ArticlePDF Available

The Primitive Thylakoid-Less Cyanobacterium Gloeobacter Is a Common Rock-Dwelling Organism

PLOS
PLOS ONE
Authors:

Abstract and Figures

Cyanobacteria are an ancient group of photosynthetic prokaryotes, which are significant in biogeochemical cycles. The most primitive among living cyanobacteria, Gloeobacter violaceus, shows a unique ancestral cell organization with a complete absence of inner membranes (thylakoids) and an uncommon structure of the photosynthetic apparatus. Numerous phylogenetic papers proved its basal position among all of the organisms and organelles capable of plant-like photosynthesis (i.e., cyanobacteria, chloroplasts of algae and plants). Hence, G. violaceus has become one of the key species in evolutionary study of photosynthetic life. It also numbers among the most widely used organisms in experimental photosynthesis research. Except for a few related culture isolates, there has been little data on the actual biology of Gloeobacter, being relegated to an "evolutionary curiosity" with an enigmatic identity. Here we show that members of the genus Gloeobacter probably are common rock-dwelling cyanobacteria. On the basis of morphological, ultrastructural, pigment, and phylogenetic comparisons of available Gloeobacter strains, as well as on the basis of three new independent isolates and historical type specimen, we have produced strong evidence as to the close relationship of Gloeobacter to a long known rock-dwelling cyanobacterial morphospecies Aphanothece caldariorum. Our results bring new clues to solving the 40 year old puzzle of the true biological identity of Gloeobacter violaceus, a model organism with a high value in several biological disciplines. A probable broader distribution of Gloeobacter in common wet-rock habitats worldwide is suggested by our data, and its ecological meaning is discussed taking into consideration the background of cyanobacterial evolution. We provide observations of previously unknown genetic variability and phenotypic plasticity, which we expect to be utilized by experimental and evolutionary researchers worldwide.
Content may be subject to copyright.
The Primitive Thylakoid-Less Cyanobacterium
Gloeobacter
Is a Common Rock-Dwelling Organism
Jan Mares
ˇ
1,2
*, Pavel Hrouzek
3
, Radek Kan
ˇ
a
3
, Stefano Ventura
4
, Otakar Strunecky
´
1,5
,Jir
ˇ
ı
´
Koma
´
rek
1,2
1 Institute of Botany ASCR, Centre for Phycology, Tr
ˇ
ebon
ˇ
, Czech Republic, 2 Department of Botany, Faculty of Science, University of South Bohemia, C
ˇ
eske
´
Bude
ˇ
jovice,
Czech Republic, 3 Institute of Microbiology ASCR, Department of Autotrophic Microorganisms - ALGATECH, Tr
ˇ
ebon
ˇ
, Czech Republic, 4 CNR-ISE Istituto per lo Studio degli
Ecosistemi, Sesto Fiorentino, Italy, 5 Centre for Polar Ecology, Faculty of Science, University of South Bohemia, C
ˇ
eske
´
Bude
ˇ
jovice, Czech Republic
Abstract
Cyanobacteria are an ancient group of photosynthetic prokaryotes, which are significant in biogeochemical cycles. The
most primitive among living cyanobacteria, Gloeobacter violaceus, shows a unique ancestral cell organization with a
complete absence of inner membranes (thylakoids) and an uncommon structure of the photosynthetic apparatus.
Numerous phylogenetic papers proved its basal position among all of the organisms and organelles capable of plant-like
photosynthesis (i.e., cyanobacteria, chloroplasts of algae and plants). Hence, G. violaceus has become one of the key species
in evolutionary study of photosynthetic life. It also numbers among the most widely used organisms in experimental
photosynthesis research. Except for a few related culture isolates, there has been little data on the actual biology of
Gloeobacter, being relegated to an ‘‘evolutionary curiosity’’ with an enigmatic identity. Here we show that members of the
genus Gloeobacter probably are common rock-dwelling cyanobacteria. On the basis of morphological, ultrastructural,
pigment, and phylogenetic comparisons of available Gloeobacter strains, as well as on the basis of three new independent
isolates and historical type specimen, we have produced strong evidence as to the close relationship of Gloeobacter to a
long known rock-dwelling cyanobacterial morphospecies Aphanothece caldariorum. Our results bring new clues to solving
the 40 year old puzzle of the true biological identity of Gloeobacter violaceus, a model organism with a high value in several
biological disciplines. A probable broader distribution of Gloeobacter in common wet-rock habitats worldwide is suggested
by our data, and its ecological meaning is discussed taking into consideration the background of cyanobacterial evolution.
We provide observations of previously unknown genetic variability and phenotypic plasticity, which we expect to be
utilized by experimental and evolutionary researchers worldwide.
Citation: Mares
ˇ
J, Hrouzek P, Kan
ˇ
a R, Ventura S, Strunecky
´
O, et al. (2013) The Primitive Thylakoid-Less Cyanobacterium Gloeobacter Is a Common Rock-Dwelling
Organism. PLoS ONE 8(6): e66323. doi:10.1371/journal.pone.0066323
Editor: John R. Battista, Louisiana State University and A & M College, United States of America
Received January 12, 2013; Accepted May 3, 2013; Published June 18, 2013
Copyright: ß 2013 Mares
ˇ
et al. This is an open-access artic le distributed under the terms of the Creative Commons Attribution License, which permits
unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Funding: This study was supported as a long-term research development project no. RVO 67985939, by the grant no. GA JU 135/2010/P, GAC
ˇ
R P506/12/1818
and by the Center for Algal Biotechnology Tr
ˇ
ebon
ˇ
- ALGATECH (CZ. 1.05/21.00/03.0110). The authors acknowledge MetaCentrum v.o. for providing
supercomputing facilities under the research agreement MSM6383917201. The funders had no role in study design, data collection and analysis, decision to
publish, or preparation of the manuscript.
Competing Interests: The authors have declared that no competing interests exist.
* E-mail: mares@butbn.cas.cz
Introduction
Cyanobacteria are the most significant group of photosynthetic
prokaryotes, generating global impact effecting biogeochemical
cycles since ancient Earth history [1,2]. As one of the most
important sources of atmospheric oxygen and crucial carbon
fixers, they have been intensively studied by experimental and
evolutionary science [3,4,5]. Gloeobacter violaceus Rippka et al. 1974,
the most primitive among living cyanobacteria, has been subjected
to both of these approaches. In its original description [6], the
authors studied a single cyanobacterial strain (PCC 7421) isolated
from the surface of a limestone rock in Kernwald (Switzerland). A
peculiar simple cell organization with complete absence of
thylakoids and an unusual structure of the photosynthetic
apparatus supported the description of a new, separate monotypic
genus Gloeobacter [6,7,8]. Following studies of the phylogenetic
comparison of SSU rRNA gene and other loci [9,10,11]
demonstrated that G. violaceus diverged very early during the
cyanobacterial radiation, in an ancient lineage preceding the
cyanobacterial chloroplast ancestors [12]. These findings were in
accordance with its primitive morphology and cell ultrastructure.
Since then, G. violaceus PCC 7421 has become one of the key
species in evolutionary studies of (cyano)bacteria [13,14,15] and
plant life in general [16,17,18]. Gloeobacter was also among the first
cyanobacterial strains having its complete genome sequenced [19].
Besides its significance in evolutionary research, G. violaceus PCC
7421 has been frequently used as a model organism for
experimental studies of oxygenic photosynthesis [20]. A unique
molecular structure of photosystems I and II [21,22,23] and an
unusual morphology of its phycobilisomes (PBS) [24] enable
Gloeobacter to harvest light and transfer energy in a manner, which
is different from other photosynthetic organisms. Unlike in other
cyanobacteria, its PBSs are composed of six peripheral phycocy-
anin/phycoerythrin rods bound as a bundle to five horizontal rods
of an allophycocyanin core, allowing atypical energy transfer
pathways [25]. Apart from the field of photosynthesis research, a
pentameric ligand-gated ion channel (GLIC) was cloned from G.
violaceus and has become an important molecular model of
membrane receptors in general biological and clinical studies
[26,27,28].
PLOS ONE | www.plosone.org 1 June 2013 | Volume 8 | Issue 6 | e66323
In contrast to detailed knowledge of cell structure, physiology
and genetics of the experimental model strain PCC 7421, data on
the ecology, distribution, life strategy and overall variability in the
genus Gloeobacter have, until now, remained extremely limited.
Apart from PCC 7421, a few more populations were collected and
isolated from nearby localities (PCC 9601, PCC 8105). The only
cyanobacterium morphologically and phylogenetically corre-
sponding to G. violaceus found, aside from the original locality,
was reported to be from a surface of a fountain in Florence, Italy
[29]. The unexplained biological identity of Gloeobacter and its
relationship to other cyanobacteria has been under discussion ever
since the establishment of the genus. Already in its original
description, Rippka et al. [6] suggested a close relationship to a
little known botanical species Gloeothece coerulea Geitler 1927. This
latter cyanobacterium has a similar, simple cell morphology and
life cycle, and contains polar granules clearly resembling those of
G. violaceus [30]. On the basis of this assumption and the
observation of fossil cyanobacterial remnants, Golubic & Camp-
bell [31] hypothesized that the Gloeobacter/Gloeothece coerulea-like
morphotype has colonized epilithic habitats since the Precambri-
an. Another rock-inhabiting species, Aphanothece caldariorum Richter
1880, was studied in detail by Hansgirg [32] and it has quite often
been found in samples of (sub-)aerophytic microalgal biofilms since
then. As noted by Koma´rek & Anagnostidis [33], this cyanobac-
terium is almost identical to G. coerulea in both morphology and life
strategy and thus potentially related to G. violaceus.
Under the current cyanobacterial taxonomy [33], the genus
Gloeobacter can be relatively easily distinguished from the morpho-
logically similar genera Aphanothece, Anathece, Cyanobium and
Gloeothece by its phylogenetic position and by the absence of
thylakoids. However, until now, Gloeobacter was studied exclusively
by observing its cultures, while similar morphospecies from other
genera (e.g., Gloeothece coerulea, Aphanothece caldariorum) have never
been isolated into cultures, making it impossible to study them
using electron microscopy and molecular analysis to provide the
necessary evidence concerning their phylogeny and cell ultra-
structure.
In this study, we provide extensive evidence for three original
isolates of Gloeobacter violaceus, a botanical type specimen of A.
caldariorum, and two reference strains G. violaceus PCC 7421 and
PCC 9601 to uncover their taxonomic identity, phylogeny, and
biology in natural populations.
Methods
Strains and Samples
Two strains of Gloeobacter violaceus (PCC 7421 and PCC 6901)
were obtained from the Pasteur Culture Collection of Cyanobac-
teria, Paris, France. Another strain (VP3-01) was purified from a
biofilm growing on a surface of a fountain in Florence, Italy, as
previously described [29]. Two new strains were isolated from
samples dominated by cyanobacteria corresponding to the
morphospecies Aphanothece caldariorum from wet rocks in artificial
waterfalls in tropical greenhouses of botanical gardens in Liberec
and Teplice, the Czech Republic, in the years 2009–2010. The
collection of samples of cyanobacteria in the greenhouses of the
public Botanical Gardens in Liberec and Teplice, the Czech
Republic was approved by their directors Miloslav Studnic
ˇ
ka
(Liberec) and Jir
ˇ
ı
´
R. Haager (Teplice). No additional permissions
were required by the legal regulations of the Czech Republic. No
protected species were sampled. For isolation, a portion of an
environmental sample was spread on the surface of an agar plate
enriched with BG 11 medium [34]. Colonies that emerged from
individual Aphanothece caldariorum-like cells (checked by optical
microscopy) were then sequentially transferred to fresh plates until
unicyanobacterial strains were obtained. The holotype of A.
caldariorum var. cavernarum Hansgirg 1889 was provided by the WU
(Institute of Botany, University of Vienna) herbarium. The type
material of the nominate variety of this species was not found in
the respective herbaria, as it was probably lost.
The strains were cultivated in both liquid and agar BG11
medium for morphological, ultrastructural and molecular analysis.
For the purpose of pigment analysis and statistical analysis of cell
dimensions, the strains were grown in 50 mL liquid batch cultures
(inoculated with 1 mL of a previous batch at growth maximum),
under constant temperature (23uC) and irradiance (2065
mmol
m
22
s
21
). Our original cultures were deposited in Culture
Collection of Autotrophic Organisms (CCALA) of the Institute
of Botany ASCR, which is accessible to the public, under accession
codes CCALA 979 ( = G. violaceus VP3-01 from Florence), CCALA
980 ( = G. violaceus [A. caldariorum morphotype] from Teplice) and
CCALA 981 ( = G. violaceus [A. caldariorum morphotype] from
Liberec).
Morphology and Ultrastructure
Fresh cyanobacteria were observed under 4006 and 10006
magnifications using Olympus BX 51 microscope equipped with
differential interference contrast, an Olympus DP71 camera, and
the QuickPhoto Micro v. 2.3 image analysis software. Statistical
analysis of cell dimensions in cultures was based on photographs
taken at 10006magnification. Length and width of 100 randomly
selected mature cells (excluding the stage of fission or just after it)
from each sample were measured with an accuracy of 60.1
mm.
Statistical differences in cell length and length:width ratio between
the batch cultures of individual strains was assessed, using
Statistica v. 9.1 [35], by a Kruskal-Wallis test; pair-wise differences
among the batches were evaluated by standard 2-tailed t-tests.
For ultrastructural studies, biological material of cyanobacteria
was fixed with 6% glutaraldehyde and kept at room temperature.
Samples were washed with 0.05 M phosphate buffer (pH 7.2) and
postfixed with 2% osmium tetroxide in the same buffer at room
temperature for 2 hours, then repeatedly washed with 0.05 M
phosphate buffer. Finally, cells were dehydrated with a graded
isopropanol series and embedded in Spurr’s resin [36] using
propylene oxide as an intermediate stage. Thin sections were
stained with uranyl acetate and lead citrate and observed in a Jeol
JEN 1010 transmission electron microscope at 80 kV.
Table 1. List of PCR and sequencing primers.
Primer Name Sequence (59to 39)
359F
1
[39] GGG GAA TYT TCC GCA ATG GG
23S30R
1
[38] CTT CGC CTC TGT GTG CCT AGG T
Cyano6r
2*
GAC GGG CCG GTG TGT ACA
T7
2
TAA TAC GAC TCA CTA TAG GG
SP6r
2
TAT TTA GGT GAC ACT ATA G
rpc/MF
1
[9] GGT GAR GTN ACN AAR CCA GAR AC
rpc/CR-1
1
[9] CCA GAR TAG TCN ACC CGT TTA CC
1
PCR primers,
2
sequencing primers,
*reverse complement to primer 14 [38].
doi:10.1371/journal.pone.0066323.t001
Gloeobacter violaceus Is a Common Organism
PLOS ONE | www.plosone.org 2 June 2013 | Volume 8 | Issue 6 | e66323
Molecular Analysis
The biomass was dried for 48 hours over silica gel and crushed
to powder in a Retsch MM200 laboratory mill with wolfram
carbide beads (3 minutes, 30? s
21
). Total genomic DNA was
isolated following the modified xanthogenate-SDS buffer extrac-
tion protocol with addition of 3% PVPP and PEG-MgCl
2
precipitation [37]. Alternatively, for the historical type specimen
of A. caldariorum var. cavernarum, a small piece (1 mm
2
) of the
herbarium material was pulverized as previously, rehydrated in
50
mL of TE buffer, and 2 mL of the suspension were directly
added to the PCR mix. A section of the rRNA operon containing
the partial SSU rRNA gene and the ITS region was amplified with
primers 359F and 23S30R [38,39] (Table 1). Ten ng of template
DNA was mixed with 6 pmol of each primer in a commercial
PCR mix with Taq polymerase (Plain PP Master Mix, Top Bio,
the Czech Republic), and amplified with an initial denaturation
Figure 1. Morphology of
Gloeobacter violaceus
and
Aphanothece caldariorum
-like samples. (A) and (B) A. caldariorum-like samples from the
botanical gardens in Liberec and Teplice, respectively, showing typical rod-shaped cells with polar granules and layered mucilaginous envelopes; (C)
cell morhology in the perfectly preserved herbarium type specimen of A. caldariorum var. cavernarum; (D) ‘‘nanocytes‘‘ in the environmental sample
from Liberec dominated by A. caldariorum-like morphotype; (E) and (F) batch cultures ordered by increasing age (from the left) of A. caldariorum-like
strain CCALA 981 and G. violaceus PCC 7421, respectively, showing a gradual color shift from grey-violet to yellow-orange; (G–J) and (K–N) change in
cell morphology from subspherical nanocyte-like cells to rod-shaped cells with occasional mucilaginous envelopes in the batch cultures of CCALA
981 and PCC 7421, respectively (the batches are the same as in panels E and F). Scale bars, 10
mm.
doi:10.1371/journal.pone.0066323.g001
Gloeobacter violaceus Is a Common Organism
PLOS ONE | www.plosone.org 3 June 2013 | Volume 8 | Issue 6 | e66323
step (5 minutes at 95uC), 35 cycles of denaturation (1 min at
94uC), primer annealing (45 s at 55uC) and elongation (2 min at
72uC), and final elongation for 10 min at 72uC. The PCR product
was cloned using the standard pGEM-T Easy (Promega Corp.,
WI, USA) vector system according to supplier instructions. The
plasmid containing the required insert was purified from the
bacterial culture using Zyppy Plasmid Miniprep kit (Zymo
Research Corp., CA, USA). The rpoC1 gene fragment was
amplified using primers rpc/MF and rpc/CR-1 (Table 1)
following the published protocol [9] (initial denaturation for
5 min at 94uC, 35 cycles of 1 min denaturation at 94uC, primer
annealing for 1 min at 52uC, elongation for 2 min at 72uC, and
final elongation step for 10 min at 72uC), and cloned as above.
PCR gel pictures are given in Figure S4. The clones were
sequenced using primers T7 and SP6r (Table 1) in Laboratory of
Genomics, Biology Centre of the Academy of Sciences of the
Czech Republic, C
ˇ
eske´ Bude
ˇ
jovice (with an ABI PRISM
3130 XL, Applied Biosystems, Life Technologies Corp., CA,
USA). Sequences were deposited in GenBank under accession
numbers KC004017–KC004023 and KC866356.
Phylogenetic Analysis
Sequences of the SSU rRNA and rpoC1 gene from thirty-eight
cyanobacteria were collected from published studies and mined
from the whole genome database available in GenBank. The
sequence matrices were assembled to include equal number of
strains from all cyanobacterial orders. Three bacterial strains, E.
coli K12, Salmonella enterica str. UK-1, and Cronobacter turicensis
z3032, were used as out-group taxa. The sequences were aligned
via MAFFT v. 6 [40] using the FFT-NS-I strategy, and manually
corrected. For the final analysis of the SSU rRNA gene data, a
region of alignment that is common to all collected sequences was
used. It spanned 1041 positions from 377 to 1432 (Escherichia coli
numbering) after ambiguous gap columns were removed. Phylo-
genetic analysis was conducted employing Bayesian inference in
MrBayes 3.1.2 [41], maximum likelihood analysis in RAxML
7.3.2 [42], and maximum parsimony analysis in PAUP* 4.0b10
Figure 2. Morphometric analysis of cells in
Aphanothece caldariorum
-like and
Gloeobacter violaceus
cultures. (A) and (B) increasing cell
length corresponding to batch cultures of increasing age of A. caldariorum-like strain CCALA 981 and G. violaceus PCC 7421 (Kruskal-Wallis test,
p,0.001), difference between individual pairs of batches was also significant (two-tailed t-test, p,0.02 except for a pair of batches 41 and 50 days
old in PCC 7421, for which the difference was at the edge of statistical significance, p = 0.045); (C) and (D) increasing cell length:width ratio in the
same batch cultures as previously (Kruskal-Wallis test, p,0.001), difference between individual pairs of batches was also significant (two-tailed t-tests,
p,0.001) except for a pair of batches 41 and 50 days old in PCC 7421. One hundred cells were measured in each sample.
doi:10.1371/journal.pone.0066323.g002
Gloeobacter violaceus Is a Common Organism
PLOS ONE | www.plosone.org 4 June 2013 | Volume 8 | Issue 6 | e66323
[43]. For the Bayesian analysis, two runs of four Markov chains
were executed for 1 000 000 generations with default parameters,
sampling every 100 generations (the final average standard
deviation of split frequencies was lower than 0.01). The maximum
likelihood calculation was executed upon the generalized time-
reversible (GTR) substitution model with discrete gamma distri-
bution in six categories. The gamma shape parameter a as well as
the proportion of invariable sites were estimated from the data set
(GTR+G+ G model), and 1000 bootstrap replicates were
calculated to evaluate the relative support of branches. A
maximum parsimony analysis involved one hundred replicate
searches with starting trees obtained by random stepwise addition,
using the tree bisection-reconnection (TBR) branch swapping
algorithm; one thousand nonparametric bootstrap replications
were run with the same settings to evaluate the relative branch
support. All bases and base changes were equally weighted, and
gaps were coded as missing data. MetaCentrum (www.
metacentrum.cz) and CIPRES (www.phylo.org) supercomputing
facilities were used for fast calculation of Bayesian and likelihood
trees.
For the combined analysis of both loci, a 778 bp long region of
partial rpoC1 data was merged with an aligned SSU rRNA gene
from corresponding strains into a final concatenated alignment of
1988 bp. Phylogenetic analysis of the concatenated alignment was
conducted using Bayesian inference, maximum likelihood and
maximum parsimony methods as mentioned previously. Phyloge-
netic trees were visualized using FigTree v. 1.3. (http://tree.bio.
ed.ac.uk/software/figtree/). The alignments were uploaded to the
TreeBase web (http://purl.org/phylo/treebase/phylows/study/
TB2:S14106 ).
Pigment Analysis
The relative concentrations of the two phycobiliproteins,
phycocyanin (PC) and allophycocyanin (APC) were estimated
from the mean absorption of light between 616–624 nm and 650–
658 nm respectively according to Krogmann et al. [24]. The
relative concentration of phycoerythrin (PE) was calculated at
662 nm, by utilizing an equation, which was used in the same
study, as [PE] = A
662
/e
662
with e
662
(extinction coefficient) at
456 mM cm
21
. Absorption spectra were recorded in vivo with a
Unicam UV 550 (Thermospectronic, UK) following the method
described by [44].
Fluorescence emission spectra of living cells at room temper-
Figure 3. Light spectroscopy analysis of photosynthetic
pigments in
Aphanothece caldariorum
-like strain and
Gloeobacter
violaceus
. (A) fluorescence emission spectra of A. caldariorum-like strain
CCALA 981 in comparison with G. violaceus PCC 7421 for excitation to
phycobilisomes. Higher content of phycoerythrin in CCALA 981 is
documented by a relative increase in its fluorescence emission at
574nm. (B) and (C) whole cell absorption spectra of CCALA 981 and PCC
7421 normalized to chlorophyll a content. Clear accumulation of
carotenoids (wide absorbance peak between 450–500 nm) at the
stationary phase of growth (third and fourth batch) is obvious in both
strains. The absorbance of 41-day old strain PCC 7421 is raised due to
the beginning of accumulation of carotenoids in this batch, but
phycobilin peaks are still recognizable. Individual absorbance peaks are
as noted; CAR, carotenoids; PE, phycoerythrin; PC+APC, phycocyanin
and allophycocyanin; Chl a, chlorophyll a.
doi:10.1371/journal.pone.0066323.g003
Table 2. Relative concentration of particular
phycoerythrobilins in phycobilisomes of Aphanothece
caldariorum-like and Gloeobacter violaceus strains calculated
from absorption spectra.
PE/APC PC/APC PE/PC
A. caldariorum
-like CCALA 981 1.7160.16 1.3460.21 1.2860.08
G. violaceus
PCC 7421 1.2660.16 1.0660.22 1.260.11
PE, phycoerythrin; APC, allophycocyanin; PC, phycocyanin. See Metho ds for
precise description of calculation of the values.
doi:10.1371/journal.pone.0066323.t002
Table 3. Number of phycoerythrin and phycocyanin trimers
in a single phycobilisome rod of Aphanothece caldariorum-like
and Gloeobacter violaceus strains.
n(PE) n(PC)
A. caldariorum
-like CCALA 981 4.560.4 3.660.6
G. violaceus
PCC 7421 3.460.4 2.860.6
PE, phycoerythrin; PC, phycocyanin. Values were calculated on the basis
absorbance changes taking into account a model of the phycobilisome for G.
violaceus PCC 7421 consisting of the phycobilisome core (16 allophycocyanins
trimers) with 6 attached rods with different amount of PE and PC trimers [24].
doi:10.1371/journal.pone.0066323.t003
Gloeobacter violaceus Is a Common Organism
PLOS ONE | www.plosone.org 5 June 2013 | Volume 8 | Issue 6 | e66323
ature were measured using an Aminco-Bowman Series 2
spectrofluorometer, in the standard instrument geometry with
excitation at 492 nm. Fluorescence emission was scanned between
550–750 nm with a 4 nm bandwidth.
To assess carotenoid composition, the biomasses of CCALA 981
and PCC 7421 were extracted with an acetone/methanol mixture
(7/3 v/v) in the dark into Eppendorf tubes and centrifuged. The
samples were subjected to HPLC analysis on the Agilent
Technologies 1200 series chromatographic system with a diode-
array detector. Separation was performed on a Luna C8 column
(3
mm, 100A, 10064.6 mm –00D-4248-E0, Phenomenex, USA)
using methanol (A) and 28 mM amonium acetate in 80%
methanol (B) as solvents (from 30% to 100% of solvent A in
30 min.). The pigments were detected on the basis of their
retention times (internal database) and absorption spectra.
Results and Discussion
Natural Morphology and Habitat
Material collected from artificial waterfalls in greenhouses of
two botanical gardens revealed the presence of a morphotype
exactly matching the species Aphanothece caldariorum as described in
botanical literature [32,33]. In both cases, it was a component of
an epilithic gelatinous cyanobacterial mass. In the native state, the
collected material (from both greenshouses) showed all of the
morphological characteristics typically described for A. caldariorum
[33], such as dimensions and shape of cells, concentrically
lamellated mucilaginous envelopes, and polar granules
(Figure 1A, B). Compared to the typical properties of Gloeobacter
violaceus strains [6,33], the A. caldariorum-like cells were significantly
longer (up to more than 10
mm versus 2–3 mminG. violaceus), often
somewhat bent or arcuate (straightly rod-shaped in G. violaceus),
with broader and more conspicuously lamellated sheaths. We also
recorded frequent occurrence of small subspherical cells,
(Figure 1D) clearly corresponding to ‘‘nanocytes‘‘ [33], which
were not reported for G. violaceus. According to our observations,
Figure 4. Comparison of cell ultrastructure in
Aphanothece caldariorum
-like strains and
Gloeobacter violaceus
. Typical cell ultrastructure
of G. violaceus PCC 9601(A), G. violaceus CCALA 979 (B), A. caldariorum-like strain CCALA 981(C) and A. caldariorum-like strain CCALA 980 (D),
respectively. Cells did not contain any thylakoid, the photosynthetic pigments accumulated in an electron-dense layer near the multi-layered cell wall.
Cells typically contained two large polyphosphate granules in polar positions. Observed ultrastructure was identical to the reference strain G.
violaceus PCC 7421 [6]. Scale bars, 500 nm.
doi:10.1371/journal.pone.0066323.g004
Gloeobacter violaceus Is a Common Organism
PLOS ONE | www.plosone.org 6 June 2013 | Volume 8 | Issue 6 | e66323
Gloeobacter violaceus Is a Common Organism
PLOS ONE | www.plosone.org 7 June 2013 | Volume 8 | Issue 6 | e66323
these cells do not seem to represent any specialized reproductive
stage. Apparently, they were produced by fast serial binary fission
under favorable conditions. A. caldariorum-like morhotype was
found on wet rocks in waterfalls, which corresponded well to the
original description of this species from botanical gardens in
Prague and wet rocks in Bohemia [32], and also to other reports of
this species from numerous similar localities worldwide
[45,46,47,48,49,50,51,52,53]. The similarity of the morphotype
to A. caldariorum was further confirmed through microscopic
examination of the well-preserved botanical type specimen of A.
caldariorum var. cavernarum (Figure 1C). Unfortunately, the type
specimen of the nominate variety caldariorum was lost or destroyed
during the Second World War, and is not available for study.
Thus, a definite proof of identity (by molecular methods) of our
material to A. caldariorum cannot be given.
Strain Morphology and Pigment Analysis
Very high similarity in basic morphometric characteristics was
found between typical Gloeobacter PCC 7421 strain and all the
studied Apahnothece caldariorum morphotype samples. The observed
cyanobacterial isolates also exhibited almost identical changes in
morphology and pigmentation during their life cycle, as described
below.
In cultures grown in nutrient-rich media (BG 11), cells
resembling Aphanothece caldariorum, immediately after inoculation,
started accelerated proliferation (Figure S1), resulting in compact
clusters of small subspherical ‘‘nanocyte-like‘‘ cells, mostly with
two distinct granules inside (Figure 1G). As long as it was cultured
in fresh media, this morphotype was maintained. Upon consulting
available literature [6,34], this strain morphology obviously
matched that of the reference strain Gloeobacter violaceus PCC
7421. Our optical microscopy observations verified identical
appearance of all gathered strains (PCC 7421, PCC 9601,
CCALA 979, CCALA 980, CCALA 981– Figure 1G–N, Figure
S2).
During the growth of batch cultures on both solid and liquid
media, a dramatic color shift was observed in all strains (Figure 1E,
F). The color changed from various shades of grey, greyish-blue-
green and greyish-violet in young cultures, through bright violet or
pinkish-violet at growth maxima, to green or yellow-green in older
cultures, ending with yellow and orange in their stage of
senescence. These observations were in accordance with reports
of an extreme color variability of A. caldariorum in natural habitats
[32,33].
In order to determine the principles and symptoms of these
changes, we compared series of batch cultures grown in
standardized conditions for one representative of typical G.
violaceus (PCC 7421) and one representative of a strain derived
from the A. caldariorum-like morphotype (CCALA 981). Four
batches of different ages spanning the color range from young and
thriving (grey, violet) to senescent cultures (yellow, orange) were
chosen for morphometric and pigment analysis for each of the two
strains (Figure 1E, F). As shown by the statistical analysis of cell
dimensions (Figure 2, p,0.001), the cells of both strains changed
their shape from subspherical to short cylindrical of ‘‘nanocytes‘‘
in young cultures, to rod-shaped cells in old cultures (Figure 1G–
N). Cultivated cells never reached a length of over 5
mm, which is
a dimension frequently exceeded in natural populations of
A.caldariorum (Figure 1A–C, [33]). A probable reason for such a
difference lies in the unnatural chemical composition of the culture
medium, and, possibly, also in the different physical properties of
the artificial medium as compared to natural substrates. However,
in the oldest batches the cells were often somewhat arcuate and
enclosed in gelatinous envelopes (Figure 1J, N) clearly resembling
the typical A. caldariorum morphotype.
To elucidate the process of color shift, we performed a light
spectroscopy analysis of the main photosynthetic pigments
(phycobilins and chlorophyll a) and HPLC analysis of carotenoids
in the same batch cultures that were used for morphometry.
In general, the pigment composition of CCALA 981 and PCC
7421 was similar. Nevertheless, the A. caldariorum-like strain had
slightly different molar ratios of particular phycobiliproteins within
the PBS in comparison to G. violaceus PCC 7421 (Table 2). The
molar ratio (measured in young cultures) in PCC 7421 was about
1APC: 1.3PC: 1.2 PE, while in CCALA 981 we found relatively
higher contents of phycoerythrins and phycocyanins in compar-
ison to allophycocyanins (1APC: 1.3 PC: 1.7 PE). The higher
content of phycoerythrin in the latter strain was also confirmed by
a relative increase in its fluorescence emission at 574nm
(Figure 3A). The observed discrepancy in the phycobilin ratios
corresponded well to a slightly different color of the cultures
(violet-grey vs. bright violet). This finding suggests a relative
increase in the length of PBS rods in CCALA 981 (Table 3). The
PBS in G. violaceus PCC 7421 was previously shown to consist of
16 APCs in the core, with 6 attached rods of different length and
PE/PC ratio [24,54] depending on growth conditions. Assuming
an identical core structure, the PBS rods of the A. caldariorum-like
strain CCALA 981 were almost 40% longer in comparison to
PCC 7421 (In Table 3, almost 8 PC+PC trimers per single rod of
CCALA 981 and only five PC+PC trimers in PCC 7421).
Alternatively, some PBSs without the APC core could occur
similarly to the so called Cpc-G2 phycobilisomes recently
described for Synechocystis sp. [55].
Our data clearly suggests that the color variability in young
cultures (grey/blue-green/violet hues) can be explained by
different ratios of individual phycobilins depending on actual
growth conditions and physiological state of the cultures
(Figure 3A, Figure S3). The versatile phycobilin ratio can also
explain the greyish color of young PCC 7421 (Figure 1F), which
differs from bright violet color typically described for this strain
[6]. As documented by absorption spectra (Figure 3B, C), in older
batch cultures the PBSs degraded while carotenoids accumulated.
The decrease, and finally the total absence of phycoerytrin
(565 nm) and phycocyanin (620 nm) peak, could be seen in the
whole cell spectra of older cultures in both CCALA 981 and PCC
7421, where the phycobilins were replaced by an intensively
absorbing band in the wavelength range of 400–550 nm,
corresponding mainly to carotenoid absorption. This process was
reflected by a gradual shift in color from greenish (chlorophyll) to
yellowish and orange (carotenoids).
The carotenoid composition of the A. caldariorum-like strain
CCALA 981 was identical to that of G. violaceus PCC7421 as
assessed by HPLC analysis. The dominant carotenoids were b-
Carotene and (2S,29S)-oscillol 2,29-di(a-l-fucoside), while echine-
none was found as a minor component. Our results matched those
in the report for G. violaceus PCC7421 by Tsuchiya and co-workers
Figure 5. Phylogenetic position of
Gloeobacter violaceus
and
Aphanothece caldariorum
. (A) Phylogenetic tree based on a SSU rRNA gene
alignment. (B) Phylogenetic tree based on a concatenated SSU rRNA gene+rpoC1 alignment. Sequences generated in this study are printed in bold
font. Branch support values (%) are given at nodes in this format: Bayesian inference/maximum likelihood/maximum parsimony. A well supported
basal clade of cyanobacteria consisting of G. violaceus and A. caldariorum is highlighted by blue colour.
doi:10.1371/journal.pone.0066323.g005
Gloeobacter violaceus Is a Common Organism
PLOS ONE | www.plosone.org 8 June 2013 | Volume 8 | Issue 6 | e66323
[16]. Presence of the rarely occurring carotenoid (2S,29S)-oscillol
2,29-di(a-l-fucoside) in the studied strains further supported their
isolated position in phylogeny of cyanobacteria. Carotenoids
belonging to oscillol 2,2-diglycosides were found only in limited
number of cyanobacteria and bacteria [56,57].
Ultrastructure
Perhaps the most remarkable feature in Gloeobacter, that
separates it from other cyanobacteria, is the complete absence of
thylakoid membranes [6,58]. Hence, the study of cell ultrastruc-
ture by TEM was one of the crucial components of our analysis.
Our results unambiguously showed an identical cellular structure
in all of the studied strains (Figure 4A–D), exactly matching that of
the reference strain PCC 7421 [6]. The multi-layered cell wall was
fringed by a band of electron-dense material (photosynthetic
pigments). No thylakoid membranes were registered but usually
two large polyphosphate granules were present, one at each pole of
the cell.
Phylogeny and Taxonomy
The phylogenetic reconstruction based on both, the SSU rRNA
gene and a combination of SSU rRNA gene with a protein-coding
housekeeping gene like rpoC1, provided congruent results regard-
ing the position of Gloeobacter violaceus and Aphanothece caldariorum-
like strains. All studied samples clustered in a single, distinct, fully
supported basal clade, which clearly corresponded to the genus
Gloeobacter (Figure 5). All the cultured strains of G. violaceus and A.
caldariorum-like isolates, including the reference strain PCC 7421,
formed an extremely tight cluster, which has to be regarded as the
single species - G. violaceus (SSU rRNA gene similarity over 99%).
This conspicuous taxon is characterized by a specific combination
of characters, i.e. phylogenetic position, absence of thylakoids,
pigment composition, life cycle, and ecology (life strategy). Given
the fact that direct comparison with the type material of A.
caldariorum var. caldariorum by molecular methods is impossible, the
decision whether this species is truly identical with G. violaceus
depends on interpretation of indirect evidence. In our opinion, at
least the assignment of A. caldariorum to the genus Gloeobacter is
clear, and the identity of the two species is quite possible.
Unfortunately, the name Gloeobacter violaceus was never validly
published under the rules of Bacteriological Code, and if identity
with A. caldariorum was assumed, the epithet ‘‘caldariorum‘‘ (and
probably also ‘‘coerulea‘‘ from Gloeothece coerulea) would have priority
over ‘‘violaceus‘‘ under the Botanical Code. Thus, the nomencla-
toric status of the species Gloeobacter violaceus is rather unclear and
has to be amended by a dedicated study. Interestingly, the SSU
rRNA gene sequence of the A. caldariorum var. cavernarum type
specimen was slightly different from the rest of Gloeobacter
sequences. Considering the SSU rRNA gene similarity (, 96%),
it could be regarded as a separate species of Gloeobacter. However,
our observations did not reveal any obvious difference (other than
DNA sequence) when compared to the material of A. caldariorum
from the greenhouses in Liberec and Teplice. The original
description [32] distinguished between the varieties on the basis of
slight differences in cell dimensions and color, characteristics,
which were proven in this study to show great variability. In our
opinion, a decisive taxonomic conclusion would require analysis of
fresh samples, isolated strains of the var. cavernarum, and collection
of more data. Nevertheless, based on the relatively common
occurrence of the A. caldariorum morphotype (as mentioned in
scientific reports worldwide), it is quite probable that members of
the genus Gloeobacter are much more common than previously
thought. This hypothesis is yet to be tested by careful study of
epilithic cyanobacterial communities in future.
While the picture of Gloeobacter as an isolated ancient lineage
agrees with most published results [4,12,13,59,60], a recent study
[61] proposed a possible relationship to other simple coccoid
cyanobacteria. In their report, Courdeau et al. [61] described a
phylogenetic cluster with moderate branch support, consisting of
Synechococcus-like morphotypes related to Gloeobacter, which they
called ’Gloeobacterales’. It also included the peculiar candidatus
Gloeomargarita lithophora that forms intracellular calcite precipitates.
In a manner similar to
Gloeobacter, members of this group diverged
earlier in evolution than the chloroplast ancestors. In our
evolutionary trees, this cluster, represented by the Synechococcus
strains isolated from the Yellowstone National Park hot springs
(C9, Ja-3-3Ab and Ja-2-3B9a), branched separately (Figure 5B).
Considering the major differences in cell ultrastructure and
photosynthetic apparatus, we propose that the order Gloeobacter-
ales should be reserved for the genus Gloeobacter, which is primarily
defined by the absence of thylakoids [62]. This is further supported
by a considerable SSU rRNA gene sequence distance; for
comparison, there is approximately 88% similarity between G.
violaceus PCC 7421 and Synechococcus sp. Ja-3-3Ab while the
similarity between PCC 7421 and a heterocytous cyanobacterium,
Nostoc sp. PCC 7120, is 87%. On the other hand, in all known
cyanobacteria from these ancient lineages, there is a clear common
tendency to colonize wet or submerged, mostly calcite rocks [61].
This evidence reflects a possible first appearance of cyanobacteria
in rock-associated, calcifying biofilm habitats, such as stromatolites
or travertine spring mats.
Conclusions
Our results brought new clues to solving a 40 years old puzzle
about the true biological identity of Gloeobacter violaceus,an
important model organism with a great value in several biological
disciplines. In the first place, we showed that the genus Gloeobacter is
a commonly occurring terrestrial cyanobacterium. On the basis of
detailed morphological, ultrastructural, biochemical, and phylo-
genetic comparisons of two available Gloeobacter strains, three new
independent isolates, and a botanical type specimen, we generated
complementary evidence of the identity of Gloeobacter with a long
known rock-dwelling cyanobacterial morphospecies Aphanothece
caldariorum. The life strategy of Gloeobacter/A. caldariorum is
congruent with that of other primitive coccoid cyanobacteria,
suggesting a possible origin of their cyanobacterial ancestors in
alkaline rock-associated biofilms. In this paper we provided
observations of previously unknown genetic variability and
phenotypic plasticity, which we expect to be utilized by
experimental and evolutionary researchers worldwide.
Supporting Information
Figure S1
A. caldariorum-like morphology: formation of nanocyte-
like cells in culture. The cells of A. caldariorum CCALA 981
started rapid successive binary fission shortly after inoculation on
fresh media. (A) Cell division into multiple small spherical cells. (B)
Nanocyte-like daughter cells forming clusters on solid medium.
Formation of nanocyte-like cells was observed in the initial stages
of cultivation directly on the agar plate when there was still some
contamination by bacteria and fungi. Scale bars, 50
mm.
(PDF)
Figure S2
Identical morphology of Aphanothece caldariorum-like
and Gloeobacter violaceus isolates in culture. (A) G.
violaceus PCC 9601; (B) G. violaceus CCALA 979; (C) A. caldariorum-
Gloeobacter violaceus Is a Common Organism
PLOS ONE | www.plosone.org 9 June 2013 | Volume 8 | Issue 6 | e66323
like strain CCALA 980. Strains PCC 7421 and CCALA 981 are
documented in Figure 1. Scale bars, 10
mm.
(PDF)
Figure S3
Proportion of phycobiliproteins in Aphanothece caldar-
iorum and Gloeobacter violaceus during the culture
senescence. (A) A. caldariorum CCALA 981, (B) G. violaceus PCC
7421. Both PE and PC were degraded as the culture aged. Thus,
the blue-green/violet colour at the beginning of the cultivation was
replaced by yellow-orange colour (carotenoids). Interestingly, the
PE/PC ratio was increased in old cultures at the end of the
cultivation in both strains. This was due to major decrease in PC,
as seen from the PC/APC curve. The relatively high PE
proportion at the end of cultivation also agreed with the orange
colour. PE, phycoerytrin; PC, phycocyanin; APC, allophycocya-
nin.
(TIF)
Figure S4
PCR products of SSU rRNA gene region and partial
rpoC1 visualized on 1.5% agarose gels. (A) and (B) SSU
rRNA gene region PCR products; (C) and (D) Partial rpo C1 gene
PCR products; (E) SSU rRNA gene region products amplified
from A. caldariorum var. cavernarum type specimen by direct PCR.
Sample names are indicated at loading wells. A standard 100 bp
DNA ladder with fragment sizes corresponding to 100, 200, 300,
400, 500, 600, 700, 800, 900, 1000, 1200 and 1517 bp was used in
all gels. The samples were stained by GelRed Nucleic Acid Dye
(Biotium, Hayward, USA). C, negative control (blank); X,
unsuccessful PCR.
(TIF)
Acknowledgments
The skillful technical assistance of Cristina Macalchi and Dana S
ˇ
vehlova´is
gratefully acknowledged. We greatly appreciate the opportunity to study
cyanobacteria from the Botanical Gardens in Liberec and Teplice, which
was provided by Miloslav Studnic
ˇ
ka and Jir
ˇ
ı
´
R. Haager.
Author Contributions
Conceived and designed the experiments: JM PH RK JK. Performed the
experiments: JM PH RK. Analyzed the data: JM PH RK SV OS.
Contributed reagents/materials/analysis tools: SV. Wrote the paper: JM
PH SV OS JK.
References
1. Tomitani A, Knoll AH, Cavanaugh CM, Ohno T (2006) The evolutionary
diversification of cyanobacteria: Molecular-phylogenetic and paleontological
perspectives. Proc Natl Acad Sci U S A 103: 5442–5447.
2. Blank CE, Sanchez-Baracaldo P (2010) Timing of morphological and ecological
innovations in the cyanobacteria - a key to understanding the rise in atmospheric
oxygen. Geobiology 8: 1–23.
3. Koksharova OA (2010) Application of molecular genetic and microbiological
techniques in ecology and biotechnology of cyanobacteria. Microbiology 79:
721–734.
4. Schirrmeister BE, Antonelli A, Bagheri HC (2011) The origin of multicellularity
in cyanobacteria. BMC Evol Biol 11: 45.
5. Beck C, Knoop H, Axmann IM, Steuer R (2012) The diversity of cyanobacterial
metabolism: genome analysis of multiple phototrophic microorganisms. BMC
Genomics 13: 56.
6. Rippka R, Waterbury J, Cohen-Bazire G (1974) Cyanobacterium which lacks
thylakoids. Arch Microbiol 100: 419–436.
7. Guglielmi G, Cohen-Bazire G, Bryant DA (1981) The structure of Gloeobacter
violaceus and its phycobilisomes. Arch Microbiol 129: 181–189.
8. Bryant DA, Cohen-Bazire G, Glazer AN (1981) Characterization of the
biliproteins of Gloeobacter violaceus chromophore content of a cyanobacterial
phycoerythrin carrying phycourobilin chromophore. Arch Microbiol 129: 190–
198.
9. Seo PS, Yokota A (2003) The phylogenetic relationships of cyanobacteria
inferred from 16S rRNA, gyrB, rpoC1 and rpoD1 gene sequences. J Gen Appl
Microbiol 49: 191–203.
10. Hoffmann L, Kasˇtovsky´ J, Koma´rek J (2005) System of cyanoprokaryotes
(cyanobacteria) state in 2004. Arch Hydrobiol Suppl Algol Stud 117: 21.
11. Gupta RS, Mathews DW (2010) Signature proteins for the major clades of
Cyanobacteria. BMC Evol Biol 10.
12. Criscuolo A, Gribaldo S (2011) Large-Scale Phylogenomic Analyses Indicate a
Deep Origin of Primary Plastids within Cyanobacteria. Mol Biol Evol 28: 3019–
3032.
13. Turner S, Pryer KM, Miao VPW, Palmer JD (1999) Investigating deep
phylogenetic relationships among cyanobacteria and plastids by small submit
rRNA sequence analysis. J Eukaryot Microbiol 46: 327–338.
14. Zhaxybayeva O, Gogarten JP, Charlebois RL, Doolittle WF, Papke RT (2006)
Phylogenetic analyses of cyanobacterial genomes: Quantification of horizontal
gene transfer events. Genome Res 16: 1099–1108.
15. Gupta RS (2009) Protein signatures (molecular synapomorphies) that are
distinctive characteristics of the major cyanobacterial clades. Int J Syst Evol
Microbiol 59: 2510–2526.
16. Tsuchiya T, Takaichi S, Misawa N, Maoka T, Miyashita H, et al. (2005) The
cyanobacterium Gloeobacter violaceus PCC 7421 uses bacterial-type phytoene
desaturase in carotenoid biosynthesis. Febs Letters 579: 2125–2129.
17. Mimuro M, Tomo T, Tsuchiya T (2008) Two unique cyanobacteria lead to a
traceable approach of the first appearance of oxygenic photosynthesis.
Photosynth Res 97: 167–176.
18. Williamson A, Conlan B, Hillier W, Wydrzynski T (2011) The evolution of
Photosystem II: insights into the past and future. Photosynth Res 107: 71–86.
19. Nakamura Y, Kaneko T, Sato S, Mimuro M, Miyashita H, et al. (2003)
Complete genome structure of Gloeobacter violaceus PCC 7421, a cyanobacterium
that lacks thylakoids. DNA Res 10: 137–145.
20. Bernat G, Schreiber U, Sendtko E, Stadnichuk IN, Rexroth S, et al. (2012)
Unique properties vs. common themes: The atypical cyanobacterium Gloeobacter
violaceus PCC 7421 is capable of state transitions and blue-light-induced
fluorescence quenching. Plant Cell Physiol 53: 528–542.
21. Inoue H, Tsuchiya T, Satoh S, Miyashita H, Kaneko T, et al. (2004) Unique
constitution of photosystem I with a novel subunit in the cyanobacterium
Gloeobacter violaceus PCC 7421. Febs Letters 578: 275–279.
22. Dreher C, Hielscher R, Prodohl A, Hellwig P, Schneider D (2010)
Characterization of two cytochrome b(6) proteins from the cyanobacterium
Gloeobacter violaceus PCC 7421. J Bioenerg Biomembr 42: 517–526.
23. Mimuro M, Yokono M, Akimoto S (2010) Variations in photosystem I
properties in the primordial cyanobacterium Gloeobacter violaceus PCC 7421.
Photochem Photobiol 86: 62–69.
24. Krogmann DW, Perez-Gomez B, Gutierrez-Cirlos EB, Chagolla-Lopez A, de la
Vara LG, et al. (2007) The presence of multidomain linkers determines the
bundle-shape structure of the phycobilisome of the cyanobacterium Gloeobacter
violaceus PCC 7421. Photosynt Res 93: 27–43.
25. Yokono M, Akimoto S, Koyama K, Tsuchiya T, Mimuro M (2008) Energy
transfer processes in Gloeobacter violaceus PCC 7421 that possesses phycobilisomes
with a unique morphology. Biochim Biophys Acta Bioenergetics1777: 55–65.
26. Bocquet N, Nury H, Baaden M, Le Poupon C, Changeux JP, et al. (2009) X-ray
structure of a pentameric ligand-gated ion channel in an apparently open
conformation. Nature 457: 111–114.
27. Weng Y, Yang LY, Corringer PJ, Sonner JM (2010) Anesthetic sensitivity of the
Gloeobacter violaceus proton-gated ion channel. Anesth Analg 110: 59–63.
28. Nury H, Van Renterghem C, Weng Y, Tran A, Baaden M, et al. (2011) X-ray
structures of general anaesthetics bound to a pentameric ligand-gated ion
channel. Nature 469: 428–431.
29. Cuzman OA, Ventura S, Sili C, Mascalchi C, Turchetti T, et al. (2010)
Biodiversity of phototrophic biofilms dwelling on monumental fountains. Microb
Ecol 60: 81–95.
30. Geitler L (1927) Neue Blaualgen aus Lunz. Archiv fu¨r Protistenkunde 60: 440–
448.
31. Golubic S, Campbell SE (1979) Analogous microbial forms in recent subaerial
habitats and in precambrian cherts Gloeothece coerulea Geitler and Eosynechococcus
moorei Hofmann. Precambrian Res 8: 201–217.
32. Hansgirg A (1892) Prodromus der Algenflora von Bo¨hmen. 2. Arch Naturwiss
Landesdurchforsch Bo¨hmen 8: 1–268.
33. Koma´rek J, Anagnostidis K (1999) Cyanoprokaryota. 1. Teil: Chroococcales.
Su¨sswasserflora von Mitteleuropa. Heidelberg, Berlin: Spektrum Akademischer
Verlag GmbH. 548.
34. Rippka R, Deruelles J, Waterbury JB, Herdman M, Stanier RY (1979) Generic
assignments, strain histories and properties of pure cultures of cyanobacteria.
J Gen Microbiol 111: 1–61.
35. Statsoft Inc. (2012) STATISTICA (data analysis software system), version 9.1.
www.statsoft.com.
36. Spurr AR (1969) A low-viscosity epoxy resin embedding medium for electron
microscopy. J Ultrastruct Res 26: 31–43.
Gloeobacter violaceus Is a Common Organism
PLOS ONE | www.plosone.org 10 June 2013 | Volume 8 | Issue 6 | e66323
37. Yilmaz M, Phlips EJ, Tillett D (2009) Improved methods for the isolation of
cyanobacterial DNA from environmental samples. J Phycol 45: 517–521.
38. Wilmotte A, Van der Auwera G, Dewachter R (1993) Structure of the 16S
ribosomal-RNA of the thermophilic cyanobacterium Chlorogloeopsis HTF
(Mastigocladus laminosus HTF) strain PCC7518, and phylogenetic analysis. FEBS
Letters 317: 96–100.
39. Nu¨ bel U, Garcia-Pichel F, Muyzer G (1997) PCR primers to amplify 16S rRNA
genes from cyanobacteria. Appl Env Microbiol 63: 3327–3332.
40. Katoh K, Asimenos G, Toh H (2009) Multiple alignment of DNA sequences
with MAFFT. In: Posada D, editor. Bioinformatics for DNA Sequence Analysis.
Totowa, New Jersey: Humana Press. 39–64.
41. Ronquist F, Huelsenbeck JP (2003) MrBayes 3: Bayesian phylogenetic inference
under mixed models. Bioinformatics 19: 1572–1574.
42. Stamatakis A (2006) RAxML-VI-HPC: Maximum likelihood-based phylogenetic
analyses with thousands of taxa and mixed models. Bioinformatics 22: 2688–
2690.
43. Swofford DL (2002) PAUP*. Phylogenetic Analysis Using Parsimony (*and other
methods). Version 4. Sunderland, Massachusetts: Sinauer Associates.
44. Kan
ˇ
aR,Pra´sˇil O, Koma´rek O, Papageorgiou GC, Govindjee (2009) Spectral
characteristic of fluorescence induction in a model cyanobacterium, Synechococcus
sp. (PCC 7942). Biochim Biophys Acta Bioenergetics 1787: 1170–1178.
45. Hauer T (2008) Epilithic cyanobacterial flora of Mohelenska´ hadcova´ steppe
Nature Reserve (western Moravia, Czech Republic) 70 years ago and now.
Fottea 8: 129–132.
46. Halda J, Hauer T, Kocia´nova´M,Mu¨hlsteinova´R,R
ˇ
eha´kova´ K, et al. (2011)
Biodiverzita ce´vnaty´ch rostlin, lisˇejnı
´
ku
˚
, sinic a r
ˇ
as na skala´ch s ledopa´dy v
Labske´m dole. Opera Corcontica 48: 45–67.
47. Johansen JR, Lowe RL, Carty S, Fuc
ˇ
ı
´
kova´ K, Olsen CE, et al. (2007) New algal
species records for Great Smoky Mountains National Park, with an annotated
checklist of all reported algal taxa for the park. Southeast Nat 6: 99–134.
48. Lamprinou V, Danielidis DB, Economou-Amilli A, Pantazidou A (2012)
Distribution survey of Cyanobacteria in three Greek caves of Peloponnese.
Int J Speleol 41: 267–273.
49. Matuła J, Pietryka M, Richter D, Wojtun
´
B (2007) Cyanoprokaryota and algae
of Arctic terrestrial ecosystems in the Hornsund area, Spitsbergen. Pol Polar Res
28: 283–315.
50. Uher B, Kova´c
ˇ
ik L (2002) Epilithic cyanobacteria of subaerial habitats in
National Park Slovak Paradise (1998–2000). Bull Slov Bot Spoloc
ˇ
n, Bratislava
24: 25–29.
51. Alvarez-Cobelas M, Gallardo T (1988) Catalogo de las algas continentales
Espan˜olas v. Cyanophyceae Schaffner 1909. Acta Bot Malacit 13: 53–76.
52. Singh SM, Singh P, Thajuddin N (2008) Biodiversity and distribution of
cyanobacteria at Dronning Maud Land, East Antarctica. Acta Bot Malacit 33:
17–28.
53. Skinner S, Entwisle TJ (2001) Non-marine algae of Australia: 1. Survey of
colonial gelatinous blue-green macroalgae (Cyanobacteria). Telopea 9: 573–599.
54. Gutierrez-Cirlos EB, Perez-Gomez B, Krogmann DW, Gomez-Lojero C (2006)
The phycocyanin-associated rod linker proteins of the phycobilisome of
Gloeobacter violaceus PCC 7421 contain unusually located rod-capping domains.
Biochim Biophys Acta Bioenergetics 1757: 130–134.
55. Kondo K, Ochiai Y, Katayama M, Ikeuchi M (2007) The membrane-associated
CpcG2-phycobilisome in Synechocystis: A new photosystem I antenna. Plant
Physiol 144: 1200–1210.
56. Takaichi S, Maoka T, Takasaki K, Hanada S (2010) Carotenoids of
Gemmatimonas aurantiaca (Gemmatimonadetes): identification of a novel caroten-
oid, deoxyoscillol 2-rhamnoside, and proposed biosynthetic pathway of oscillol
2,2-dirhamnoside. Microbiology 156: 757–763.
57. Foss P, Skulberg OM, Kilaas L, Liaaen-Jensen S (1986) The carbohydrate
moieties bound to the carotenoids myxol and oscillol and their chemosystematic
applications. Phytochemistry 25: 1127–1132.
58. Rexroth S, Mullineaux CW, Ellinger D, Sendtko E, Rogner M, et al. (2011) The
plasma membrane of the cyanobacterium Gloeobacter violaceus contains segregated
bioenergetic domains. Plant Cell 23: 2379–2390.
59. Herrero A, Flores E (2008) The Cyanobacteria: Molecular biology, genomics
and evolution. Wymondham, Norfolk, UK: Caister Academic. 484 p.
60. Wu DY, Hugenholtz P, Mavromatis K, Pukall R, Dalin E, et al. (2009) A
phylogeny-driven genomic encyclopaedia of Bacteria and Archaea. Nature 462:
1056–1060.
61. Couradeau E, Benzerara K, Gerard E, Moreira D, Bernard S, et al. (2012) An
early-branching microbialite cyanobacterium forms intracellular carbonates.
Science 336: 459–462.
62. Cavalier-Smith T (2002) The neomuran origin of archaebacteria, the
negibacterial root of the universal tree and bacterial megaclassification.
Int J Syst Evol Microbiol 52: 7–76.
Gloeobacter violaceus Is a Common Organism
PLOS ONE | www.plosone.org 11 June 2013 | Volume 8 | Issue 6 | e66323
... The photosynthetic apparatus in organisms that operate OP consists of two protein-pigment complexes, namely photosystem I (PSI) and II (PSII), which are in the thylakoid membranes (Hohmann-Marriott & Blankenship, 2011;Shevela et al., 2023). In the cyanobacteria of the genus Gloeobacter, for example, the thylakoids are absent and the photosystems are located within the cell membrane (Rippka et al., 1974;Mareš et al., 2013). OP is inextricably linked to the appearance of several key genes and their protein products. ...
Article
Chloroplasts are the result of endosymbiosis of cyanobacterial organisms with proto‐eukaryotes. The psbA , psbD and psbO genes are present in all oxyphototrophs and encode the D1/D2 proteins of photosystem II (PSII) and PsbO, respectively. PsbO is a peripheral protein that stabilizes the O 2 ‐evolving complex in PSII. Of these genes, psbA and psbD remained in the chloroplastic genome, while psbO was transferred to the nucleus. The genomes of selected cyanobacteria, chloroplasts and cyanophages carrying psbA and psbD , respectively, were analysed. The highest density of genes and coding sequences (CDSs) was estimated for the genomes of cyanophages, cyanobacteria and chloroplasts. The synonymous mutation rate ( r S ) of psbA and psbD in chloroplasts remained almost unchanged and is lower than that of psbO . The results indicate that the decreasing genome size in chloroplasts is more similar to the genome reduction observed in contemporary endosymbiotic organisms than in streamlined genomes of free‐living cyanobacteria. The r S of atpA , which encodes the α‐subunit of ATP synthase in chloroplasts, suggests that psbA and psbD , and to a lesser extent psbO , are ancient and conservative and arose early in the evolution of oxygenic photosynthesis. The role of cyanophages in the evolution of oxyphototrophs and chloroplastic genomes is discussed.
... Their origin postdates the emergence of crown cyanobacteria. Indeed, the earliest diverging extant cyanobacterial lineage, the genus Gloeobacter, characterized by the absence of thylakoids 6,30 , is placed at the most basal position in phylogenetic reconstructions 17,31 , which suggests that this lineage may have conserved traits inherited from the ancestral lineage of cyanobacteria 32 . The recent discovery of a second thylakoid-less cyanobacterial genus, Anthocerotibacter, was used to estimate divergence with Gloeobacter around 1.4 Ga (ref. ...
Article
Full-text available
Today oxygenic photosynthesis is unique to cyanobacteria and their plastid relatives within eukaryotes. Although its origin before the Great Oxidation Event is still debated1–4, the accumulation of O2 profoundly modified the redox chemistry of the Earth and the evolution of the biosphere, including complex life. Understanding the diversification of cyanobacteria is thus crucial to grasping the coevolution of our planet and life, but their early fossil record remains ambiguous⁵. Extant cyanobacteria include the thylakoid-less Gloeobacter-like group and the remainder of cyanobacteria that acquired thylakoid membranes6,7. The timing of this divergence is indirectly estimated at between 2.7 and 2.0 billion years ago (Ga) based on molecular clocks and phylogenies8–11 and inferred from the earliest undisputed fossil record of Eoentophysalis belcherensis, a 2.018–1.854 Ga pleurocapsalean cyanobacterium preserved in silicified stromatolites12,13. Here we report the oldest direct evidence of thylakoid membranes in a parallel-to-contorted arrangement within the enigmatic cylindrical microfossils Navifusa majensis from the McDermott Formation, Tawallah Group, Australia (1.78–1.73 Ga), and in a parietal arrangement in specimens from the Grassy Bay Formation, Shaler Supergroup, Canada (1.01–0.9 Ga). This discovery extends their fossil record by at least 1.2 Ga and provides a minimum age for the divergence of thylakoid-bearing cyanobacteria at roughly 1.75 Ga. It allows the unambiguous identification of early oxygenic photosynthesizers and a new redox proxy for probing early Earth ecosystems, highlighting the importance of examining the ultrastructure of fossil cells to decipher their palaeobiology and early evolution.
... Cyanobacteria are one of the most ancient organisms [1], which arose around 3500 million years ago [2]. They are present in a wide diversity of habitats, in terrestrial and aquatic ecosystems [3][4][5], and are common inhabitants of extreme environments [5][6][7]. ...
Article
Full-text available
Coccoid cyanobacteria represent an important part of cyanobacterial freshwater diversity, with many studied strains in public databases identified as Synechococcus. This is a diverse genus, both morphologically and ecologically, with a global distribution. However, many of the so-called Synechococcus-like cyanobacteria strains could represent several independent genera that require further studies. In this work, four strains of a Synechococcus-like cyanobacteria isolated from freshwater lakes and terrestrial atmophytic habitats on São Miguel and Flores Islands (Azores archipelago) were studied genetically using the 16S rRNA and 16S–23S rRNA ITS, morphologically with light and transmission electron microscopy, and ecologically. A draft genome was produced from the reference strain by Illumina sequencing, which allowed a more complete phylogenetic study and a deeper taxonomic analysis, revealing a divergent phylogenetic evolution and low ANI and AAI values (69.4% and 66.3%, respectively) to Thermosynechococcus, the closest phylogenetic genus. Although morphologically similar to Synechococcus, the 16S rRNA and genome phylogenetic analysis placed the studied strains in a clade sister to Thermosynechococcus, inside the Thermosynechococcaceae. Thus, Pseudocalidococcus azoricus gen. sp. nov. is described as a new coccoid freshwater genus and species from the Azores archipelago. A detailed comparison with similar morphological taxa is provided, supporting the separation of the new genus. The 16S rRNA with a high genetic similarity to other strains from several continents identified as Synechococcus sp. suggests that the new genus probably has a worldwide distribution. Future studies should be performed to clarify the taxonomic identity of those strains.
... Gloeobacter violaceus PCC7421 is an early unicellular cyanobacterium that lacks the thylakoid membrane that usually contains the photosystem. It is believed that a rhodopsin-encoding gene in this compact system assists in light absorption and bioenergy production [24][25][26]. Multiple studies have shown the interaction of carotenoids with Gloeobacter rhodopsin (GR), including salinixanthin (SAL), a carotenoid from Salinibacter ruber (SAL) [27][28][29]. ...
... [5][6][7] Indeed, fossils of cyanobacteria have been reported from late Mesoproterozoic shales interpreted as non-marine, 8,9 although a marine influence has been suggested for at least some of these deposits. 10,11 In any event, Gloeobacter, the sister of all other extant cyanobacteria, lives on rock surfaces, 12 and a close relative of Gloeobacter, recently identified from metagenomic data, was also discovered in a non-marine environment. 13 Moreover, the cyanobacterial endosymbiont that gave rise to photosynthesis in eukaryotes is also inferred to have lived in a low salinity habitat. ...
Article
Full-text available
Cyanobacteria have a long evolutionary history, well documented in marine rocks. They are also abundant and diverse in terrestrial environments; however, although phylogenies suggest that the group colonized land early in its history, paleontological documentation of this remains limited. The Rhynie chert (407 Ma), our best preserved record of early terrestrial ecosystems, provides an opportunity to illuminate aspects of cyanobacterial diversity and ecology as plants began to radiate across the land surface. We used light microscopy and super-resolution confocal laser scanning microscopy to study a new population of Rhynie cyanobacteria; we also reinvestigated previously described specimens that resemble the new fossils. Our study demonstrates that all are part of a single fossil species belonging to the Hapalosiphonaceae (Nostocales). Along with other Rhynie microfossils, these remains show that the accommodation of morphologically complex cyanobacteria to terrestrial ecosystems transformed by embryophytes was well underway more than 400 million years ago.
... Despite their importance for the study of the evolution of oxygenic photosynthesis, little is known about the ecology of the early branching Gloeobacterales compared to the other Cyanobacteria [88,97,105,114]. Gloeobacter, which was for many decades the only described genus in this order, is typically found in low-light, wet rock habitats [102,115,116]. Amplicon sequencing studies have also reported 16S rRNA gene sequences loosely related to Gloeobacter spp. in Arctic [18,22] and temperate [117] soil crusts, and in Arctic [83,118] and Antarctic [119] microbial mats. ...
Article
Full-text available
Benthic microbial mats dominated by Cyanobacteria are important features of polar lakes. Although culture-independent studies have provided important insights into the diversity of polar Cyanobacteria, only a handful of genomes have been sequenced to date. Here, we applied a genome-resolved metagenomics approach to data obtained from Arctic, sub-Antarctic and Antarctic microbial mats. We recovered 37 metagenome-assembled genomes (MAGs) of Cyanobacteria representing 17 distinct species, most of which are only distantly related to genomes that have been sequenced so far. These include (i) lineages that are common in polar microbial mats such as the filamentous taxa Pseudanabaena, Leptolyngbya, Microcoleus/Tychonema and Phormidium; (ii) the less common taxa Crinalium and Chamaesiphon; (iii) an enigmatic Chroococcales lineage only distantly related to Microcystis; and (iv) an early branching lineage in the order Gloeobacterales that is distributed across the cold biosphere, for which we propose the name Candidatus Sivonenia alaskensis. Our results show that genome-resolved metagenomics is a powerful tool for expanding our understanding of the diversity of Cyanobacteria, especially in understudied remote and extreme environments.
Article
Full-text available
Molecular sequence data have transformed research on cryptogams (e.g., lichens, microalgae, fungi, and symbionts thereof) but methods are still strongly hampered by the small size and intermingled growth of the target organisms, poor cultivability and detrimental effects of their secondary metabolites. Here, we aim to showcase examples on which a modified direct PCR approach for diverse aspects of molecular work on environmental samples concerning biocrusts, biofilms, and cryptogams gives new options for the research community. Unlike traditional approaches, this methodology only requires biomass equivalent to colonies and fragments of 0.2 mm in diameter, which can be picked directly from the environmental sample, and includes a quick DNA lysis followed by a standardized PCR cycle that allows co-cycling of various organisms/target regions in the same run. We demonstrate that this modified method can (i) amplify the most widely used taxonomic gene regions and those used for applied and environmental sciences from single colonies and filaments of free-living cyanobacteria, bryophytes, fungi, and lichens, including their mycobionts, chlorobionts, and cyanobionts from both isolates and in situ material during co-cycling; (ii) act as a tool to confirm that the dominant lichen photobiont was isolated from the original sample; and (iii) optionally remove inhibitory secondary lichen substances. Our results represent examples which highlight the method’s potential for future applications covering mycology, phycology, biocrusts, and lichenology, in particular. IMPORTANCE Cyanobacteria, green algae, lichens, and other cryptogams play crucial roles in complex microbial systems such as biological soil crusts of arid biomes or biofilms in caves. Molecular investigations on environmental samples or isolates of these microorganisms are often hampered by their dense aggregation, small size, or metabolism products which complicate DNA extraction and subsequent PCRs. Our work presents various examples of how a direct DNA extraction and PCR method relying on low biomass inserts can overcome these common problems and discusses additional applications of the workflow including adaptations.
Article
Full-text available
In this study, we used microscopic, spectroscopic, and molecular analysis to characterize endolithic colonization in gypsum (selenites and white crystalline gypsum) from several sites in Sicily. Our results showed that the dominant microorganisms in these environments are cyanobacteria, including: Chroococcidiopsis sp., Gloeocapsopsis pleurocapsoides , Gloeocapsa compacta , and Nostoc sp., as well as orange pigmented green microalgae from the Stephanospherinia clade. Single cell and filament sequencing coupled with 16S rRNA amplicon metagenomic profiling provided new insights into the phylogenetic and taxonomic diversity of the endolithic cyanobacteria. These organisms form differently pigmented zones within the gypsum. Our metagenomic profiling also showed differences in the taxonomic composition of endoliths in different gypsum varieties. Raman spectroscopy revealed that carotenoids were the most common pigments present in the samples. Other pigments such as gloeocapsin and scytonemin were also detected in the near-surface areas, suggesting that they play a significant role in the biology of endoliths in this environment. These pigments can be used as biomarkers for basic taxonomic identification, especially in case of cyanobacteria. The findings of this study provide new insights into the diversity and distribution of phototrophic microorganisms and their pigments in gypsum in Southern Sicily. Furthemore, this study highlights the complex nature of endolithic ecosystems and the effects of gypsum varieties on these communities, providing additional information on the general bioreceptivity of these environments.
Article
Cyanobacteria are a lineage of Eubacteria that have long captured the attention of scientists. Approximately 5310 species of cyanobacteria have been hitherto described and new species are continually being found, named and described according to established rules. The correct determination of cyanobacteria strains concerns new biotechnological applications as well as ecological studies. There are many situations where it is crucial to recognise distinct algae species, however methods for doing so vary greatly. The aim of this review is to summarize the state of the art of the main and most recent molecular studies focusing on the phylum Cyanobacteria, with particular attention to the most frequently used gene markers. For a long time, the classification method used for cyanobacteria as well as traditionally described species was mainly based on morphology. Over time, integrative taxonomy, which involves the inclusion of many characters and comprehensive taxa sampling, has become the rule as it provides a better resolution of species relationships. For a better resolution of the phylogenies of the phylum Cyanobacteria, it is usually necessary to focus on different genetic markers: from the most common, like the 16S and 23S rRNA, ITS, rbcLXS and rpoC genes, to genes not so widely used, such as hetR, psbA, tufA, gyp and cpcBA. Also, the highly repetitive sequences often used for the symbiotic cyanobacteria represent an important factor in the inference of the phylogenetic relationships.
Article
Full-text available
Se refiere aquí un catálogo de las Cianoficeas/Cianobacteriasespañolas citadas hasta julio de 1981, las cuales ascienden a 554 táxones.
Article
Full-text available
In 72 samples collected from various types of habitats of West Spitsbergen 150 algal taxa have been identified, including 100 taxa of Cyanoprokaryota, 40 of Chlorophyceae, and 10 of Xanthophyceae. Seventy-two species, mainly blue-green algae (55 taxa) are considered as new for Svalbard flora.
Article
Full-text available
Skinner, S. and Entwisle, T.J. (Royal Botanic Gardens Sydney, Mrs Macquaries Road, Sydney NSW 2000, Australia. e-mail: tim.entwisle@rbgsyd.nsw.gov.au) 2001. Non-marine algae of Australia: 1. Survey of colonial gelatinous blue-green macroalgae (Cyanobacteria). Telopea 9(3): 573–599. Non-planktonic freshwater and terrestrial blue-green algae (Cyanobacteria) large enough to be noticeable even to casual observers, are frequently encountered in Australia, but appear only occasionally in the literature. Sixteen species of gelatinous colonial blue-green algae (Cyanobacteria) from Australia, are documented here. Two species — Nostoc borzioides and Rivularia concentrica — are new to science, and nine (Aphanothece caldariorum Richter, A. pallida (Kützing) Rabenhorst; Nostochopsis lobatus H.C. Wood ex Bornet & Flahault; Rivularia aquatica de Wilde; Gloeotrichia pilgeri Schmidle; Dichothrix gypsophila (Kützing) Bornet & Flahault; Nostoc flagelliforme Berkley & Curtis; Nostoc pruniforme (L.) C. Agardh ex Bornet & Flahault; Nostoc gelatinosum Schousboe in Bornet) are newly recorded from Australia. The others are Rivularia beccariana; Cylindrospermum licheniforme; Cylindrospermum stagnale; Nostoc commune and Nostoc verrucosum. Keys to the genera and species occurring in Australia are provided.
Article
Full-text available
Species composition of subaerial cyanobacteria in 8 gorges of National Park Slovenský raj (Prielom Hornádu, Kláštorská roklina, Kyseľ, Suchá Belá, Piecky, Sokol, Zelená dolina and Stratenský kaňon), was investigated over a period of three years with samples collected and monospecific cultures cultivated. A total of 41 genera and 102 species of cyanobacteria were determined. The coccal cyanobacteria (Chroococcales) with 54% are dominating in these subaerial calcareous habitats.
Article
The nucleotide sequence of the entire genome of a cyanobacterium Gloeobacter violaceus PCC 7421 was determined. The genome of G. violaceus was a single circular chromosome 4,659,019 bp long with an average GC content of 62%. No plasmid was detected. The chromosome comprises 4430 potential protein-encoding genes, one set of rRNA genes, 45 tRNA genes representing 44 tRNA species and genes for tmRNA, B subunit of RNase P, SRP RNA and 6Sa RNA. Forty-one percent of the potential protein-encoding genes showed sequence similarity to genes of known function, 37% to hypothetical genes, and the remaining 22% had no apparent similarity to reported genes. Comparison of the assigned gene components with those of other cyanobacteria has unveiled distinctive features of the G. violaceus genome. Genes for PsaI, PsaJ, PsaK, and PsaX for Photosystem I and PsbY, PsbZ and Psb27 for Photosystem II were missing, and those for PsaF, PsbO, PsbU, and PsbV were poorly conserved. cpcG for a rod core linker peptide for phycobilisomes and nblA related to the degradation of phycobilisomes were also missing. Potential signal peptides of the presumptive products of petJ and petE for soluble electron transfer catalysts were less conserved than the remaining portions. These observations may be related to the fact that photosynthesis in G. violaceus takes place not in thylakoid membranes but in the cytoplasmic membrane. A large number of genes for sigma factors and transcription factors in the LuxR, LysR, PadR, TetR, and MarR families could be identified, while those for major elements for circadian clock, kaiABC were not found. These differences may reflect the phylogenetic distance between G. violaceus and other cyanobacteria.
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
The mid-Precambrian microbial fossil Eosynechococcus moorei Hofmann is remarkably similar in morphology to the modern subaerial cyanophyte Gloeothece coerulea Geitler. Eosynechococcus moorei was discovered and described by Hofmann (1976) in cherts of the approximately 1.9 Ga old Belcher Island Formation, Canada, as a member of a microbial assemblage in stromatolites that were interpreted as intertidal. Gloeothece coerulea is a member of a lithophytic microbial assemblage that forms thin crusts on periodically wetted terrestrial rocks. The fossil-to-Recent comparison is based on study of the Eosynechococcus moorei type slides and fresh collections of Gloeothece coerulea from the mountains of Norway.
Article
On the basis of a comparative study of 178 strains of cyanobacteria, representative of this group of prokaryotes, revised definitions of many genera are proposed. Revisions are designed to permit the generic identification of cultures, often difficult through use of the field-based system of phycological classification. The differential characters proposed are both constant and readily determinable in cultured material. The 22 genera recognized are placed in five sections, each distinguished by a particular pattern of structure and development. Generic descriptions are accompanied by strain histories, brief accounts of strain properties, and illustrations; one or more reference strains are proposed for each genus. The collection on which this analysis was based has been deposited in the American Type Culture Collection, where strains will be listed under the generic designations proposed here.
Article
The present distribution and ecology of New Zealand plants is discussed from a historical viewpoint.It is suggested that during the Miocene a southern extension of the New Zealand archipelago supported a cool temperste flora, which gave rise to the present mountain flora after the onset of orogeny and climatic cooling in the Pliocene.As there was scarcely any simultaneous development of a distinctive flora adapted to the dry conditions which prevail to the east of the mountain axis, Cockayne§ opinion that extremely arid Pleistocene climates evoked certain characteristic life forms—notably the divaricating juvenile form of some trees—is considered to be substantially incorrect. That these life forms are adapted to still–existing conditions seems more probable.The broader features of present distribution are explllined in terms of Pleistocene glaciation and subsequent climatic amelioration.From the evidence of endemism and discontinuous distribution, it is concluded that Otago and Southland,Nelson and Marlborough, Auckland,the subantarctic regions and the Chatham Islands are areas where much of the present flora survived during the glaciation,whereas the middle portion of the South Island and the south of the North Island were characterised by extinction.Adjustment of the vegetation to post Pleistocene conditions is still incomplete, and complicated by the effect of continuing climatic fluctuations