersley (1987), Ajisaka (1989), Tanaka (in Hori 1993), Draisma and Prud’homme van Reine<br />

(2001), Pohnert and Boland (2002), and Lee at al (2003).<br />

and Syringodermatales are distributed from tropical to<br />

Antarctic waters (Prud’homme van Reine 1982, Henry<br />

1984). The distributions of living taxa of these basal<br />

groups speculate that the brown algae may have originated<br />

in tropical waters during a geological period<br />

with warm climatic conditions, whereas most brown<br />

algae in the crown group appear to exhibit the greatest<br />

diversity in terms of species and morphology in cold<br />

waters (Bold and Wynne 1985). The Paleozoic may<br />

have been the period of the earliest brown algae, as<br />

suggested by Clayton (1984), or the Mesozoic, which<br />

roughly corresponds to 155 million years ago based on<br />

rDNA SSU sequence divergence (Medlin et al. 1997)<br />

or approximately 200 million years ago based on 5S<br />

rRNA sequences (Lim et al. 1986).<br />

Our molecular study shows conclusively that the<br />

family Ishigeaceae produces a basal brown algal group<br />

with the Choristocarpaceae, Dictyotales, Sphacelariales,<br />

Onslowiaceae, and Syringodermatales. These early<br />

diverging brown algae have apical growth and<br />

pyrenoidless discoid plastids as morphological<br />

symplesiomorphy. In addition, the Ishigeaceae is similar<br />

to the Dictyotales, Sphacelariales, Onslowiaceae,<br />

and probably Choristocarpaceae in its isomorphic life<br />

history. However, our rbcL, psaA, and psbA phylogenies<br />

indicate that the family Ishigeaceae makes an independent<br />

lineage from the early diverging brown algae<br />

in having oligostichous organization of thallus, cryptostomata,<br />

and uni- and plurilocular sporangia within<br />

cortical layer (Table 4). Our studies on the three protein<br />

coding plastid genes and ultrastructure of plastids<br />

do not confirm the current classification of the family<br />

Ishigeaceae within the Ectocarpales s.l. or Chordariaceae<br />

s. str. Although we did not examine the Ascoseirales,<br />

which is only order of the Phaeophyceae not<br />

included in the present study because of no collection,<br />

it does not appear that the inclusion of the taxon would<br />

change our finding that the Ishigeaceae is a distinct<br />

basal group clearly separated from other brown algal<br />

orders or families. The unique molecular and ultrastructural<br />

evidence that defines the Ishigeaceae as an<br />

early diverging but independent brown algal group is<br />

best expressed by placement of the family in a separate<br />

order, Ishigeales ord. nov.<br />

Ishigeales ord. nov. G. Y. Cho et Boo<br />

Diagnosis: Novus ordo Phaeophycearum. Plantae<br />

isomorphicae, ad 20 cm altae, ex haptero parvo aut basi<br />

crustosa extenta crescentes, epiphyticae vel epilithicae.<br />

Augmen per cellulas apicales. Frondes ramosae teretes vel<br />

foliosae, medulla corticeque instructae. Cortex pseudoparenchymatus,<br />

filamentis assimilantibus. Cellulae plastides<br />

aliquot discoideos sine pyrenoide continentes. Medulla cum<br />

hyphis. Pili phaeophycei caespitosi, e cryptostomatibus<br />

orientes. Sporangia unilocularia e cellulis corticalibus<br />

transformatis facta, terminalia. Sporangia plurilocularia<br />

e filamentis assimilantibus transformatis facta, uniseriata,<br />

sine cellula terminali.<br />

Type family: Ishigeaceae Okamura in Segawa with<br />

the same characteristics as the order.


Type genus: Ishige Yendo<br />

New order of Phaeophyceae. Plants isomorphic, to<br />

20 cm high, growing from a small holdfast, or an<br />

extended crustose base, epiphytic or epilithic.<br />

Growth from apical cells. Fronds branched, terete,<br />

or foliose, with medulla and cortex. Cortex pseudoparenchymatous,<br />

with assimilatory filaments. Cells<br />

containing several discoid plastids without pyrenoids.<br />

Medulla with hyphae. Phaeophycean hairs clustered,<br />

growing from cryptostomata. Unilocular sporangia<br />

transformed from cortical cells, terminal. Plurilocular<br />

sporangia transformed from assimilatory filaments,<br />

uniseriate, lacking a terminal cell.<br />

We appreciate Drs. Joana Kain, Nina Klotchkova, Wendy Nelson,<br />

and Nathalie Simon for help with the collection of samples<br />

and their identification; Drs. Christos Kataros, Tae Jun<br />

Han, and Wook Jae Lee for sharing their samples; Drs.<br />

Suzanne Fredericq, Giovanni Furnari, and Mark A. Garland<br />

for Latin translation; and Dr. Willem F. Prud’homme van<br />

Reine for valuable comments. We thank W. J. Lee for providing<br />

his unpublished sequences and Dr. Yeun Sim Keum for<br />

sharing DNA stocks. We sincerely thank Drs. Debashish<br />

Bhattacharya and Hwan Su Yoon for providing information<br />

on the psaA and psbA regions. The director of the Marine Station<br />

in Chiba, Japan arranged accommodation of the collection.<br />

This study was supported by KRF grant 2002-070-<br />

C00083 to S. M. B.<br />

Ajisaka, T. 1989. Plurilocular sporangia and the development of<br />

plurispores in Ishige okamurae Yendo (Phaeophyceae) from the<br />

Kada coast, Wakayama Prefecture in Japan. Jpn. J. Phycol.<br />

37:17–22.<br />

Arasaki, S. 1943. On the life history of Ishige foliacea Okamura. Bot.<br />

Mag. Tokyo 57:34–41.<br />

Assali, N. E., Mache, R. & de Göer, S. L. 1990. Evidence for a<br />

composite phylogenetic origin of the plastid genome of the<br />

brown alga Pylaiella littoralis (L.) Kjellm. Plant Mol. Biol.<br />

15:307–15.<br />

Bailey, J. C. & Andersen, R. A. 1998. Phylogenetic relationships<br />

among nine species of the Xanthophyceae inferred from rbcL<br />

and 18S rRNA gene sequences. Phycologia 37:458–66.<br />

Bailey, J. C., Bidigare, R. R., Christensen, S. J. & Andersen, R. A.<br />

1998. Phaeothamniophyceae classis nova: a new lineage of<br />

Chromophytes based upon photosynthetic pigments, rbcL<br />

sequence analysis and ultrastructure. Protist 149:245–63.<br />

Bold, H. C. & Wynne, M. J. 1985. Introduction to the Algae: Structure<br />

and Reproduction. Prentice Hall, Englewood Cliffs, NJ, 720 pp.<br />

Bourne, V. L. & Cole, K. 1968. Some observations on the fine<br />

structure of the marine brown alga Phaeostrophion irregulare.<br />

Can. J. Bot. 46:1369–79.<br />

Bull, J. J., Huelsenbeck, J. P., Cunningham, C. W., Swofford, D. L.<br />

& Waddell, P. J. 1993. Partitioning and combining data in<br />

phylogenetic analysis. Syst. Biol. 42:384–97.<br />

Burrowes, R., Rousseau, F., Müller, D. G. & de Reviers, B. 2003.<br />

Taxonomic placement of Microzonia (Phaeophyceae) in<br />

Syringodermatales based on rbcL and 28S nrDNA sequences.<br />

Cryptog. Algol. 24:63–73.<br />

Chi, E. Y. 1971. Brown algal pyrenoids: brief report. Protoplasma<br />

72:101–4.<br />

Chihara, M. 1969. Pseudogloiophloea okamurai (Setchell) comb. nov.<br />

and Ishige sinicola (Setchell and Gardner) comb. nov. Bull. Jpn.<br />

Soc. Phycol. 17:1–4.<br />

Cho, T. O., Cho, G. Y., Yoon, H. S., Boo, S. M. & Lee, W. J. 2003.<br />

New records of Myelophycus cavus (Scytosiphonaceae, Phaeophyceae)<br />

in Korea and the taxonomic position of the genus on<br />

the basis of a plastid DNA phylogeny. Nova Hedw. 76:381–97.<br />

Clayton, M. N. 1984. Evolution of the Phaeophyta with particular<br />

reference to the Fucales. Prog. Phycol. Res. 3:11–46.<br />

<strong>ISHIGEALES</strong> ORD. NOV. 935<br />

Clayton, M. N. 1985. A critical investigation of the vegetative anatomy,<br />

growth and taxonomic affinities of Adenocystis, Scytothamnus,<br />

and Splachnidium (Phaeophyta). Br. Phycol. J. 20:285–96.<br />

de Reviers, B. & Rousseau, F. 1999. Towards a new classification of<br />

the brown algae. Prog. Phycol. Res. 13:107–201.<br />

Draisma, S. G. A., Peters, A. F. & Fletcher, R. L. 2003. Evolution<br />

and taxonomy in the Phaeophyceae: effects of the molecular<br />

age on brown algal systematics. In Norton, T. A. [Ed.] Out of the<br />

Past: Collected Reviews to Celebrate the Jubilee of the British Phycological<br />

Society. British Phycology Society, Belfast, pp. 87–102.<br />

Draisma, S. G. A. & Prud’homme van Reine, W. F. 2001. Onlsowiaceae<br />

fam.nov. (Phaeophyceae). J. Phycol. 37:647–9.<br />

Draisma, S. G. A., Prud’homme van Reine, W. F., Stam, W. T. &<br />

Olsen, J. L. 2001. A reassessment of phylogenetic relationships<br />

within the Phaeophyceae based on RUBISCO large subunit<br />

and ribosomal DNA sequences. J. Phycol. 37:586–603.<br />

Evans, L. V. 1966. Distribution of pyrenoids among some brown<br />

algae. J. Cell Sci. 1:449–54.<br />

Evans, L. V. 1968. Chloroplast morphology and fine structure in<br />

British fucoids. New Phytol. 67:173–8.<br />

Farris, J. S., Källersjö, M., Kluge, A. G. & Bult, C. 1994. Testing<br />

significance of incongruence. Cladistics 10:315–9.<br />

Fritsch, F. E. 1945. The Structure and Reproduction of the Algae. II.<br />

Foreword, Phaeophyceae, Rhodophyceae, Myxophyceae. University<br />

Press, Cambridge, 939 pp.<br />

Gatesy, J., Milinkovitch, M., Waddell, V. & Stanhope, M. 1999a.<br />

Stability of cladistic relationships between Cetacea and higherlevel<br />

Artiodactyl taxa. Syst. Biol. 48:6–20.<br />

Gatesy, J., O’Grady, P. & Baker, R. H. 1999b. Correlation among<br />

data sets in simultaneous analysis: hidden support for phylogenetic<br />

relationships among higher level artiodactyl taxa. Cladistics<br />

15:271–313.<br />

Gilbert, D. G. 1995. SeqPup, a Biological Sequence Editor and Analysis<br />

Program for Macintosh Computers. Biology Department, Indiana<br />

University, Bloomington.<br />

Graham, L. E. & Wilcox, L. W. 2000. Algae. Prentice Hall, NJ,<br />

640 pp.<br />

Hasegawa, M., Kishino, H. & Yano, T. A. 1985. Dating of the<br />

human-ape splitting by a molecular clock of mitochondrial<br />

DNA. J. Mol. Evol. 22:160–74.<br />

Henry, E. C. 1984. Syringodermatales ord. nov. and Syringoderma<br />

floridana sp. nov. (Phaeophyceae). Phycologia 23:419–26.<br />

Henry, E. C. 1987. Morphology and life histories of Onslowia<br />

bahamensis sp. nov. and Verosphacela ebrachia gen. et sp. nov.,<br />

with a reassessment of the Choristocarpaceae (Sphacelariales,<br />

Phaeophyceae). Phycologia 26:182–91.<br />

Hori, T. 1971. Survey of pyrenoid distribution in brown algae. Bot.<br />

Mag. Tokyo 84:231–42.<br />

Hori, T. 1993. An Illustrated Atlas of the Life History of Algae. Vol. 2.<br />

Brown and Red Algae. Uchida Rokakuho, Tokyo, 345 pp.<br />

Hori, T. & Ueda, R. 1975. The fine structure of algal chloroplasts<br />

and algal phylogeny. In Tokida, J. & Hirose, H. [Eds.] Advance<br />

of Phycology in Japan. Dr. W. Junk, The Hague, pp. 11–42.<br />

Huelsenbeck, J. P. & Ronquist, F. 2001. MRBAYES: Bayesian<br />

inference of phylogeny. Bioinformatics 17:754–5.<br />

Inagaki, K. 1958. A systematic study of the order Chordariales<br />

from Japan and its vicinity. Sci. Pap. Inst. Algol. Res. Hokkaido<br />

Univ. 4:87–197.<br />

Kawai, H. 1992. A summary of the morphology of chloroplasts<br />

and flagellated cells in the Phaeophyceae. Korean J. Phycol. 7:<br />

33–43.<br />

Kawai, H., Maeba, S., Sasaki, H., Okuda, K. & Henry, E. C. 2003.<br />

Schizocladia ischiensis: a new filamentous marine chromophyte<br />

belonging to a new class, Schizocladiophyceae. Protist 154:<br />

211–28.<br />

Kawai, H. & Sasaki, H. 2000. Molecular phylogeny of the brown<br />

algal genera Akkesiphycus and Halosiphon (Laminariales),<br />

resulting in the circumscription of the new families Akkesiphycaceae<br />

and Halosiphonaceae. Phycologia 39:416–28.<br />

Kogame, K., Horiguchi, T. & Masuda, M. 1999. Phylogeny of<br />

the order Scytosiphonales (Phaeophyceae) based on DNA<br />

sequences of rbcL, partial rbcS, and partial LSU nrDNA.<br />

Phycologia 38:496–502.


936<br />

Lavoué, S., Sullivan, J. P. & Hopkins, C. D. 2003. Phylogenetic<br />

utility of the first two introns of the S7 ribosomal protein gene<br />

in African electric fishes (Mormyroidea: Teleostei) and congruence<br />

with other molecular markers. Biol. J. Linn. Soc.<br />

78:273–92.<br />

Lee, E. Y., Lee, I. K. & Choi, H. G. 2003. Morphology and nuclear<br />

small-subunit rDNA sequences of Ishige (Ishigeaceae, Phaeophyceae)<br />

and its phylogenetic relationship among selected<br />

brown algal orders. Bot. Mar. 46:193–201.<br />

Lim, B. L., Kawai, H., Hori, H. & Ozawa, S. 1986. Molecular evolution<br />

of 5S ribosomal RNA from red and brown algae. Jpn. J.<br />

Genet. 61:169–76.<br />

Medlin, L. K., Kooistra, W. H. C. F., Potter, D., Saunders, G. W. &<br />

Andersen, R. A. 1997. Phylogenetic relationships of the ‘‘golden<br />

algae’’ (haptophytes, heterokont chromophytes) and their<br />

plastids. Plant Syst. Evol. 11(suppl):187–216.<br />

Morden, C. W. & Sherwood, A. R. 2002. Continued evolutionary<br />

surprises among dinoflagellates. Proc. Natl. Acad. Sci. USA<br />

99:11558–60.<br />

Müller, D. G., Parodi, E. R. & Peters, A. F. 1998. Asterocladon lobatum<br />

gen. et sp. nov., a new brown alga with stellate chloroplast arrangement,<br />

and its systematic position judged from nuclear<br />

rDNA sequences. Phycologia 37:425–32.<br />

Okamura, K. 1936. Nihon Kaiso-shi. Uchida Rokakuho, Tokyo, 964<br />

pp.<br />

Pearson, G., Serrão, E. & Cancela, M. L. 2001. Suppression subtractive<br />

hybridization for studying gene expression during<br />

aerial exposure and desiccation in fucoid algae. Eur. J. Phycol.<br />

36:359–66.<br />

Peters, A. F. & Clayton, M. N. 1998. Molecular and morphological<br />

investigation of three brown algal genera with stellate plastids:<br />

evidence for Scytothamnales ord. nov. (Phaeophyceae). Phycologia<br />

37:106–13.<br />

Peters, A. F. & Ramírez, M. E. 2001. Molecular phylogeny of small<br />

brown algae, with special reference to the systematic position<br />

of Caepidium antarcticum (Adenocystaceae, Ectocarpales). Cryptog.<br />

Algol. 22:187–200.<br />

Pohnert, G. & Boland, W. 2002. The oxylipin chemistry of attraction<br />

and defense in brown algae. Nat. Prod. Rep. 19:108–22.<br />

Posada, D. & Crandall, K. A. 1998. Modeltest: Testing the model of<br />

DNA substitution. Bioinformatics 14:817–8.<br />

Prud’homme van Reine, W. F. 1982. A Taxonomic Revision of the<br />

European Sphacelariaceae (Sphacelariales, Phaeophyceae). Leiden<br />

University Press, Leiden, 293 pp.<br />

Reynolds, E. S. 1963. The use of lead citrate at high pH as an<br />

electron-opaque stain in electron microscopy. J. Cell. Biol.<br />

17:208–12.<br />

Rousseau, F., Burrowes, R., Peters, A. F., Kuhlenkamp, R. & de<br />

Reviers, B. 2001. A comprehensive phylogeny of the Phaeophyceae<br />

based on nrDNA sequences resolves the earliest<br />

divergences. C. R. Acad. Sci. Paris Sci. Life Sci. 324:305–19.<br />

Rousseau, F. & de Reviers, B. 1999. Circumscription of the order<br />

Ectocarpales (Phaeophyceae): bibliographical synthesis and<br />

molecular evidence. Cryptog. Algol. 20:5–18.<br />

Rousseau, F., de Reviers, B., Leclerc, M.-C., Asensi, A. & Delépine,<br />

D. 2000. Adenocystaceae fam. nov. (Phaeophyceae), a new<br />

GA YOUN CHO ET AL.<br />

family based on morphological and molecular evidence. Eur.<br />

J. Phycol. 35:35–40.<br />

Sasaki, H., Flores-Moya, A., Henry, E. C., Müller, D. G. & Kawai, H.<br />

2001. Molecular phylogeny of Phyllariaceae, Halosiphonaceae<br />

and Tilopteridales (Phaeophyceae). Phycologia 40:123–34.<br />

Segawa, S. 1935. On the marine algae of Susaki, Prov. Idzu and its<br />

vicinity. Sci. Pap. Inst. Alg. Res., Fac. Sci., Hokkaido Univ. 1:59–90.<br />

Setchell, W. A. & Gardner, N. L. 1924. New marine algae from the<br />

Gulf of California. Proc. Calif. Acad. Sci. 12:659–949.<br />

Shimabukuro-Dias, C. K., Oliviera, C., Reis, R. E. & Foresti, F.<br />

2004. Molecular phylogeny of the armored catfish family Callichthyidae<br />

(Ostariophysi, Siluriformes). Mol. Phylogenet. Evol.<br />

32:152–63.<br />

Shimodaira, H. & Hasegawa, M. 1999. Multiple comparisons of<br />

log-likelihoods with applications to phylogenetic inference.<br />

Mol. Biol. Evol. 16:1114–6.<br />

Siemer, B. L., Stam, W. T., Olsen, J. L. & Pedersen, P. M. 1998.<br />

Phylogenetic relationships of the brown algal orders Ectocarpales,<br />

Chordariales, Dictyosiphonales, and Tilopteridales<br />

(Phaeophyceae) based on RUBISCO large subunit and spacer<br />

sequences. J. Phycol. 34:1038–48.<br />

Spurr, A. R. 1969. A low-viscosity epoxy resin-embedding medium<br />

for electron microscopy. J. Ultrastr. Res. 26:31–43.<br />

Swofford, D. L. 2002. PAUP*. Phylogenetic Analysis Using Parsimony (*<br />

And Other Methods). Version 4.0b8. Sinauer, Sunderland, MA.<br />

Tan, I. H. & Druehl, L. D. 1994. A molecular analysis of Analipus<br />

and Ralfsia (Phaeophyceae) suggests the order Ectocarpales is<br />

polyphyletic. J. Phycol. 30:721–9.<br />

Tseng, C. K. 1983. Common Seaweeds of China. Science Press,<br />

Beijing, 316 pp.<br />

van den Hoek, C., Mann, D. G. & Jahns, H. M. 1995. Algae,<br />

An Introduction to Phycology. Cambridge University Press,<br />

Cambridge, 627 pp.<br />

Winhauer, T., Jager, S., Valentin, K. & Zetsche, K. 1991. Structural<br />

similarities between psbA gene from red and brown algae. Curr.<br />

Genet. 20:177–80.<br />

Womersley, H. B. S. 1987. The Marine Benthic Flora of Southern Australia.<br />

Part II. South Australia Government, Adelaide, 484 pp.<br />

Yang, E. C. & Boo, S. M. 2004. Evidence for two independent<br />

lineages of Griffithsia (Ceramiaceae, Rhodophyta) based on<br />

plastid protein-coding psaA, psbA, and rbcL gene sequences.<br />

Mol. Phylogen. Evol. 31:680–8.<br />

Yendo, K. 1907. The Fucaceae of Japan. J. Coll. Sci. Tokyo Univ.<br />

21:1–174.<br />

Yoon, H. S. & Boo, S. M. 1999. Phylogeny of Alariaceae (Phaeophyta)<br />

with the special reference to Undaria based on sequences<br />

of the RuBisCo spacer region. Hydrobiologia 398/399:47–55.<br />

Yoon, H. S., Hackett, J. D. & Bhattacharya, D. 2002a. A single<br />

origin of the peridinin- and fucoxanthin-containing plastids<br />

in dinoflagellates through tertiary endosymbiosis. Proc. Natl.<br />

Acad. Sci. USA 99:11724–9.<br />

Yoon, H. S., Hackett, J. D., Pinto, G. & Bhattacharya, D. 2002b. A<br />

single, ancient origin of chromist plastids. Proc. Natl. Acad. Sci.<br />

USA 99:15507–12.<br />

Yoshida, T. 1998. Marine Algae of Japan. Uchida Rokakuho, Tokyo,<br />

1222 pp.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!