Download Ecology of aerobic anoxygenic phototrophs in aquatic environments

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Effects of global warming on oceans wikipedia , lookup

Marine debris wikipedia , lookup

The Marine Mammal Center wikipedia , lookup

Marine pollution wikipedia , lookup

Marine habitats wikipedia , lookup

Marine biology wikipedia , lookup

Ecosystem of the North Pacific Subtropical Gyre wikipedia , lookup

Marine life wikipedia , lookup

Marine microorganism wikipedia , lookup

Transcript
FEMS Microbiology Reviews, fuv032, 39, 2015, 854–870
doi: 10.1093/femsre/fuv032
Advance Access Publication Date: 2 July 2015
Review Article
REVIEW ARTICLE
Ecology of aerobic anoxygenic phototrophs
in aquatic environments
Michal Koblı́žek1,2,∗
1
Institute of Microbiology CAS, Center Algatech, 379 81 Třeboň, Czech Republic and 2 Faculty of Science,
University of South Bohemia, Branišovská 31, 370 05 České Budějovice, Czech Republic
∗ Corresponding author: Institute of Microbiology CAS, Center Algatech, 379 81 Třeboň, Czech Republic. Tel: +420-384340432; Fax: +420-384340415;
E-mail: [email protected]
One sentence summary: The review covers current knowledge about the distribution and diversity of aerobic anoxygenic phototrophs in aquatic
environments and their role in the aquatic food webs and the microbial loop.
Editor: Corina Brussaard
ABSTRACT
Recognition of the environmental role of photoheterotrophic bacteria has been one of the main themes of aquatic
microbiology over the last 15 years. Aside from cyanobacteria and proteorhodopsin-containing bacteria, aerobic anoxygenic
phototrophic (AAP) bacteria are the third most numerous group of phototrophic prokaryotes in the ocean. This functional
group represents a diverse assembly of species which taxonomically belong to various subgroups of Alpha-, Beta- and
Gammaproteobacteria. AAP bacteria are facultative photoheterotrophs which use bacteriochlorophyll-containing reaction
centers to harvest light energy. The light-derived energy increases their bacterial growth efficiency, which provides a
competitive advantage over heterotrophic species. Thanks to their enzymatic machinery AAP bacteria are active, rapidly
growing organisms which contribute significantly to the recycling of organic matter. This chapter summarizes the current
knowledge of the ecology of AAP bacteria in aquatic environments, implying their specific role in the microbial loop.
Keywords: photoheterotrophs; microbial loop; aquatic food webs; bacteriochlorophyll; bacterial growth
INTRODUCTION
Phototrophic bacteria are one of the oldest life forms on Earth.
The ability to use light represents a major evolutionary innovation which secured a continuous supply of energy to sustain
life on this planet. While oxygenic phototrophs—cyanobacteria
and algae—provide most of the fixed carbon in the ocean,
there exists a broad phylogenetic spectrum of anoxygenic
(non-evolving oxygen) phototrophic bacteria containing bacteriochlorophyll (BChl). Anoxygenic phototrophs are divided into six
major phyla—Proteobacteria (purple photosynthetic bacteria),
Chlorobi (green sulfur bacteria), Chloroflexi (green non-sulfur
bacteria), Firmicutes (heliobacteria), Acidobacteria and Gemmatimonadetes (Fig. 1). Each of these lineages contain a unique
photosynthetic apparatus differing in the light-harvesting com-
plex architecture, pigment composition and function of their reaction centers (Hohmann-Marriott and Blankenship 2011; Zeng
et al. 2014). In addition, there are a large number of phototrophic
bacteria that contain special proteins called proteorhodopsins
(Béjà et al. 2000). The proteorhodopsin-containing bacteria are
highly abundant in both marine and freshwater environments
(Finkel, Béjà and Belkin 2013), but these organisms will not be
covered here.
Anoxygenic phototrophs probably evolved during the
Archean period some 3 Gyr ago, when the Earth’s atmosphere
was largely anoxic. Therefore, many anoxygenic phototrophs
evolved as anaerobic species, which till now grow and photosynthesize only under anoxic conditions. At the beginning
of the Proterozoic era (2.4–2.0 Gyr ago), oxygen produced
by cyanobacteria started to gradually oxygenate the Earth’s
Received: 14 January 2015; Accepted: 5 June 2015
C FEMS 2015. All rights reserved. For permissions, please e-mail: [email protected]
854
Koblı́žek
s
rmu
ter
ia
tro
Ni
flex
oid
et
es
Ba
cte
r
i
pro
s
lta
ale
De ibrion
fov
sul
eae
De
urellac
Desulf
lonproteobacteria
a
oro
ia
ter
ac
b
teo
ir
sp
Chl
eres
Fibrobact
tes
nade
ac
mo
mati
Gem
ob
The
us−
occ
noc
tin
Planctomycetes
C
Lentis hlamydiae
ph
Ver aerae
r
Ch ucomicr
lor
obia
ob
Dei
Ac
855
i
Epsi
Firmicutes
Acidobacteria
ae
s
Archaea
ae
og
ot
m
ia
te
ae
er
ter
ch
Th
iro
ria
cte
ba
teo
ro
ap
ph
Al
aproteobacteria
Beta- and Gamm
0.1
Ch
Def rysiogen
errib
e
acte tes
res
Fu
so
ba
Cy
ct
an
er
ob
ia
ac
Sp
a
teri
bac
o
f
l
u
es
des
tet
mo
s
r
i
e
g
r
Th
ne
Sy
Aquific
Figure 1. 16S rRNA phylogenetic tree of main phyla within domain Bacteria. AAP bacteria belong to phylum Proteobacteria (marked in red). The other phyla containing
chlorophototrophic species are shown in orange.
atmosphere, reaching the present concentrations approximately 500 Myr ago (Bekker et al. 2004; Holland 2006). Under the
new conditions, many anaerobic species probably disappeared
or retreated to the remaining anoxic habitats. However, some
groups of purple non-sulfur bacteria adapted to the new aerobic
conditions and gradually embarked on an aerobic life style
(Koblı́žek et al. 2013).
DISCOVERY OF AEROBIC ANOXYGENIC
PHOTOTROPHS
The first aerobic species of anoxygenic phototrophs were discovered by Tsuneo Shiba (see Fig. 1, Supporting Information)
and Keiji Harashima in the 1970s (Harashima, Shiba and Murata 1989). They isolated several pigmented aerobic strains from
the surface of the green seaweed Enteromorpha linza collected
at Aburatsubo Inlet, Miura Peninsula near Tokyo. During the
thin layer chromatography analyses of the pigment composition they noticed a blue spot, which turned out to be BChl a.
This finding came as a surprise since BChl a had, until then, only
been reported in photosynthetic bacteria grown under anaerobic
or semiaerobic conditions. One of the first isolates, the orange
strain OCh 101 (Harashima et al. 1978; Shiba, Simidu and Taga
1979a), was later formally described as Erythrobacter longus establishing the first genus containing aerobic photosynthetic bacteria (Shiba and Shimidu 1982). Another organism isolated from
E. linza was a pink bacterium originally called Erythrobacter sp.
OCh 114. This strain, later reclassified as Roseobacter denitrificans,
established (together with R. litoralis) the second new genus—
Roseobacter (Shiba 1991). Subsequently, many BChl a-containing
strains were isolated from various seaweeds, algal mats, stromatolites, sand or coastal waters (Shiba, Simidu and Taga 1979b;
Shiba et al. 1991; Nishimura et al. 1994). All these strains contained BChl a as a main light-harvesting pigment, but in contrast to purple non-sulfur photosynthetic bacteria, they were obligate aerobes, which grew and produced BChl a under fully oxic
conditions. The newly established group of phototrophic organisms was originally named aerobic photosynthetic bacteria (Harashima, Shiba and Murata 1989), but later the terminology settled on the term aerobic anoxygenic phototrophic (AAP) bacteria (Yurkov and Beatty 1998a) or aerobic anoxygenic phototrophs
(Yurkov and Csotonyi 2009).
During the last two decades of the 20th century, AAP bacteria remained a domain of classical microbiologists who kept describing novel AAP genera from various aquatic habitats (Fig. 2,
Supporting Information). The first freshwater AAP species were
isolated by Yurkov and Gorlenko from the surface of alkaline
cyanobacterial mats growing in and around an alkaline thermal
spring near Lake Baikal (Yurkov and Gorlenko 1990, 1992; Yurkov
et al. 1994). Later, acidophilic species of the genus Acidobacterium containing a zinc form of BChl a were isolated from mine
drainage water (Wichlacz, Unz and Langworthy 1986; Wakao
et al. 1993; Hiraishi and Shimada 2001). Several Roseobacterrelated AAP species were isolated from saline lake Clifton in
856
FEMS Microbiology Reviews, 2015, Vol. 39, No. 6
Figure 2. Colorful cultures of AAP bacteria. From left to right: Roseobacter litoralis, Roseococcus thiosulfatophilus, Erythrobacter longus and Sphingomonas sp. AAP2.
UTILIZATION OF LIGHT ENERGY
One of the issues discussed from the very beginning was the
photosynthetic competence of newly isolated organisms. All
AAP species contain photosynthetic complexes with BChl a as
the main light-harvesting pigment. However, their BChl a content is significantly lower when compared to purple non-sulfur
bacteria (Harashima et al. 1980; Shiba 1987; Biebl and WagnerDöbler 2006; Koblı́žek et al. 2010; Fig. 3, Supporting Information).
Interestingly, BChl a synthesis in most AAP species is inhibited
by light (Shiba 1987; Iba and Takamiya 1989). The reason for this
physiological feature is probably the need to avoid the simultaneous presence of light, BChl a biosynthesis intermediates and
oxygen, which could lead to the production of harmful reactive
oxygen species (Nishimura et al. 1996).
The visible pigmentation of AAP species is usually orange,
red, pink, brown or yellow due to the presence of various
carotenoids (see Fig. 2). The carotenoid composition differs
among various AAP groups (Yurkov and Csotonyi 2009; Zheng
et al. 2013). Carotenoids serve as auxiliary pigments which extend absorption to the blue-green part of the spectrum (see
Fig. 3). This is important in aquatic habitats, where the UV-A
and near infrared part of the spectrum available for BChl a are
rapidly absorbed by the water and only the blue-green part of
the solar spectrum penetrates into the deeper parts of the water column. The excitation energy captured by the carotenoids
is transferred within picoseconds to the BChl a molecules (Šlouf
et al. 2013) and used for primary charge separation in the reaction center. In addition, some AAP species (i.e. genera belonging to Rhodospirillales and Sphingomonadales) contain large
amounts of carotenoids, which are not bound to photosynthetic
complexes (Takaichi et al. 1991; Yurkov, Gad’on and Drews 1993).
Erythrobacter sp. NAP1
Dinoroseobacter shibae
Congregibacter litoralis
Absorbance [rel. units]
West Australia (Shiba et al. 1991) and lake Ekho in Antarctica
(Labrenz et al. 1999, 2000). The first AAP bacterium belonging to
the Betaproteobacteria Roseateles depolymerans was isolated by
Suyama et al. (1999) from Hanamuro River in Japan. The classical
culture work on AAP bacteria has been excellently reviewed in
the following references (Yurkov and Beatty 1998a; Yurkov and
Csotonyi 2009; Yurkov and Hughes 2013).
300
U.V.
400
500
600
700
800
900
N.I.R
Wavelength [nm]
Figure 3. In vivo absorption spectra of AAP species recorded using Shimadzu
UV3000. The cells were resuspended in 70% glycerol to reduce scattering. Notice
the presence of peripheral light-harvesting complexes in D. shibae absorbing at
804 nm and the large amount of non-photosynthetic carotenoids in Erythrobacter
sp. NAP1.
These carotenoids do not have any light-harvesting function and
probably only serve for photoprotection.
The functionality of the photosynthetic reaction centers in
AAP bacteria has been demonstrated by spectroscopic (Harashima et al. 1987; Stadnichuk et al. 2009; Rathgeber et al. 2012)
and IR kinetic fluorescence measurements (Kolber et al. 2001;
Sato-Takabe, Hamasaki and Suzuki 2012, 2014). On the other
hand, AAP bacteria are unable to grow photoautotrophically as
they lack carbon fixation pathways (Fuchs et al. 2007; Swingley
et al. 2007; Koblı́žek et al. 2011). (The weak light-mediated CO2
incorporation observed in some AAP species (Shiba 1984; Kishimoto 1995; Koblı́žek et al. 2003) originates from anaplerotic carboxylation reactions connected to the citric acid cycle (Tang et al.
2009). The anaplerotic carboxylation may contribute 0.6–11% of
Koblı́žek
total cellular carbon (Hauruseu and Koblı́žek 2012). The obligatory requirement for organic substrates and the ability to grow
in the dark (Shioi 1986; Yurkov et al. 1999; Koblı́žek et al. 2003;
Spring et al. 2013) indicate that AAP bacteria utilize light energy
as an additional source of energy for their mostly heterotrophic
metabolism (Harashima et al. 1987). Light exposure of AAP cells
inhibits respiration (Harashima et al. 1987; Koblı́žek et al. 2010)
and increases their cellular ATP concentration (Okamura et al.
1986; Candela, Zaccherini and Zannoni 2001). This documents
that AAP bacteria are able to replace a large part of oxidative
phosphorylation with photophosphorylation. The light-derived
energy may cover a large portion of cellular metabolic needs. It
has been demonstrated that light exposure increases the survival of AAP cells under starvation (Shiba 1984; Soora and Cypionka 2013). Utilization of light energy also saves organic substrates, which would otherwise have to be respired. Indeed,
it has been repeatedly shown that exposure of AAP bacteria
to light increases biomass yields (Shioi 1986; Yurkov and van
Gemerden 1993; Biebl and Wagner-Döbler 2006; Spring et al.
2009). Experiments using chemostat cultures of Erythrobacter
sp. NAP1 and Roseobacter sp. COL2P grown on defined carbon
sources show that under light–dark cycles these bacteria accumulate 25–110% more carbon when compared to cultures grown
in the dark (Hauruseu and Koblı́žek 2012). In summary, AAP bacteria can be classified as facultative photoheterotrophs—they
can grow in the dark on organic carbon substrates; on the other
hand, they are able to derive a significant portion of their energy
requirements from light (Hauruseu and Koblı́žek 2012; Kirchman
and Hanson 2013).
DISCOVERY OF AAP BACTERIA IN THE OCEAN
In spite of the accumulating knowledge regarding the diversity and physiology of AAP bacteria, nothing was known about
the distribution or the role of AAP bacteria in the natural environment. This situation changed dramatically in 1998, when
Vladimir V. Yurkov and his colleagues reported the isolation
of a novel AAP bacterium Citromicrobium bathyomarinum. This
organism was collected from black smoker plume waters of
the Juan de Fuca Ridge (Yurkov and Beatty 1998b; Yurkov
et al. 1999). It was speculated that hydrothermal vents emanating infrared radiation and supplying an ample amount of
reduced metals or sulfides may have represented an environment where the first anoxygenic phototrophs originally evolved
(Nisbet, Cann and Van Dover 1995). This news caught the attention of marine biologists Falkowski and Kolber who then
decided to explore the presence of novel species directly in
the vicinity of the hydrothermal vents. They modified their
fast repetition rate fluorometer, originally developed for phytoplankton measurements, extending its sensitivity to the infrared region. The measurements were conducted during the
cruise exploring the hydrothermal vents at East Pacific Rise
in the spring of 1999 and the samples were collected using
the deep-ocean submersible Alvin (DSV-2). Unfortunately, the
new instrument failed to record any BChl a signals in the water samples collected around the vents. After several futile attempts, Kolber decided to test the instrument with samples
collected from the surface. Unexpectedly, the surface samples
revealed weak infrared fluorescence transients, signaling the
presence of BChl-containing organisms. The infrared fluorescence was then recorded along a more than 1000 km long
transect, demonstrating that the observed signals were distinct
from phytoplankton emission and that they indeed originated
857
from AAP bacteria (Kolber et al. 2000). The article of Kolber
et al. (2000) together with a study describing proteorhodopsincontaining bacteria (Béjà et al. 2000) immediately attracted the
attention of marine microbiologists. (Interestingly, by coincidence these articles describing the two main groups of marine
photoheterotrophs were published almost in the same moment.
The article of Kolber et al. published in Nature on 14 September
2000 preceeded the study of Béjà et al. (2000) in Science just by 1
day.) The reports of large amounts of photoheterotrophic organisms in the ocean (Eiler 2006; Béjà and Suzuki 2008; Zubkov 2009)
challenged the classical view of marine bacteria as heterotrophic
organisms fully dependent on recycling dissolved organic matter (DOM) produced by photoautotrophic phytoplankton.
A controversy surrounded the first reports regarding AAP
abundance. In the follow-up study using infrared epifluorescence microscopy, Kolber et al. (2001) found that AAP bacteria
made up 11% of total bacteria in the North East Pacific. These
reports were questioned by Schwalbach and Fuhrman (2005),
who reported that the fraction of AAP bacteria in the San Pedro Channel in California was much lower, on average 1.66%
of total bacteria. Using an alternative quantification by qPCR,
the same authors found that AAP bacteria may represent 0.01–
18% of total bacteria in various marine environments (Schwalbach and Fuhrman 2005). Similar criticism was raised by Goericke (2002) who based on the HPLC analyses of BChl a concentration off the shores of California (2–41 ng L−1 ) asserted that
anoxygenic photosynthesis in the ocean may not be significant.
Following investigations calmed down the initial issues documenting that, despite large differences, AAP bacteria on average
constitute typically 1–7% of total prokaryotes in the oligotrophic
areas of the oceans. In shelf seas or river estuaries, AAP bacteria may represent 2–15% of total prokaryotes (see Table 1). These
numbers were also confirmed by an independent qPCR enumeration (Du et al. 2006) as well as by an analysis of Venter’s metagenomic dataset (Global Ocean Sampling Expedition I), which documented that AAP bacteria made up 1–10% of total prokaryotes
in the surface ocean (Yutin et al. 2007).
DIVERSITY OF MARINE AAP BACTERIA
During the first two decades of AAP research, all information
on diversity of marine AAP bacteria was gained from culturedependent studies. The first of Shiba’s isolates, strains OCh 101
(Shiba and Shimidu 1982) and OCh 114 (Shiba 1991), established
the first two AAP genera Erythrobacter (class Alphaproteobacteria, order Sphingomonadales) and Roseobacter (class Alphaproteobacteria, order Rhodobacterales). Continuing cultivation efforts led to the description of a number of AAP genera belonging
to the Roseobacter clade which represents a large environmentally important subgroup of marine bacteria within the order
Rhodobacterales (Wagner-Döbler and Biebl 2006; Luo and Moran
2014; Pujalte et al. 2014).
Despite the growing number of AAP taxa, culture-dependent
studies cannot provide reliable information on species composition and diversity in environmental samples. This problem was
overcome with the application of culture-independent molecular techniques. The majority of environmental studies employ
the pufM gene (encoding the M subunit of the bacterial reaction
center) as a convenient marker for anoxygenic phototrophs harboring type-2 reaction centers. (The first PCR primers for pufM
genes were designed by Nagashima et al. (1997) and Achenbach,
Carey and Madigan (2001). Most of the later studies use various
modifications of these two primer sets (see Table S1, Supporting
858
FEMS Microbiology Reviews, 2015, Vol. 39, No. 6
Table 1. AAP abundance determined in various marine environments using infrared epifluorescence microscopy.
Environment
Period
North East Pacific
San Pedro Channel
NW Atlantic
Sargasso Sea
Mid Atlantic Bight
North Pacific Gyre
Baltic Sea∗
South Pacific
Sargasso Sea
East China Sea∗
Shelf seas∗
Open Oceans∗ ∗ ∗
North Atlantic
Baltic Sea
Mediterranean Sea
Chesapeake Bay
Mediterranean Sea
July 2000
Apr–Sept 2002
Oct 2001, Mar 2002
Oct 2001, Mar 2002
Aug 2003
Feb 2004
Sept 2004–Oct 2005
Oct–Dec 2004
May 2006
Apr 2002–Mar 2003
2003–05
2003, 2005–06
May–July 2005
July 2006
Spring 2007
June 2006
Sept 2005, June and Sept
2007, May 2009
June 2008
Summer 2009
2010–11
Mediterranean Sea
Western Arctic Ocean
Blanes Bay∗
AAP
abundance
103 cells mL−1
Fraction of
total
Prokaryotes
Samples
References
15–90
n.a.
13–66
7–15
7–150
0.5–67
0–383
5–194
6–19
2.2–79
n.a.
n.a.
13–120
8–310
4.4–65
200–800
2.7–87.7
11.3 ± 1.7%
1.66 ± 0.55%
0.8–9.4%
0.8–2.6%
0.8–18%
0–7%
0–11.6%∗ ∗
0.3–24.4%
1.9–4.3%
0.5–11.6%
4.46 ± 2.41%
1.52 ± 1.30%
1.5–12.9%
1.0–11.1%
1.4–6.5%
2–12%
0.6–11.1%
30
19
16
10
22
30
71
27
12
∼80
147
135
24
55
18
8
72
Kolber et al. (2001)
Schwalbach and Fuhrman (2005)
Sieracki et al. (2006)
Sieracki et al. (2006)
Cottrell, Mannino and Kirchman (2006)
Cottrell, Mannino and Kirchman (2006)
Mašı́n et al. (2006)
Lami et al. (2007)
Koblı́žek et al. (2007)
Zhang and Jiao (2007)
Jiao et al. (2007)
Jiao et al. (2007)
Michelou, Cottrell and Kirchman (2007)
Salka et al. (2008)
Lami et al. (2009)
Cottrell, Ras and Kirchman (2010)
Hojerová et al. (2011)
0.3–35
n.a.
1.12–50.2
0.1–4%
0.1–14.8%
0.2–6.3%
60
n.a.
16
Lamy et al. (2011b)
Boeuf et al. (2013)
Ferrera et al. (2014)
∗
Seasonal studies.
Whole year average 2.37 ± 2.50% (mean ± st.dev.).
∗∗∗
Refers to surface collected open ocean samples. AAP percentage averages in individual oceans were: 3.79 ± 1.72% for the Indian, 1.57 ± 0.68% for the Atlantic and
1.08 ± 0.74% for the Pacific.
∗∗
Information). The first analyses of AAP composition in marine
environments were conducted by Béjà et al. (2002) using bacterial
artificial chromosome (BAC) libraries constructed from environmental DNA collected from Monterey Bay in California. One of
the fosmids (BAC 60D04) contained a complete set of genes encoding bacterial photosynthesis, documenting the presence of
Roseobacter-related organisms. Surprisingly, two other fosmids
(BAC65D09 and BAC29C02) contained completely novel pufM sequences only distantly related to at that time known organisms. The new sequences were first tentatively ascribed to phototrophic Beta- or Gammaproteobacteria (Béjà et al. 2002). The
ambiguous phylogenetic assignment was later clarified after
successful isolation of several gammaproteobacterial strains belonging to the NOR5/OM6 cluster, which contained photosynthetic genes related to the originally retrieved sequences (Cho
et al. 2007; Fuchs et al. 2007).
Some uncertainty remains regarding the composition of
AAP communities in various marine habitats. The first pufM
clone libraries constructed from Mediterranean and Red Sea waters indicated that the local AAP communities are dominated
by members of the Roseobacter clade (Oz et al. 2005) with a
smaller proportion of Gammaproteobacteria (Yutin, Suzuki and
Béjà 2005). Analysis of Craig Venter’s Global Ocean Sampling
metagenomic data indicated that the AAP community in the
Atlantic and Pacific Oceans were formed mostly of Roseobacterrelated species and only a small amount of Gammaproteobacteria species (Yutin et al. 2007). Similar results were also obtained
in the Baltic Sea, where the local community was dominated
by Roseobacter-related pufM sequences with a small contribution of Sphingomonadales and Betaproteobacteria (Salka et al.
2008). In contrast, a combination of IR epifluorescence and FISH
analyses documented the majority of AAP cells in the Baltic Sea
were Gammaproteobacteria (Mašı́n et al. 2006). The larger pro-
portion of Gammaproteobacteria was also indicated from pufM
analyses conducted in the East China Sea and the North Pacific Gyre where more than one half of the sequences were related to Gammaproteobacteria (Hu et al. 2006). Similarly, analyses conducted along a large Mediterranean Sea transect documented that up to 80% of the obtained pufM sequences came
from Gammaproteobacteria (Lehours et al. 2010; Jeanthon et al.
2011).
The situation in coastal habitats is much more variable.
The study conducted in Delaware Estuary documented that
the composition of the AAP community was very diverse with
its composition changing along the salinity gradient (Waidner and Kirchman 2008). The upper part of the estuary with
a salinity of <5 was dominated by Rhodoferax-related AAP
species (Betaproteobacteria), while the lower part of the bay with
higher salinity was dominated by Gammaproteobacteria. pufM
sequences related to Rhodobacterales were distributed throughout the estuary exhibiting some minor seasonal changes (Waidner and Kirchman 2008). A somewhat different situation was
observed at the coastal observatory in the Blanes Bay (western
Mediterranean Sea). The pyrosequencing analysis of seasonal
changes of AAP communities documented that gammaproteobacterial sequences dominated throughout the year except
in spring when Roseobacter-related sequences prevailed (Ferrera
et al. 2014).
Less information is available from polar regions. Diversity of
AAP communities in the Chukchi and the Beaufort Sea (western Arctic Ocean) coastal waters seemed to be mostly influenced
by riverine inputs (Boeuf et al. 2014). They were mostly composed of Betaproteobacteria and Rhodobacterales; Gammaproteobacteria were not registered (Cottrell and Kirchman 2009).
Similar results were also reported by Boeuf et al. (2013) who also
registered smaller amount of Sphingomonadales. A somewhat
Koblı́žek
different composition was also found in sea ice and seawater
AAP communities in the Ross Sea near Antarctica, which consisted of various members of order Rhodobacterales whereas Betaproteobacteria were absent (Koh, Phua and Ryan 2011).
AAP BACTERIA IN THE INLAND WATERS
Freshwaters
In contrast to the accumulating information regarding marine AAP bacteria, their freshwater counterparts have received
less attention. The first freshwater AAP species Erythrobacter
(now Sandaracinobacter) sibiricus and Erythromicrobium (now Erythromonas) ursincola were isolated by Yurkov and Gorlenko (1990,
1992) from the surface of alkaline cyanobacterial mats growing
in and alkaline thermal spring nearby Lake Baikal. Later, Porphyrobacter neustonensis was cultured from neuston layers of freshwater lakes in Australia (Fuerst et al. 1993), and Sandarakinorhabdus limnophila was retrieved from the mountain lake Starnberger
See in Germany (Gich and Overmann 2006). Similarly to the first
marine species, all these isolates belonged to the order Sphingomonadales within Alphaproteobacteria. A completely new
AAP species Roseateles depolymerans belonging to Betaproteobacteria was isolated by Suyama et al. (1999) from Hanamuro River
in Japan when screening for poly(hexamethylene) carbonatedegrading bacteria. Later, another phototrophic betaproteobacterium HTCC528 was isolated from an oligotrophic Crater lake in
Oregon; unfortunately, its physiology has not been tested (Page,
Connon and Giovannoni 2004).
Culture-independent approaches were first applied by Waidner and Kirchman (2005) using DNA collected from the Delaware
River, USA. In their fosmid library, they identified two fragments of bacterial photosynthesis gene clusters. While one originated from a phototrophic betaproteobacterium, the second
one was related to the purple non-sulfur bacterium Rhodobacter which documents that phototrophic communities in freshwater systems may be composed of both AAP bacteria and purple non-sulfur species even in oxic environments. The common
occurence of AAP bacteria in Swedish lakes was confirmed using bchL clone libraries targeting both oxygenic and anoxygenic
species (Eiler et al. 2009). Photosynthesis genes indicating the
presence of AAP species were also found in several single-cell
genomes collected from freshwater lakes in Maine and Wisconsin (Martinez-Garcia et al. 2011).
The systematic survey of AAP abundance in various limnic
systems in central Europe revealed that these organisms during
summer made up 2–21% of total bacteria in many oligotrophic
and mesotrophic lakes (Mašı́n et al. 2008, 2012). Similar numbers (3–22%) were also found in mountain lakes in Tyrol, Austria
(Čuperová et al. 2013), Ebro river (5–14%) in Spain (Ruiz-González
et al. 2013) or temperate and boreal lakes (1–37%) in Canada
(Fauteux et al. 2015). Interestingly, the highest abundances were
found in the acidified mountain lakes Plešné and Čertovo in the
Šumava Mountains (Czech Rep., 1090 and 1028 m a.s.l.) where
AAP bacteria formed, during the summer maximum, more than
half of the bacterial biomass (Mašı́n et al. 2008; see Fig. 7). It has
been repreatedly found that in temperate freshwater habitats
AAP bacterial contributions undergo large seasonal changes oscillating from almost nil in the late winter period to up to 20%
during the summer or early autumn maxima (Čuperová et al.
2013; Lew et al. 2015).
A large survey of AAP diversity was conducted in eight lakes
in the Meckleburg lake district (Germany) by Salka et al. (2011).
This analysis documented that most of the pufM sequences in-
859
deed originated from clades which contain AAP species, despite the fact that four lakes also contained smaller fractions
of Rhodobacter-related sequences (purple non-sulfur bacteria).
The majority of the obtained sequences was related to Betaproteobacterium Rhodoferax fermentans, with a smaller contribution
of various subgroups of Alphaproteobacteria (Salka et al. 2011,
Fig. 4). AAP bacteria were also surveyed in several mountain
lakes in the Tyrolean Alps. A large difference in both abundance, morphology and diversity was observed between lakes
located above and below the treeline. While AAP populations
inhabiting the lakes at the highest elevation were composed of
large rods probably members of Sphingomonadales (Alphaproteobacteria), the lakes under the treeline were inhabited by a
variety of smaller morphotypes representing both Alpha- and
Betaproteobacteria (Čuperová et al. 2013; Fig. 5). In contrast,
a recent study of ultraoligotrophic cold high mountain lakes
in Central Pyrenees, Spain, documented that local AAP communities were mostly composed of Betaproteobacteria. Unlike
in the Alps, conductivity, pH and nitrate concentration were
the main factors influencing the AAP community composition
(Caliz and Casamayor 2014). The common theme of all these
above-mentioned analyses is the widespread occurence of phototrophic Betaproteobacteria. The recent report of the presence
of photosynthesis genes in the genomes of two Limnohabitans
species (Zeng et al. 2012) signalizes that the obtained environmental sequences may have originated from similar organisms.
Limnohabitans species are a common component of freshwater
bacterioplankton in temperate regions, but till recently it was
considered to contain only heterotrophic species (Hahn et al.
2010).
Saline lakes
The presence of AAP bacteria in brackish and saline lakes was
first demonstrated by culture-dependent work in Clifton and
Heyward lakes, Australia (Shiba et al. 1991); Ekho Lake, Antarctica (Labrenz et al. 1999); Mahoney Lake, Canada (Yurkova et al.
2002); and East German Creek system (Csotonyi et al. 2008).
AAP bacteria were reported from Mediterranean coastal lagoons
where they made up 0.1–15% of total bacteria (Lamy et al. 2011a).
Similar numbers were also found in high altitude saline lakes
in Tibet (pH > 9), where AAP bacteria represented between 0.5
and 9% of total prokaryotes (Jiang et al. 2009). The conducted
pufM analyses documented that local AAP communities have a
similar composition as found earlier in marine environments.
Lakes Gahai and Qinghai with higher salinity were dominated by
Roseobacter-related species, whereas the brakish lake Erhai was
dominated by Gammaproteobacteria (Jiang et al. 2009).
One of the main problems of ecological studies on inland
lakes is that available IR epifluorescence or pigment analysis
techniques are not able to distinguish between various groups
of BChl a-containing bacteria, such as purple sulfur, purple nonsulfur bacteria and AAP bacteria. In larger lakes where the oxic
epilimnion and anoxic hypolimnion and sediment are physically
well separated, it is relatively safe to assume that observed BChl
a positive cells are predominantly AAP species, since standard
purple bacteria synthesize only minimum amounts of BChl a
under aerobic conditions. Shallow, dynamic water bodies such
as coastal lagoons (Lamy et al. 2011a) or soda lakes (Medová
et al. 2011) represent a much more challenging situation, where
anaerobic species may be mixed up from soil or sediment. Here,
the separation of different functional groups is impossible without parallel genetic analyses (Tank, Blümel and Imhoff 2011).
On the other hand, in many ecological studies, the separation
860
FEMS Microbiology Reviews, 2015, Vol. 39, No. 6
α
vim
ans
ium radiotoler
Methylobacter
m
rubru um
hilium crypt
m
Acidip
u
i
l
i1
iphi
TA
Acid
mB
biu
izo
gi
ela yrh
a p Brad
Ful
arin
dus
rhab
kino
dara
Ci
San
● ●
■
■
92
83
97
■
91
●
■
99
■
99
■
99
■
■■
59
▲
M
Ro eth
■
se ylo
ate v
les ers
■
a
de tili
po s u
lym ni
■
era ver
ns sali
s
Nev
skia
■ ■98
ram
osa
■
99
■
▲
78
β
limn
op h
ila
tro
9
9
mi
Bl crob
as iu
to m
E
m
on bath ryth
yo rob
as
ur mar acte
sin
inu r l
co
m o ng
la
us
0.2
62
■
59
99
53
52
76
51
79
9
9
cl G a m m a p
on
roteob
a
cteriu
e
m HT
eB
AC
env. clone eB CC2080
AC 65D09 99
re
59
99
99
99
95
▲
▲▲
0
Rhodobacterales sp. BS9
e
a
ib
h
rs
bacte
oseo
Dinor
99
99
05
var
iu
deni s TM1
trific 03
ans 5
Roseiba
cterium
elongatu
m
cter
Ros
eoba
seo
Ro
v.
d2
5D
99
02
29C
AC
eB
one
en
▲
69
75
47
sp
.R
im
28
90
. cl
75
env
ns
■
▲
▲
Mediterranean Sea
Coastal lagoon
■ Mecklenburg lake district
● Alpine lakes
▲
4
D0
60
AC
S2
is
rii
ica
us
CC rmis
eB
ens
ctic
nd ph
e
ter
ifo antar estfold
xa otro
lon bac gutt
v
s
ale hot
. c eo ter alinu nella
zia a p
nv Ros bac seis Lokta
e
en fle
lfito Ro
99
br oe
Su
La H
99
De
lRiverF
L no
os06H
■ imno habita n
03
ss
ha
p. R
bit
im
a
93
γ
■
■
■en
v. clon
■■Lim
e
65
Congregibacter litoralis ▲
▲ 98
▲
9
bra is 9
a ru tens▲
e
i
l
Ha s syl
hilu
inip
m
Lu
62 5
7
81
67
71
■
■
50
β
●
P
▲ lan
kt
om
ar
in
a
te
m
pe
ra
ta
▲
Roseobacter clade
Figure 4. Molecular phylogenetic analysis of partial pufM aminoacid sequences (∼220 aminoacid). The sequences for Mediterranean Sea were obtained from Oz et al.
(2005), the Baltic Sea coastal lagoon from Tank, Blümel and Imhoff (2011), freshwater lakes in northen Germany from Salka et al. (2011) and from Alpine lakes in
Austria from Čuperová et al. (2013). In addition, the tree includes environmental pufM sequences identified in fosmid libraries published by Béjà et al. (2002), Waidner
and Kirchman (2005) and Oz et al. (2005). The phylogeny was inferred by using the maximum likelihood method using the LG evolution model. The tree with the highest
log likelihood (−11297.4615) is shown. A discrete Gamma distribution was used to model evolutionary rate differences among sites (five categories (+G, parameter =
0.8231)). The rate variation model allowed for some sites to be evolutionarily invariable ([+I], 11.7036% sites). The tree is drawn to scale, with branch lengths measured
in the number of substitutions per site. All positions containing gaps and missing data were eliminated. Only bootstrap values higher than 50 are shown.
Koblı́žek
861
Figure 5. Various AAP morphotypes observed in fresh waters (Čuperová et al. 2013).
of functional groups may not be necessary, and one may report
the observed organisms simply as purple phototrophic bacteria
or anoxygenic phototrophs without specifying their physiology.
NEW DEFINITION OF AAP BACTERIA
The original definition of AAP bacteria as organisms was based
on the first obtained isolates namely Erythrobacter and Roseobacter species. The AAP bacteria are defined as aerobic species
which synthesize BChl a and conduct photosynthesis under aerobic conditions. The increasing knowledge on the ecology, physiology and genomic of these organisms urges for an update of
the current definition.
In many freshwater habitats, we encounter species belonging to phototrophic methylotrophs (i.e. genus Methylobacterium)
or aerobic phototrophic rhizobia (i.e. genus Bradyrhizobium).
These organisms are typically found in soils, associated with
vegetation (Atamna-Ismaeel et al. 2012) or plant roots (Sato 1978;
Fleischman and Kramer 1998). The relationship of these two
groups to AAP bacteria is a matter of dispute. While Shiba and
Harashima included chapters about BChl a-containing facultative methylotrophs in their book ‘Aerobic Photosynthetic Bacteria’ (Harashima, Shiba and Murata 1989), Yurkov and Csotonyi
(2009) excluded these two subgroups from AAP bacteria. In spite
of some subtle differences, both phototrophic methylotrophs
and phototrophic rhizobia share the same basic physiology
with standard AAP bacteria (Fleischman and Kramer 1998). Both
groups are mostly aerobic species, which synthetize BChl a and
perform photosynthesis under aerobic conditions. Moreover,
laboratory experiments with Bradyrhizobium sp. BTAi showed
that photosynthetic electron transport in this organism is oxygen dependent, (similarly to Roseobacter denitrificans) and cannot
be conducted under anaerobic conditions (Kramer, Kanazawa
and Fleischman 1997).
Another source of uncertainty are species which contain
photosynthetic genes, but express no or only a minimum
amount of BChl a under laboratory conditions. Many such
species were originally considered heterotrophic, and the presence of photosynthesis genes was identified only after full
genome sequencing. A typical example of such an organism is
Planktomarina temperata representing an abundant RCA lineage
of the Roseobacter cluster (Selje, Simon and Brinkhoff 2004).
This bacterium contains functional photosynthetic genes but
expresses only miniscule amounts of the pigment under laboratory conditions (Giebel et al. 2013). However, recent metatranscriptomic analysis documented that these organisms express its photosynthetic genes under in situ condition (Voget
et al. 2015). A similar situation was found among the representatives of the freshwater genus Limnohabitans, which was originally established for heterotrophic species. While pigment expression has not been observed under laboratory conditions, the
full genome sequencing documented the presence of the complete photosynthesis gene clusters among two Limnohabitans sp.
strains (Zeng et al. 2012). An example of another cryptic phototroph is the common neuston bacterium Nevskia ramosa. This
organism was already described in the 19th century (Famintzin
et al. 1892), but the presence of photosynthetic genes was revealed only after its full genome sequencing in 2013. These organisms do not fit the strict definition of an AAP bacterium since
they do not synthesize BChl a under tested laboratory conditions. They presumably do express photosynthetic apparatus
under some specific environmental in situ conditions, but the information on their pigment expression under natural conditions
is so far missing.
FEMS Microbiology Reviews, 2015, Vol. 39, No. 6
Facing all these uncertainties, I suggest to expand the term
AAP bacteria to all mostly aerobic species, which express (at
least facultatively) purple bacterial photosynthetic reaction centers under fully oxic in situ conditions. An alternative would
be the term aerobic BChl-containing bacteria (introduced by JF
Imhoff), which may serve as an umbrella group for AAP bacteria
(sensu stricto), phototrophic methylotrophs, phototrophic rhizobia and all the other mostly aerobic photoheterotrophs without
indicating any specific physiology, ecology or taxonomical classification.
5
3
2
1
INFLUENCE OF THE MAIN ENVIRONMENTAL
FACTORS
0
0.01
0.1
1
10
100
-3
Chlorophyll a [mg m ]
12
AAP bacteria [%]
The ecology of AAP bacteria is not completely understood. In his
pioneering paper, Kolber et al. (2000) speculated that the ability to utilize light may be especially beneficial in nutrient-poor
marine environments. The hypothesis was repeatedly disproved
based on the data documenting that AAP bacteria are more numerous in more productive regions (Schwalbach and Fuhrman
2005; Sieracki et al. 2006; Jiao et al. 2007; Lamy et al. 2011b). In
part, the controversy was caused by the different backgrounds of
the individual research groups. Phytoplankton ecologist Kolber
routinely used normalization per chlorophyll which resulted in
higher bacteriochlorophyll a/chlorophyll a ratios in oligotrophic
areas (Kolber et al. 2000) and smaller ratios in meso and eutrophic regions. The same trends were observed on the pigment
data from the Black and Baltic Seas (Koblı́žek et al. 2005; Koblı́žek,
Falkowski and Kolber 2006; see Fig. 6 upper panel) and the global
dataset by Jiao, Zhang and Hong (2010). In contrast, normalization of the AAP abundance per total bacteria (DAPI) used by microbiologists frequently revealed higher proportions of AAP bacteria in more productive marine regions or coastal waters (see
Table 1 and Fig. 6 lower panel) documenting that AAP bacteria
prefer more productive habitats.
Most of the surveys of AAP bacteria in marine environments
have shown a positive correlation between AAP abundance and
chlorophyll concentration (Jiao et al. 2007; Hojerová et al. 2011;
Ritchie and Johnson 2012). A similar relationship was also documented for saline (Medová et al. 2011) and freshwater lakes
(Mašı́n et al. 2012). However, it is not clear whether the relationship reflects the direct association with primary producers or the
same dependence on limiting nutrients such as phosphate or nitrogen. Indeed, the common relationship of AAP abundance, total bacteria, chlorophyll and total phosphorus was documented
in freshwater lakes (Mašı́n et al. 2012; Fauteux et al. 2015).
Surprisingly, the most enigmatic factor in the AAP ecology is
the influence of light. While the positive influence of light has
been repeatedly demonstrated in laboratory experiments, its effect on natural AAP communities remains inconclusive. A DNA
fingerprinting study indicated a strong influence of light over a
large part of the marine bacterial community in the North Pacific
Ocean (Van Mooy, Devol and Keil 2004). In contrast, experiments
conducted in San Pedro Channel (California) did not show any
effect of light on bacterioplankton structure (Schwalbach, Brown
and Fuhrman 2005). Ambiguous results were also obtained in
experiments directly analysing natural AAP assemblages. While
AAP bacteria in Chesapeake Bay had higher leucine incorporation activities when compared to average bacteria, no effect was
found between light and dark treatments (Stegman, Cottrell and
Kirchman 2014). Also, in our experiments conducted in freshwater lakes we failed to observe any clear stimulation of AAP
growth by light. It seems that the effect of light on natural AAP
Atlantic Ocean
Baltic Sea
Black Sea
Mediterranean Sea
4
BChl a/Chl a ratio [%]
862
9
6
3
0
0.01
0.1
1
10
100
Chlorophyll a [mg m -3 ]
Figure 6. Upper panel: bacteriochlorophyll a/chlorophyll a ratios determined in
the Baltic Sea (Koblı́žek et al. 2005), Black Sea (Koblı́žek, Falkowski and Kolber
2006), Mediterranean Sea (Koblı́žek, Falkowski and Kolber 2006; Lami et al. 2009;
Hojerová et al. 2011) and the Atlantic Ocean (Koblı́žek et al. 2007). Bacteriochlorophyll was estimated from IR fluorescence measurements, chlorophyll in pigment
extracts. Lower panel: relative AAP abundances (percentage of total DAPI counts)
determined by IR epifluorescence microscopy in the Baltic Sea (Mašı́n et al. 2006),
Sargasso Sea (Koblı́žek et al. 2007) and the Mediterranean Sea (Hojerová et al.
2011).
consortia is probably too small to be easily recognized during
short experiments. On the other hand, AAP bacteria are typically observed in the euphotic zones of the oceans or lakes (see
Fig. 7). Below the euphotic zone, AAP numbers fall to nil. A positive effect of light was also be documented from seasonal studies conducted in the Mediterranean coastal waters and coastal
lagoons showing a positive correlation of AAP numbers and day
length (Lamy et al. 2011a; Ferrera et al. 2014). Clearly, the environmental data signalize that light has a profound influence in
the long term. The light-derived energy increases the bacterial
growth efficiency of AAP species. Also, it has been documented
that light enhances survival of AAP bacteria under nutrient starvation (Shiba 1984; Soora and Cypionka 2013). Both the effects
become apparent only in the long term, but represent a strong
competitive advantage and increase fitness of AAP bacteria over
heterotrophic species.
Among other environmental factors influencing AAP abundance seems to be temperature. The positive effect of temperature was observed in all seasonal studies from both coastal
waters (Mašı́n et al. 2006; Zhang and Jiao 2007; Lamy et al.
Koblı́žek
863
AAP abundance [103 cells mL-1 ]
0
10
20
30
40
50
0
20
Depth [m]
40
60
80
100
AAP abundance
Chlorophyll a
Temperature
120
140
0.0
0.1
0.2
0.3
0.4
0.5
Chlorophyll concentration [mg m-3 ]
10
15
20
25
30
Temperature [ºC]
Figure 7. Depth profiles of AAP abundance, chlorophyll a concentration and temperature in the Balearic Sea, station ‘D’, September 2007 (Hojerová et al. 2011).
8
6
150
4
100
2
50
0
0
Mar
May
Jul
Sep
Nov
Jan
Mar
May
Jul
Sep
25
20
15
10
Temperature [ºC]
AAP biomass
Chlorophyll
Temperature
Chlorophyll a [mg m -3 ]
AAP biomass [mg Cm -3 ]
200
5
0
Nov
2004 - 2005
Figure 8. Seasonal changes in microbial community in Čertovo Lake during 2004–05. The abundance of AAPs was determined by IR epifluorescence microscopy as the
surface area of BChl a positive cells (Mašı́n et al. 2008).
2011a; Ferrera et al. 2014) and freshwater lakes (Mašı́n et al. 2008;
Čuperová et al. 2013; see Fig. 8). However, the correlations of AAP
abundance and temperature are relatively weak and overlapping
with other factors. Interestingly, a seasonal study from the East
China Sea documented higher AAP percentages during the winter period, despite the fact that the total AAP abundance was
higher in summer (Zhang and Jiao 2007). Till now, it is not clear
whether temperature itself stimulates the growth of AAP bacteria or if the effect is only indirect. In many inland lakes and some
marine environments, the water temperature reflects the inte-
grated sum of received solar radiation during the season (Ferrera
et al. 2014). As such the observed correlations may simply reflect
the long-term positive effect of light on both photoheterotrophic
organisms and on autochtonous production of organic carbon by
phytoplankton.
An intensively discussed issue represents a specific relationship between AAP species and phytoplankton. AAP bacteria
are frequently present in algal cultures (Prof. Michael Sieracki,
pers. comm.). Several AAP strains were isolated from cultures of
the dinoflagellates Alexandrium ostenfeldii and Prorocentrum lima
864
FEMS Microbiology Reviews, 2015, Vol. 39, No. 6
(Allgaier et al. 2003). Four of the obtained isolates were later formally described as Dinoroseobacter shibae, Roseovarius mucosus,
Hoeflea phototrophica and Labrenzia alexandrii (Biebl et al. 2005a,b;
Biebl, Tindall and Wagner-Döbler 2006; Biebl et al. 2007). The exact mode of coexistence of AAP bacteria and algae is not clear.
Both species may live freely in the absence of their partner, but
it seems that there is a mutual benefit living in coculture. One
of the possible explanations is that algae provide organic carbon for AAP bacteria in exchange for vitamins, and that the algae are not able to produce themselves. This hypothesis seem
to be confirmed by the experiments where D. shibae stimulated
the growth of the dinoflagellate P. minimum, probably by providing vitamins B1 and B12 (Wagner-Döbler et al. 2010). Surprisingly,
at later stages of the cocultivation experiments this mutualistic
relationship broke and the bacterium killed its algal host (Wang
et al. 2014).
While laboratory data on the relationship of AAP bacteria and
algae are accumulating, the information from the field is very
limited. A study investigating the influence of algal blooms on
AAP communities in the East China Sea documented that AAP
abundance was frequently higher at stations with diatom or dinoflagelate blooms (Chen, Zhang and Jiao 2011). An exception
was the bloom of dinoflagellate Akashiwo sanguinea where the
AAP numbers were low (Chen, Zhang and Jiao 2011). The results
indicate that the association between AAP bacteria and algae
is likely to be complex and species specific. Clearly, the phenomenon deserves more attention.
An important ecological issue is whether AAP bacteria are
planktonic free-living organisms or mostly particle-attached organisms. Interestingly, most of the first cultured AAP species
originated from seaweeds, algal mats or stromatolites (Shiba,
Simidu and Taga 1979b; Yurkov and Gorlenko 1990; Nishimura
et al. 1994). AAP species were detected in epilithic river biofilms
(Hirose et al. 2012) and plant surfaces (Atamna-Ismaeel et al.
2012). Many cultured AAP species also form aggregates and adhere to the culture vessels. The situation in the field is a matter
of dispute. Kolber et al. (2000) reported that most of the AAP bacteria in the subtropical Pacific pass through the GF/F filter, which
suggests that the community is composed mostly of free-living
single cells smaller than 0.7 μm. A very different picture was
found in Chesapeake and Delaware estuaries where 30–94% of
total AAP bacteria were particle attached (Waidner and Kirchman 2007; Cottrell, Ras and Kirchman 2010). A larger survey conducted by Lami et al. (2009) documented that the proportion of
free-living and particle-attached AAP bacteria varied among the
environments. In off-shore and open ocean environments, the
majority of the AAP communities were small (<0.8 μm) freeliving cells whereas in coastal regions up to one third of the AAP
community was particle attached. A large fraction of particleattached AAP cells was also observed in freshwater lakes (Mašı́n
et al. 2012). It seems that AAP bacteria do not exhibit any enhanced tendency to attach to particles when compared to heterotroph species. On the other hand, metabolically active AAP
bacteria may benefit from the nutrient-rich microenvironment
in the particles. The particle-attached lifestyle may also provide
protection against grazing.
THE ROLE OF AAP BACTERIA
IN THE MICROBIAL LOOP
The microbial activity of AAP bacteria and their role in biogeochemical cycles remains an open question. Since the AAP bacteria lack carbon fixation capacity and depend on organic carbon
substrates, they have to be regarded as secondary producers due
to their role in recycling DOM. On the other hand, the capacity of utilizing light energy may help these organisms to better
utilize available resources. Laboratory experiments documented
that AAP bacteria have significantly higher bacterial growth efficiency when grown in a light–dark regime, i.e. they accumulate more biomass per unit of supplied substrates (Biebl and
Wagner-Döbler 2006; Hauruseu and Koblı́žek 2012). This gives
AAP bacteria (when grown in light) a clear competitive advantage over chemoheterotrophs with lower growth efficiencies.
One can speculate that additional energy obtained from light
may even help AAP bacteria to utilize some low energetic or recalcitrant carbon sources. However, the direct proof of such capacity is so far missing.
Despite all the uncertainties, AAP bacteria appear to be a
highly active part of the microbial community. Experiments
conducted in the Delaware estuary documented that AAP cells
incorporated approximately two times more 3 H-leucine when
compared to average bacteria (Stegman, Cottrell and Kirchman
2014). Similarly, it was shown that AAP bacteria have approximately three times higher frequency of dividing cells when compared to heterotrophic bacteria (Liu, Zhang and Jiao 2010). High
activity of AAP bacteria also signalized high levels of puf and
bch genes registered in the metatranscriptome of the microbial
community collected at the ALOHA station in the Pacific Ocean
(Frias-Lopez et al. 2008), despite the fact that AAP bacteria are
not very abundant in these oligotrophic waters (Cottrell, Mannino and Kirchman 2006; Ritchie and Johnson 2012).
A question arises why there are so few AAP bacteria given
that they are so metabolically active? Why do they not dominate the bacterial communities in the euphotic zones? The answer is more complex. Firstly, the euphotic zones of the oceans
actually seem dominated by photoheterotrophic species. Over
the last 15 years, it has become established that large numbers
of bacteria within the euphotic zone contain proteorhodopsin
molecules which seem to serve for light harvesting (Béjà et al.
2000; Eiler 2006; Béjà and Suzuki 2008). Moreover, Prochlorococcus,
the dominant phytoplankton species in the oligotrophic ocean,
may also employ photoheterotrophic assimilation of various
organic carbon sources (Zubkov 2009). Thus, photoheterotrophy and mixotrophy seems to be widespread in many aquatic
environments and AAP bacteria are just one of many photoheterotrophic groups using this strategy.
Secondly, the ability to use light energy helps AAP bacteria
to better utilize the available substrate. This provides a competitive advantage mainly under carbon-limited conditions. However, the microbial growth is, in most of the upper ocean environments, limited by the availability of other nutrients: nitrogen,
phosphorus and in some areas also iron (Church 2008). Under
such conditions, AAP bacteria do not seem to have any strong
advantage above regular chemoheterotrophic species. On the
contrary, their larger cell sizes restrict the diffusion of nutrients,
and their larger genomes, requiring more phosphorus, may (under phosphorus starvation) actually represent a disadvantage.
Lastly, the abundance of AAP bacteria may be significantly reduced by grazing. Microscopic analyses of open ocean (Sieracki
et al. 2006) and estuary (Kirchman et al. 2014; Stegman, Cottrell
and Kirchman 2014) bacterial communities documented that
AAP bacteria were on average larger than regular heterotrophic
bacteria. The larger size likely makes AAP bacteria more vulnerable targets for protist grazing. It has been suggested that BChl a
can be used as a natural in situ tracer, as its decay during the day
reflects AAP mortality (Koblı́žek et al. 2005). Using this approach,
it has been found that in the Atlantic oligotrophic gyres BChl a
Koblı́žek
865
4
BChl a turnover [days
-1
]
Villefranche Bay
3
North Atlantic
2
Equatorial Atlantic
W Mediterranean
1
Sargasso Sea
0
0.0
0.1
0.2
0.3
0.4
0.5
0.6
-3
Chlorophyll a [mg m ]
Figure 9. Relationship between bacteriochlorophyll a turnover and chlorophyll concentration. Each datapoint represents an average value from two to five individual
diel measurements. The Atlantic Ocean and Villefranche Bay data were taken from Koblı́žek et al. (2007). Western Mediterranean data from Hojerová et al. (2011).
decayed at rate of 0.7 d−1 which should be proportional to AAP
mortality. Under stable steady-state conditions, the high mortality rates have to be balanced by equally rapid growth rates. This
signalizes that AAP bacteria in oligotrophic regions divide at a
rate of one division per day. In more productive marine regions,
such as the Equatorial upwelling and North Atlantic, AAP bacteria decayed at rates of two to three divisions per day (Koblı́žek
et al. 2007). A similar situation was also observed in the coastal
regions of the Mediterranean Sea (Hojerová et al. 2011, see Fig. 9).
The higher growth rates of AAP bacteria were also confirmed
during manipulation experiments conducted in coastal Mediterranean waters. In treatments where grazing pressure was reduced by filtration or dilution, the proportion of AAP bacteria increased rapidly, reaching up to 20% of total bacteria (Ferrera et al.
2011), documenting that the relatively low numbers of these organisms in nature are caused by a strong grazing pressure. The
calculated AAP growth rates were about two times higher than
those determined for average bacteria (Ferrera et al. 2011).
The available data document an important contribution of
AAP to the carbon cycle. AAP bacteria represent 1–7% of total
bacteria in the euphotic zones of the oceans (see Table 1), but
their cells are on average two times larger than average bacteria
(Sieracki et al. 2006). Moreover, in some environments AAP bacteria exhibit faster growth rates when compared to other bacteria
(Ferrera et al. 2011). As a result, the contribution of AAP bacteria
to the total bacterial (secondary) production may be significantly
higher than their abundance alone would indicate.
All the presented data indicate that AAP bacteria are
metabolically active rapidly growing species with many enzymatic activities (in spite of the fact that even among AAP
bacteria there may exist significant heterogeneity with more
or less active species). The increasing number of genomic sequences (see Table S3, Supporting Information) document that
AAP species possess relatively large genomes, 3.1–4.8 Mb for
Roseobacter-related AAP species, 3.6 Mb for gammaproteobacterium HTCC2080 and 4.3 Mb for Congregibacter litoralis (Fuchs
et al. 2007), 3.3 Mb for Erythrobacter sp. NAP1 (Koblı́žek et al. 2011),
3.41 and 3.03 Mb for phototrophic Limnohabitans species (Zeng
et al. 2012), encoding a large variety of enzymatic functions. In
contrast, the genome of Pelagibacter ubique, a cultivated member
of the ubiquitous SAR11 clade containing the proteorhodopsin
gene, is only 1.3 Mb (Giovannoni et al. 2005). Interestingly, the
bacterial groups like Roseobacter clade, gammaproteobacterial
clade NOR5/OM6 or Limnohabitans, which are known to harbor
AAP species, are all very metabolically active and conduct a
more generalist life strategy. The larger genome sizes in these
species come from their large enzyme inventory in general, and
not from photosynthesis genes. Actually, the cell machinery for
anoxygenic phototrophy is only encoded by 30–40 genes and
represents about 1% of the total genome (Swingley et al. 2009;
Yurkov and Hughes 2013; Zheng et al. 2013). From this perspective, the ability to use light appears as only one of many alternative enzymatic functions in the metabolic portfolio of these
versatile organisms.
CONCLUDING REMARKS
Since the discovery of AAP bacteria by Harashima and Shiba in
1970s, these organisms have become commonly known among
aquatic microbiologists studying both marine and limnic environments. Interestingly, it seems that AAP bacteria are not restricted to only aquatic environments as they are being found
in many unexpected habitats such as soil crusts (Csotonyi et al.
2010), phyllospheres (Atamna-Ismaeel et al. 2012; Stiefel, Zambelli and Vorholt 2013), biofilms (Hirose et al. 2012) or Antarctic Sea Ice (Koh, Phua and Ryan 2011). The effort of many researchers clarified most of the basic questions regarding their
physiology, diversity and distribution. The initial exploratory
phase of AAP research is almost over. To further expand our
understanding of this interesting group, it is necessary to apply
new approaches and ask novel questions.
One of the crucial questions remains how and to what extend
AAP bacteria actually benefit from light under natural (field) conditions. It is clear that the capacity to generate energy through
anoxygenic photosynthesis is beneficial in many different habitats. Many questions however remain. Do we find AAP bacteria
866
FEMS Microbiology Reviews, 2015, Vol. 39, No. 6
in permanently dark habitats such as the deep ocean? Does light
only provide ATP for the cell maintenance and metabolism or
does the anoxygenic photosynthesis makes it possible to utilize
substrates and nutrients which would otherwise not be accessible? Does light help AAP bacteria to use recalcitrant organic
carbon? Does light energy stimulate phosphorus or nitrogen acquisition? To answer these questions, it is necessary to conduct
carefully designed experiments in the field. The impact of experimental treatments has to be studied using all available tools,
such as epifluorescence microscopy and pigment analyses. The
community structure can be analyzed using next generation sequencing tools.
There are still some novel AAP species to be isolated. Cultivation on agar plates has proven to be a suitable strategy for the
isolation of novel species, but novel approaches such as serial dilutions (Salka et al. 2008), micromanipulation (Stiefel, Zambelli
and Vorholt 2013) or fluorescence-activated cell sorting in combination with appropriate optical screening may allow the isolation of species which defy growth on solid media.
The high diversity of AAP bacteria and their obvious relationship to purple non-sulfur bacteria raises the question of
their phylogenetic origin. The current explosion of genome sequencing data may help to explore these questions and put
to test various evolutionary scenarios. It seems that the evolution of phototrophy in Proteobacteria is a very complex phenomenon (Nagashima et al. 1997; Koblı́žek et al. 2013) and has
to be explored step by step divided in better managable pieces
(Boldareva-Nuianzina et al. 2013; Koblı́žek et al. 2015).
Another field which still has to be explored is AAP physiology. Under which conditions do AAP bacteria express their phototrophic apparatus? Many species do not express pigments under laboratory conditions, but they probably use photosynthesis in the field under specific conditions (Voget et al. 2015). This
fact has an important implication for the role of these organisms
in the geochemical cycles, since it signalizes that, in the natural environment they can act as either heterotrophic or photoheterotrophic organisms depending on the actual environmental or physiological conditions. In Roseateles depolymerans, it has
been suggested that bacteriochlorophyll expression is expressed
under low carbon conditions (Suyama et al. 2002). A similar phenomenon was also observed in Hoeflea phototrophica (Biebl and
Wagner-Döbler 2006), but more details are so far missing. Careful laboratory experiments with cultured species with known
genomes will allow us to study changes in total gene expresion
(Tomasch et al. 2011) and the influence of various environmental stimuli (light, oxygen concentration, nutrient status). When
integrated with protein and metabolic profiles, these modern
techniques will provide a once unthinkable wealth of data which
will help to understand the physiology and metabolism in all its
complexity.
In summary, all the available data show that AAP bacteria are
common in many aquatic environments. They are metabolically
active organisms, rapidly utilizing a variety of potential substrates and energy sources, including light. This enables them to
maintain higher growth rates when compared to more specialized organisms. In result, large and rapidly growing AAP cells
contribute more significantly to the secondary production and
aquatic carbon cycling than it would appear from cell numbers
alone.
SUPPLEMENTARY DATA
Supplementary data are available at FEMSRE online.
ACKNOWLEDGEMENTS
MK thanks Prof. David L. Kirchman, Dr Yonghui Zeng, Dr Isabel Ferrera and the anonymous reviewers for their valuable
comments and suggestions on this text. Thanks also to Dr
Marcus Tank for providing his pufM sequences from the Baltic
Sea lagoon.
FUNDING
This research was supported by Czech Science Foundation
project 13-11281S and project Algatech plus (LO1416).
Conflict of interest. None declared.
REFERENCES
Achenbach LA, Carey J, Madigan MT. Photosynthetic and phylogenetic primers for detection of anoxygenic phototrophs in
natural environments. Appl Environ Microb 2001;67:2922–6.
Allgaier M, Uphoff H, Felske A, et al. Aerobic anoxygenic photosynthesis in Roseobacter clade bacteria from diverse marine
habitats. Appl Environ Microb 2003;69:5051–9.
Atamna-Ismaeel N, Finkel O, Glaser F, et al. Bacterial anoxygenic
photosynthesis on plant leaf surfaces. Environ Microbiol Rep
2012;4:209–16.
Béjà O, Aravind L, Koonin EV, et al. Bacterial rhodopsin: evidence for a new type of phototrophy in the sea. Science
2000;289:1902–6.
Béjà O, Suzuki MT. Photoheterotrophic marine prokaryotes. In:
Kirchman DL (ed). Microbial Ecology of the Oceans. New York,
NY, USA: Wiley-Blackwell, 2008, 131–57.
Béjà O, Suzuki MT, Heidelberg JF, et al. Unsuspected diversity among marine aerobic anoxygenic phototrophs. Nature
2002;415:630–3.
Bekker A, Holland HD, Wang PL, et al. Dating the rise of atmospheric oxygen. Nature 2004;427:117–20.
Biebl H, Allgaier M, Lünsdorf H, et al. Roseovarius mucosus sp. nov.,
a member of the Roseobacter clade with trace amounts of
bacteriochlorophyll a. Int J Syst Evol Micr 2005b;55:2377–83.
Biebl H, Allgaier M, Tindall BJ, et al. Dinoroseobacter shibae gen.
nov., sp. nov., a new aerobic phototrophic bacterium isolated
from dinoflagellates. Int J Syst Evol Micr 2005a;55:1089–96.
Biebl H, Pukall R, Lünsdorf H, et al. Description of Labrenzia
alexandrii gen. nov., sp. nov., a novel alphaproteobacterium
containing bacteriochlorophyll a, and a proposal for reclassification of Stappia aggregata as Labrenzia aggregata comb.
nov., of Stappia marina as Labrenzia marina comb. nov. and of
Stappia alba as Labrenzia alba comb. nov., and emended descriptions of the genera Pannonibacter, Stappia and Roseibium,
and of the species Roseibium denhamense and Roseibium hamelinense. Int J Syst Evol Micr 2007;57:1095–107.
Biebl H, Tindall BJ, Wagner-Döbler I. Hoeflea phototrophica sp. nov.,
a novel marine aerobic alphaproteobacterium that forms
bacteriochlorophyll a. Int J Syst Evol Micr 2006;56:821–6.
Biebl H, Wagner-Döbler I. Growth and bacteriochlorophyll a
formation in taxonomically diverse aerobic anoxygenic
phototrophic bacteria in chemostat culture: influence
of light regiment and starvation. Proc Biochem 2006;41:
2153–9.
Boeuf D, Cottrell MT, Kirchman DL, et al. Summer community
structure of aerobic anoxygenic phototrophic bacteria in the
western Arctic Ocean. FEMS Microbiol Ecol 2013;85:417–32.
Koblı́žek
Boeuf D, Humily F, Jeanthon C. Diversity of Arctic pelagic Bacteria with an emphasis on photoheterotrophs: a review. Biogeosciences 2014;11:3309–22.
Boldareva-Nuianzina, Bláhová Z, Sobotka R, et al. Distribution
and origin of oxygen-dependent and oxygen-independent
forms of Mg-protoporphyrin monomethylester cyclase
among phototrophic proteobacteria. Appl Environ Microb
2013;79:2596–604.
Caliz J, Casamayor EO. Environmental controls and composition of anoxygenic photoheterotrophs in ultraoligotrophic
high-altitude lakes (Central Pyrenees). Environ Microbiol Rep
2014;6:145–51.
Candela M, Zaccherini E, Zannoni D. Respiratory electron transport and light-induced energy transduction in membranes
from the aerobic photosynthetic bacterium Roseobacter denitrificans. Arch Microbiol 2001;175:168–77.
Chen Y, Zhang Y, Jiao N. Responses of aerobic anoxygenic phototrophic bacteria to algal blooms in the East China Sea. Hydrobiologia 2011;661:435–43.
Cho J-C, Stapels MD, Morris RM, et al. Polyphyletic photosynthetic reaction centre genes in oligotrophic marine
Gammaproteobacteria. Environ Microbiol 2007;9:1456–63.
Church MJ. Resource control of bacterial dynamics in the sea. In:
Kirchman DL (ed). Microbial Ecology of the Oceans. New York,
NY, USA: Wiley-Blackwell, 2008, 335–82.
Cottrell MT, Kirchman DL. Photoheterotrophic microbes in the
Arctic Ocean in summer and winter. Appl Environ Microb
2009;75:4958–66.
Cottrell MT, Mannino A, Kirchman DL. Aerobic anoxygenic phototrophic bacteria in the Mid-Atlantic Bight and the North
Pacific Gyre. Appl Environ Microb 2006;72:557–64.
Cottrell MT, Ras J, Kirchman DL. Bacteriochlorophyll and community structure of aerobic anoxygenic phototrophic bacteria in a particle-rich estuary. ISME J 2010;4:945–54.
Csotonyi JT, Swiderski J, Stackebrandt E, et al. Novel halophilic
aerobic anoxygenic photrotrophs from a Canadian hypersaline spring system. Extremophiles 2008;12:529–39.
Csotonyi JT, Swiderski J, Stackebrandt E, et al. A new environment for aerobic anoxygenic phototrophic bacteria: biological soil crusts. Environ Microbiol 2010;2:651–56.
Čuperová Z, Holzer E, Salka I, et al. Temporal changes and altitudinal distribution of aerobic anoxygenic phototrophs in
mountain lakes. Appl Environ Microb 2013;79:6439–46.
Du H, Jiao N, Hu Y, et al. Real-time PCR for quantification of aerobic anoxygenic phototrophic bacteria based on pufM gene in
marine environment. J Exp Mar Biol Ecol 2006;329:113–21.
Eiler A. Evidence for the ubiquity of mixotrophic bacteria in the
upper ocean: implications and consequences. Appl Environ
Microb 2006;72:7431–7.
Eiler A, Beier S, Sawstrom C, et al. High ratio of bacteriochlorophyll biosynthesis genes to chlorophyll biosynthesis genes
in bacteria of humic lakes. Appl Environ Microb 2009;75:
7221–8.
Famintzin A. Eine neue Bakterienform: Nevskia ramosa. Bull Acad
Imp Sci St Petersburg, New Ser 2. 1892;34:481–6.
Fauteux L, Cottrell MT, Kirchman DL, et al. Patterns in abundance, cell size and pigment content of aerobic anoxygenic phototrophic bacteria along environmental gradients
in northern lakes. PLoS One 2015;10:e0124035.
Ferrera I, Borrego CM, Salazar G, et al. Marked seasonality of
aerobic anoxygenic phototrophic bacteria in the coastal NW
Mediterranean Sea as revealed by cell abundance, pigment
concentration and pyrosequencing of pufM gene. Environ Microbiol 2014;16:2953–65.
867
Ferrera I, Gasol JM, Sebastián M, et al. Growth rates of aerobic
anoxygenic phototrophic bacteria as compared to other bacterioplankton groups in coastal Mediterranean waters. Appl
Environ Microb 2011;77:7451–8.
Finkel OM, Béjà O, Belkin S. Global abundance of microbial
rhodopsins. ISME J 2013;7:448–51.
Fleischman D, Kramer D. Photosynthetic rhizobia. Biochim Biophys Acta 1998;1364:17–36.
Frias-Lopez J, Shi Y, Tyson GW, et al. Microbial community gene
expression in ocean surface waters. P Natl Acad Sci USA
2008;105:3805–10.
Fuchs BM, Spring S, Teeling H, et al. Characterization of
a marine gammaproteobacterium capable of aerobic
anoxygenic photosynthesis. P Natl Acad Sci USA 2007;104:
2891–6.
Fuerst JA, Hawkins JA, Holmes A, et al. Porphyrobacter neustonensis gen. nov., sp. nov., an aerobic bacteriochlorophyllsynthesizing budding bacterium from fresh water. Int J Syst
Bacteriol 1993;43:125–34.
Gich F, Overmann J. Sandarakinorhabdus limnophila gen. nov., sp.
nov., a novel bacteriochlorophyll a-containing, obligately aerobic bacterium isolated from freshwater lakes. Int J Syst Evol
Micr 2006;56:847–54.
Giebel H-A, Kalhoefer D, Gahl-Janssen R, et al. Planktomarina temperata gen. nov., sp. nov., belonging to the globally distributed
RCA cluster of the marine Roseobacter clade, isolated from
the German Wadden Sea. Int J Syst Evol Micr 2013;63:4207–17.
Giovannoni SJ, Bibbs L, Cho JC, et al. Proteorhodopsin in the ubiquitous marine bacterium SAR11. Nature 2005;438:82–5.
Goericke R. Bacteriochlorophyll a in the ocean: is anoxygenic bacterial photosynthesis important? Limnol Oceanogr
2002;47:290–5.
Hahn MW, Kasalický V, Jezbera J, et al. Limnohabitans australis
sp nov., isolated from a freshwater pond, and emended
description of the genus Limnohabitans. Int J Syst Evol Micr
2010;60:2946–50.
Harashima K, Hayasaki J-I, Ikari T, et al. O2 -stimulated synthesis of bacteriochlorophyll and carotenoids in marine bacteria. Plant Cell Physiol 1980;21:1283–94.
Harashima K, Kawazoe K, Yoshida I, et al. Light stimulated aerobic growth of Erythrobacter species OCh 114. Plant Cell Physiol
1987;28:365–74.
Harashima K, Shiba T, Murata N. Aerobic Photosynthetic Bacteria.
Tokyo: Japan Scientific Societies Press, 1989.
Harashima K, Shiba T, Totsuka T, et al. Occurence of bacteriochlorophyll a in a strain of an aerobic heterotrophic bacterium. Agr Biol Chem 1978;42:1627–8.
Hauruseu D, Koblı́žek M. The influence of light on carbon utilization in aerobic anoxygenic phototrophs. Appl Environ Microb
2012;78:7414–9.
Hiraishi A, Shimada K. Aerobic anoxygenic photosynthetic
bacteria with zinc-bacteriochlorophyll. J Gen Appl Microbiol
2001;47:161–80.
Hirose S, Nagashima KV, Matsuura K, et al. Diversity of purple phototrophic bacteria, inferred from pufM gene, within
epilithic biofilm in Tama River, Japan. Microbes Environ
2012;27:327–9.
Hohmann-Marriott MF, Blankenship RE. Evolution of photosynthesis. Annu Rev Plant Biol 2011;62:515–48.
Hojerová E, Mašı́n M, Brunet C, et al. Distribution and growth of
aerobic anoxygenic phototrophs in the Mediterranean Sea.
Environ Microbiol 2011;13:2717–25.
Holland HD. The oxygenation of the atmosphere and oceans.
Philos T Roy Soc B 2006;361:903–15.
868
FEMS Microbiology Reviews, 2015, Vol. 39, No. 6
Hu YH, Du HL, Jiao NZ, et al. Abundant presence of the gammalike proteobacterial pufM gene in oxic seawater. FEMS Microbiol Lett 2006;263:200–6.
Iba K, Takamiya K. Action spectra for light-inhibition of
bacteriochlorophyll and carotenoid accumulation during
aerobic growth of photosynthetic bacteria. Plant Cell Physiol
1989;30:471–7.
Jeanthon C, Boeuf D, Dahan O, et al. Diversity of cultivated and
metabolically active aerobic anoxygenic phototrophic bacteria along an oligotrophic gradient in the Mediterranean Sea.
Biogeosciences 2011;8:1955–70.
Jiao N, Zhang F, Hong N. Significant roles of bacteriochlorophyll a supplemental to chlorophyll a in the ocean. ISME J
2010;4:595–7.
Jiao N, Zhang Y, Zeng Y, et al. Distinct distribution pattern of
abundance and diversity of aerobic anoxygenic phototrophic
bacteria in the global ocean. Environ Microbiol 2007;9:
3091–9.
Jiang HC, Dong HL, Yu BS, et al. Abundance and diversity of aerobic anoxygenic phototrophic bacteria in saline lakes on the
Tibetan plateau. FEMS Microbiol Ecol 2009;67:268–78.
Kirchman DL, Hanson TE. Bioenergetics of photoheterotrophic
bacteria in the oceans. Environ Microbiol Rep 2013;5:188–99.
Kirchman DL, Stegman MR, Nikrad MP, et al. Abundance, size,
and activity of aerobic anoxygenic phototrophic bacteria in
coastal waters of the West Antarctic Peninsula. Aquat Microb
Ecol 2014;73:41–9.
Kishimoto N. Distribution of bacteriochlorophyll a among aerobic and acidophilic bacteria and light-enhanced CO2 incorporation in Acidiphillium rubrum. FEMS Microbiol Ecol
1995;16:291–6.
Koblı́žek M, Béjà O, Bidigare RR, et al. Isolation and characterization of Erythrobacter sp. strains from the upper ocean. Arch
Microbiol 2003;180:327–38.
Koblı́žek M, Falkowski PG, Kolber ZS. Diversity and distribution
of photosynthetic bacteria in the Black Sea. Deep-Sea Res Pt II
2006;53:1934–44.
Koblı́žek M, Janouškovec J, Obornı́k M, et al. Genome sequence
of the marine photoheterotrophic bacterium Erythrobacter sp.
strain NAP1. J Bacteriol 2011;193:5881–2.
Koblı́žek M, Mašı́n M, Ras J, et al. Rapid growth rates of aerobic anoxygenic phototrophs in the ocean. Environ Microbiol
2007;9:2401–6.
Koblı́žek M, Mlčoušková J, Kolber Z, et al. On the photosynthetic
properties of marine bacterium COL2P belonging to Roseobacter clade. Arch Microbiol 2010;192:41–9.
Koblı́žek M, Moulisová V, Muroňová M, et al. Horizontal gene
transfer of puf operon in Roseobacter clade. Folia Microbiol
2015;60:37–43.
Koblı́žek M, Ston-Egiert J, Sagan S, et al. Diel changes in bacteriochlorophyll a concentration suggest rapid bacterioplankton
cycling in the Baltic Sea. FEMS Microbiol Ecol 2005;51:353–61.
Koblı́žek M, Zeng Y, Horák A, et al. Regressive evolution of photosynthesis in the Roseobacter clade. Adv Bot Res 2013;66:385–
405.
Koh EY, Phua W, Ryan KG. Aerobic anoxygenic phototrophic bacteria in Antarctic sea ice and seawater. Environ Microbiol Rep
2011;3:710–6.
Kolber ZS, Plumley FG, Lang AS, et al. Contribution of aerobic
photoheterotrophic bacteria to the carbon cycle in the ocean.
Science 2001;292:2492–5.
Kolber ZS, Van Dover CL, Niederman RA, et al.Bacterial photosynthesis in surface waters of the open ocean. Nature
2000;407:177–9.
Kramer DM, Kanazawa A, Fleischman D. Oxygen dependence of
photosynthetic electron transport in a bacteriochlorophyllcontaining rhizobium. FEBS Lett 1997;417:275–8.
Labrenz M, Collins MD, Lawson PA, et al. Roseovarius tolerans gen.
nov., sp. nov., a budding bacterium with variable bacteriochlorophyll a production from hypersaline Ekho Lake. Int J
Syst Bacteriol 1999;49:137–47.
Labrenz M, Tindall BJ, Lawson PA, et al. Staleya guttiformis
gen. nov., sp. nov. and Sulfitobacter brevis sp. nov., alpha3-Proteobacteria from hypersaline, heliothermal and
meromictic antarctic Ekho Lake. Int J Syst Evol Micr
2000;50:303–13.
Lami R, Cottrell MT, Ras J, et al. High abundances of aerobic anoxygenic photosynthetic bacteria in the South Pacific
Ocean. Appl Environ Microb 2007;73:4198–205.
Lami R, Čuperová Z, Ras J, et al. Distribution of free-living versus particle-attached aerobic anoxygenic phototrophic bacteria in marine environments. Aquat Microbial Ecol 2009;
55:31–8.
Lamy D, De Carvalho-Maalouf P, Cottrell MT, et al. Seasonal dynamics of aerobic anoxygenic phototrophs in a
Mediterranean coastal lagoon. Aquat Microbial Ecol 2011a;62:
153–63.
Lamy D, Jeanthon C, Cottrell MT, et al. Ecology of aerobic anoxygenic phototrophic bacteria along an oligotrophic gradient in
the Mediterranean Sea. Biogeosciences 2011b;8:973–85.
Lehours A-C, Cottrell MT, Dahan O, et al. Summer distribution
and diversity of aerobic anoxygenic phototrophic bacteria
in the Mediterranean Sea in relation to environmental variables. FEMS Microbiol Ecol 2010;74:397–409.
Lew S, Koblı́žek M, Lew M, et al. Seasonal changes of microbial
communities in two shallow peat-bog lakes. Folia Microbiol
2015;60:165–75.
Liu R, Zhang Y, Jiao N. Diel variations in frequency of dividing cells and abundance of aerobic anoxygenic phototrophic
bacteria in a coral reef system of the South China Sea. Aquat
Microb Ecol 2010;58:303–10.
Luo H, Moran MA. Evolutionary ecology of the marine Roseobacter
clade. Microbiol Mol Biol R 2014;78:573–87.
Martinez-Garcia M, Swan BK, Poulton NJ, et al. High-throughput
single-cell sequencing identifies photoheterotrophs and
chemoautotrophs in freshwater bacterioplankton. ISME J
2011;6:113–23.
Mašı́n M, Čuperová Z, Hojerová E, et al. Distribution of aerobic
anoxygenic phototrophic bacteria in glacial lakes of northern
Europe. Aquat Microbial Ecol 2012;66:77–86.
Mašı́n M, Nedoma J, Pechar L, et al. Distribution of aerobic anoxygenic phototrophs in temperate freshwater systems. Environ
Microbiol 2008;10:1988–96.
Mašı́n M, Zdun A, Stoń-Egiert J, et al. Seasonal changes and diversity of aerobic anoxygenic phototrophs in the Baltic Sea.
Aquat Microbial Ecol 2006;45:247–54.
Medová H, Boldareva E, Hrouzek P, et al. High abundances of aerobic anoxygenic phototrophs in saline steppe lakes. FEMS Microbiol Ecol 2011;76:393–400.
Michelou VK, Cottrell MT, Kirchman DL. Light-stimulated organic matter assimilation by cyanobacteria and other microbes in the North Atlantic Ocean. Appl Environ Microb
2007;73:5539–46.
Nagashima KVP, Hiraishi A, Shimada K, et al. Horizontal transfer
of genes coding for the photosynthetic reaction centers of
purple bacteria. J Mol Evol 1997;45:131–6.
Nisbet EG, Cann JR, Van Dover CL. Origins of photosynthesis. Nature 1995;373:479–80.
Koblı́žek
Nishimura Y, Muroga Y, Saito S, et al. DNA relatedness and
chemotaxonomic feature of aerobic bacteriochlorophyllcontaining bacteria isolated from coast of Australia. J Gen
Appl Microbiol 1994;40:287–96.
Nishimura K, Shimada H, Ohta H, et al. Expression of the puf
operon in an aerobic photosynthetic bacterium, Roseobacter
denitrificans. Plant Cell Physiol 1996;37:153–9.
Okamura K, Mitsumori F, Ito O, et al. Photophosphorylation
and oxidative phosphorylation in intact cells and chromatophores of an aerobic photosynthetic bacterium, Erythrobacter sp. strain OCh 114. J Bacteriol 1986;168:1142–6.
Oz A, Sabehi G, Koblı́žek M, et al. Roseobacter-like bacteria in Red
and Mediterranean Sea aerobic anoxygenic photosynthetic
populations. Appl Environ Microb 2005;71:344–53.
Page KA, Connon SA, Giovannoni SJ. Representative freshwater
bacterioplancton isolated from Crater Lake, Oregon. Appl Environ Microb 2004;70:6542–50.
Pujalte MJ, Lucena T, Ruvira MA, et al. The family Rhodobacteraceae. In: Rosenberg G, DeLong EF, Lory S, et al. (eds).
Prokaryotes: Alphaproteobacteria and Betaproteobacteria. Heidelberg, Berlin: Springer, 2014, 439–512.
Rathgeber C, Alric J, Hughes E, et al. The photosynthetic apparatus and photoinduced electron transfer in the aerobic phototrophic bacteria Roseicyclus mahoneyensis and Porphyrobacter
meromictius. Photosynth Res 2012;110:193–203.
Ritchie AE, Johnson ZI. Abundance and genetic diversity of
aerobic anoxygenic phototrophic bacteria of coastal regions of the Pacific Ocean. Appl Environ Microb 2012;78:
2858–66.
Ruiz-González C, Proia L, Ferrera I, et al. Effects of large river dam
regulation on bacterioplankton community structure. FEMS
Microbiol Ecol 2013;84:316–31.
Salka I, Čuperová Z, Mašı́n M, et al. Rhodoferax-related pufM
gene cluster dominates the aerobic anoxygenic phototrophic
communities in German freshwater lakes. Environ Microbiol
2011;13:2865–75.
Salka I, Moulisová V, Koblı́žek M, et al. Abundance, depth distribution, and composition of aerobic bacteriochlorophyll aproducing bacteria in four deeps of the central Baltic Sea.
Appl Environ Microb 2008;74:4398–404.
Sato K. Bacteriochlorophyll formation by facultative methylotrophs, Protaminobacter ruber and Pseudomonas AM1. FEBS
Lett 1978;85:207–10.
Sato-Takabe Y, Hamasaki K, Suzuki K. Photosynthetic characteristics of marine aerobic anoxygenic phototrophic bacteria Roseobacter and Erythrobacter strains. Arch Microbiol
2012;194:331–41.
Sato-Takabe Y, Hamasaki K, Suzuki K. Photosynthetic competence of the marine aerobic anoxygenic phototrophic bacterium Roseobacter sp. under organic substrate limitation. Microbes Environ 2014;29:100–3.
Schwalbach MS, Brown M, Fuhrman JA. Impact of light on marine bacterioplankton community structure. Aquat Microb
Ecol 2005;39:235–45.
Schwalbach MS, Fuhrman JA. Wide-ranging abundances of aerobic anoxygenic phototrophic bacteria in the world ocean revealed by epifluorescence microscopy and quantitative PCR.
Limnol Oceanogr 2005;50:620–8.
Selje N, Simon M, Brinkhoff T. A newly discovered Roseobacter cluster in temperate and polar oceans. Nature 2004;427:
445–8.
Shiba T. Utilization of light energy by the strictly aerobic
bacterium Erythrobacter sp. Och 114. J Gen Appl Microbiol
1984;30:239–44.
869
Shiba T. O2 regulation of bacteriochlorophyll synthesisin the
aerobic bacterium Erythrobacter. Plant Cell Physiol 1987;28:
1313–20.
Shiba T. Roseobacter litoralis gen. nov., sp. nov. and Roseobacter denitrificans sp. nov., aerobic pink-pigmented bacteria
which contain bacteriochlorophyll a. Syst Appl Microbiol
1991;14:140–5.
Shiba T, Shioi Y, Takamiya K-I, et al. Distribution and physiology of aerobic bacteria containing bacteriochlorophyll a on
the east and west coasts of Australia. Appl Environ Microb
1991;57:295–300.
Shiba T, Simidu U. Erythrobacter longus gen. nov., sp. nov., an aerobic bacterium which contains bacteriochlorophyll a. Int J Syst
Bacteriol 1982;32:211–7.
Shiba T, Simidu U, Taga N. Another aerobic bacterium which
contains bacteriochlorophyll a. B Jpn Soc Sci Fish 1979a;45:801.
Shiba T, Simidu U, Taga N. Distribution of aerobic bacteria
which contain bacteriochlorophyll a. Appl Environ Microb
1979b;38:43–5.
Shioi Y. Growth characteristics and substrate specificity of aerobic photosynthetic bacterium Erythrobacter sp. (OCh 114).
Plant Cell Physiol 1986;27:567–72.
Sieracki ME, Gilg IC, Thier EC, et al. Distribution of planktonic aerobic anoxygenic photoheterotrophic bacteria in the northwest Atlantic. Limnol Oceanogr 2006;51:38–46.
Šlouf V, Fuciman M, Dulebo A, et al. Carotenoid charge transfer states and their role in energy transfer processes in LH1RC complexes from aerobic anoxygenic phototrophs. J Phys
Chem 2013;117:10987–99.
Soora M, Cypionka H. Light enhances survival of Dinoroseobacter
shibae during long-term starvation. PLOS One 2013;8:e83960.
Spring S, Lunsdorf H, Fuchs BM, et al. The photosynthetic apparatus and its regulation in the aerobic gammaproteobacterium Congregibacter litoralis gen. nov., sp. nov. PLoS One
2009;4:e4866.
Spring S, Riedel T, Spröer C, et al. Taxonomy and evolution of bacteriochlorophyll a-containing members of the OM60/NOR5
clade of marine gammaproteobacteria: description of Luminiphilus syltensis gen. nov., sp. nov., reclassification of
Haliea rubra as Pseudohaliea rubra gen. nov., comb. nov., and
emendation of Chromatocurvus halotolerans. BMC Microbiol
2013;13:118
Stadnichuk IN, Yanyushin MF, Boychenko VA, et al. Photosynthetic activity and components of the electron transport
chain in the aerobic bacteriochlorophyll a-containing
bacterium. Roseinatronobacter thiooxidans. Mikrobiologiya
2009;78:7–15 (in Engl. transl.).
Stegman MR, Cottrell MT, Kirchman DL. Leucine incorporation
by aerobic anoxygenic phototrophic bacteria in the Delaware
estuary. ISME J 2014;8:2339–48.
Stiefel P, Zambelli T, Vorholt JA. Isolation of optically targeted
single bacteria by application of fluidic force microscopy to
aerobic anoxygenic phototrophs from the phyllosphere. Appl
Environ Microb 2013;79:4895–905.
Suyama T, Shigematsu T, Suzuki T, et al. Photosynthetic apparatus in Roseateles depolymerans 61A is transcriptionally
induced by carbon limitation. Appl Environ Microb 2002;68:
1665–73.
Suyama T, Shigematsu T, Takaichi S, et al. Roseateles depolymerans
gen. nov., sp. nov., a new bacteriochlorophyll a-containing
obligate aerobe belonging to the β-subclass of the Proteobacteria. Int J Syst Bacteriol 1999;49:449–57.
Swingley WD, Blankenship RE, Raymond J. Evolutionary relationship among purple photosynthetic bacteria and the
870
FEMS Microbiology Reviews, 2015, Vol. 39, No. 6
origin of Proteobacterial photosynthetic systems. In: Hunter
CN, Daldal F, Thurnauer et al. (eds). The Purple Phototrophic
Bacteria. Dordrecht: Springer, 2009, 17–29.
Swingley WD, Sadekar S, Mastrian SD, et al. The complete genome sequence of Roseobacter denitrificans reveals a
mixotrophic rather than photosynthetic metabolism. J Bacteriol 2007;189:683–90.
Takaichi S, Furihata K, Ishidsu J-I, et al. Carotenoid sulphates
from the aerobic photosynthetic bacterium Erythrobacter
longus. Phytochemistry 1991;30:3411–5.
Tank M, Blümel M, Imhoff JF. Communities of purple sulfur bacteria in a Baltic Sea coastal lagoon analyzed by pufLM gene
libraries and the impact of temperature and NaCl concentration in experimental enrichment cultures. FEMS Microbiol
Ecol 2011;78:428–38.
Tang K, Feng X, Tang YJ, et al. Carbohydrate metabolism and
carbon fixation in Roseobacter denitrificans OCh 114. PLoS One
2009;4:e7233.
Tomasch J, Gohl R, Bunk B, et al. Transcriptional response of
the photoheterotrophic marine bacterium Dinoroseobacter
shibae to changing light regimes. ISME J 2011;5:1957–68.
Van Mooy BAS, Devol AH, Keil RG. Relationships between
bacterial community structure, light, and carbon cycling
in the eastern subarctic North Pacific. Limnol Oceanogr
2004;49:1056–62.
Voget S, Wemheuer B, Brinkhoff T, et al. Adaptation of an abundant Roseobacter RCA organism to pelagic systems revealed
by genomic and transcriptomic analyses. ISME J 2015;9:371–
84.
Wagner-Döbler I, Ballhausen B, Berger M, et al. The complete
genome sequence of the algal symbiont Dinoroseobacter
shibae: a hitchhiker’s guide to life in the sea. ISME J 2010;4:61–
77.
Wagner-Döbler I, Biebl H. Environmental biology of the marine
Roseobacter lineage. Annu Rev Microbiol 2006;60:255–80.
Waidner LA, Kirchman DL. Aerobic anoxygenic photosynthesis
genes and operons in uncultured bacteria in the Delaware
River. Environ Microbiol 2005;7:1896–908.
Waidner LA, Kirchman DL. Aerobic anoxygenic phototrophic
bacteria attached to particles in turbid waters of the
Delaware and Chesapeake estuaries. Appl Environ Microb
2007;73:3936–44.
Waidner LA, Kirchman DL. Diversity and distribution of ecotypes of the aerobic anoxygenic phototrophy gene pufM in
the Delaware estuary. Appl Environ Microb 2008;74:4012–21.
Wakao N, Shiba T, Hiraishi A, et al. Distribution of bacteriochlorophyll a in species of the genus Acidiphilium. Curr Microbiol 1993;27:277–9.
Wang H, Tomasch J, Jarek M, et al. A dual-species co-cultivation
system to study the interactions between Roseobacters and
dinoflagellates. Front Microbiol 2014;5:311.
Wichlacz PL, Unz RF, Langworthy TS. Acidiphilium angustum sp.,
Acidiphilium facilis sp. nov., and Acidiphilium rubrum sp. nov.:
acidophilic heterotrophic bacteria isolated from acidic coal
mine drainage. Int J Syst Bacteriol 1986;36:197–201.
Yurkov V, Hughes E. Genes associated with the peculiar phenotypes of the aerobic anoxygenic phototrophs. Adv Bot Res
2013;66:327–58.
Yurkov VV, Beatty JT. Aerobic anoxygenic phototrophic bacteria.
Microbiol Mol Biol R 1998a;62:695–724.
Yurkov VV, Beatty JT. Isolation of aerobic anoxygenic photosynthetic bacteria from Black Smoker plume waters of the
Juan de Fuca Ridge in the Pacific Ocean. Appl Environ Microb
1998b;64:337–41.
Yurkov VV, Csotonyi JT. New light on aerobic anoxygenic phototrophs. In: Hunter CN, Daldal F, Thurnauer MC, et al. (eds).
The Purple Phototrophic Bacteria.Advances in Photosynthesis and
Respiration, Vol. 28. Dordrecht, the Netherlands: Springer,
2009, 31–55.
Yurkov VV, Gad’on N, Drews G. The major part of polar carotenoids of the aerobic bacteria Roseococcus thiosulfatophilus RB3 and Erythromicrobium ramosum E5 is not
bound to the bacteriochlorophyll a-complexes of the photosynthetic apparatus. Arch Microbiol 1993;160:372–6.
Yurkov VV, Gorlenko VM. Erythrobacter sibiricus sp. nov., a new
freshwater aerobic species containing bacteriochlorophyll a.
Mikrobiologiya 1990;59:120–6. (engl. transl.).
Yurkov VV, Gorlenko VM. New species of aerobic bacteria from
the genus Erythromicrobium containing bacteriochlorophyll a.
Mikrobiologiya 1992;61:248–55. (engl. transl.).
Yurkov VV, Krieger S, Stackebrandt E, et al. Citromicrobium bathyomarinum, a novel aerobic bacterium isolated from deep-sea
hydrothermal vent plume waters that contains photosynthetic pigment-protein complexes. J Bacteriol 1999;181:4517–
25.
Yurkov VV, Stackebrandt E, Holmes A, et al. Phylogenetic positions of novel aerobic, bacteriochlorophyll a-containing bacteria and description of Roseococcus thiosulfatophilus gen. nov.,
sp. nov., Erythromicrobium ramosum gen. nov., sp. nov., and
Erythrobacter litoralis sp. nov. Int J Syst Bacteriol 1994;44:427–
34.
Yurkov VV, van Gemerden H. Impact of light/dark regimen
on growth rate, biomass formation and bacteriochlorophyll
synthesis in Erythromicrobium hydrolyticum. Arch Microbiol
1993;159:84–9.
Yurkova N, Rathgeber C, Swiderski J, et al. Diversity, distribution and physiology of the aerobic phototrophic bacteria in
the mixolimnion of a meromictic lake. FEMS Microbiol Ecol
2002;40:191–204.
Yutin N, Suzuki MT, Béjà O. Novel primers reveal wider diversity
among marine aerobic anoxygenic phototrophs. Appl Environ
Microb 2005;71:8958–62.
Yutin N, Suzuki MT, Teeling H, et al. Assessing diversity and
biogeography of aerobic anoxygenic phototrophic bacteria
in surface waters of the Atlantic and Pacific Oceans using
the Global Ocean Sampling expedition metagenomes. Environ Microbiol 2007;9:1464–75.
Zeng Y, Feng F, Medová H, et al. Functional type-2 photosynthetic
reaction centers found in the rare bacterial phylum Gemmatimonadetes. P Natl Acad Sci USA 2014;111:7795–800.
Zeng Y, Kasalický V, Šimek K, et al. Genome sequences of
two freshwater betaproteobacterial isolates, Limnohabitans
species strains Rim28 and Rim47, indicate their capabilities
as both photoautotrophs and ammonia oxidizers (Genome
Announcement). J Bacteriol 2012;194:6302–3.
Zhang Y, Jiao N. Dynamics of aerobic anoxygenic phototrophic bacteria in the East China Sea. FEMS Microbiol Ecol
2007;61:459–69.
Zheng Q, Koblı́žek M, Beatty JT, et al. Evolutionary divergence
of marine aerobic anoxygenic phototrophic bacteria as seen
from diverse organizations of their photosynthesis gene
clusters. Adv Bot Res 2013;66:359–83.
Zubkov M. Photoheterotrophy in marine prokaryotes. J Plankton
Res 2009;31:933–8.