ArticlePDF Available

Biology of Porphyra pulchella sp. nov. from Australia and New Zealand

Authors:

Abstract and Figures

Porphyra pulchella sp. nov. Ackland, West, Scott and Zuccarello was obtained at Mimosa Rock National Park, New South Wales; Westgate Bridge, Victoria, Australia; and Waihau Bay, North Island, New Zealand. It occurs mainly in mangrove habitats and is very small (± 1 mm) in field collections. In laboratory culture at 21 ± 2°C tiny blades (0.5- 3.0 mm) reproduced exclusively by archeospores liberated from vegetative cells of the upper sector of the blades. The archeospores displayed amoeboid and gliding motility once discharged. At 14 ± 2°C the blades grew to 25 mm and produced longitudinal spermatangial streaks mixed with 'phyllosporangial' streaks. The discharged 'phyl- lospores' showed amoeboid motility and germinated forming asexual blades. A conchocelis phase with typical ban- giophycidean pit connections was observed in blade cultures after 8-10 weeks at 14 ± 2°C. Conchocelis filaments produced conchosporangia and these released amoeboid conchospores that developed into archeosporangiate blades. Molecular data indicate that all 3 isolates are genetically identical.
Content may be subject to copyright.
INTRODUCTION
Porphyra is a red algal genus that comprises over 130
species, many of which form a major component of rocky
intertidal habitats (Yoshida 1997). Representatives of this
genus are cosmopolitan, spanning temperate and cold
water systems, with particularly high abundance record-
ed in the North Pacific Ocean (Mumford and Cole 1977).
Whilst most Porphyra species are thought to be regionally
confined, distribution of these species has not been thor-
oughly established (e.g. Nelson et al. 2001, 2003; Broom et
al. 2004).
The commercial importance of Porphyra has generated
much interest in its life history and systematics. Research
has focused on the effects of environmental factors such
as photoperiod, light intensity, temperature and nutri-
ents, on the morphological and reproductive develop-
ment of the species in culture (e.g. Frazer and Brown
1995; Nelson and Knight 1996; Kim and Notoya 1997;
Knight and Nelson 1999; Ruangchuay and Notoya 2003;
Niwa et al. 2004). The establishment of optimal develop-
mental conditions for Porphyra species has not only
expanded our knowledge on how to enhance cultivar
yields, but has provided a means of assessing the
impacts of harvesting wild Porphyra populations (Schiel
and Nelson 1990). Furthermore, these studies have
improved our understanding of the diversity within this
genus.
Species of Porphyra are typically recognized by a
biphasic life cycle in which there is an alternation
between a macroscopic, gametophytic leafy blade phase
and a microscopic, sporophytic filamentous phase called
the ‘conchocelis’ (Drew 1949; Kurogi 1972; Nelson et al.
2000). It is, however, well established that many varia-
tions in the life history of Porphyra species exist. Conway
and Wylie (1972) described P. subtumens (now
Pyrophyllon subtumens (J. Agardh ex R.M. Laing) W.A.
Nelson) as an asexual species, lacking the alternation of
the diploid conchocelis phase. Similarly, populations of
P. sanjuanensis V. Krishnamurthy lack the ability to sexu-
ally reproduce (Lindstrom and Cole 1990).
Algae
Volume 21(2): 1-10, 2006
Biology of Porphyra pulchella sp. nov. from
Australia and New Zealand
Jillian C. Ackland1, John A. West1, Joseph Scott2,
Giuseppe C. Zuccarello3and Judy Broom4
1School of Botany, University of Melbourne, Parkville, Victoria 3010, Australia
2Department of Biology, PO Box 8795, College of William and Mary, Williamsburg, VA 23187 USA
3School of Biological Sciences, Victoria University of Wellington, PO Box 600, Wellington, 6001 New Zealand
4Department of Biochemistry, University of Otago, PO Box 56, Dunedin, New Zealand
Porphyra pulchella sp. nov. Ackland, West, Scott and Zuccarello was obtained at Mimosa Rock National Park, New
South Wales; Westgate Bridge, Victoria, Australia; and Waihau Bay, North Island, New Zealand. It occurs mainly in
mangrove habitats and is very small (± 1 mm) in field collections. In laboratory culture at 21 ± 2°C tiny blades (0.5-
3.0 mm) reproduced exclusively by archeospores liberated from vegetative cells of the upper sector of the blades.
The archeospores displayed amoeboid and gliding motility once discharged. At 14 ± 2°C the blades grew to 25 mm
and produced longitudinal spermatangial streaks mixed with ‘phyllosporangial’ streaks. The discharged ‘phyl-
lospores’ showed amoeboid motility and germinated forming asexual blades. A conchocelis phase with typical ban-
giophycidean pit connections was observed in blade cultures after 8-10 weeks at 14 ± 2°C. Conchocelis filaments
produced conchosporangia and these released amoeboid conchospores that developed into archeosporangiate
blades. Molecular data indicate that all 3 isolates are genetically identical.
Key Words: Australia, molecular phylogeny, New Zealand, Porphyra pulchella sp. nov. SSU rDNA, TEM
*Corresponding author (jwest@unimelb.edu.au)
Figure 23만컬
Despite the current knowledge on Porphyra life histo-
ries, the taxonomy of this genus remains problematic.
Porphyra species are morphologically simple and subse-
quently have few morphological characters for species
identification (Kunimoto et al. 1999; Broom et al. 2002).
Moreover, morphological plasticity caused by environ-
mental change can hamper the identification of species
limits (Conway and Wylie 1972; Hannach and Waaland
1989). Species discrimination based on morphological
analysis is a traditional technique now thought to be
unreliable when used independently (Kunimoto et al.
1999; Nelson et al. 2003). Molecular techniques are accu-
rate indicators of phylogenetic diversity now commonly
used in combination with morphological data to identify
taxa (Broom et al. 1999; Broom et al. 2002; Kunimoto et al.
1999; Nelson et al. 2001; Nelson et al. 2003). These studies
have revealed that similarities in morphology do not nec-
essarily infer close relationships, yet morphological vari-
ation is not always an indicator of phylogenetic diver-
gence.
Species identifications based on nuclear SSU (18S)
rDNA have unveiled a greater diversity of Porphyra than
was previously recognised using morphological criteria
alone (Broom et al. 1999; Nelson et al. 1998, 2001, 2003;
Oliveira and Ragan 1994; Oliveira et al. 1995). The SSU
rDNA locus is a slowly evolving gene (Hillis and Dixon
1991) that has been used to infer systematic relationships
of red algal specimens at an ordinal and familial level
(Saunders and Kraft 1993; Bailey and Freshwater 1997;
Ragan et al. 1994). Whilst it is too conservative to be used
to distinguish organisms at an interspecies level, the final
third (3( end) of the SSU rDNA locus in Porphyra com-
prises variable domains which differ amongst species.
These regions have proved useful for the discrimination
of Porphyra entities at a species level (Broom et al. 1999;
Kunimoto et al. 1999). Moreover, small subunit rDNA
sequencing analyses have been used to reassess the taxo-
nomic status of Porphyra species previously identified on
the basis of morphological characters. Nelson et al. (2003)
used this technique to establish that Pyrophyllon subtu-
mens, P. cameronii (W.A. Nelson) W.A. Nelson and
Chlidophyllon kaspar (W.A. Nelson and N.M. Adams)
W.A. Nelson, originally placed in the genus Porphyra
(order Bangiales), are members of the order
Erythropeltidales. The results of this study signify that
the morphological and anatomical characters originally
used to identify these Porphyra species are a result of evo-
lutionary convergence.
In this paper, we have described a new species of
Porphyra (P. pulchella) based on morphology and life his-
tory in culture, TEM of the conchocelis phase and molec-
ular analysis of the nuclear SSU rDNA locus. Two
Australian populations of P. pulchella (isolates 3924 and
4422) were morphologically and genetically contrasted to
a morphotype from New Zealand and previously
described Porphyra entities, to confirm the taxonomic sta-
tus of this species. The Australian isolates are especially
distinct from most other Porphyra species because of the
very small blade size (1-2 mm) seen in field collections
and the unusual habitat, i.e., mangroves of temperate
coastlines.
MATERIALS AND METHODS
Field collections
Foliose thalli (1-2 mm) of Porphyra (holotype isolate
3923) were collected on 16th December, 1998 from pneu-
matophores of the mangrove Avicennia marina (Forsk.)
Vierh. at Nelson’s Lagoon, Mimosa Rock National Park,
New South Wales, Australia (36°41’ S, 149°59°E). A sec-
ond collection (paratype isolate 4422) was made 6th April
2004 from a band of Caloglossa vieillardii (Kützing)
Setchell on wood pilings at Westgate Bridge, Port Phillip
Bay, Victoria, Australia (37°49’S, 144°53’E). A morpho-
type also examined, was collected 1st August 2003 from
Waihau Bay east, North Island, New Zealand (37°37’S,
177°55’E). Terminology for reproductive structures and
life history phases of Porphyra is in accordance with
Nelson et al. (1999).
Culturing the blade phase
Thalli were given two mild osmotic shocks to kill vari-
ous colourless flagellates. Blades were placed in a 60 ×15
mm Petri dish containing deionised water for 10 seconds,
then transferred to Modified Provasoli’s Medium
[MPM/2 (West 2005), 10 ml of enrichment per litre of
sterilised 30 psu natural seawater] for one minute, before
receiving a second osmotic shock for ten seconds. Blades
were then incubated in a 60 ×15 mm Petri dish contain-
ing MPM/2 treated with 25
µ
g/ml antibacterial
ciprofloxacin hydrochloride (ciprofloxacin, Sigma
Chemical Co. Pty. Ltd, St Louis, MO) for 48 hours to
reduce bacterial contamination. Petri dishes were main-
tained in a controlled environment culture room at 21 ±
2°C, 12:12 LD photoperiod and 10-20
µ
mol photons m–2
s–1 of cool white fluorescent lighting. Clean blades were
transferred into PyrexTM (#3250, Corning Glass Works,
Corning, NY, USA) 500 ml storage dishes and 60 ×15
2 Algae Vol. 21(2), 2006
mm Petri dishes containing 250 ml and 10 ml of medium,
respectively, and maintained in the above conditions.
Archeospore discharge was induced by incubating
blades with archeosporangia in fresh medium.
Archeospores were pipetted into 500 ml storage dish-
es, 60 ×15 mm Petri dishes and 10 ml plastic six-welled
plates (Iwaki SciTech Div., Asahi Techno Glass,
Funabashi, Japan) containing MPM/2 medium. A cover-
slip (22 ×22 mm or 22 ×50 mm, #1) was placed in each
dish as a substratum for archeospore settlement.
Dishes were maintained in a controlled environment
culture room or controlled environment E-36L plant
growth chambers (Percival Scientific Inc., Perry, Iowa,
USA) set to the desired photoperiod and temperature.
Desired irradiance levels were obtained by covering fluo-
rescent lamps with dark plastic window screen (1mm
mesh) and using cardboard boxes to elevate dishes closer
to the light source. Dishes were placed on New
BrunswickTM Model G2 rotary shakers (New Brunswick
Scientific Co., New Brunswick, New Jersey, USA) set at
approximately 80 rpm.
Culturing the conchocelis phase
Conchocelis filament tufts were treated in the same
way as blades to eliminate colourless flagellates and
reduce bacterial contamination, and cultured in the same
light and temperature regime.
Sterilised mollusc shell pieces (area 1.0-1.5 cm2, thick-
ness 0.5-1.0 mm) were inoculated with five conchocelis
filament tufts (0.2-0.5 mm in diameter) and placed in a 60
×15 mm Petri dish containing MPM/2. Conchocelis fila-
ments were secured to shells using a coverslip (22 ×22
mm, #3) and 2mm glass rods. The coverslip and glass
rods were removed after conchocelis filaments attached
to the shell pieces.
All Petri dishes were sealed with parafilm to minimise
evaporation. The medium was renewed every 3-10 days.
Blades and conchocelis filaments were examined using a
Zeiss GFL bright field compound microscope and a Zeiss
dissecting microscope with a Zeiss LCD 1500 fiber optic
light source. Photomicrographs were taken with a Zeiss
MC 100 35 mm camera using Ektachrome 200 colour
film. The transparencies were scanned with an Epson
FilmScan 200 using Photoshop 5.0 software on a
Macintosh G4 computer.
Transmission electron microscopy (TEM)
Conchocelis filaments were removed from culture dish
bottoms and fixed for 90 minutes at ambient temperature
in 2.5% glutaraldehyde or 1.5% paraformaldehyde in a
0.1 M phosphate buffer solution (pH 6.8) with 0.25 M
sucrose. Following buffer rinses, samples were post-
fixed in the same buffer for 90 min in 1% OsO4at ambi-
ent temperature, rinsed thoroughly in distilled H2O, left
in 50% acetone for 30 minutes and stored in a 70% ace-
tone-2% uranyl acetate solution at ambient temperature
for 4 hours. Samples were then dehydrated in a graded
acetone series, infiltrated and embedded in EmBed 812
(Electron Microscopy Sciences, P.O. Box 251, 321 Morris
Rd., Fort Washington, PA 19034) and polymerized at
70°C for 3 days. Thin sections were cut with an RMC
MT6000-XL ultramicrotome, stained with lead acetate
and viewed with a Zeiss EM 109 electron microscope.
Molecular phylogeny
DNA extraction followed a modified Chelex extraction
method (Zuccarello et al. 1999). Amplification of an
approximately 900 bp region of the nuclear-encoded
small subunit of ribosomal RNA (SSU), corresponding to
the final third of the molecule (Saunders and Kraft
1994;Broom et al. 1999) followed the procedure in Broom
et al. (1999). All PCR products were electrophoresed in 1-
2% agarose to check product size and sequenced follow-
ing procedures in Zuccarello et al. (1999).
Sequences were assembled using the computer soft-
ware supplied with the ABI sequencer, and aligned with
Clustal X (Thompson et al. 1997). All sequences were
compiled in Se-Al version2a11 (Rambaut 1996).
Phylogenetic relationships were inferred with PAUP*
4.0b10 (Swofford 2002). Outgroups and related sequences
used were selected from GenBank deposits and the
accession numbers are indicated in Fig. 26. Outgroups
used were Erythrocladia sp., Erythrotrichia carnea
(Dillwyn) J. Agardh and Smithora naiadum (C.L.
Anderson) G.J. Hollenberg.
Maximum-parsimony trees (MP) were constructed in
PAUP*, using the heuristic search option, 500 random
sequence additions, TBR branch swapping, unordered
and unweighted characters, gaps treated as missing data.
The program Modeltest version 3.6 (Posada and
Crandall, 1998) was used to find the model of sequence
evolution that best fits each data set by a hierarchical
likelihood ratio test (
α
= 0.01) (Posada and Crandall,
2001). When the best sequence evolution model had been
determined, maximum-likelihood was performed in
PAUP* using the estimated parameters (substitution
model, gamma distribution, proportion of invariable
sites) (5 random additions).
Ackland et al.: Biology of Porphyra pulchella sp. nov. 3
Support for individual internal branches was deter-
mined by bootstrap analysis (Felsenstein 1985), as imple-
mented in PAUP*. For MP bootstrap analysis, 1000 boot-
strap data sets were generated from resampled data (5
random sequence additions), for ML bootstrap analysis
100 bootstrap data sets were generated (1 random
sequence addition).
RESULTS
All observations were made on specimens (isolates
3923 and 4422) grown in unialgal culture.
Porphyra pulchella J.C. Ackland, J.A. West, J. Scott and
G.C. Zuccarello sp. nov.
Description: Laminae roseo-rubrae, stipiti carentes,
rotundae vel ovatae (0.5-3 mm ad 21 ± 2°C) aut lineari-
lanceolatae (10-25 mm ad 14 ± 2°C), in maturitate mar-
ginibus modice undulatis praebentibus thallis aspectum
leviter undulatum; laminae monostromaticae 19-21
µ
m
crassae. Divisiones cellularum diffusae marginales.
Cellulae marginales uni- vel biseriatae, quadratae (20-24
µ
m) vel elongatae, 20-24
µ
m latae, 40-48
µ
m longae; cel-
lulae centrales vegetativae demum irregulariter disposi-
tae, polygonales (19-24
µ
m); cellulae basales irregulariter
dispositae, polygonales et demum elongatae, cellulis cen-
tralibus vegetativis plerumque majores, chloroplastis sin-
gulis stellatis pyrenoide unica. Archaeosporangia secus
marginem superiorem laminae prodientia, regulariter
disposita, aliquantum quadrata (19-22
µ
m). Archaeosporae
liberatae 10-14
µ
m in diametro, motum amoeboideum et
prolabentem exhibentes. Planta monoica; spermatangia
strias conspicuas elongatas pallidas secus laminam supe-
riorem se praebentia cum striis phyllosporangiorum
putatorum mixtas; 32 divisione [a/4, b/4, c/2] facta, 4-6
µ
m in diametro, dimissa immota. Phyllosporangia
obscure roseo-rubrae, cellulis spermatangialibus majora,
singula vel bina terna quaternave aggregata; phyllospo-
rae 14-16
µ
m in diametro, motum amoeboideum
exhibentes.
Conchocelis, praebens se caespitem densum e filamen-
tis tenuibus ramosis uniseriatis constantem, includere se
in conchas potest. Divisiones cellularum intercalares.
Cellulae breves vel elongatae, 9-12
µ
m latae 32-40
µ
m
longae, conjunctionibus primariis fovearum junctae;
chloroplasti singuli elongati torti parietales pyrenoide
unica vel pluribus. Conchosporangia obscure roseo-
rubra, filamenta uniseriata et fasciculos 2-12-cellulares
facta, a superficie visa isodiametra, brevia vel elongata,
10-27
µ
m lata 7-27
µ
m longa. Conchosporae liberatae 10-
14
µ
m in diametro, motum amoeboideum exhibentes.
Blades pinkish-red, lacking stipe, round to ovate (0.5-3
mm at 21 ± 2°C) or linear-lanceolate (10-25 mm at 14 ±
2°C). Margins of mature blades moderately undulate giv-
ing thalli a slightly ruffled appearance. Blades monostro-
matic, 19-21 (m thick. Cell divisions diffuse and margin-
al. Marginal cells in one or two rows, quadrate (20-24
µ
m) to elongate, 20-24
µ
m wide and 40-48
µ
m long; cen-
tral vegetative cells becoming irregularly arranged,
polygonal (19-24
µ
m); basal cells irregularly arranged,
polygonal and becoming elongate, mostly larger than
central vegetative cells, chloroplasts single and stellate
with one pyrenoid. Archeosporangia along the upper
blade margin, regularly arranged, relatively quadrate
(19-22
µ
m). Liberated archeospores 10-14
µ
m in diame-
ter, exhibiting amoeboid and gliding motility.
Monoecious; spermatangia conspicuous, elongate, pale
streaks along the upper blade, intermixed with streaks of
presumed phyllosporangia; 32 spermatangia formed by
division [a/4, b/4, c/2], 4-6
µ
m in diameter, immotile
when discharged. Phyllosporangia dark pinkish-red,
larger than spermatangial cells, single or divided into
groups of 2-4; phyllospores 14-16
µ
m in diameter and
exhibiting amoeboid motion. Conchocelis pinkish-red
dense tuft of fine, branched, uniseriate filaments with the
ability to embed in mollusc shells. Cell divisions inter-
calary. Cells short to elongate, 9-12
µ
m wide and 32-40
µ
m long, connected by primary pit connections; chloro-
plast single, elongate, twisted and parietal with one or
more pyrenoids. Conchosporangia dark pinkish-red,
formed as uniseriate filaments and in clusters of 2-12
cells, isodiametric in surface view, short to elongate, 10-
27
µ
m wide and 7-27
µ
m long. Liberated conchospores
10-14
µ
m in diameter, exhibiting amoeboid motion.
Holotype specimen: Blades (NSW 722255), conchocelis
(NSW722445), Royal Botanic Gardens, Mrs. Macquaries
Road, Sydney, NSW 2000, Australia. Collection data: 16
xii 1998, on pneumatophores of the mangrove Avicennia
marina at Nelson’s Lagoon, Mimosa Rock National Park,
New South Wales (36°41’ S, 149°59’E).
Holotype culture: CCAP 1379/3. Culture Collection of
Algae and Protozoa, SAMS Research Services Ltd.,
Dunstaffnage Marine Laboratory, Dunbeg, Argyll, PA37
1QA, UK. Blade and conchocelis phases of isolate 3923
are included.
Paratype: NSW 722446, Royal Botanic Gardens, Mrs.
Macquaries Road, Sydney, NSW 2000, Australia.
Collection data: 6 iv 2004, in a band of Caloglossa vieil-
lardii (Kützing) Setchell on wood pilings at Westgate
4 Algae Vol. 21(2), 2006
Bridge, Port Phillip Bay, Victoria (37°49’S, 144°53’E).
New Zealand material: Collection date. 1 viii 2003,
dense patches in sand on rocks in stream, Waihau Bay,
North Island, (37°37’S, 177°55’E), collected by Tracy Farr.
The collection number of the herbarium sheet and DNA
sample is ASD 154. It is deposited and registered at Te
Papa (with a WELT number).
Etymology: Latin pulchellus = small and beautiful
(Stern 1973).
Development of the gametophytic blade phase in cul-
ture
Young blades were round to ovate and matured into
ovate blades at 21 ± 2°C (Fig. 1) and to linear-lanceolate
blades at 14 ± 2°C (Fig. 2) that were moderately ruffled
at the margin. In culture, blades were pinkish-red in
colour. Thalli were monostromatic, each cell containing a
single, stellate chloroplast with one pyrenoid, a vacuole
and a nucleus. The small size of blades was a distin-
guishing feature of this species. They grew up to 3 mm in
diameter at 21 ± 2°C and up to 25 mm in length at 14 ±
2°C. Blades attached to a substratum by a mass of
branched rhizoids (Fig. 3). Rhizoids developed from
irregularly arranged, slightly elongated polygonal basal
cells at the base (Fig. 4). Vegetative cells were smaller
than basal cells and mostly polygonal and irregularly
arranged, becoming quadrate to rectangular and regular-
ly aligned along the thallus margin (Fig. 5). In mature
thalli, marginal vegetative cells were commonly grouped
in pairs.
The development of archeosporangia and subsequent
mass release of archeospores readily occurred in blades
maintained at 21 ± 2°C. Thalli ranging from 1-3 mm in
length and as young as 16 days developed archeosporan-
gia from up to 14 rows of vegetative cells located at the
blade apex. During sporangium development, the vac-
uoles of differentiating cells became reduced in size.
Archeosporangia thus appeared a darker pinkish-red
than vegetative cells (Fig. 6). Prior to discharge, the cell
wall enclosing each archeosporangium disintegrated.
Archeospores remained surrounded only by a plasma
membrane and were loosely positioned in the extracellu-
lar polysaccharide matrix. The entire mass release of
archeospores occurred in approximately 25 minutes,
leaving behind the colourless cell wall with depressions
where the archeospores were originally positioned (Fig.
7). Extracellular matrix was extruded from the thallus
during spore release. Discharged archeospores displayed
a pulsing amoeboid motility, whereby fat pseudopodia
randomly extended and retracted from the cell (Fig. 8).
Amoeboid archeospores travelled at < 8
µ
m min–1 and
remained dynamic for up to 24 hours. Mucilage was
secreted from all amoeboid cells (Fig. 8). Occasionally
archeospores exhibited gliding motility. These cells
remained in contact with the slide as they translocated at
a relatively constant velocity of 0.27-0.4
µ
m s–1 (Fig. 9).
Archeospores settled and developed a single digitate rhi-
zoid within 24 hours of release. All germinating spores
developed into asexual blades. Development of
archeosporangia and subsequent mass release of
archeospores was induced by transferring blades from 14
± 2°C to 21 ± 2°C, 16L:8D or 12L:12D hours and 40
µ
mol
m–2 s–1, for 48-72 hours.
Male sexual reproductive structures developed in 10-
20 mm long monoecious thalli, after 5-6 weeks at 14 ±
2°C. Spermatangial sori were conspicuous, as pale, gold-
en streaks 3-7 cells across, up to 3 mm long at the blade
apex (Fig 10). Spermatangial packets were 16-20
µ
m wide
and 22-28
µ
m long, and were divided into 32 cells (4 x 4 x
2). These packets were intermixed with larger, dark pink-
red phyllosporangia. Phyllosporangia were commonly
undivided but did occur in packets of two and four, 16-
22
µ
m wide and 20-28
µ
m long (Figs 10 and 11).
Spermatia (4-6
µ
m in diameter) and phyllospores (14-16
µ
m in diameter) were released simultaneously following
the degeneration of the surrounding cell walls (Fig. 12).
Discharged spermatia were slightly deformed in shape
but remained immotile.
Phyllospores were amoeboid following their release.
All phyllospore germlings developed into blades (Fig
13). Conchocelis filaments occasionally developed in cul-
tures of blades bearing spermatangia and phyllosporan-
gia, after 8-10 weeks at 14 ± 2°C. Whilst this indicates the
presence of female reproductive structures (carpogonia),
we did not observe any carpogonia or zygotosporangia
between phyllosporangia and spermatangia.
Development and TEM observations of the conchocelis
phase in culture
The conchocelis comprised a densely branched net-
work of narrow, uniseriate filaments in tight, spherical
tufts that were free floating or anchored to a substrate.
Vegetative filament cells were 9-12
µ
m in diameter and
32-40
µ
m in length (Fig. 14). The single chloroplast in
each cell was rose to purple coloured, elongate, twisted
and parietal with one or more pyrenoids.
Conchosporangia were variable in form. Entire vegeta-
tive filaments commonly differentiated into conchospo-
Ackland et al.: Biology of Porphyra pulchella sp. nov. 5
6 Algae Vol. 21(2), 2006
Fig. 1. Small 2 mm orbiculate blade formed at 14°C. Scale bar = 0.2 mm.
Fig. 2. Lanceolate 20 mm blade with spermatangia and phyllosporangia in linear streaks at the apical end. Scale bar = 5 mm.
Fig. 3. Basal attachment with multiple rhizoids. Scale bar = 20
µ
m.
Fig. 4. Rhizoids (arrowheads) developing from basal ends of cells near the blade base. Scale bar = 15
µ
m.
Fig. 5. Elongate meristematic cells at blade margin and more quadrate meristematic cells away from the margin. Scale bar = 15
µ
m.
rangial branches (Fig. 14). These branches were 10-27
µ
m
in diameter with transverse to oblique cross walls.
Conchosporangia also occurred in large clusters of up to
12 sporangia, irregularly distributed amongst vegetative
filaments (Fig. 14, 15). The cells were usually shorter (7-
27
µ
m) than those of vegetative cells (32-40
µ
m),
appeared a darker pinkish-red and contained reduced
vacuoles and a single chloroplast that occupied a large
portion of the cell. Prior to conchospore discharge, the
cell wall at the apex of conchosporangial filament broke
down. Conchospores appeared slightly elongated as they
successively passed through the filament before dis-
charging at the filament apex. Discharged conchospores
were pinkish-red, approximately 10-14 um in diameter
and displayed an amoeboid motion similar to that of
archeospores and phyllospores. All germinating con-
chospores developed into blades. The development of
these blades was often aberrant. Many conchospore
germlings underwent irregular cell divisions and exhibit-
ed a high cell death rate. Blades consequently developed
into irregular shapes (Fig. 16). Some thalli produced
patches of vegetative cells that were lighter in colour
Ackland et al.: Biology of Porphyra pulchella sp. nov. 7
Fig. 6. Transition zone of larger vacuolate vegetative cells below and somewhat smaller developing archeosporangial cells with
reduced vacuoles above. Scale bar = 10
µ
m.
Fig. 7. Discharged archeospores and empty sporangial walls visible. Scale bar = 30
µ
m.
Fig. 8. Freshly discharged archeospores have short fat pseudopodia (arrow) that move continuously in all directions. Faint strands of
mucilage (arrowhead) are visible. Scale bar = 8
µ
m.
Fig. 9. Archeospores in contact with a slide displaying smooth gliding movement through the medium. Scale bar = 10
µ
m.
8 Algae Vol. 21(2), 2006
Fig. 10. Linear streaks of alternating spermatangia (arrowhead) and phyllosporangia (arrow) in blade grown at 14°C. 16 spermatan-
gia are formed in two layers (32 total) from each vegetative cell. Single or double phyllospores are formed from each vegetative
cell. Scale bar = 30
µ
m.
Fig. 11. Spermatangial packet (bracket) with 16 spermatia. Phyllosporangial packets of 2 (single arrow, arrowhead) or 4 (double
arrow). Scale bar = 20
µ
m.
Fig. 12. Discharged spermatia (arrowhead) and phyllospores (arrow). Spermatia are non-motile and phyllospores are amoeboid.
Scale bar = 30
µ
m.
Fig. 13. Phyllospores still in the old blade matrix and showing the bipolar germination (arrowheads) forming new blades (not con-
chocelis). Scale bar = 12
µ
m.
compared to other vegetative cells of the same blade.
Despite this unusual development, mature blades repro-
duced asexually by archeospores when grown at 20-
22°C.
As is typical of Porphyra species, the conchocelis fila-
ments had the ability to burrow into calcium carbonate
shell matrices. Eleven days after placing the conchocelis
filaments on mollusk shell fragments, filaments had
attached and begun penetration. By six weeks, vegetative
filaments appeared on the same plane as calcium carbon-
ate crystals of the shell matrix, indicating that the fila-
ments had burrowed into the shell. After three months,
Ackland et al.: Biology of Porphyra pulchella sp. nov. 9
Fig. 14. Conchocelis filaments growing free in culture. Vegetative filaments (arrowhead), intercalary conchosporangial filaments
(arrow) and conchosporangial clusters (bracket) are seen. Scale bar = 30
µ
m.
Fig. 15. A cluster of conchosporangial cells. Scale bar = 15
µ
m.
Fig. 16. Conchosporeling blade with two colour and cell patterns in upper and lower sectors. Scale bar = 35
µ
m.
Fig. 17. Conchocelis filaments growing in a mollusk shell matrix. Scale bar = 50
µ
m.
an intricate network of filaments bearing conchosporan-
gia had developed throughout the shell matrices (Fig.
17). Some conchosporangial filaments emerged from the
shell surface and eventually liberated conchospores that
germinated, forming blades. Conchocelis filaments also
reproduced asexually by fragmentation.
TEM preparation often resulted in the disruption of
conchocelis filament cell walls. Cell wall breakage per-
mitted a greater influx of fixatives and embedding resins.
The cellular content of disrupted cells was thus better
preserved than that of cells with intact walls. TEM of
conchocelis vegetative filament cells revealed that the
cells were connected by a primary pit connection (Figs.
18 and 19). The nucleus of each cell contained a promi-
nent nucleolus and condensed chromatin (Fig. 20).
Although mitochondria and Golgi bodies were scarce the
cis-region of each Golgi complex was always associated
with a mitochondrion (Fig. 21). Floridean starch granules
were relatively small and abundant in the cytoplasm of
each cell (Fig. 21). The chloroplast thylakoids were typi-
cally unstacked and covered by disk-shaped phycobili-
somes (Fig. 22). The pyrenoids had inconspicuous, phy-
cobilisome-free thylakoids winding through their matrix
and small plastoglobuli frequently occurred at the
periphery of pyrenoids and within several regions of the
chloroplast stroma (Fig. 22).
Figure 23 represents a diagram summarising the life
cycle of Porphyra pulchella in culture as we currently
understand it.
Phylogenetic analysis
The partial sequence of the 18S gene consisted of 840
characters, 165 of which were parsimony-informative.
The evolutionary model least-rejected by the hierarchical
likelihood test was: TrN (Tamura and Nei 1993) (substi-
tution rate matrix: a = 1.0, b = 1.65, c = 1.0, d = 1.0, e =
5.07, f = 1.00) plus proportion of invariable sites set to
0.599 and a gamma distribution of 0.432. The MP analy-
sis produced 60 MP trees of 453 steps. Maximum
Likelihood (ML) analysis produced one tree of -ln L score
of 3502.2473. Both tree topologies were similar and the
MP strict consensus topology is shown in Fig. 24.
Sequences of ribosomal DNA PCR amplification prod-
ucts from the Australian and New Zealand Porphyra
samples were compared. Sequences from the blades and
‘conchocelis’ filaments of isolate 3923 were identical,
confirming that both phases were of the same species.
Blades of isolates 3923 and 4422 shared identical
sequences. These samples form a very supported clade
with a sample from Waihau Bay, North Island, New
Zealand. The tree topology is congruent with other stud-
ies showing a polyphyletic Porphyra (Broom et al. 2004;
Nelson et al. 2005). No sequences in GenBank had
sequences identical to P. pulchella. Porphyra pulchella
forms a sister group relationship with P. pseudolinealis,
although this relationship is only poorly supported (<
70%).
DISCUSSION
The conchocelis phase in the one culture isolate (3923)
of Porphyra pulchella occurred without any evidence of
sexual reproduction in the blades grown for many
months at 20 ± 2°C. Molecular data (Fig. 25) clearly veri-
fies its genetic link with the blade phase. Conchospores
developed frequently and spore germination produced
blades that reproduced by archeospores. The early stages
of conchospore germination often showed a pattern of
cell shape and size aberrant in the lower and upper sec-
tors of the blade. We have no evidence of what is occur-
ring but the pattern suggests that meiotic segregation
occurred in a manner similar to that in P. yezoensis Ueda
(Miura and Ohme-Takagi 1994).
Porphyra pulchella displayed a striking difference in
blade size and reproduction at different temperatures. At
the higher temperature (20 ± 2°C) in two weeks, small
blades (0.5-3 mm) developed that reproduced with asex-
ual archeospores. At 14 ± 2°C the blades take 5 weeks to
mature reaching 20-25 mm before reproducing with
spermatangia and phyllosporangia. This indicates that P.
pulchella may be adapted to warmer coastal waters that
allow more rapid reproduction and spore dissemination.
Kim and Notoya (2003) investigated the life history of
Porphyra koreana M.S. Hwang and I.K. Lee, observing that
archeospore formation began at 2 weeks when blades
were 1.5 mm long in 15-25°C in long days (14:10 LD) and
in short days (10:14LD) archeospore formation began at 4
weeks in 20-25°C and at 5 weeks in 15°C. At 20°C sper-
matia and zygotospores began release at 7-8 weeks in
long days. In short days at 15-25°C no sexual structures
developed for the entire culture period of 18 weeks.
Blades reached only 4 mm at 20-25°C but grew to 60 mm
in 5-10°C. Archeospores were formed in all conditions.
In contrast to P. koreana, at higher temperatures only
archeospores formed in P. pulchella and archeospores,
spermatia and phyllospores were present in lower tem-
peratures.
Notoya et al. (1993) investigated the life histories of
10 Algae Vol. 21(2), 2006
Ackland et al.: Biology of Porphyra pulchella sp. nov. 11
Figs 18-22. TEM of conchocelis phase. Fig. 18. Low magnification TEM of several cells connected by pit plugs (arrowheads). Scale bar
= 4.5
µ
m. Fig. 19. High magnification TEM of pit plug between two vegetative filament cells. Although difficult to see in this
image, pit plugs of Porphyra and Bangia possess a singe, thin cap but no cap membrane. Scale bar = 0.25
µ
m. Fig. 20. Nucleus
with prominent nucleolus (large asterisk) and electron-dense chromatin (arrowheads). Mitochondria (small asterisks). Scale bar
= 0.11
µ
m. Fig. 21. Mitochondrion (M) is associated with the cis-region of the Golgi complex (arrowhead) in conchocelis cells.
Abundant starch granules (S) can be seen in the cytoplasm of cells. Scale bar = 0.37
µ
m. Fig. 22. Unstacked chloroplast thylakoids
covered by disc-shaped phycobilisomes (arrowheads). The phycobilisome-free pyrenoid thylakoids and plastoglobuli (arrows)
are more readily apparent in this micrograph. Scale bar = 1.32
µ
m.
Porphyra lacerata A. Miura and P. suborbiculata Kjellman.
Archeospores formed within 1-3 weeks at higher temper-
atures (25°C) and blades remained small without sexual
reproduction whereas at lower temperature (15-20°C)
spermatia and zygotospores developed on blades in 4-5
weeks. These patterns appear similar to that seen for P.
pulchella.
In the paper describing P. koreana as a new species
Hwang and Lee (1994) provide a table of general charac-
ters showing the basic differences among the species, P.
koreana, P. kinositae, P. tenera, P. lacerata and P. kuniedae.
The apparent lack of carpogonia in P. pulchella sug-
gests that other conditions may be needed to have fully
functional sexual blades. Certainly the spermatia and the
phyllosporangia look normal and are released normally.
We observed over 2000 phyllospore germlings that
12 Algae Vol. 21(2), 2006
Fig. 23. Life history diagram of P. pulchella.
Blade
Gametophyte
Apica l
Spor e line
Archeospore
discharge
Archeospore
Archeospore
germlin gs
Young blade
Spermatia & phyllospore
release
Fer tilisat ion
?
Zygotospor e
Con chocelis
Spor ophyte
Conchosporangia Conchospore
sper ma tia
phyllospore
?
Car pogonia
f
Spermatangia &
intercalary phyllosporangia
formed only blades (no conchocelis filaments). These
were clearly not obtained from sexual fertililization.
Phyllospores, archeospores and conchospores display
a distinct amoeboid movement. This has been reported
briefly in Porphyra dioica (Holmes and Brodie 2004) and
in Porphyra yezoensis (Ueki et al. 2005). Spore motility is
common to most red algae that have been investigated
with time lapse videomicroscopy (Pickett Heaps et al.
2001). Ackland et al. (2006) provides evidence that
pseudopodia involved in amoeboid movement are regu-
lated by the cytoskeletal components actin and myosin.
Other workers have seen amoeboid motility of mono-
spores in Colaconema caespitosum (J. Agardh) Jackelman,
Stegenga and J.J. Bolton [as Audouinella botryocarpa
(Harvey) Woelkerling)] (Guiry et al. 1987), monospores
of Erythrotrichia (Rosenvinge 1927), carpospores and
tetraspores of Liagora harveyana Zeh and Helminthora
stackhousei (Clemente) Cremades and Pérez-Cirera
Ackland et al.: Biology of Porphyra pulchella sp. nov. 13
Fig. 24. Strict consensus of Maximum Parsimony trees based on partial 18S data of various Porphyra samples. Genbank samples given
with their accession numbers. Branches showing bootstrap support 50% shown (before diagonal MP analysis; after diagonal
ML analysis). Erythropeltidales selected as the outgroup.
(Cunningham et al. 1993, Guiry 1990). Although it is a
relatively slow process, spore movement appears neces-
sary for substrate selection and attachment prior to ger-
mination. The mechanism of amoeboid movement in red
algae has not been fully investigated.
The complete lack of spermatium motility in Porphyra
pulchella corresponds to our observations on other red
algae such as Bostrychia and Murrayella (Pickett-Heaps et
al. 1998, McBride and West 1999b, Wilson et al. 2002,
2003).
Molecular phylogeny of Porphyra is being slowly
investigated to resolve the species complexes. Porphyra
suborbiculata, P. koreana, P. lacerata and several other
smaller species have morphological and reproductive
features similar to those of P. pulchella but insufficient
molecular data are available to clearly indicate their
overall relationships.
ACKNOWLEDGEMENTS
This research is partially supported by Australian
Research Council grants SG0935526 (1994), S19812824
(1998), S19917056 (1999-2001), S0005005 (2000), a grant
from the Australian Biological Resources Study (2002-
2005) as well as a grant from the Hermon Slade
Foundation (2005-2007) to JAW and GCZ.
REFERENCES
Ackland, J. C., West, J.A and Pickett-Heaps, J.D. 2006. Actin and
myosin regulate pseudopodia of Porphyra pulchella
(Rhodophyta) archeospores. J. Phycol. submitted.
Bailey C.J. and Freshwater W.D. 1997. Molecular systematics of
the Gelidiales: inferences from separate and combined
analyses of plastid rbcL and nuclear SSU gene sequences.
Eur. J. Phycol. 32: 343-52.
Broom, J.E., Farr, T.J. and Nelson, W.A. 2004. Phylogeny of the
Bangia flora of New Zealand suggests a southern origin for
Porphyra and Bangia (Bangiales, Rhodophyta) Mol. Phyl.
Evol. 31: 1197-207.
Broom, J.E., Jones, W.A., Hill, D.F., Knight, G.A. and Nelson,
W.A. 1999. Species recognition in New Zealand Porphyra
using 18S rDNA sequencing. J. Appl. Phycol. 11: 421-8.
Broom, J.E., Nelson, W.A., Yarish, C., Jones, W.A., Aguilar-
Rosas, R. and Aguilar-Rosas, L.E. 2002. A reassessment of
the taxonomic status of Porphyra suborbiculata, Porphyra car-
olinensis, and Porphyra lilliputiana (Bangiales, Rhodophyta)
based on molecular and morphological data. Eur. J. Phycol.
37: 227-35.
Conway, E. and Wylie, A.P. 1972. Spore organisation and repro-
ductive modes in two species of Porphyra from New
Zealand. Proc. 7thIntern. Seaweed Symp., 105-108
Cunningham, E., Guiry, M.D. and Breeman, A. 1993.
Environmental regulation of development, life history and
biogeography of Helminthora stackhousei (Rhodophyta) by
daylength and temperature. J. Exp. Mar. Biol. Ecol. 171: 1-21.
Drew, K.M. 1949. Conchocelis-phase in the life history of
Porphyra umbilicalis. Nature 164: 148.
Felsenstein, J. 1985. Confidence intervals on phylogenies: an
approach using the bootstrap. Evolution 39: 783-91.
Frazer, A.W.J. and Brown, M.T. 1995. Growth of the conchocelis
phase of Porphyra columbina (Bangiales, Rhodophyta) at dif-
ferent temperatures and levels of light, nitrogen and phos-
phorus. Phycol. Res. 45: 249-53.
Guiry, M.D. 1990. The life history of Liagora harveyana
(Nemaliales, Rhodophyta) from south-eastern Australia.
Br. Phycol. J. 25: 353-362.
Guiry, M.D., Kee, W.R. and Garbary, D. 1987. Morphology,
temperature and photoperiodic responses in Audouinella
botryocarpa (Harvey) Woelkerling (Acrochaetiaceae,
Rhodophyta) from Ireland. Giorn. Bot. Ital. 121: 229-46.
Hannach, G. and Waaland J.R. 1989. Growth and morphology
of young gametophytes of Porphyra abbottae (Rhodophyta):
Effects of environmental factors in culture. J. Phycol. 25:
247-54.
Hillis D.M. and Dixon M.T. 1991. Ribosomal DNA: Molecular
evolution and phylogenetic inference. Quart. Rev. Biol. 66:
411-53.
Holmes, M.J. and Brodie, J. 2005. Morphology, seasonal phenol-
ogy and observations on some aspects of the life history in
culture of Porphyra dioica (Bangiales, Rhodophyta) from
Devon, UK. Phycologia 43: 176-88.
Hwang, M.S. and Lee, I.K. 1994. Two species of Porphyra
(Bangiales, Rhodophyta), P. koreana sp. nov. and P. lacerata
Miura from Korea. Kor. J. Phycol. 9: 169-177.
Kim, N-G. and Notoya, M. 2003. Life history of Porphyra koreana
(Bangiales, Rhodophyta) from Korea in culture. Proc. 17th
Intern. Seaweed Symp., pp. 435-42.
Kim, N-G. and Notoya, M. 1997. Life history of Porphyra lacerata
(Bangiales, Rhodophyta) from Korea in culture. Jpn. J.
Phycol. 45: 82 (in Japanese).
Knight, G.A. and Nelson, W.A. 1999. An evaluation of charac-
ters obtained from life history studies for distinguishing
New Zealand Porphyra species. J. Appl. Phycol. 11: 411-9.
Kunimoto M., Kito H., Yamamoto Y., Cheney D.P., Kaminishi
Y. and Mizukami Y. 1999. Discrimination of Porphyra
species based on small subunit ribosomal RNA gene
sequence. J. Appl. Phycol. 11: 203-9.
Kurogi, M. 1972. Systematics of Porphyra in Japan. In Abbott,
I.A. and Kurogi, M. (Eds.) Contributions to the systematics of
benthic marine algae of the north Pacific. pp 167-92.
Lindstrom S.C. and Cole K.M. 1990. An evaluation of species
relationships in the Porphyra perforata complex (Bangiales,
Rhodophyta) using starch gel electrophoresis. Hydrobiol.
204/205: 179-83.
McBride, D.L. and West, J.A. 1999a. Patterns of tetraspore dis-
charge in Caloglossa and Murrayella (Ceramiales,
Rhodophyta). Phycol. Res. 47: 21-6.
14 Algae Vol. 21(2), 2006
McBride, D.L. and West, J.A. 1999b. Salinity reduction induces
spermatial release in some cultured red algae. Algae 14:
133-137.
Miura, A. and Ohme-Tagaki, M. 1994. Mendelian inheritance of
pigmentation mutant types in Porphyra yezoensis
(Bangiophyceae, Rhodophyta). Jpn. J. Phycol. 42: 83-101.
Mumford, T.F. and Cole, K.M.. 1977. Chromosome numbers for
fifteen species in the genus Porphyra (Bangiales,
Rhodophyta) from the west coast of North America. J.
Phycol. 16: 373-7.
Nelson, W.A., Brodie, J. and Guiry, M.D. 1999. Terminology
used to describe reproduction and life history stages in the
genus Porphyra (Bangiales, Rhodophyta). J. Appl. Phycol. 11:
407-10.
Nelson, W.A., Broom, J.E., Farr, T.J. 2003. Pyrophyllon and
Chlidophyllon (Erythropeltidales, Rhodophyta): Two new
genera for obligate epiphytic species previously placed in
Porphyra, and a discussion of the orders Erythropeltidales
and Bangiales. Phycologia. 42: 308-15.
Nelson, W.A. and Knight, G.A. 1996. Life history in culture of
the obligate epiphyte Porphyra subtumens (Bangiales,
Rhodophyta) endemic to New Zealand. Phycol. Res. 44: 19-
25.
Nelson, W.A., Knight, G.A. and Hawkes, M.W. 1998. Porphyra
lilliputiana sp. nov. (Bangiales, Rhodophyta): A diminutive
New Zealand endemic with novel reproductive biology.
Phycol. Res. 46: 57-61.
Nelson, W. A., Farr, T. J. and Broom, J. E. S. 2005. Dione and
Minerva, two new genera from New Zealand circumscribed
for basal taxa in the Bangiales (Rhodophyta). Phycologia. 44:
139-45.
Nelson W.A., Broom J.E. and Farr T.J. 2001. Four new species of
Porphyra (Bangiales, Rhodophyta) from New Zealand
region described using traditional characters and 18S
rDNA sequence data. Cryptog. Algol. 22: 263-84.
Notoya, M. 1999. ‘Seed’ production of Pophyra spp. by tissue
culture. J. Appl. Phycol. 11: 105-110.
Notoya, M., Kikuchi, N. Matsuo, M., Aruga, Y. and Miura, A.
1993. Culture studies of four species of Porphyra
(Rhodophyta) from Japan. Nippon Suisan Gakkaishi 59: 431-
6.
Oliveira M.C., Kurniawan J., Bird C.J., Rice E.L., Murphy C.A.,
Singh R.K., Gutell R.R. and Ragan M.A. 1995. A prelimi-
nary investigation of the order Bangiales (Bangiophycidae,
Rhodophyta) based on sequences of nuclear small-subunit
ribosomal RNA genes. Phycol. Res. 43: 71-9.
Oliveira M.C. and Ragan M.A. 1994. Variant forms of a group 1
intron in nuclear small-subunit rRNA genes of the marine
red alga Porphyra spiralis var. amplifolia. Mol. Biol. Evol. 11:
195-207.
Pickett-Heaps, J. and West, J.A. 1998. Time-lapse video observa-
tions on sexual plasmogamy in the red alga Bostrychia. Eur.
J. Phycol. 33: 43-56.
Pickett-Heaps, J.D., West, J.A., Wilson, S.M. and McBride, D.M.
2001. Time-lapse videomicroscopy of cell (spore) move-
ment in red algae. Eur. J. Phycol. 36: 9-22.
Posada, D. and Crandall, K.A. 1998. Modeltest: testing the
model of DNA substitution. Bioinformatics 14: 817-8.
Posada, D. and Crandall, K.A. 2001. Selecting the best-fit model
of nucleotide substitution. Syst. Biol. 50: 580-601.
Ragan M.A., Bird C.J., Rice E.L., Gutell R.R., Murphey C.A. and
Singh R.K. 1994. A molecular phylogeny of the marine red
algae (Rhodophyta) based on the nuclear small-subunit
rRNA gene. Plant Biol. 91: 7276-80.
Rambaut, A. 1996. Se-Al: Sequence Alignment Editor. Available
athttp://evolve.zoo.ox.ac.uk/.
Rosenvinge, L.K. 1927. On mobility in the reproductive cells of
the Rhodophyceae. Bot. Tidsk. 40: 72-80.
Ruangchuay, R. and Notoya, M. 2003. Physiological responses
of blade and conchocelis of Porphyra vietnamensis Tanaka et
Pham-Hoang Ho (Bangiales, Rhodophyta) from Thailand
in culture. Algae 18: 21-8.
Saunders, G.W. and Kraft, G.T. 1994. Small-subunit rRNA gene
sequences from representatives of selected families of the
Gigartinales and Rhodymeniales (Rhodophyta). 1.
Evidence for the Plocamiales ord. nov. Can. J. Bot. 72: 1250-
63.
Schiel, D.R. and Nelson, W.A. 1990. The harvesting of macroal-
gae in New Zealand. Hydrobiologia 204-205: 25-33.
Stern, W.T. 1973. Botanical Latin. David and Charles, London,
xiv + 566 pp.
Swofford, D.L. 2002. PAUP* Phylogenetic Analysis Using
Parsimony (* and other methods) Version 4. Sinauer
Associates, Sunderland, MA, USA.
Tamura, K. and Nei, M. 1993. Estimation of the number of
nucleotide substitutions in the control region of mitochon-
drial DNA in humans and chimpanzees. Mol. Biol. Evol. 10:
512-26.
Thompson, J.D., Gibson, T.J., Plewniak, F., Jeanmougin, F. and
Higgins, D.G. 1997. The Clustal X windows interface: flexi-
ble strategies for multiple sequence alignment aided by
quality analysis tools. Nucl. Acids Res. 24: 4876-4882.
Ueki, C., Nagasato, C., Motomura, T. and Saga, N. 2005.
Processes of germinations of various spores in the life his-
tory of Porphyra yezoensis. Jpn. J. Phycol. 53: 101. ( in
Japanese).
Yoshida T., Notoya M., Kikuchi N. and Miyata M. 1997.
Catalogue of species of Porphyra in the world, with special
reference to the type locality and bibliography. Chiba Nat.
Hist. Mus.Inst. Nat. Hist. Res., Special Issue. 3: 5-18.
West, J.A. 2005. Long Term Macroalgal Culture Maintenance, In
Andersen, R. (ed.), Algal Culturing Techniques. Elsevier,
New York, pp. 157-63.
Wilson, S.M., Pickett-Heaps, J.D. and West, J.A. 2002.
Fertilisation and the cytoskeleton in the red alga Bostrychia
moritziana (Rhodomelaceae, Rhodophyta). Eur. J. Phycol. 37:
509-22.
Wilson, S.M., West, J.A. and Pickett-Heaps, J.D. 2003. Time-
lapse videomicroscopy of fertilisation and the actin
cytoskeleton in Murrayella periclados (Rhodomelaceae,
Rhodophyta). Phycologia 42: 638-45.
Zuccarello, G.C., West, J.A., Kamiya, M. and King, R.J. 1999. A
Ackland et al.: Biology of Porphyra pulchella sp. nov. 15
rapid method to score plastid haplotypes in red seaweeds
and its use in determining parental inheritance of plastids
in the red alga Bostrychia (Ceramiales). Hydrobiologia. 401:
207-14.
Received 28 April 2006
Accepted 20 May 2006
16 Algae Vol. 21(2), 2006
... Meanwhile, this range of the Ambon Island habitat is significantly warmer than that for Porphyra sp. living in other temperate regions such as Australia and New Zealand (14 o C), Portugal (15 o C), British Isles (18 o C) and New England of the US (10-15 o C)(Ackland et al 2006;Kim et al 2007;Knoop et al 2020;Pereira et al 2004); and for the sub-polar region such as Canada (10-15 o C)(Blouin et al 2007) and Alaska (11 o C)(Lin 1999;Stekoll & Lin 1999). Regarding salinity linked to nutrient for supporting Porphyra sp., there is a key difference between the Ambon Island habitat and the other tropical habitats. ...
Article
Full-text available
This study provides evidence on the roles of Banda Sea upwelling in supporting favorable environmental conditions for Porphyra sp. in Ambon Island, eastern Indonesia. A cross-correlation analysis using historical sea surface temperature (SST) datasets between Banda Sea upwelling hotspot in eastern Banda Sea and waters nearby Ambon Island was employed to investigate the ocean teleconnection of Banda Sea upwelled water at Ambon Island. This ocean teleconnection was also investigated via weekly fieldwork samplings, measuring SST, sea surface salinity (SSS) and surface dissolved nutrients in the Porphyra habitats of Seri and Allang in Ambon Island during the peak of Banda Sea upwelling season (June-August 2023) to detect the presence of Banda Sea upwelled water in the habitats. The fieldwork also measured the growth of Porphyra sp. The principal component analysis (PCA) was further employed to identify the correlation between the environmental parameters and the growth of Porphyra sp. The cross-correlation analysis showed a very strong connection (r≥0.90) with very low time lag (<1 week) for the SST between Banda Sea upwelling hotspot in eastern Banda Sea and waters nearby Ambon Island. This indicated an immediate arrival of cool, salty, nutrient-rich upwelled Banda Sea water from the upwelling hotspot at Ambon Island during the peak upwelling season (JuneAugust). Weekly observations in the Porphyra habitats in Ambon Island during the peak Banda Sea upwelling season confirmed the arrival of the upwelled water, showing cool SST (reaching up to 26oC), high SSS similar to typical deep-layer Banda Sea water (≥34.5 psu) and the elevated concentration of surface dissolved nitrate (reaching as much as 0.1 ppm) that were evident in the locations, coincident with the occurrence of Porphyra sp. From PCA analyses, the growth of Porphyra sp. correlated with SST cooling and the increases in SSS and nutrients. This environmental condition characterizes typical oceanic conditions in Ambon Island during Banda Sea upwelling season. The knowledge of the ecology of Porphyra sp. in Ambon Island presented here has application for being a reference in implementing successful cultivating process of the Ambon Island Porphyra sp
... An understanding of the regulatory mechanisms involved in sexual reproduction is a prerequisite for efficient breeding to improve existing plants and create new varieties. However, although influences of light and temperature on the sexual maturation of gametophytes have been reported in Pyropia/Porphyra species (Ackland et al. 2006), the mechanisms remain to be elucidate. ...
Article
Full-text available
Plant growth regulators (PGRs) play a pivotal role in vascular plants, regulating growth, development, and stress responses; however, the role of PGRs in algae remains largely unexplored. Here, the role of ethylene, a simple plant growth regulator, was demonstrated in sexual reproduction of the marine red alga Pyropia yezoensis. Application of the ethylene precursor 1-aminocylopropane-1-carboxylic acid (ACC) promoted the formation of spermatia and zygotospores in the gametophytes as well as ethylene production, whereas the growth rate was repressed in comparison to gametophytes not treated with ACC. In addition, gametophytes treated with ACC and mature gametophytes showed enhanced tolerance to oxidative stress. Gene expression profiles revealed upregulation of genes involved in cell division and stress response in gametophytes treated with ACC and in mature gametophytes. These results indicate that ethylene plays an important role in the regulation of gamete formation and protection against stress-induced damage during the sexual reproductive stage. Considered together, these findings demonstrate that ethylene is involved in regulating the switching from a vegetative to a sexual reproductive phase in P. yezoensis.
... The ribosomal region (SSU) has been used to differentiate species in several studies (e.g. Broom et al. 1999; Ackland et al. 2006; Zuccarello et al. 2008). We believe that this molecular identity with these markers and the lack of robust characters to separate the two genera should lead to them being placed in taxonomic synonymy. ...
Article
Full-text available
Molecular analyses of the 18S ribosomal RNA gene and the psbA gene of Membranella nitens and Smithora naiadum from Pacific North America show that both taxa are genetically identical. The variability of the characters distinguishing these two genera has not been well documented. The generic name Smithora Hollenberg has priority over Membranella Hollenberg & Abbott therefore, M. nitens is placed in taxonomic synonymy as S. naiadum.
Article
Cyanobacteria and chloroplasts of algae and plants harbor specialized thylakoid membranes that convert sunlight into chemical energy. These membranes house photosystems II and I, the vital protein-pigment complexes that drive oxygenic photosynthesis. In the course of their evolution, thylakoid membranes have diversified in structure. However, the core machinery for photosynthetic electron transport remained largely unchanged, with adaptations occurring primarily in the light-harvesting antenna systems. Whereas thylakoid membranes in cyanobacteria are relatively simple they become more complex in algae and plants. The chloroplasts of vascular plants contain intricate networks of stacked grana and unstacked stroma thylakoids. This review provides an in-depth view of thylakoid membrane architectures in phototrophs, and the determinants that shape their forms, as well as presenting recent insights into the spatial organization of their biogenesis and maintenance. Its overall goal is to define the underlying principles that have guided the evolution of these bioenergetic membranes.
Article
The cosmopolitan red algal genus Pyropia sensu lato is the most speciose of the bladed Bangiales genera. In a major revision of the Bangiales, Pyropia was resurrected from Porphyra, although there was evidence at the time that species of Pyropia could be separated into several genera. Subsequent global phylogenetic analyses continued to resolve species assigned to Pyropia into several major clades with strong support, and the latest biogeographic analyses indicated that species distribution was also a pointer to the underlying phylogeny of Pyropia sensu lato. Therefore, in the present study, we have redefined the genus Pyropia, resurrected Porphyrella, and proposed four new genera: Calidia, Neoporphyra, Neopyropia and Uedaea. Based on a molecular phylogenetic study of the bladed Bangiales of China, a species which did not match any known taxa, was resolved in the new genus Calidia. The species, Calidia pseudolobata sp. nov., is described based on both morphological and molecular data. Molecular sequence data for rbcL, 18S and COI‐5P were amplified for fifteen samples in the present study. All the obtained rbcL sequences were identical to each other except for one (LYCN117) with one base pair difference. Two haplotypes of 18S (V9 region) were observed with one base pair difference (C/T30). All the obtained COI‐5P sequences were identical. Morphological comparisons were conducted not only with species in Calidia, but also with generically uncertain species currently assigned to Porphyra.
Article
Adenosine 5’-triphosphate (ATP) is a versatile extracellular signal along the tree of life, whereas cAMP plays a major role in vertebrates as an intracellular messenger for hormones, transmitters, tastants and odorants. Since red algal spore coalescence may somehow be considered analogous to the congregation process of social amoeba, which is stimulated by cAMP, we ascertained whether exogenous applications of ATP, cAMP, adenine, or adenosine modified spore survival and motility, spore settlement and coalescence. Concentration-response studies were performed with carpospores of Mazzaella laminarioides (Gigartinales), incubated with and without added purines. Stirring of algae blades released ADP/ATP to the cell media in a time-dependent manner. 10-300 μM ATP significantly increased spore survival; however, 1500 μM ATP, cAMP or adenine induced 100% mortality within less than 24 h; the exception was adenosine which up to 3000 μM did not alter spore survival. ATP exposure elicited spore movement with speeds of 2.2 to 2.5 μm. s−1. 14 d after 1000 μM ATP addition, spore abundance in the central zone of the plaques was increased 2.7 fold as compared with parallel controls. Likewise, 1-10 μM cAMP or 30-100 μM adenine also increased central zone spore abundance, albeit these purines were less efficacious than ATP; adenosine up to 3000 μM did not influence settlement. Moreover, 1000 μM ATP markedly accelerated coalescence, the other purines caused a variable effect. We conclude that exogenous cAMP, adenine, but particularly ATP, markedly influence red algae spore physiology; effects are compatible with the expression of one or more membrane purinoceptor(s), discarding adenosine receptor participation.This article is protected by copyright. All rights reserved.
Article
Thylakoids mediate photosynthetic electron transfer and represent one of the most elaborate energy-transducing membrane systems. Despite our detailed knowledge of its structure and function, much remains to be learned about how the machinery is put together. The concerted synthesis and assembly of lipids, proteins and low-molecular-weight cofactors like pigments and transition metal ions requires a high level of spatiotemporal coordination. While increasing numbers of assembly factors are being functionally characterized, the principles that govern how thylakoid membrane maturation is organized in space are just starting to emerge. In both cyanobacteria and chloroplasts, distinct production lines for the fabrication of photosynthetic complexes, in particular photosystem II, have been identified. This article is part of a Special Issue entitled: Chloroplast Biogenesis. Copyright © 2015. Published by Elsevier B.V.
Article
Full-text available
Low molecular weight carbohydrates, phycobilin pigments and cell structure using light and transmission electron microscopy were used to describe a new genus of unicellular red algae, Erythrolobus coxiae (Porphyridiales, Porphyrideophyceae, Rhodophyta). The nucleus of Erythrolobus is located at the cell periphery and the pyrenoid, enclosed by a cytoplasmic starch sheath, is in the cell center. The pyrenoid matrix contains branched tubular thylakoids and four or more chloroplast lobes extend from the pyrenoid along the cell periphery. A peripheral encircling thylakoid is absent. The Golgi apparatus faces outward at the cell periphery and is always associated with a mitochondrion. Porphyridium and Flintiella, the other members of the Porphyrideophyceae, also lack a peripheral encircling thylakoid and have an ER-mitochondria-Golgi association. The low molecular weight carbohydrates digeneaside and floridoside are present, unlike both Porphyridium and Flintiella, which have only floridoside. The phycobilin pigments B-phycoerythrin, R-phycocyanin and allophycocyanin are present, similar to Porphyridium purpureum. The cells have a slow gliding motility without changing shape and do not require substrate contact. The ultrastructural features are unique to members of the Porphyrideophyceae and recent molecular analyses clearly establish the validity of this new red algal class and the genus Erythrolobus.
Article
Full-text available
Pulvinus veneticus gen. et sp. nov. is a small (< 500 [Lm in diameter), cushion shaped epiphyte isolated into culture from Caloglossa vieillardii that was collected from Vanuatu. Molecular analysis places Pulvinus veneticus in the Compsopogonales. It has prostrate-adherent filaments or free filaments with one or more discoid to spiral bluish-green plastids without pyrenoids per cell. Pulvinus is euryhaline, growing and reproducing in salinities of 2 to 30 practical salinity units (psu). Monosporangia are formed successively by terminal branch cells or when a sector of a thallus gelatinises, releasing masses of spores. Monospores are round (6-10 mu m diameter) and glide at speeds up to 4.5 mu m s(-1). Some monospores have an extracellular polysaccharide tail and move more slowly, 0.25 to 1.0 mu m s-1. Spores avoid contact with other objects and do not require substrate contact during movement. The ultrastructure of Pulvinus is very similar to that of Compsopogon and Boldia in that the cis-region of Golgi bodies is not associated with a mitochondrion and plastids have a peripheral thylakoid. No pit connections occur between derivative cells.
Article
Full-text available
The red algal order Bangiales has been revised as a result of detailed regional studies and the development of expert local knowledge of Bangiales floras, followed by collaborative global analyses based on wide taxon sampling and molecular analyses. Combined analyses of the nuclear SSU rRNA gene and the plastid RUBISCO LSU (rbcL) gene for 157 Bangiales taxa have been conducted. Fifteen genera of Bangiales, seven filamentous and eight foliose, are recognized. This classification includes five newly described and two resurrected genera. This revision constitutes a major change in understanding relationships and evolution in this order. The genus Porphyra is now restricted to five described species and a number of undescribed species. Other foliose taxa previously placed in Porphyra are now recognized to belong to the genera Boreophyllum gen. nov., Clymene gen. nov., Fuscifolium gen. nov., Lysithea gen. nov., Miuraea gen. nov., Pyropia, and Wildemania. Four of the seven filamentous genera recognized in our analyses already have generic names (Bangia, Dione, Minerva, and Pseudobangia), and are all currently monotypic. The unnamed filamentous genera are clearly composed of multiple species, and few of these species have names. Further research is required: the genus to which the marine taxon Bangia fuscopurpurea belongs is not known, and there are also a large number of species previously described as Porphyra for which nuclear SSU ribosomal RNA (nrSSU) or rbcL sequence data should be obtained so that they can be assigned to the appropriate genus.
Article
In the red alga Bostrychia moritziana, release of spermatia is triggered by slight osmotic shock; they emerge under pressure apparently generated by swelling of the mucilaginous sheath. Spermatia adhere tenaciously to trichogynes of the carpogonium. Adhesion triggers spermatial mitosis, which is complete in about 30–45 min; there is no cytokinesis and the binucleate spermatium becomes vacuolated. The delicate, dynamic trichogyne cytoplasm contains complex membranous components and vacuoles. At the contact zone, the trichogyne and spermatial wall erode, forming a pore, and cytoplasmic continuity (plasmogamy) is achieved after about 50–70 min. Many trichogynes rupture during these events because of inadequate structural connection with the spermatia. Normally, both spermatial nuclei enter the trichogyne in sequence; rarely, both nuclei enter together. Entrance is rapid, and the nuclei often become thin and greatly elongated as each squeezes through the narrow pore into the trichogyne. Once inside, each nucleus resumes its normal shape as it starts to move steadily along the trichogyne, often with irregular pauses. One nucleus of each pair (not necessarily the first out of the spermatium) migrates along the trichogyne towards the carpogonium base to fertilize the female nucleus; the other moves in the opposite direction, away from the entry site, and it often ends up near the tip of the trichogyne. This same scenario was observed for each of several spermatia contributing nuclei to one trichogyne. Thus, our observations indicate that the two nuclei in each spermatium are differentiated so that only one is capable of fertilization, differentiation being visibly expressed in the direction that the nucleus moves as it enters the trichogyne.
Article
The red algae generally are considered to have no motile stages. We have confirmed sporadic reports of motile spores in a few red algae by recording freshly released live spores with time-lapse videomicroscopy. Of the 25+ species investigated, only a few had spores that appeared to be immobile. Cells of unicellular species often displayed active movement although were otherwise indistinguishable from nearby immobile cells, and these may be equivalent to the motile spores released after differentiation in more complex multicellular species. There was considerable variation among members of the Bangiophycidae. Cells of Porphyridium often moved at around 0·66 μm s¹ and had conspicuous mucilage tails. Flintiella and Rhodospora showed no cellular movement. Rhodosorus sometimes moved a little and all four strains displayed continuous cytoplasmic rotation within the wall. Glaucosphaera was inert but displayed formation and transport of vesicles from the perinuclear region to the surface where they were discharged. Spores (whether monospores, tetraspores or carpospores) of almost all other species or strains available displayed various forms of motility. Some (e.g. Batrachospermum-Chantransia monospores) showed smooth, directional and continuous gliding (c. 2·2 μm s¹) for shorter or longer distances, clearly detectable to the naked eye through the microscope. In others, movement could be almost as fast or slower, non-continuous and undirectional. Certain spores (e.g. monospores of the filamentous Audouinella sp.) were amoeboid, and in some cases (e.g. Sahlingia) cells actively squeezed through the irregular interstices of a crustose colony. The mechanism of movement is unknown; polysaccharide secretion may be involved in some cases. These forms of motility may be significant as dispersal mechanisms, by moving spores out of the stagnant boundary layer, and in settling processes, by allowing spores limited ability to optimize their site of germination.
Article
The laboratory culture study of Porphyra seriata Kjellman from Korea was conducted at different conditions of temperatures (5, 10, 15, 20, 25 and 30{^{\circ}C}), photon flux densities (10, 20, 40 and 80 \mumol ^{-2}s^{-1}) and photoperiods (14L: 10D and 10L:14D). Conchocelis filaments grew fast at 15-20{^{\circ}C} and 20-80 \mumol ^{-2}s^{-1} under both photoperiods. Concho sporangial branches were produced at 5-25{^{\circ}C}, and abundant when the conchocelis filaments were cultured at higher temperatures of 20-25{^{\circ}C} under both photoperiods. Foliose thalli grew well at 15-20{^{\circ}C} under 10L:14D and at 20{^{\circ}C} under 14L:10D. At 30{^{\circ}C}, the foliose thallus failed to survive. No archespores were observed at any culture conditions. Spermatangia and zygotosporangia were formed in squarish patches at the upper marginal portion of mature thalli. Anatomical examination revealed that the mature spermatangia were 64 (a/4, b/2, c/8) and 128 (a/4, b/4, c/8), and that of zygotosporangium was 16 (a/2, b/2, c/4) according to the Hus` formula.
Article
A group IC1 intron occurs in nuclear small-subunit (18S) ribosomal RNA (SSU rRNA) genes of the marine red alga Porphyra spiralis var. amplifolia. This intron occurs at the same position as the self-splicing group IC1 introns in nuclear SSU rDNAs of the fungus Pneumocystis carinii and in the green alga Chlorella ellipsoidea and shares sequence identity with the Pneumocystis carinii intron in domains L1, P1, P2, and L2, outside the conserved core. Three size variants, differing in amount of sequence in L1, exist and are differentially distributed in geographically distinct populations. Preliminary data suggest that the largest variant can self-splice in vitro. Short open reading frames are present but do not correspond to known genes. Repeated nucleotide motifs, reminiscent of duplicated target sites of transposons or Alu elements, are associated with the intron and with one of the variant forms of L1. Insertions are present in nuclear SSU rDNAs of several other Porphyra species and of the red alga Bangia atropurpurea; insertionless rDNA variants also occur in several Porphyra species. Our observations are most readily explained by intron mobility, although it remains unclear how transfer could have been mediated between genomes of organisms as ecologically diverse as marine red algae, freshwater green algae, and a mammalian-pathogenic fungus.
Book
— We studied sequence variation in 16S rDNA in 204 individuals from 37 populations of the land snail Candidula unifasciata (Poiret 1801) across the core species range in France, Switzerland, and Germany. Phylogeographic, nested clade, and coalescence analyses were used to elucidate the species evolutionary history. The study revealed the presence of two major evolutionary lineages that evolved in separate refuges in southeast France as result of previous fragmentation during the Pleistocene. Applying a recent extension of the nested clade analysis (Templeton 2001), we inferred that range expansions along river valleys in independent corridors to the north led eventually to a secondary contact zone of the major clades around the Geneva Basin. There is evidence supporting the idea that the formation of the secondary contact zone and the colonization of Germany might be postglacial events. The phylogeographic history inferred for C. unifasciata differs from general biogeographic patterns of postglacial colonization previously identified for other taxa, and it might represent a common model for species with restricted dispersal.
Article
Porphyra lacerata, P. tenuipedalis, P. suborbiculata, and P. dentata were cultured under various photon flux densities (10-80 μmol m-2s-1), temperatures (10-25°C), and photoperiods (14L: 10D, 10L: 14D) to observe their life cycles. P. lacerata and P. suboribculata both had a similar type of life cycle with a monoecious, foliose thallus producing monospores at an early stage at 10-25°C and a conchocelis-phase which produced only conchosporangial branches. P. dentata had a dioecious, foliose thallus which produced no monospores. P. tenuipedalis had a monoecious, foliose thallus which produced no conchosporangial branches. In these four species, the optimum temperature and photoperiod for growth of the conchocelis- and foliose-phases were similar; 20-25°C under 14L: 10D for the conchocelis-phase and 15-20°C under 1OL: 14D for foliose thalli.
Article
A list of chromosome numbers for fifteen species of Porphyra from the west coast of North America is presented. The basic haploid number found in vegetative cells of the foliose thallus and beta-spores is three, although numbers from two to five are found in some species. Diploid chromosome numbers were determined during divisions leading to alpha-spore formation in four species. Chromosome numbers have been recorded for the vegetative cells of the conchocelis phase of seven species; four are haploid and three diploid. Implications of these data in systematics are discussed.