ArticlePDF Available

Nutrient Acquisition and the Metabolic Potential of Photoferrotrophic Chlorobi

Frontiers
Frontiers in Microbiology
Authors:

Abstract and Figures

Anoxygenic photosynthesis evolved prior to oxygenic photosynthesis and harnessed energy from sunlight to support biomass production on the early Earth. Models that consider the availability of electron donors predict that anoxygenic photosynthesis using Fe(II), known as photoferrotrophy, would have supported most global primary production before the proliferation of oxygenic phototrophs at approximately 2.3 billion years ago. These photoferrotrophs have also been implicated in the deposition of banded iron formations, the world’s largest sedimentary iron ore deposits that formed mostly in late Archean and early Proterozoic Eons. In this work we present new data and analyses that illuminate the metabolic capacity of photoferrotrophy in the phylum Chlorobi. Our laboratory growth experiments and biochemical analyses demonstrate that photoferrotrophic Chlorobi are capable of assimilatory sulfate reduction and nitrogen fixation under sulfate and nitrogen limiting conditions, respectively. Furthermore, the evolutionary histories of key enzymes in both sulfur (CysH and CysD) and nitrogen fixation (NifDKH) pathways are convoluted; protein phylogenies, however, suggest that early Chlorobi could have had the capacity to assimilate sulfur and fix nitrogen. We argue, then, that the capacity for photoferrotrophic Chlorobi to acquire these key nutrients enabled them to support primary production and underpin global biogeochemical cycles in the Precambrian.
Content may be subject to copyright.
fmicb-08-01212 July 4, 2017 Time: 16:2 # 1
ORIGINAL RESEARCH
published: 06 July 2017
doi: 10.3389/fmicb.2017.01212
Edited by:
Trinity L. Hamilton,
University of Cincinnati, United States
Reviewed by:
Katja Laufer,
Aarhus University, Denmark
Erin Field,
East Carolina University, United States
*Correspondence:
Sean A. Crowe
sean.crowe@ubc.ca
Specialty section:
This article was submitted to
Microbiological Chemistry
and Geomicrobiology,
a section of the journal
Frontiers in Microbiology
Received: 01 April 2017
Accepted: 14 June 2017
Published: 06 July 2017
Citation:
Thompson KJ, Simister RL,
Hahn AS, Hallam SJ and Crowe SA
(2017) Nutrient Acquisition
and the Metabolic Potential
of Photoferrotrophic Chlorobi.
Front. Microbiol. 8:1212.
doi: 10.3389/fmicb.2017.01212
Nutrient Acquisition and the
Metabolic Potential of
Photoferrotrophic Chlorobi
Katharine J. Thompson1, Rachel L. Simister1, Aria S. Hahn1, Steven J. Hallam1and
Sean A. Crowe1,2*
1Department of Microbiology and Immunology, University of British Columbia, Vancouver, BC, Canada, 2Departments of
Earth, Ocean and Atmospheric Sciences, University of British Columbia, Vancouver, BC, Canada
Anoxygenic photosynthesis evolved prior to oxygenic photosynthesis and harnessed
energy from sunlight to support biomass production on the early Earth. Models that
consider the availability of electron donors predict that anoxygenic photosynthesis
using Fe(II), known as photoferrotrophy, would have supported most global primary
production before the proliferation of oxygenic phototrophs at approximately 2.3 billion
years ago. These photoferrotrophs have also been implicated in the deposition of
banded iron formations, the world’s largest sedimentary iron ore deposits that formed
mostly in late Archean and early Proterozoic Eons. In this work we present new data
and analyses that illuminate the metabolic capacity of photoferrotrophy in the phylum
Chlorobi. Our laboratory growth experiments and biochemical analyses demonstrate
that photoferrotrophic Chlorobi are capable of assimilatory sulfate reduction and nitrogen
fixation under sulfate and nitrogen limiting conditions, respectively. Furthermore, the
evolutionary histories of key enzymes in both sulfur (CysH and CysD) and nitrogen
fixation (NifDKH) pathways are convoluted; protein phylogenies, however, suggest
that early Chlorobi could have had the capacity to assimilate sulfur and fix nitrogen.
We argue, then, that the capacity for photoferrotrophic Chlorobi to acquire these
key nutrients enabled them to support primary production and underpin global
biogeochemical cycles in the Precambrian.
Keywords: Chlorobi, Archean ocean, nitrogen, photoferrotrophy, sulfur
INTRODUCTION
Modern global primary production is supported through oxygenic photosynthesis, which converts
sunlight and CO2into biomass, fuelling the biosphere and driving fluxes of matter and energy
at global scales (Field et al., 1998). Primary production is limited by the availability of nutrients
that are essential for growth such as phosphorus, nitrogen, and sulfur (Howarth, 1988). Primary
producers thus expend valuable energy to meet their nutrient quotas. In the modern oceans, for
example, cyanobacteria can fix nitrogen in the photic zone to support their nitrogen requirements
(Karl et al., 1997). This in turn provides a competitive advantage that frequently allows nitrogen-
fixing cyanobacterial species like Trichodesmium to outcompete non-nitrogen fixing species and
can lead to cyanobacterial blooms (Capone and Carpenter, 1982;Capone et al., 2005). In addition
to their role as primary producers in the modern oceans, cyanobacteria play a key role in
the acquisition and redistribution of nutrients (Carpenter and Romans, 1991), driving global
biogeochemical cycles since their evolution and proliferation in the Precambrian Eons.
Frontiers in Microbiology | www.frontiersin.org 1July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 2
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
Oxygenic photosynthesis and cyanobacteria emerged early in
the Archean Eon (Crowe et al., 2013; Planavsky et al., 2014),
evolving from anoxygenic phototrophs (Xiong et al., 2000), which
arose as early as 3.8 Ga (Czaja et al., 2013). Like oxygenic
phototrophs, anoxygenic phototrophs fix carbon dioxide into
biomass, but instead of water as the electron donor they use
a diverse set of inorganic species [e.g., H2, H2S, and Fe(II)]
to replace electrons transferred from the photosystem to CO2
(Blankenship et al., 2006). Most anoxygenic phototrophs that
grow in illuminated anoxic waters today use reduced sulfur
species as their electron donors. During much of Earth’s early
history, however, reduced sulfur species were likely scarce and
the chemistry of marine sediments suggests that the oceans
were overwhelmingly iron-rich (ferruginous) for long stretches
of both the Archean and Proterozoic Eons (Canfield et al.,
2008;Planavsky et al., 2011;Poulton and Canfield, 2011). Under
these ferruginous conditions, ferrous iron would have been the
most abundant and available inorganic electron donor (Canfield
et al., 2006). Models for primary production in these ferruginous
oceans suggest that anoxygenic phototrophs using Fe(II) as
their electron donor—photoferrotrophs—could have supported
up to 10% of modern day primary production before the
proliferation of cyanobacteria (Canfield et al., 2006;Jones et al.,
2015). Together, the evolutionary history of the photosystem and
current knowledge on the history of ocean redox states imply
that photoferrotrophs could have played a key role in driving
global fluxes of matter and energy throughout the Precambrian
Eons.
Compelling, but indirect, evidence for photoferrotrophy
during Archean and Paleoproterozoic times comes from the
deposition of banded iron formations (BIFs) (Garrels et al.,
1973;Konhauser et al., 2002;Kappler et al., 2005). BIFs are
massive iron ore deposits that were mostly deposited toward the
end of the Neoarchean, though their deposition spans from the
Eoarchean through to the Neoproterozoic Eras (Klein, 2005).
Classical models for the deposition of iron from seawater to
form BIF invoke large-scale oxidation of seawater Fe(II) by
oxygen produced as a by-product of cyanobacterial growth
and the subsequent precipitation and sedimentation of ferric
iron minerals (Cloud, 1973;Garrels et al., 1973;Walker,
1987). Oxygen levels through the Archean, however, appear
too low to support oxidation of Fe(II) at rates sufficient to
sustain the rapid ferric Fe deposition needed to form even
some of the apparently small BIFs like the Isua Greenstone
belt in Greenland (Czaja et al., 2013). Instead, Fe(III) could
have come from abiotic photochemical iron oxidation through
UV photolysis (Garrels et al., 1973;Hartman, 1984), but this
also appears too slow to support ferric iron deposition at
rates recorded in BIFs (Konhauser et al., 2007). Alternatively,
direct photosynthetic iron oxidation through photoferrotrophy
could supply ferric Fe to form BIFs (Widdel et al., 1993;
Konhauser et al., 2002). Accepting that oxygen levels were too
low to drive Fe(II) oxidation and that UV photolysis appears
similarly ineffective, photoferrotrophy may be the only viable
mechanism to support appreciable ferric iron deposition and
BIF formation. Nevertheless, the role of photoferrotrophs in
BIF deposition remains controversial since direct evidence,
like lipid biomarkers in BIFs, to diagnose photoferrotrophy,
remain elusive. Extant cultures of photoferrotrophic bacteria
are thus employed in efforts to further test the possible role
of photoferrotrophs in BIF deposition and to identify signals
that might be used to diagnose photoferrotrophy in the rock
record.
A total of eight enrichments and isolates of photoferrotrophic
bacteria have been brought into laboratory collections over the
last 30 years. These cultures were largely obtained from a variety
of benthic environments, such as marine mud flats and freshwater
sediments (Widdel et al., 1993;Ehrenreich and Widdel, 1994;
Heising and Schink, 1998;Heising et al., 1999;Straub et al.,
1999), with a single isolate originating from a ferruginous
water column (Llirós et al., 2015). Laboratory cultures of
photoferrotrophs are distributed across the Alphaproteobacteria,
the Gammaproteobacteria, and the Chlorobi and experiments
conducted with these cultures reveal diverse physiological traits
that translate to differential growth rates across a wide range of
culture conditions (Kappler and Newman, 2004;Hegler et al.,
2008;Posth et al., 2010). Notably, under modest light availability,
many of these cultures grow sufficiently fast to oxidize Fe(II)
at rates that would support the deposition of some of the
largest BIFs (Konhauser et al., 2002;Kappler et al., 2005).
This gives confidence in the capacity of photoferrotrophs to
deposit BIFs, but laboratory culturing media are notoriously
nutrient rich. Natural settings, on the other hand, are typically
nutrient poor in comparison (Moore et al., 2013), and thus
the role of photoferrotrophs in both BIF deposition and
primary production would have depended on their capacity
to grow and acquire nutrients from Precambrian seawater at
concentrations almost certainly much lower than typical culture
media.
Many laboratory experiments have been conducted
with photoferrotrophs from the Alphaproteobacteria and
Gammaproteobacteria (Kappler and Newman, 2004;Hegler
et al., 2008;Posth et al., 2010;Bose and Newman, 2011;Pereira
et al., 2012), but the ecological relevance of these groups
in natural ferruginous settings is uncertain. In all modern
ferruginous environments supporting photoferrotrophy,
members of the Chlorobi appear to dominate (Crowe et al.,
2008;Walter et al., 2014;Llirós et al., 2015). Furthermore,
most or many extant photosynthetic communities dominated
by anoxygenic phototrophs are comprised mostly of Chlorobi
(Garrity et al., 2001). While anoxygenic photosynthesis by
the Proteobacteria likely evolved early (Xiong et al., 2000),
more recent phylogenomic analyses imply that the original
phototrophs belonged to the Chlorobi (Sadekar et al., 2006;
Bryant et al., 2012;Satoh et al., 2013). The reason for the
apparent prevalence of the Chlorobi in modern environments
is uncertain, but it is likely related to their ability to grow
under environmentally relevant conditions including low
nutrient availability and low light (Borrego and Garciagil,
1995;Manske et al., 2005;Gregersen et al., 2009;Romero-
Viana et al., 2010;Biderre-Petit et al., 2011;Crowe et al.,
2014a). Thus, despite the fact that photoferrotrophy by
Proteobacteria may be relevant to Precambrian ecosystems,
here, we focus our analyses on the Chlorobi because of
Frontiers in Microbiology | www.frontiersin.org 2July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 3
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
their apparent ecological prominence in many modern
systems and their deeper ancestry compared to phototrophic
Proteobacteria.
Both phosphorus and nitrogen often limit photosynthetic
activities and primary production in the modern oceans and
in freshwater environments (Howarth, 1988). Phosphorus is
generally considered the ultimate limiting nutrient on geological
time scales as nitrogen can be fixed from the atmosphere
when phosphorus is available (Tyrrell, 1999). Phosphorus is
essential for life and is required in phospholipid, nucleic acid,
and adenosine tri-phosphate (ATP) biosynthesis. Phosphorus
throughout the Precambrian Eons was scarce with seawater
concentrations orders of magnitude lower than today (Jones et al.,
2015). This phosphorus scarcity would have led to low primary
production, influencing the ecology and elemental stoichiometry
of the photosynthetic primary producers (Reinhard et al., 2016).
While phosphorus scarcity likely played an outsized role in
shaping the Precambrian biosphere, nitrogen scarcity may have
developed locally and transiently throughout the Precambrian
Eons (Canfield et al., 2010;Michiels et al., 2017). Nitrogen is
required to build essential cellular components such as DNA
and amino acids. Biologically available nitrogen is supplied
to the oceans through rock weathering and volcanism, but
ammonium uptake and ultimate burial, however, would have
eventually depleted the oceanic bioavailable nitrogen reservoir
(Johnson and Goldblatt, 2015). In the modern ocean, biological
fixation of atmospheric nitrogen keeps pace with phosphate
supplies over geologic time scales (Moore et al., 2009). Nitrogen
fixation is one of the most energetically expensive processes
in the metabolic repertoire of life and yet it is distributed
across distantly related groups of microorganisms (Meyer et al.,
1978;Raymond et al., 2004). This underscores the importance
of nitrogen fixation to microbial growth and production, is
consistent with the early evolution and radiation of nitrogen
fixation (Boyd et al., 2015;Stüeken et al., 2015;Weiss et al.,
2016), and exemplifies how the distribution of core metabolic
machinery across diverse lineages and functional guilds ensures
survival of essential biogeochemical functions over geologic
time (Falkowski et al., 2008). While the genomic potential for
nitrogen fixation exists within the Chlorobi (Bryant et al., 2012),
the capacity of photoferrotrophic Chlorobi to conduct nitrogen
fixation and thus support Precambrian marine nitrogen quotas
remains untested and unsubstantiated. This leaves our knowledge
of the possible ecological role that photoferrotrophs may have
played in the acquisition and redistribution of nitrogen and
its attendant biogeochemical cycling in the Precambrian oceans
entirely unknown.
In addition to phosphorus and nitrogen, sulfur is also essential
for life and can limit biological production and growth when
scarce (Da Silva and Williams, 2001). Sulfur on the modern
Earth is abundantly available as the fully oxidized sulfate ion
due to high concentrations of oxygen in the atmosphere and
oceans, which promotes oxidative sulfur weathering and the
recycling of sulfur from anoxic marine sediments. During the
Precambrian Eons, however, marine sulfate concentrations were
much lower (Walker and Brimblecombe, 1985;Crowe et al.,
2014b) likely due to limited oxidative weathering and recycling
under low O2atmospheres (Walker and Brimblecombe, 1985;
Habicht et al., 2002;Crowe et al., 2014b;Paris et al., 2014;
Zhelezinskaia et al., 2014). Instead, sulfur was likely scarce
and biologically available as low concentrations of sulfate,
very low concentrations of sulfide, and possibly organic sulfur
(Crowe et al., 2014b). Assimilatory sulfate reduction (ASR),
therefore, would have been a key nutrient acquisition pathway,
supporting primary production under low sulfur conditions.
The genomic potential for ASR has been detected within
two members the Chlorobi [Chlorobium ferrooxidans and
Chlorobium luteolum (Frigaard and Bryant, 2008)], yet the
role of ASR in photoferrotrophic growth remains uncertain.
Photosynthetic growth of C. ferrooxidans on ferrous iron and
without reduced sulfur compounds implies that the genomic
potential for ASR translates into physiological capacity to convert
sulfate into biomass sulfur (Heising et al., 1999). Given the
likely low sulfate and extremely low sulfide concentrations
perceived for the Precambrian oceans, ASR may have been
absolutely critical for photoferrotrophs to operate as primary
producers and contribute to a reservoir of biologically available
reduced sulfur compounds in the ocean. The evolutionary
history of ASR in the photoferrotrophic Chlorobi has not been
explored, nor has sulfate uptake been quantitatively assessed.
The role of photoferrotrophs in driving sulfur cycling during the
Precambrian remains underappreciated and untested creating
another gap in our knowledge of nutrient acquisition and
redistribution in the Precambrian oceans.
To address the response of photoferrotrophy to nitrogen and
sulfur scarcity, and to create new knowledge relevant to nitrogen
and sulfur acquisition and redistribution in the Precambrian
oceans, we examined two extant demonstrably photoferrotrophic
Chlorobi: benthic C. ferrooxidans (grown in co-culture with
Geospirillum sp. KoFum) (Heising et al., 1999), and pelagic
Chlorobium phaeoferrooxidans (Llirós et al., 2015;Crowe et al.,
2017). We also examined putative benthic photoferrotroph
C. luteolum, postulated to grow through photoferrotrophy
because of its genomic potential for ASR (Frigaard and Bryant,
2008). We verified the capacity of photoferrotrophic Chlorobi to
fix inorganic nitrogen and sulfur, and constrained the antiquity
of this capacity in the Chlorobi through phylogenetic analyses.
RESULTS AND DISCUSSION
Nitrogen Fixation
The process of fixing dinitrogen is kinetically challenging and
energetically expensive as it involves overcoming the activation
energy required in breaking the triple bond between the two
nitrogen molecules. The enzyme necessary for nitrogen fixation,
nitrogenase, is a multi-subunit protein that is assembled and
regulated by a series of other related proteins. All nitrogenases
require a metal ion cofactor – molybdenum, iron, or vanadium –
with each cofactor being recruited and incorporated into the
nitrogenase by a different set of proteins, Nif, Anf, and Vnf,
respectively. Current studies indicate that the majority of
nitrogenases depend on the molybdenum ion cofactor for their
enzymatic activity (reviewed in Rubio and Ludden, 2008), while
Frontiers in Microbiology | www.frontiersin.org 3July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 4
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
FIGURE 1 | Nitrogenase gene cassettes of the photoferrotrophic Chlorobi, detailing the position of each gene and the differences and similarities between the gene
cassettes.
the iron and vanadium dependant nitrogenases may play a role
in molybdenum limiting environments (Joerger et al., 1988).
Phylogenetic evidence suggests that the molybdenum-dependant
version of the enzyme evolved first (Boyd et al., 2011), which is
further supported by the observation that organisms identified as
having an iron or vanadium dependant nitrogenase all contain
a copy of the molybdenum-dependant nitrogenase (Raymond
et al., 2004;Soboh et al., 2010). There are up to 25 proteins,
depending on the species, required to assemble and regulate
the nitrogenase including three conserved structural proteins:
NifD, NifK, and NifH. NifH is often used as the marker gene
for nitrogen fixation in natural environments, due to its role
in the main enzyme structure and in cofactor recruitment.
Further phylogenetic information, however, can be obtained
when all three structural proteins (NifDKH) are concatenated
due to increased sequence information and the conserved nature
of all three proteins. Here we explored these key structural
proteins to test for the metabolic potential for nitrogen fixation
in the photoferrotrophic Chlorobi. We compare nitrogen fixation
in photoferrotrophic Chlorobi to the other members of the
phylum Chlorobi and to representatives from all phyla capable of
nitrogen fixation to assess the evolutionary history of nitrogenase
in relevant to photoferrotrophy in the Chlorobi and to place
constraints on the possible role of photoferrotrophs in supplying
fixed nitrogen to the Precambrian oceans.
Distribution of Nitrogen Fixation Pathways within
Chlorobi
Previous analyses of Chlorobi genomes identified that the
metabolic capacity for nitrogen fixation is distributed across the
phylum with the exception of the Ignavibacterium sp. (Bryant
et al., 2012;Liu et al., 2012;Hiras et al., 2016). Ignavibacterium
sp. is the deepest branching member of the Chlorobi and the only
class of non-photosynthetic organisms in the phylum. Here we
show that genes coding for the proteins required for nitrogen
fixation are present in the genomes of the photoferrotrophic
Chlorobi C. ferrooxidans and C. phaeoferrooxidans, putative
photoferrotroph C. luteolum (Figure 1), and in genomes of all
other members of the Chlorobi (data not shown). Specifically, we
identified one homolog of each of the molybdenum-dependant
nitrogenase proteins in all three photoferrotrophic Chlorobi. No
homologs of the alternative vanadium or iron-only nitrogenase
proteins were detected (PSI-Blast, expect threshold 10). These
results indicate that the photoferrotrophic Chlorobi have the
genomic capacity to fix nitrogen. Furthermore, nitrogen fixation
is wide spread among the Chlorobi, with all available Chlorobi
genome sequences coding the necessary proteins apart from
Ignavibacterium album.
Biochemical Verification of Nitrogen Fixation
To test for the biochemical capacity to fix nitrogen during
photosynthetic growth on Fe(II), nitrogen free (below limit
of detection ammonium, ammonia, nitrate, or nitrite) media
was inoculated with C. phaeoferrooxidans or C. ferrooxidans.
Both species were also grown in the standard growth medium
containing 5.6 mM ammonium (Hegler et al., 2008), for
comparison. Both species were able to fix nitrogen while
growing through photosynthetic Fe(II) oxidation with
doubling times of 45 and 36 h for C. phaeoferrooxidans and
C. ferrooxidans, respectively (Figures 2A,B). Fe(II) oxidation
Frontiers in Microbiology | www.frontiersin.org 4July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 5
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
FIGURE 2 | Fe(II) concentrations and cell counts over time for both C. phaeoferrooxidans (A,C) and C. ferrooxidans (B,D) under two sets of media: no bioavailable
nitrogen – N2as sole nitrogen source (A,B) and ammonium rich (C,D). Data points used to calculate growth rates and Fe(II) oxidation rates are highlighted in each
panel.
rates, during exponential growth phase, were 4.8 ±0.33 µM/hour
(C. phaeoferrooxidans) and 16 ±0.56 µM/hour (C. ferrooxidans)
(Figures 2A,B). Growth under ammonium-rich conditions
supported shorter doubling times (15 and 27 h) and higher rates
of Fe(II) oxidation (50 ±2.4 µM/hour and 23 ±0.7 µM/hour)
for C. phaeoferrooxidans and C. ferrooxidans, respectively
(Figures 2C,D). These results indicate that both pelagic
C. phaeoferrooxidans and benthic C. ferrooxidans are capable
of using dinitrogen gas as their sole source of nitrogen during
growth, but that the need to fix N decreases growth rates.
To further explore the metabolic capacity of photoferrotrophic
Chlorobi under both sets of conditions, cell specific Fe(II)
oxidation rates were calculated for each species. C. phaeoferro
oxidans oxidized Fe(II) at 21.2 ±1.4 fmol/cell while fixing
nitrogen and 47.8 ±2.3 fmol/cell under ammonium-rich
conditions. Conversely, C. ferrooxidans oxidized Fe(II) at
30.0 ±0.9 fmol/cell and 29.4 ±1.0 fmol/cell in ammonium free
and ammonium-rich media, respectively, with no appreciable
difference during N-fixation. The apparent insensitivity of
C. ferrooxidans to N-availability may be related to the presence
of its co-culture partner, Geosprillum sp. KoFum. Further
experiments with KoFum could help constrain its possible
role in N metabolism within the co-culture The observation
that C. phaeoferrooxidans has lower cell specific growth rates
under N scarcity, however, implies lower growth yields during
N fixation. Both species are ultimately capable of growth
Frontiers in Microbiology | www.frontiersin.org 5July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 6
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
TABLE 1 | Gene length (bp), codon adaptation index (CAI), and GC content (%) for each of the genes in the nitrogenase cassette.
Gene C. phaeoferrooxidans C. ferrooxidans C. luteolum
Length (bp) CAI GC content (%) Length (bp) CAI GC content (%) Length (bp) CAI GC content (%)
NifB 1275 0.80 52.63 1275 0.76 53.18 1263 0.75 60.89
NifN 1353 0.79 53.22 1350 0.75 53.48 1353 0.74 59.42
NifE 1362 0.79 50.07 1362 0.75 49.63 1362 0.74 57.34
NifK 1383 0.81 53.51 1383 0.73 52.78 1380 0.76 60.14
NifD 1635 0.79 49.54 1635 0.76 49.66 1641 0.77 57.22
PII regulator 378 0.81 50.00 378 0.72 50.00 378 0.72 59.26
PII regulator 357 0.76 49.30 357 0.73 48.74 357 0.70 56.30
NifH 825 0.83 49.58 825 0.79 49.21 825 0.81 59.03
Each parameter was calculated for all three photoferrotrophic Chlorobi, whose whole genome GC contents are: 49.72% for Chlorobium phaeoferrooxidans, 49.9% for
C. ferrooxidans, and 58.1% for C. luteolum.
FIGURE 3 | Phylogenies of the Chlorobi and Bacteroidetes using (A) the concatenated NifDKH proteins and (B) 16S rRNA with bootstrap values shown at each
node (maximum likelihood/maximum parsimony). The blue colors delineate the organisms of the Phylum Chlorobi, with each shade representing a different genus,
while the purple color delineates the Phylum Bacteroidetes. The orange lines indicate the position of the photoferrotrophic Chlorobi. The trees were rooted with four
cyanobacterial organisms. Note: Azobacteroides pseudotrichonymphae CFP2 is abbreviated from Candidatus Azobacteroides pseudotrichonymphae genomovar.
CFP2.
and Fe(II) oxidation while fixing nitrogen but the differential
response of cell specific iron oxidation rates to N-scarcity
implies that nutrient availability can influence the ecology of
photoferrotrophs in the environment.
Evolutionary History of Nitrogen Fixation in the
Chlorobi
To assess the evolutionary history of nitrogen fixation in the
Chlorobi we tested for horizontal gene transfer (HGT) within
the photoferrotrophic Chlorobi and conducted phylogenetic
analyses of Nif proteins, which we compared to small subunit
16S ribosomal RNA (SSU rRNA) genes. Deviations in the
branching orders between these phylogenies would indicate non-
vertical inheritance and HGT. To test for horizontal transfer
of Nif genes in the photoferrotrophic Chlorobi, we looked
for characteristic signatures of HGT within nif gene cassettes.
Codon adaptation index (CAI) values, a metric used to describe
differences in codon usage between specific genes and the
genomic background, were calculated for all individual nif
genes belonging to C. ferrooxidans,C. phaeoferrooxidans, and
C. luteolum. All CAI values were greater than the threshold
value, 0.70, below which HGT is indicated (Table 1). In addition,
GC contents of nif genes were very similar to GC contents of
genomic backgrounds providing no evidence for HGT (Table 1).
Our analyses also failed to identify tRNAs, transposases, or
other genetic elements commonly associated with gene mobility
in close proximity (within 5000 bp) to the nitrogenase gene
cassette in any of the photoferrotrophic Chlorobi. The general
lack of tRNAs or transposases near the nif cassettes in the
photoferrotrophic Chlorobi, combined with super threshold CAI
values and nif gene GC contents that are homogenous against
genomic backgrounds, imply nif gene acquisition through
vertical decent.
To test the evolutionary history of the nif genes in the
Chlorobi, we conducted phylogenetic analyses of concatenated
NifDKH proteins of all cultured and sequenced Chlorobi that
have the genomic potential to fix dinitrogen. The Chlorobi
sequences were aligned with selected sequences from the
next closest phylum – Bacteroidetes – and the tree was
rooted using four Cyanobacterial species as an out-group
(Figure 3A). The genus Chlorobium, which includes all three
photoferrotrophic Chlorobi, the genus Chlorobaculum, and
Frontiers in Microbiology | www.frontiersin.org 6July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 7
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
FIGURE 4 | Phylogenies of (A) the concatenated NifDKH proteins and (B) 16S rRNA for two to four representatives of several nitrogen-fixing phyla with bootstrap
values shown at each node (maximum likelihood/maximum parsimony). The green color delineates the Chlorobi/Bacteroidetes monophyletic grouping, while the
orange line indicates the position of the photoferrotrophic Chlorobi. Note: Azobacteroides pseudotrichonymphae CFP2 is abbreviated from Candidatus
Azobacteroides pseudotrichonymphae genomovar. CFP2.
the genus Prosthecochloris all form monophyletic groups that
are collectively part of the phylum Chlorobi clade. Likewise,
the Bacteroidetes form a monophyletic group and share a
common ancestor with the members of the phylum Chlorobi.
Furthermore, when the NifDKH tree is compared to a 16S
rRNA tree of the same organisms (Figure 3B), all of the
genera within the phylum Chlorobi branch in an identical
order to those in the 16S rRNA phylogeny. The phylogenetic
relationship between the Chlorobi and Bacteroidetes is the same
for both NifDKH and 16S rRNA sequences, indicating that
the common ancestor to the phyla Chlorobi and Bacteroidetes
likely contained a nitrogenase and therefore the ability to fix
dinitrogen. Ignavibacterium sp., the sole members of the phylum
Chlorobi who do not posses a nitrogenase, likely lost the
capability to fix nitrogen as the remainder of the Chlorobi and
the phylum Bacteroidetes bracket the phylogenetic position of the
Ignavibacterium sp. Taken together, available data imply vertical
decent.
Accepting largely vertical descent of NifDKH from the
common ancestor of the Chlorobi and Bacteroidetes, NifDKH
must have emerged within this line of descent before the
divergence of the Chlorobi and Bacteroidetes. The timing of
this divergence has been estimated using a whole genome
molecular clock (David and Alm, 2011) to between 3 and
1.6 Gya, which implies the capacity to fix N in the ancestors
of the Chlorobi before this time. Independent N isotope data
from metasedimentary kerogen implies N fixation by at least
3.2 Gy (Stüeken et al., 2015). Combined, the evidence for vertical
inheritance of NifDKH in the Chlorobi on the taxonomic levels
of genus and phylum, the timing of divergence between the
Chlorobi and the Bacteroidetes, and the N isotope record, imply
that ancestors of modern Chlorobi likely had capacity to fix
nitrogen in the iron-rich oceans of the paleoproterozoic and
perhaps as early as the mesoarchean eras.
To place N fixation in the Chlorobi, and Bacteroidetes, within
the broader context of nitrogenase evolution in general, we
conducted further phylogenetic analyses using a greater diversity
of organisms. We analyzed the NifDKH phylogeny using two
to four representatives from every phylum that had a cultured
and sequenced species with previously documented genomic
potential for nitrogen fixation (Figure 4A). This phylogeny places
the Nif proteins found in the Chlorobi and Bacteroidetes in
a single clade, supporting their emergence from a common
ancestor and the vertical inheritance of NifDKH from this
ancestor. The phylogeny of the NifDKH protein is, however,
incongruent with that of the 16S rRNA gene from the same
organisms (Figure 4B). While the Chlorobi and Bacteroidetes
group together in both phylogenies, the Spirochetes, Chloroflexi,
and Firmicutes also group with the Chlorobi in the NifDKH
phylogeny, but belong to distinct clades in the 16S rRNA gene
phylogeny. The differences between these phylogenies confound
further constraints on the evolutionary history of nitrogenase
within the Chlorobi based on phylogeny and add to the
overwhelming evidence for horizontal transfer of NifDKH genes
(Raymond et al., 2004;Falkowski et al., 2008;Boyd and Peters,
2013).
Ecology of Nitrogen Fixation in Chlorobi, Past and
Present
Members of the phylum Chlorobi underpin biological
production in many modern anoxic environments, both
sulfidic (Overmann, 1997;Tonolla et al., 2004;Gregersen et al.,
2009;Kondo et al., 2009;Meyer et al., 2011) and ferruginous
(Walter et al., 2014;Llirós et al., 2015), through their ability to
Frontiers in Microbiology | www.frontiersin.org 7July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 8
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
harness light energy and fix inorganic carbon into biomass, even
at low light intensities (Manske et al., 2005). Chlorobi further
contribute to biogeochemical cycling in these systems through
the acquisition and redistribution of essential nutrients, such
as nitrogen. This ecological role would have extended to global
scales in the low oxygen Precambrian oceans. Our analyses
confirm the genomic potential to fix N in all but one of the
Chlorobi lineages and directly demonstrate the capacity of the
photoferrotrophic Chlorobi to fix dinitrogen as their sole source
of nitrogen while oxidizing Fe(II). Rates of Fe(II) oxidation
are, however, slower when photosynthetic growth is supported
through N-fixation rather than ammonium assimilation. To test
the impact of slower rates of Fe(II) oxidation, and therefore
growth, on the deposition of BIFs, we ran our cell counts
and Fe(II) oxidation rates through the calculation outlined
by Konhauser et al. (2002). Our data indicates that both
photoferrotrophic strains would be capable of generating even
the largest BIFs (i.e., the Hamersley BIF) with maximum of
2.44 ×103photoferrotrophic cells/mL required in the basin.
Thus, photoferrotrophic growth coupled to N-fixation could
support BIF deposition, even in the face of nitrogen scarcity.
Chlorobium phaeoferrooxidans, and C. ferrooxidans exhibit
differential responses to N scarcity that manifest in different
cell specific Fe(II) oxidation rates and different ratio’s between
microbial growth (cell doubling times) and Fe(II) oxidation.
C. phaeoferrooxidans has a cell doubling time to Fe(II)
oxidation ratio of 9.4 under N-fixing conditions compared to
0.3 when there is ample ammonium, whereas C. ferrooxidans
has comparable ratio’s of 2.3 and 1.2 for N-fixing and
ammonium-rich conditions comparatively. This differential
response indicates that under ammonium-rich conditions
C. phaeoferrooxidans grows more efficiently (i.e., with a
higher growth yield) whereas when N-fixation is required
C. ferrooxidans grows more efficiently. This creates niches for
each microorganism defined by N availability. The differential
response also implies that the stoichiometry of Fe-oxidation to
biomass production and cell growth is partly decoupled and
depends on N availability. Essentially, this decoupling means
that more Fe(II) is oxidized to produce an individual cell during
growth supported by N-fixation than by ammonium assimilation.
Such a decoupling thus requires either the diversion of reducing
equivalents (NADH) produced during photosynthesis into
compounds not used directly in cell growth, or that cell growth
and division requires more fixed carbon during N-fixation. The
former could include conversion of N2to ammines and the
biosynthesis of cell exudates, and the latter might include the
biosynthesis of cellular proteins needed to conduct N-fixation.
Such a decoupling would influence the overall biogeochemical
functioning and ecology of ecosystems supported through
primary production by photoferrotrophy. The overall activity
of the marine biosphere through the Precambrian Eons may
thus have been influenced by the availability of fixed N to
photoferrotrophs.
Assimilatory Sulfate Reduction (ASR)
Sulfate ions are biologically inert and organisms expend
tremendous energy ‘activating’ sulfate for three main functions:
(1) reduction and incorporation into amino acids; (2)
condensation and incorporation into sulfolipids and other
small molecules; and (3) for dissimilatory sulfate respiration.
In addition to the reduction of sulfate, organisms can acquire
organic sulfur compounds like amino acids, and hydrogen sulfide
directly from the environment. Acquisition of these reduced
sulfur compounds can considerably reduce the expenditure of
energy on sulfur acquisition. Here we focus on the first two
assimilatory pathways and the capacity for reductive sulfur
assimilation in the photoferrotrophic Chlorobi. The proteins
required to complete an entire ASR pathway include: CysD, the
sulfateadenyl transferase that activates sulfate to form APS; CysN
which catalyzes GTP hydrolysis providing the energy needed to
adenylate imported sulfate; CysC (a domain of CysN), the APS
kinase that phosphorylates APS to PAPS; and CysH, the APS
reductase which reduces the sulfur in APS to sulfite. We have
explored the metabolic potential for sulfate assimilation in the
genomes of the photoferrotrophic Chlorobi and directly tested
sulfate incorporation into biomass.
Distribution of ASR Pathways within Chlorobi
Previous analyses of Chlorobi genomes identified the metabolic
capacity for ASR in C. ferrooxidans and C. luteolum (Frigaard
and Bryant, 2008). Using all currently available genomic
information we identified components of the ASR pathways
distributed throughout the Chlorobi (Table 2). We find
that the photoferrotrophic Chlorobi, C. ferrooxidans and
C. phaeoferrooxidans, as well as putative photoferrotroph
C. luteolum, all possess the necessary proteins for ASR – CysD,
CysN/C, CysH – and therefore have the potential capacity
to synthesize amino acids from exogenous sulfate (Figure 5).
Notably, the presence of both CysD and CysN indicate that
sulfate activation to APS in these Chlorobi is coupled to
GTP hydrolysis. Sulfate assimilation in the Chlorobi, therefore,
offsets the energetic expense associated with sulfate activation.
The presence of CysN/C indicates the metabolic potential to
phosphorylate APS to PAPS implying that these strains might
have capacity to synthesize sulfate-containing compounds like
sulfolipids. Finally, while components of assimilatory sulfate
metabolisms are more broadly distributed throughout the
Chlorobi, genes coding for key components of the pathway are
mostly missing implying a lack of capacity for sulfate assimilation
outside the photoferrotrophic Chlorobi (Table 2). Given the
metabolic potential for ASR in the photoferrotrophic Chlorobi,
we sought to biochemically verify this process.
Biochemical Verification of ASR
Chlorobium phaeoferrooxidans and C. ferrooxidans are both
known to grow in media where sulfur is supplied exclusively
in the form of sulfate, which directly demonstrates the
physiological capacity for sulfate assimilation. We quantitatively
tested this capacity by measuring the uptake of 35S labeled
sulfate in low sulfate growth media. C. phaeoferrooxidans,
indeed took up 35S labeled sulfate into TCA extractable
biomass, demonstrating assimilatory reduction of sulfate and its
incorporation into amino acids. Over the course of these sulfate
uptake experiments, C. phaeoferrooxidans oxidized 3070 µM
Frontiers in Microbiology | www.frontiersin.org 8July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 9
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
TABLE 2 | Assimilatory (orange) and dissimilatory (blue) sulfur proteins present in the genomes of green sulfur bacteria.
Organism CysD Sat CysNC CysH AprAb DsrAB
Chlorobium phaeoferrooxidans KB01 + − + +
Chlorobium ferrooxidans DSM 13101 + − + +
Chlorobium luteolum DSM 273 + − + + +
Chlorobium phaeovibrioides DSM 265 + − + +
Prosthecochloris aestuarii DSM 271 + − + +
Chlorobium phaeobacteroides BS1 − + + + +
Chlorobium chlorochromatii CaD3 − + + + +
Pelodictyon phaeoclathratiforme − + + +
Chlorobium tepidum TLS − + + +
Chlorobium phaeobacteroides DSM 266 − − +
Chlorobium limicola DSM 245 − − +
Chlorobaculum parvum NCIB 8327 − − +
Chloroherpeton thalassium − −
Ignavibacterium album − −
FIGURE 5 | Assimilatory sulfate reduction (ASR) gene cassettes for the photoferrotrophic Chlorobi, detailing the position of each gene and the differences and
similarities between the gene cassettes.
Fe(II). This implies the fixation of 770 µM C, based on
the 4:1 stoichiometry between Fe(II) oxidation and C fixation
observed for C. phaeoferrooxidans during growth on Fe(II),
and for photoferrotrophic organisms, more generally (Widdel
et al., 1993). A corresponding total of 3 µM S was fixed
demonstrating a ratio of 260:1 C to S, which we take as
approximately indicative of the S content of C. phaeoferrooxidans.
There are few data to compare with, but our results suggest
that C. phaeoferrooxidans has relatively low S quotas compared
to aquatic and cultured bacteria (C:S from 10–60) (Fagerbakke
et al., 1996) and particulate organic matter from the North
Pacific (C:S of 50) (Chen et al., 1996). By analogy to
C. phaeoferrooxidans, photoferrotrophic Chlorobi likely have
capacity to fix sulfate into biomass under low sulfate conditions,
Frontiers in Microbiology | www.frontiersin.org 9July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 10
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
which they appear well adapted to do based on minimal cellular
sulfur quotas in comparison to other bacteria and marine organic
material.
Evolutionary History of ASR in the Chlorobi
To test for horizontal transfer of ASR genes to the
photoferrotrophic Chlorobi, we searched for characteristic
signatures of HGT within the ASR cassettes and conducted
phylogenetic analyses of ASR genes, which we compared to 16S
rRNA gene phylogenies. The CAI value for each of the ASR
genes belonging to C. ferrooxidans and C. phaeoferrooxidans
were all greater than the threshold value, 0.70, below which HGT
is indicated (Table 3). ASR genes in C. luteolum, however, had
sub-threshold CAI values, as low as 0.53, indicating possible
ASR gene acquisition through horizontal transfer. The GC
contents of ASR genes for all three species were very similar
to GC contents of their respective genomic backgrounds,
providing no evidence for HGT (Table 3). Collectively, these
data provide little evidence for the lateral acquisition of ASR
gene cassettes in the photoferrotrophic Chlorobi, although the
evidence for vertical descent is greater in C. phaeoferrooxidans
and C. ferrooxidans than in C. luteolum. A single transposase
(Figure 5) was found on a contig adjacent to that hosting the ASR
gene cassette in C. phaeoferrooxidans. The general lack of tRNAs
or transposases near the ASR cassettes in the photoferrotrophic
Chlorobi combined with super threshold CAI values and ASR
gene GC contents that are homogenous against the genomic
backgrounds, implies ASR gene acquisition through vertical
decent.
To further test the evolutionary history of ASR, the CysH
protein was analyzed to examine the phylogenetic relationship
between the proteins used in the photoferrotrophic Chlorobi
and ASR in other organisms. The photoferrotrophic Chlorobi
grouped together forming a monophyletic clade within the
CysH phylogeny (Figure 6A). The photoferrotrophic Chlorobi
exhibit congruent phylogenies between the CysH protein and
16S rRNA gene (Figure 6B), providing further evidence in
support of vertical inheritance of the ASR pathway in the
photoferrotrophic Chlorobi. The more general evolutionary
history of the CysH protein, however, is convoluted given
abundant incongruences between the CysH protein and 16S
rRNA gene phylogenies. Accepting vertical inheritance of the
ASR pathway in the photoferrotrophic Chlorobi and the early
divergence of the Chlorobi from other organisms, we hypothesize
that gene loss explains the lack of a complete ASR pathways in
other photosynthetic Chlorobi and this hypothesis is supported
by the partial presence of ASR pathway components across the
phylum Chlorobi (Table 2).
Ecology of ASR in Chlorobi, Past and Present
The presence of an ASR pathway in all known photoferrotrophic
Chlorobi implies that ASR is advantageous to growth under
ferruginous conditions. The lack of the ASR pathway in the
canonically sulfur oxidizing Chlorobi makes sense in light of
the availability of reduced sulfur compounds in their preferred
habitats. The energetic expense of ASR would tend to favor
assimilation of reduced compounds when available. Conversely,
ferruginous environments are by definition sulfur poor and the
availability of reduced sulfur compounds can be limited by the
solubility of FeS. Sulfate, therefore, is likely the most abundant
and available sulfur source in modern ferruginous environments.
The rock record also demonstrates that ferruginous marine
conditions persisted throughout much of the Precambrian
Eons and reduced sulfur species were likely scarce with the
exception of in the apparently ephemeral developments of costal
euxinia. ASR may thus have supported sulfur requirements
of photoferrotrophic primary producers over long stretches of
Earth’s history.
The apparent role of ASR in supporting primary production
through photoferrotrophy implies that sulfate availability could
have been an important control on global productivity. At
28 mM, sulfate is the principle anion in modern seawater, but
sulfate concentrations could have been as low as a few µM
in the Archean oceans (Crowe et al., 2014b). Nutrients like
phosphorus and nitrogen are known to become limiting at such
low concentrations. The apparently low sulfur quotas of the
photoferrotrophic Chlorobi (260:1, C:S) thus seem well adapted
to growth in the low sulfate oceans of the Archean, which would
have enhanced productivity in the face of sulfur scarcity.
Under low sulfate conditions dissimilatory sulfate reduction
(DSR) would have played a comparatively small role in the
remineralization of organic matter in Archean oceans (Crowe
et al., 2014b). Qualitatively then ASR would have played an
outsized role in the reduction of sulfur and the global sulfur
cycle in the Archean oceans, relative to today. We therefore
hypothesize that primary production through photoferrotrophy
was a key pathway in the production of an organic reduced
sulfur pool, which would have provided an important vector
for sulfur to Archean sediments. We further hypothesize that
ASR may have predated DSR. Earliest evidence for DSR comes
from S-isotope fractionation recorded in 3.47Gya barites (Shen
et al., 2001), whereas photoferrotrophy likely operated as early
as 3.8Gya (Czaja et al., 2013) and presumably required ASR. The
idea that ASR predates DSR could be tested if homology could
be established in enzymes involved in both pathways. Although
ASR and DSR serve different functions – sulfate acquisition
versus energy transduction, respectively – both pathways actively
transport sulfate into the cell and the first enzymes in the
pathways are thus analogous. Comparison of amino acid
sequences of the first enzymes (CysD and Sat, respectively) in the
two pathways indicates a strong degree of homology implying
evolutionary relationships between components of ASR and
DSR.
To examine the phylogenetic relationships between enzymes
that transport sulfate for use in ASR and DSR, we aligned
CysD and Sat proteins from the majority of the Chlorobi and a
selection of representative microorganisms from diverse phyla.
The resulting phylogeny clearly separated amino acid sequences
annotated as CysD from those annotated as Sat (Figure 7). Both
CysD and Sat appear to support sulfate transport in relation to
multiple sulfur metabolisms, but the phylogenetic relationships
appear complicated and likely require more detailed analyses.
Nevertheless, homology between the two proteins implies a
possible evolutionary relationship between ASR and DSR that
Frontiers in Microbiology | www.frontiersin.org 10 July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 11
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
TABLE 3 | Gene length (bp), CAI, and GC content (%) for each of the genes in the ASR cassette.
Gene C. phaeoferrooxidans C. ferrooxidans C. luteolum
Length (bp) CAI GC content (%) Length (bp) CAI GC content (%) Length (bp) CAI GC content (%)
CysH 714 0.78 54.34 714 0.78 54.62 753 0.59 56.97
CsyD 882 0.80 56.12 882 0.78 55.56 915 0.61 57.38
CsyN/C 1800 0.81 54.28 1800 0.77 53.67 1800 0.69 58.06
Siroheme synthase 453 0.81 54.08 453 0.71 52.98 453 0.53 57.17
Uroporphyrin-III
C-methyltransferase
1287 0.73 57.96 1287 0.73 55.40 258 0.62 57.36
CysA 1074 0.77 53.26 1074 0.80 52.42 1074 0.67 59.22
CysW 870 0.79 53.22 870 0.79 52.76 870 0.64 58.74
CysT 834 0.79 52.64 834 0.81 53.36 834 0.66 58.03
Sulfate transporter 1020 0.82 53.14 1020 0.79 52.65 1008 0.75 59.52
Each parameter was calculated for all three photoferrotrophic Chlorobi, whose whole genome GC contents are: 49.72% for C. phaeoferrooxidans, 49.9% for
C. ferrooxidans, and 58.1% for C. luteolum.
FIGURE 6 | Phylogenies of (A) the CysH protein and (B) 16S rRNA with bootstrap values shown at each node (maximum likelihood/maximum parsimony). The
orange line indicates the position of the photoferrotrophic Chlorobi. The trees are rooted with two Archaeal species.
may inform evolutionary histories and should be tested in the
future.
OUTLOOK
Photoferrotrophy links the C and Fe biogeochemical cycles
through coupled CO2fixation and Fe(II) oxidation and has
likely done so since the early Archean Eon. Models for
photoferrotrophic growth in the Archean oceans remain poorly
constrained as they are extrapolated from growth rates in
nutrient rich laboratory culture media. Here we demonstrate
that photoferrotrophic Chlorobi have the physiological capacity
to fix inorganic N and S into biomass when availability of
these nutrients is low and have likely had this capacity since
the Archean Eon. Thus, under N and S limited ferruginous
conditions, photoferrotrophy underpins biogeochemical cycling
of C, N, S, and Fe. Nutrient availability, however, influences
growth and Fe(II) oxidation rates and has consequences for
the stoichiometric relationships between C, N, S, and Fe
transformations. Undoubtedly, these relationships should be
assessed and studied in more detail with additional physiological
experimentation and should be applied to further constrain
models of photoferrotrophy, biological production, and global
biogeochemical cycling in the Archean Eon.
MATERIALS AND METHODS
Strains and Growth Medium
Media was prepared after Hegler et al. (2008), and allocated into
serum bottles (100 mL media and 160 mL total volume), with
0.3 g/L NH4Cl, 0.5g/L MgSO4·7H2O, 0.1g/L CaCl2·2H2O, and
0.6g/L KH2PO4. After autoclaving, 22 mmol L1bicarbonate,
trace elements, mixed vitamin solution, selenate-tungstate,
vitamin B12, and FeCl2were added and the pH was adjusted
Frontiers in Microbiology | www.frontiersin.org 11 July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 12
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
FIGURE 7 | Phylogeny of the Sat/CysD protein with bootstrap values shown at each node (maximum likelihood/maximum parsimony), to compare the ASR and
DSR pathway among a diverse set of organisms. The dashed line delineates the organisms with CysD versus those with Sat. The orange line indicates the position
of the photoferrotrophic Chlorobi.
to 6.8–6.9 under an N2/CO2atmosphere (80:20). 10 mmol L1
FeCl2was added to all media (regular, NH4+deplete, and SO4
poor) – Fe(II) concentrations from 200 µmol L1to 10 mmol
L1have been shown to produce the same growth rates under
nutrient rich conditions. The 10 mmol L1media was filtered
after being made to remove any precipitates, which resulted in a
final Fe(II) concentration of 2 mmol L1for the standard media
and 4mmol L1for the NH4+deplete media. The low SO4
media was left unfiltered with an Fe(II) concentration of 10 mmol
L1. In the ammonium free media, NH4Cl was replaced with
0.3 g/L KCl and an additional 10 mL of N2gas was injected into
the headspace. In the low sulfate media, 0.0025 g/L MgSO4and
0.4 g/L MgCl2were added instead of the usual 0.5 g/L MgSO4.
Furthermore, approximately 10 kBq of carrier-free 35S was added
to all of the low sulfate cultures. The cultures for the N-fixation
experiments were grown in ammonium free conditions once and
then transferred into the final experimental bottles. The culture
for the S35 experiment was grown up in standard media, spun
down and decanted to avoid adding extra sulfate, before the cells
were inoculated into the final experimental bottles. All cultures
were grown under a constant light intensity of 14 µE m2s1.
Analytical Techniques
Spectrophotometric analysis of Fe(II) and Fe(III) concentrations
were performed using the ferrozine method; samples were
measured directly as well as after being fixed in 1 N
HCl – after Viollier et al. (2000). Pigments were measured
spectrophotometrically after 24 h extractions of 1 mL of pelleted
cells in acetone:methanol (7:2 v/v) (Frigaard et al., 1996).
Cells numbers were then obtained using a pigment to cell
count conversion factor of 6.3 ×1010 pigment/cell/mL for
C. phaeoferrooxidans and 5.8 ×1010 pigment/cell/mL for
C. ferrooxidans. The cells from the 35S experiment were collected
via filtration along with a liquid sample as a background
measurement. The filtered samples were subsequently washed
with 5% Trichloroacetic acid (TCA) in order to kill, wash, and
dissolve cellular material. TCA precipitates DNA and proteins,
leaving only these cellular components on the filter and therefore
any counts associated with the filtered samples would indicate 35S
that had been incorporated into this cellular biomass (Cuhel et al.,
1981). Five milliliter of scintillation fluid were added to the 35S
samples (1 mL of liquid or the filter) and all samples were counted
using a scintillation counter.
Bioinformatics
Genomes of Chlorobi stains used in this paper were retrieved
from NCBI under the following accession numbers with
the completion percentage of each genome in brackets after
the number: NC_008639.1 (99.45%), NZ_AASE00000000.1
(90.71%), NC_007514.1 (97.8%), NC_009337.1 (98.91%),
NC_010803.1 (99.98%), NC_002932.3 (97.8%), NC_011027.1
(98.89%), and NC_007512.1(98.91%). Genomes were
analyzed using MetaPathways V2.5.1, an open source
pipeline for predicting reactions and pathways using default
Frontiers in Microbiology | www.frontiersin.org 12 July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 13
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
settings1(Konwar et al., 2013, 2015) and using the following
databases: MetaCyc-v4-11-07-03 (Caspi et al., 2012), Kyoto
Encyclopedia of Genes and Genomes (KEGG-11-06-18)
(Kanehisa et al., 2008), SEED-14-01-302, Clusters of Orthologous
Groups (COG-13-12-27) (Kaufmann, 2006), Carbohydrate-
Active enZYmes (CAZY-14-09-04) (Cantarel et al., 2009), and
RefSeq-nr-14-01-18 (Agarwala et al., 2015) databases. Initially,
we identified all sequences with a functional assignment affiliated
with nitrogen fixation and assimilatory sulfur reduction using the
MetaPathways functional annotation table output.
Phylogenetic Trees for Nitrogen Fixation
Individual NifDKH gene sequences from all organisms outside
of the phylum Chlorobi were retrieved from NCBI searches
from described strains, concatenated, and then aligned using
the package software ClustalX2.1 (Larkin et al., 2007). To
rigorously test the evolutionary history of nitrogen fixation
multiple tree construction methods [Maximum likelihood (ML)
and Maximum parsimony (MP)] were employed. ML and MP
trees were constructed in MEGA version 7 (Tamura et al.,
2013;Kumar et al., 2016) and all trees bootstrapped 500 times.
Bootstrap values are indicated at the nodes.
Phylogenetic Trees for Assimilatory
Sulfate reduction
CysH and CysD/Sat gene sequences from all organisms outside
of the phylum Chlorobi were retrieved from NCBI searches
from described strains, and then aligned using the package
software ClustalX2.1 (Larkin et al., 2007). To rigorously test the
evolutionary history of ASR multiple tree construction methods
1https://github.com/hallamlab/metapathways2/wiki
2http://www.theseed.org/
(ML and MP) were employed. ML and MP trees were constructed
in MEGA version 7 (Tamura et al., 2013;Kumar et al., 2016)
and bootstrapped 500 times. Bootstrap values are indicated at the
nodes.
Phylogenetic Trees for 16S rRNA
16S rRNA sequences were retrieved from strains used in Nif
and ASR gene trees from the Silva online database – version
128 (Pruesse et al., 2007;Quast et al., 2012). Only full-length
(>1400 bp) sequences were selected, and these were aligned,
using the package software ClustalX2.1 (Larkin et al., 2007). To
rigorously test the evolutionary history of nitrogen fixation and
ASR multiple tree construction methods (ML and MP) were
employed. ML and MP trees were constructed in MEGA version
7 (Tamura et al., 2013) and all trees bootstrapped 500 times.
Bootstrap values are indicated at the nodes.
AUTHOR CONTRIBUTIONS
KT performed all laboratory work, except for biochemical
verification of assimilatory sulfate reduction, which was
performed by SC; KT, RS, and SC interpreted and analyzed the
data with bioinformatic data analysis conducted by AH; KT and
SC wrote the paper with input from RS. SH contributed to data
interpretation and insights, SC supervised the group.
FUNDING
This work was supported by Agouron Institute and NSERC
discovery grants to SC.
REFERENCES
Agarwala, R., Barrett, T., Beck, J., Benson, D. A., Bollin, C., Bolton, E., et al. (2015).
Database resources of the national center for biotechnology information.
Nucleic Acids Res. 43, D6–D17. doi: 10.1093/nar/gku1130
Biderre-Petit, C., Boucher, D., Kuever, J., Alberic, P., Jézéquel, D., Chebance, B.,
et al. (2011). Identification of sulfur-cycle prokaryotes in a low-sulfate lake
(Lake Pavin) using aprA and 16S rRNA gene markers. Microb.Ecol. 61, 313–327.
doi: 10.1007/s00248-010- 9769-4
Blankenship, R. E., Madigan, M. T., and Bauer, C. E. (2006). Anoxygenic
Photosynthetic Bacteria. Berlin: Springer Science & Business Media.
Borrego, C. M., and Garciagil, L. J. (1995). Rearrangement of light-harvesting
bacteriochlorophyll homologs as a response of green sulfur bacteria to low-light
intensities. Photosynth. Res. 45, 21–30. doi: 10.1007/BF00032232
Bose, A., and Newman, D. K. (2011). Regulation of the phototrophic iron oxidation
(pio) genes in Rhodopseudomonas palustris TIE-1 is mediated by the global
regulator, FixK. Mol. Microbiol. 79, 63–75. doi: 10.1111/j.1365-2958.2010.
07430.x
Boyd, E., and Peters, J. W. (2013). New insights into the evolutionary history of
biological nitrogen fixation. Front. Microbiol. 4:201. doi: 10.3389/fmicb.2013.
00201
Boyd, E. S., Costas, A. M. G., Hamilton, T. L., Mus, F., and Peters, J. W.
(2015). Evolution of molybdenum nitrogenase during the transition from
anaerobic to aerobic metabolism. J. Bacteriol. 197, 1690–1699. doi: 10.1128/JB.
02611-14
Boyd, E. S., Hamilton, T. L., and Peters, J. W. (2011). An alternative path for the
evolution of biological nitrogen fixation. Front. Microbiol. 2:205. doi: 10.3389/
fmicb.2011.00205
Bryant, D. A., Liu, Z., Li, T., Zhao, F., Costas, A. M. G., Klatt, C. G., et al. (2012).
“Comparative and functional genomics of anoxygenic green bacteria from
the taxa Chlorobi, Chloroflexi, and Acidobacteria,” in Functional Genomics
and Evolution of Photosynthetic Systems, eds R. L. Burnap and W. Vermaas
(Dordrecht: Springer).
Canfield, D. E., Glazer, A. N., and Falkowski, P. G. (2010). The evolution and future
of Earth’s nitrogen cycle. Science 330, 192–196. doi: 10.1126/science.1186120
Canfield, D. E., Poulton, S. W., Knoll, A. H., Narbonne, G. M., Ross, G.,
Goldberg, T., et al. (2008). Ferruginous conditions dominated
later Neoproterozoic deep-water chemistry. Science 321, 949–952.
doi: 10.1126/science.1154499
Canfield, D. E., Rosing, M. T., and Bjerrum, C. (2006). Early anaerobic
metabolisms. Philos. Trans. R. Soc. B Biol. Sci. 361, 1819–1834. doi: 10.1098/
rstb.2006.1906
Cantarel, B. L., Coutinho, P. M., Rancurel, C., Bernard, T., Lombard, V., and
Henrissat, B. (2009). The Carbohydrate-Active EnZymes database (CAZy):
an expert resource for glycogenomics. Nucleic Acids Res. 37, D233–D238.
doi: 10.1093/nar/gkn663
Capone, D. G., Burns, J. A., Montoya, J. P., Subramaniam, A., Mahaffey, C.,
Gunderson, T., et al. (2005). Nitrogen fixation by Trichodesmium spp.: an
important source of new nitrogen to the tropical and subtropical North Atlantic
Ocean. Glob. Biogeochem. Cycles 19:GB2024.
Frontiers in Microbiology | www.frontiersin.org 13 July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 14
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
Capone, D. G., and Carpenter, E. J. (1982). Nitrogen fixation in the
marine environment. Science 217, 1140–1142. doi: 10.1126/science.217.456
5.1140
Carpenter, E. J., and Romans, K. (1991). Major role of the cyanobacterium
Trichodesmium in nutrient cycling in the North Atlantic Ocean. Science 254,
1356–1358. doi: 10.1126/science.254.5036.1356
Caspi, R., Altman, T., Dreher, K., Fulcher, C. A., Subhraveti, P., Keseler, I. M.,
et al. (2012). The MetaCyc database of metabolic pathways and enzymes and
the BioCyc collection of pathway/genome databases. Nucleic Acids Res. 40,
D742–D753. doi: 10.1093/nar/gkr1014
Chen, C.-T. A., Lin, C.-M., Huang, B.-T., and Chang, L.-F. (1996). Stoichiometry
of carbon, hydrogen, nitrogen, sulfur and oxygen in the particulate matter of
the western North Pacific marginal seas. Mar. Chem. 54, 179–190. doi: 10.1016/
0304-4203(96)00021- 7
Cloud, P. (1973). Paleoecological significance of banded iron formation. Econ. Geol.
68, 1135–1143. doi: 10.2113/gsecongeo.68.7.1135
Crowe, S. A., Døssing, L. N., Beukes, N. J., Bau, M., Kruger, S. J., Frei, R., et al.
(2013). Atmospheric oxygenation three billion years ago. Nature 501, 535–538.
doi: 10.1038/nature12426
Crowe, S. A., Hahn, A. S., Morgan-Lang, C., Thompson, K. J., Simister, R. L.,
Lliros, M., et al. (2017). Draft genome sequence of the pelagic photoferrotroph
Chlorobium phaeoferrooxidans.Genome Announc. 5:e01584-16. doi: 10.1128/
genomeA.01584-16
Crowe, S. A., Jones, C., Katsev, S., Magen, C., O’neill, A. H., Sturm, A.,
et al. (2008). Photoferrotrophs thrive in an Archean Ocean analogue.
Proc. Natl. Acad. Sci. U.S.A. 105, 15938–15943. doi: 10.1073/pnas.080
5313105
Crowe, S. A., Maresca, J., Jones, C., Sturm, A., Henny, C., Fowle, D. A., et al.
(2014a). Deep-water anoxygenic photosythesis in a ferruginous chemocline.
Geobiology 12, 322–339. doi: 10.1111/gbi.12089
Crowe, S. A., Paris, G., Katsev, S., Jones, C., Kim, S.-T., Zerkle, A. L., et al. (2014b).
Sulfate was a trace constituent of Archean seawater. Science 346, 735–739.
doi: 10.1126/science.1258966
Cuhel, R. L., Taylor, C. D., and Jannasch, H. W. (1981). Assimilatory sulfur
metabolism in marine microorganisms: sulfur metabolism, protein synthesis,
and growth of Pseudomonas halodurans and Alteromonas luteo-violaceus
during unperturbed batch growth. Arch. Microbiol. 130, 8–13. doi: 10.1007/
BF00527064
Czaja, A. D., Johnson, C. M., Beard, B. L., Roden, E. E., Li, W., and Moorbath, S.
(2013). Biological Fe oxidation controlled deposition of banded iron formation
in the ca. 3770Ma Isua Supracrustal Belt (West Greenland). Earth Planet. Sci.
Lett. 363, 192–203. doi: 10.1016/j.epsl.2012.12.025
Da Silva, J. F., and Williams, R. J. P. (2001). The Biological Chemistry of the
Elements: The Inorganic Chemistry of Life. New York, NY: Oxford University
Press.
David, L. A., and Alm, E. J. (2011). Rapid evolutionary innovation during
an Archaean genetic expansion. Nature 469, 93–96. doi: 10.1038/nature
09649
Ehrenreich, A., and Widdel, F. (1994). Anaerobic oxidation of ferrous iron
by purple bacteria, a new type of phototrophic metabolism. Appl. Environ.
Microbiol. 60, 4517–4526.
Fagerbakke, K. M., Heldal, M., and Norland, S. (1996). Content of carbon, nitrogen,
oxygen, sulfur and phosphorus in native aquatic and cultured bacteria. Aquat.
Microb. Ecol. 10, 15–27. doi: 10.3354/ame010015
Falkowski, P. G., Fenchel, T., and Delong, E. F. (2008). The microbial engines
that drive Earth’s biogeochemical cycles. Science 320, 1034–1039. doi: 10.1126/
science.1153213
Field, C. B., Behrenfeld, M. J., Randerson, J. T., and Falkowski, P. (1998). Primary
production of the biosphere: integrating terrestrial and oceanic components.
Science 281, 237–240. doi: 10.1126/science.281.5374.237
Frigaard, N.-U., and Bryant, D. A. (2008). “Genomic insights into the sulfur
metabolism of phototrophic green sulfur bacteria,” in Sulfur Metabolism in
Phototrophic Organisms, eds R. Hell, C. Dahl, D. Knaff, and T. Leustek
(New York, NY: Springer).
Frigaard, N.-U., Larsen, K. L., and Cox, R. P. (1996). Spectrochromatography
of photosynthetic pigments as a fingerprinting technique for microbial
phototrophs. FEMS Microbiol. Ecol. 20, 69–77. doi: 10.1111/j.1574-6941.1996.
tb00306.x
Garrels, R., Perry, E., and Mackenzie, F. (1973). Genesis of Precambrian iron-
formations and the development of atmospheric oxygen. Econ. Geol. 68,
1173–1179. doi: 10.2113/gsecongeo.68.7.1173
Garrity, G. M., Holt, J. G., Overmann, J., Pfennig, N., Gibson, J., and Gorlenko,
V. M. (2001). Phylum BXI. Chlorobi phy. nov. Bergey’s Manual( of Systematic
Bacteriology. New York, NY: Springer.
Gregersen, L. H., Habicht, K. S., Peduzzi, S., Tonolla, M., Canfield, D. E., Miller,M.,
et al. (2009). Dominance of a clonal green sulfur bacterial population in a
stratified lake. FEMS Microbiol. Ecol. 70, 30–41. doi: 10.1111/j.1574-6941.2009.
00737.x
Habicht, K. S., Gade, M., Thamdrup, B., Berg, P., and Canfield, D. E. (2002).
Calibration of sulfate levels in the Archean ocean. Science 298, 2372–2374.
doi: 10.1126/science.1078265
Hartman, H. (1984). “The evolution of photosynthesis and microbial mats: a
speculation on the banded iron formations,” in Microbial Mats: Stromatolites,
eds Y. Cohen, R. W. Castenholz, and H. O. Halvorson (New York, NY: Alan R.
Liss, Inc).
Hegler, F., Posth, N. R., Jiang, J., and Kappler, A. (2008). Physiology of phototrophic
iron(II)-oxidizing bacteria: implications for modern and ancient environments.
FEMS Microbiol. Ecol. 66, 250–260. doi: 10.1111/j.1574-6941.2008.00592.x
Heising, S., Richter, L., Ludwig, W., and Schink, B. (1999). Chlorobium ferrooxidans
sp nov., a phototrophic green sulfur bacterium that oxidizes ferrous iron
in coculture with a “Geospirillum” sp strain. Arch. Microbiol. 172, 116–124.
doi: 10.1007/s002030050748
Heising, S., and Schink, B. (1998). Phototrophic oxidation of ferrous iron by a
Rhodomicrobium vannielii strain. Microbiology 144, 2263–2269. doi: 10.1099/
00221287-144- 8-2263
Hiras, J., Wu, Y.-W., Eichorst, S. A., Simmons, B. A., and Singer, S. W. (2016).
Refining the phylum Chlorobi by resolving the phylogeny and metabolic
potential of the representative of a deeply branching, uncultivated lineage. ISME
J. 10, 833–845. doi: 10.1038/ismej.2015.158
Howarth, R. W. (1988). Nutrient limitation of net primary production in marine
ecosystems. Annu. Rev. Ecol. Syst. 19, 89–110. doi: 10.1146/annurev.es.19.
110188.000513
Joerger, R. D., Bishop, P. E., and Evans, H. J. (1988). Bacterial alternative
nitrogen fixation systems. CRC Crit. Rev. Microbiol. 16, 1–14. doi: 10.3109/
10408418809104465
Johnson, B., and Goldblatt, C. (2015). The nitrogen budget of Earth. Earth Sci. Rev.
148, 150–173. doi: 10.1016/j.earscirev.2015.05.006
Jones, C., Nomosatryo, S., Crowe, S. A., Bjerrum, C. J., and Canfield, D. E. (2015).
Iron oxides, divalent cations, silica, and the early earth phosphorus crisis.
Geology 43, 135–138. doi: 10.1130/G36044.1
Kanehisa, M., Araki, M., Goto, S., Hattori, M., Hirakawa, M., Itoh, M., et al. (2008).
KEGG for linking genomes to life and the environment. Nucleic Acids Res. 36,
D480–D484. doi: 10.1093/nar/gkm882
Kappler, A., and Newman, D. K. (2004). Formation of Fe(III)-minerals by Fe(II)-
oxidizing photoautotrophic bacteria. Geochim. Cosmochim. Acta 68, 1217–
1226. doi: 10.1016/j.gca.2003.09.006
Kappler, A., Pasquero, C., Konhauser, K. O., and Newman, D. K. (2005).
Deposition of banded iron formations by anoxygenic phototrophic Fe(II)-
oxidizing bacteria. Geology 33, 865–868. doi: 10.1130/G21658.1
Karl, D., Letelier, R., Tupas, L., Dore, J., Christian, J., and Hebel, D. (1997). The role
of nitrogen fixation in biogeochemical cycling in the subtropical North Pacific
Ocean. Nature 388, 533–538. doi: 10.1038/41474
Kaufmann, M. (2006). The role of the COG database in comparative and functional
genomics. Curr. Bioinform. 1, 291–300. doi: 10.2174/157489306777828017
Klein, C. (2005). Some Precambrian banded iron-formations (BIFs) from
around the world: their age, geologic setting, mineralogy, metamorphism,
geochemistry, and origins. Am. Mineral. 90, 1473–1499. doi: 10.2138/am.2005.
1871
Kondo, R., Nakagawa, A., Mochizuki, L., Osawa, K., Fujioka, Y., and Butani, J.
(2009). Dominant bacterioplankton populations in the meromictic Lake
Suigetsu as determined by denaturing gradient gel electrophoresis of 16S rRNA
gene fragments. Limnology 10, 63–69. doi: 10.1007/s10201-009-0261-0
Konhauser, K. O., Amskold, L., Lalonde, S. V., Posth, N. R., Kappler, A., and
Anbar, A. (2007). Decoupling photochemical Fe (II) oxidation from shallow-
water BIF deposition. Earth Planet. Sci. Lett. 258, 87–100. doi: 10.1016/j.epsl.
2007.03.026
Frontiers in Microbiology | www.frontiersin.org 14 July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 15
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
Konhauser, K. O., Hamade,T., Raiswell, R., Morris, R. C., Ferris, F. G., Southam, G.,
et al. (2002). Could bacteria have formed the Precambrian banded iron
formations? Geology 30, 1079–1082. doi: 10.1130/0091-7613(2002)030< 1079:
CBHFTP>2.0.CO;2
Konwar, K. M., Hanson, N. W., Bhatia, M. P., Kim, D., Wu, S.-J., Hahn, A. S., et al.
(2015). MetaPathways v2. 5: quantitative functional, taxonomic and usability
improvements. Bioinformatics 31, 3345–3347. doi: 10.1093/bioinformatics/
btv361
Konwar, K. M., Hanson, N. W., Pagé, A. P., and Hallam, S. J. (2013).
MetaPathways: a modular pipeline for constructing pathway/genome databases
from environmental sequence information. BMC Bioinformatics 14:202.
doi: 10.1186/1471-2105- 14-202
Kumar, S., Stecher, G., and Tamura, K. (2016). MEGA7: molecular evolutionary
genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evolut. 33,
1870–1874. doi: 10.1093/molbev/msw054
Larkin, M. A., Blackshields, G., Brown, N., Chenna, R., Mcgettigan, P. A.,
Mcwilliam, H., et al. (2007). Clustal W and Clustal X version 2.0. Bioinformatics
23, 2947–2948. doi: 10.1093/bioinformatics/btm404
Liu, Z., Frigaard, N.-U., Vogl, K., Iino, T., Ohkuma, M., Overmann, J., et al.
(2012). Complete genome of Ignavibacterium album, a metabolically versatile,
flagellated, facultative anaerobe from the phylum Chlorobi. Front. Microbiol.
3:185. doi: 10.3389/fmicb.2012.00185
Llirós, M., García–Armisen, T., Darchambeau, F., Morana, C., Triadó–
Margarit, X., Inceoˇ
glu, Ö, et al. (2015). Pelagic photoferrotrophy and iron
cycling in a modern ferruginous basin. Sci. Rep. 5:13803. doi: 10.1038/srep
13803
Manske, A. K., Glaeser, J., Kuypers, M. A. M., and Overmann, J. (2005). Physiology
and phylogeny of green sulfur bacteria forming a monospecific phototrophic
assemblage at a depth of 100 meters in the Black Sea. Appl. Environ. Microbiol.
71, 8049–8060. doi: 10.1128/AEM.71.12.8049-8060.2005
Meyer, J., Kelley, B. C., and Vignais, P. M. (1978). Nitrogen fixation and hydrogen
metabolism in photosynthetic bacteria. Biochimie 60, 245–260. doi: 10.1016/
S0300-9084(78)80821- 9
Meyer, K., Macalady, J., Fulton, J., Kump, L., Schaperdoth, I., and Freeman, K.
(2011). Carotenoid biomarkers as an imperfect reflection of the anoxygenic
phototrophic community in meromictic Fayetteville Green Lake. Geobiology 9,
321–329. doi: 10.1111/j.1472-4669.2011.00285.x
Michiels, C. C., Darchambeau, F., Roland, F. A., Morana, C., Llirós, M., García-
Armisen, T., et al. (2017). Iron-dependent nitrogen cycling in a ferruginous
lake and the nutrient status of Proterozoic oceans. Nat. Geosci. 10, 217–221.
doi: 10.1038/ngeo2886
Moore, C., Mills, M., Arrigo, K., Berman-Frank, I., Bopp, L., Boyd, P., et al. (2013).
Processes and patterns of oceanic nutrient limitation. Nat. Geosci. 6, 701–710.
doi: 10.1038/ngeo1765
Moore, C. M., Mills, M. M., Achterberg, E. P., Geider, R. J., Laroche, J., Lucas, M. I.,
et al. (2009). Large-scale distribution of Atlantic nitrogen fixation controlled by
iron availability. Nat. Geosci. 2, 867–871. doi: 10.1038/ngeo667
Overmann, J. (1997). “Mahoney Lake: a case study of the ecological significance of
phototrophic sulfur bacteria,” in Advances in Microbial Ecology, ed. J. G. Jones
(Berlin: Springer).
Paris, G., Adkins, J., Sessions, A. L., Webb, S., and Fischer, W. (2014). Neoarchean
carbonate–associated sulfate records positive 133S anomalies. Science 346,
739–741. doi: 10.1126/science.1258211
Pereira, L., Saraiva, I., Coelho, R., Newman, D., Louro, R., and Frazão, C.
(2012). Crystallization and preliminary crystallographic studies of FoxE
from Rhodobacter ferrooxidans SW2, an FeII oxidoreductase involved in
photoferrotrophy. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 68,
1106–1108. doi: 10.1107/S174430911203271X
Planavsky, N. J., Asael, D., Hofmann, A., Reinhard, C. T., Lalonde, S. V.,
Knudsen, A., et al. (2014). Evidence for oxygenic photosynthesis half a billion
years before the great oxidation event. Nat. Geosci. 7, 283–286. doi: 10.1038/
ngeo2122
Planavsky, N. J., Mcgoldrick, P., Scott, C. T., Li, C., Reinhard, C. T., Kelly, A. E.,
et al. (2011). Widespread iron-rich conditions in the mid-Proterozoic ocean.
Nature 477, 448–451. doi: 10.1038/nature10327
Posth, N. R., Huelin, S., Konhauser, K. O., and Kappler, A. (2010). Size, density
and composition of cell–mineral aggregates formed during anoxygenic
phototrophic Fe (II) oxidation: impact on modern and ancient environments.
Geochim. Cosmochim. Acta 74, 3476–3493. doi: 10.1016/j.gca.2010.
02.036
Poulton, S. W., and Canfield, D. E. (2011). Ferruginous conditions: a dominant
feature of the ocean through Earth’s history. Elements 7, 107–112. doi: 10.2113/
gselements.7.2.107
Pruesse, E., Quast, C., Knittel, K., Fuchs, B. M., Ludwig, W., Peplies, J., et al.
(2007). SILVA: a comprehensive online resource for quality checked and aligned
ribosomal RNA sequence data compatible with ARB. Nucleic Acids Res. 35,
7188–7196. doi: 10.1093/nar/gkm864
Quast, C., Pruesse, E., Yilmaz, P., Gerken, J., Schweer, T., Yarza, P., et al. (2012).
The SILVA ribosomal RNA gene database project: improved data processing
and web-based tools. Nucleic Acids Res. 41:gks1219. doi: 10.1093/nar/gk
s1219
Raymond, J., Siefert, J. L., Staples, C. R., and Blankenship, R. E. (2004). The natural
history of nitrogen fixation. Mol. Biol. Evol. 21, 541–554. doi: 10.1093/molbev/
msh047
Reinhard, C. T., Planavsky, N. J., Gill, B. C., Ozaki, K., Robbins, L. J., Lyons, T. W.,
et al. (2016). Evolution of the global phosphorus cycle. Nature 541, 386–389.
doi: 10.1038/nature20772
Romero-Viana, L., Keely, B. J., Camacho, A., Vicente, E., and Miracle, M. R.
(2010). Primary production in Lake La Cruz (Spain) over the last four centuries:
reconstruction based on sedimentary signal of photosynthetic pigments.
J. Paleolimnol. 43, 771–786. doi: 10.1007/s10933-009-9367-y
Rubio, L. M., and Ludden, P. W. (2008). Biosynthesis of the iron-molybdenum
cofactor of nitrogenase. Annu. Rev. Microbiol. 62, 93–111. doi: 10.1146/
annurev.micro.62.081307.162737
Sadekar, S., Raymond, J., and Blankenship, R. E. (2006). Conservation of distantly
related membrane proteins: photosynthetic reaction centers share a common
structural core. Mol. Biol. Evol. 23, 2001–2007. doi: 10.1093/molbev/msl079
Satoh, S., Mimuro, M., and Tanaka, A. (2013). Construction of a phylogenetic tree
of photosynthetic prokaryotes based on average similarities of whole genome
sequences. PLoS ONE 8:e70290. doi: 10.1371/journal.pone.0070290
Shen, Y., Buick, R., and Canfield, D. E. (2001). Isotopic evidence for microbial
sulphate reduction in the early Archaean era. Nature 410, 77–81. doi: 10.1038/
35065071
Soboh, B., Boyd, E. S., Zhao, D., Peters, J. W., and Rubio, L. M. (2010). Substrate
specificity and evolutionary implications of a NifDK enzyme carrying NifB-co
at its active site. FEBS Lett. 584, 1487–1492. doi: 10.1016/j.febslet.2010.02.064
Straub, K. L., Rainey, F. A., and Widdel, F. (1999). Rhodovulum iodosum sp. nov.
and Rhodovulum robiginosum sp. nov., two new marine phototrophic ferrous-
iron-oxidizing purple bacteria. Int. J. Syst. Evolut. Microbiol. 49, 729–735.
doi: 10.1099/00207713-49- 2-729
Stüeken, E. E., Buick, R., Guy, B. M., and Koehler, M. C. (2015). Isotopic evidence
for biological nitrogen fixation by molybdenum-nitrogenase from 3.2 Gyr.
Nature 520, 666–669. doi: 10.1038/nature14180
Tamura, K., Stecher, G., Peterson, D., Filipski, A., and Kumar, S. (2013). MEGA6:
molecular evolutionary genetics analysis version 6.0. Mol. Biol. Evolut. 30,
2725–2729. doi: 10.1093/molbev/mst197
Tonolla, M., Peduzzi, S., Demarta, A., Peduzzi, R., and Dittmar, H. (2004).
Phototropic sulfur and sulfate-reducing bacteria in the chemocline of
meromictic Lake Cadagno, Switzerland. J. Limnol. 63, 161–170. doi: 10.4081/
jlimnol.2004.161
Tyrrell, T. (1999). The relative influences of nitrogen and phosphorus on oceanic
primary production. Nature 400, 525–531. doi: 10.1038/22941
Viollier, E., Inglett, P., Hunter, K., Roychoudhury, A., and Van Cappellen, P. (2000).
The ferrozine method revisited: Fe (II)/Fe (III) determination in natural waters.
Appl. Geochem. 15, 785–790. doi: 10.1016/S0883-2927(99)00097-9
Walker, J. C. (1987). Was the Archaean biosphere upside down? Nature 329,
710–712. doi: 10.1038/329710a0
Walker, J. C., and Brimblecombe, P. (1985). Iron and sulfur in the pre-
biologic ocean. Precambrian Res. 28, 205–222. doi: 10.1016/0301-9268(85)9
0031-2
Walter, X. A., Picazo, A., Miracle, M. R., Vicente, E., Camacho, A., Aragno, M.,
et al. (2014). Phototrophic Fe (II)-oxidation in the chemocline of a ferruginous
meromictic lake. Front. Microbiol. 5:713. doi: 10.3389/fmicb.2014.00713
Weiss, M. C., Sousa, F. L., Mrnjavac, N., Neukirchen, S., Roettger, M., Nelson-
Sathi, S., et al. (2016). The physiology and habitat of the last universal common
ancestor. Nat. Microbiol. 1:16116. doi: 10.1038/nmicrobiol.2016.116
Frontiers in Microbiology | www.frontiersin.org 15 July 2017 | Volume 8 | Article 1212
fmicb-08-01212 July 4, 2017 Time: 16:2 # 16
Thompson et al. Metabolic Potential of Photoferrotrophic Chlorobi
Widdel, F., Schnell, S., Heising, S., Ehrenreich, A., Assmus, B., and Schink, B.
(1993). Ferrous iron oxidation by anoxygenic phototrophic bacteria. Nature
362, 834–836. doi: 10.1038/362834a0
Xiong, J., Fischer, W. M., Inoue, K., Nakahara, M., and Bauer,
C. E. (2000). Molecular evidence for the early evolution of
photosynthesis. Science 289, 1724–1730. doi: 10.1126/science.289.5485.
1724
Zhelezinskaia, I., Kaufman, A. J., Farquhar, J., and Cliff, J. (2014).
Large sulfur isotope fractionations associated with Neoarchean
microbial sulfate reduction. Science 346, 742–744. doi: 10.1126/science.
1256211
Conflict of Interest Statement: The authors declare that the research was
conducted in the absence of any commercial or financial relationships that could
be construed as a potential conflict of interest.
Copyright © 2017 Thompson, Simister, Hahn, Hallam and Crowe. This is an open-
access article distributed under the terms of the Creative Commons Attribution
License (CC BY). The use, distribution or reproduction in other forums is permitted,
provided the original author(s) or licensor are credited and that the original
publication in this journal is cited, in accordance with accepted academic practice.
No use, distribution or reproduction is permitted which does not comply with these
terms.
Frontiers in Microbiology | www.frontiersin.org 16 July 2017 | Volume 8 | Article 1212
... Photoferrotrophs achieve the highest volume-specific productivity rate of 50 μmol C/L/yr among microbial groups. With a photoferrotroph cellular carbon content of 3.8 x10 -14 g (Laufer et al., 2017) and a conservative estimate of nutrient-limited doubling times of 50 h (Thompson, Simister, Hahn, Hallam, & Crowe, 2017), this productivity rate translates to photoferrotroph cell densities of around 2 x 10 5 cells/ml, which is similar to the volume-specific abundance of photosynthetic organisms in the modern oligotrophic ocean, but lower than that measured in modern ferruginous basins (Crowe, Jones, et al., 2008) or photoferrotroph cultures (Laufer et al., 2017;Thompson et al., 2017). The estimated peak cell densities of oxygenic phototrophs and microaerophilic IOB are on the order of 1 x 10 4 and 1 x 10 3 cells/ml, respectively. ...
... Photoferrotrophs achieve the highest volume-specific productivity rate of 50 μmol C/L/yr among microbial groups. With a photoferrotroph cellular carbon content of 3.8 x10 -14 g (Laufer et al., 2017) and a conservative estimate of nutrient-limited doubling times of 50 h (Thompson, Simister, Hahn, Hallam, & Crowe, 2017), this productivity rate translates to photoferrotroph cell densities of around 2 x 10 5 cells/ml, which is similar to the volume-specific abundance of photosynthetic organisms in the modern oligotrophic ocean, but lower than that measured in modern ferruginous basins (Crowe, Jones, et al., 2008) or photoferrotroph cultures (Laufer et al., 2017;Thompson et al., 2017). The estimated peak cell densities of oxygenic phototrophs and microaerophilic IOB are on the order of 1 x 10 4 and 1 x 10 3 cells/ml, respectively. ...
Preprint
Full-text available
Banded Iron Formations (BIFs) are both the world’s largest ore deposits and important geological archives that record the early evolution of the Earth-Life system. BIFs were likely deposited as the result of ferrous iron [Fe(II)] oxidation, precipitation, and sedimentation from iron-rich (ferruginous) seawater, mostly during the Archean Eon. Proposed mechanisms for iron oxidation include abiotic reactions with photosynthetic oxygen, reaction with oxygen catalyzed by iron-oxidizing bacteria (IOB), and anoxic oxidation by anoxygenic iron-oxidizing phototrophic bacteria (photoferrotrophs). These iron oxidation processes may have operated concurrently, but their relative contributions to BIF deposition have not been considered. Here, we developed a 1-D ferruginous ocean model incorporating abiotic iron cycling and the physiology of oxygenic phototrophs, microaerophilic IOB, photoferrotrophs, and iron-reducing bacteria. Our model shows that, under Archean ocean conditions, most iron oxidation and precipitation would have been driven by photoferrotrophy, with a small fraction by microaerophilic IOB and a negligible contribution from abiotic reactions. The combined activities of these pathways led to BIF deposition at rates in line with geological records and, importantly, allowed the development of an Fe(II)-free surface ocean conducive to the formation of oxygen oases and the proliferation of oxygenic phototrophs. Teaser Archean ocean simulation shows that photoferrotrophs dominated the precipitation of BIFs and promoted the formation of marine oxygen oases.
... Chlorobi members are known to undergo photoferrotrophy, which executes anoxygenic photosynthesis using Fe 2+ . Chlorobi can also reduce assimilatory sulfate and nitrogen fixation (Thompson et al., 2017). These functions of Chlorobi can be presumed by the significant positive association observed with Fe 2+ , NH 4 + , and S 2− (r = 0.59 to 0.83, p < 0.05-0.01). ...
Article
Metal contamination and other geochemical alterations affect microbial composition and functional activities, disturbing natural biogeochemical cycles. Therefore, it is essential to understand the influences of multi-metal and geochemical interactions on microbial communities. This work investigated the distributions of total mercury (THg), methylmercury (MeHg), and trace metals in the anthropogenically affected sediment. The microbial communities and functional genes profiles were further determined to explore their association with Hg-methylation and geochemical features. The levels of THg and MeHg in sediment cores ranged between 10 and 40 mg/kg and 0.01-0.16 mg/kg, respectively, with an increasing trend toward bottom horizons. The major metals present at all depths were Al, Fe, Mn, and Zn. The enrichment and contamination indices confirmed that the trace metals were highly enriched in the anthropogenically affected sediment. Various functional genes were detected in all strata, indicating the presence of active microbial metabolic processes. The microbial community profiles revealed that the phyla Proteobacteria, Bacteroidetes, Bathyarchaeota, and Euryarchaeota, and the genera Thauera, Woeseia, Methanomethylovorans, and Methanosarcina were the dominant microbes. Correlating major taxa with geochemical variables inferred that sediment geochemistry substantially affects microbial community and biogeochemical cycles. Furthermore, archaeal methanogens and the bacterial phyla Chloroflexi and Firmicutes may play crucial roles in enhancing MeHg levels. Overall, these findings shed new light on the microbial communities potentially involved in Hg-methylation process and other biogeochemical cycles.
... Green sulfur bacteria (GSB) of the order Chlorobiales are anaerobes able to use reduced sulfur compounds, ferrous iron or hydrogen as electron donors for anoxygenic photosynthesis, driving carbon fixation using the rTCA pathway 6 . As such, GSB have been suggested to be among the most primitive phototrophs, having evolved under anoxic early ocean conditions, prior to the advent of oxygenic photosynthesis 7,8 . Their photoautotrophic physiology requires two ATP per pyruvate produced, consistent with the energy limitations of a faint young sun scenario and a largely anoxic early ocean-atmosphere system 1 . ...
Article
Full-text available
The reverse tricarboxylic acid (rTCA) cycle is touted as a primordial mode of carbon fixation due to its autocatalytic propensity and oxygen intolerance. Despite this inferred antiquity, however, the earliest rock record affords scant supporting evidence. In fact, based on the chimeric inheritance of rTCA cycle steps within the Chlorobiaceae, even the use of the chemical fossil record of this group is now subject to question. While the 1.64-billion-year-old Barney Creek Formation contains chemical fossils of the earliest known putative Chlorobiaceae-derived carotenoids, interferences from the accompanying hydrocarbon matrix have hitherto precluded the carbon isotope measurements necessary to establish the physiology of the organisms that produced them. Overcoming this obstacle, here we report a suite of compound-specific carbon isotope measurements identifying a cyanobacterially dominated ecosystem featuring heterotrophic bacteria. We demonstrate chlorobactane is ¹³C-depleted when compared to contemporary equivalents, showing only slight ¹³C-enrichment over co-existing cyanobacterial carotenoids. The absence of this diagnostic isotopic fingerprint, in turn, confirms phylogenomic hypotheses that call for the late assembly of the rTCA cycle and, thus, the delayed acquisition of autotrophy within the Chlorobiaceae. We suggest that progressive oxygenation of the Earth System caused an increase in the marine sulfate inventory thereby providing the selective pressure to fuel the Neoproterozoic shift towards energy-efficient photoautotrophy within the Chlorobiaceae.
... The taxa we found represent a range of aerobic and anaerobic microbes. Many of these phyla are known to contain extremophilic members and members utilizing chemolithotrophy (Fujitani et al., 2020;Slobodkin et al., 2019;Thompson et al., 2017;Waite et al., 2020). Of note, all samples contain Deinococcus radiodurans, an extremophilic species known to withstand high levels of radiation, desiccation, and cold temperatures (Ho et al., 2016). ...
Article
Full-text available
Lava tubes are key targets in the search for life on Mars. Their basaltic walls provide protection from radiation and changing environmental conditions, which could enable life or preservation of previous life in an otherwise harsh environment. We can understand the potential for Martian life in lava tubes by studying the habitability of analog environments on Earth. In this study, we present the first characterization of the microbial life inside a pristine Mauna Loa lava tube. This study is the first to combine 16S SSU rRNA sequencing and whole genome shotgun sequencing to map the taxonomic makeup and functional potential of any lava tube community in Hawaii, enabling a deep understanding of the types of microbes that thrive in this unique environment and the metabolisms they use. We find a surprisingly high degree of niche partitioning over small spatial scales and discuss implications for life detection strategies. Based on recent bioinformatic advancements in metagenomics, we also assemble dozens of high‐quality metagenome assembled genomes from the microbes living in the lava tubes, including several novel species.
Article
The relationships among eutrophication, anoxia, and microbial distribution were investigated for Nagatsura-Ura Lagoon on the northeastern Pacific coast of Japan. In September 2017, the bottom environment in a small area of the inner part of the lagoon (which has a basin-shaped bottom topology) was eutrophic and anoxic, with high carbon, nitrogen, phosphate, acid-volatile sulfide, and low dissolved oxygen and oxidation-reduction potential. Dissolved oxygen levels improved during the winter. Bacillariophyta (diatoms) were the main organic component according to pigment analysis and next-generation sequencing of nucleic acids in seawater samples. Phylum Proteobacteria was dominant among the bacterial flora in the sediment but the proportions of Class Epsilon-proteobacteria and Chlorobium (a green sulfur-utilizing bacterium) were high in the inner part of the lagoon compared to other stations, and these groups were also present in winter. Apparently groups able to thrive in both anoxic and aerobic conditions were predominant in the inner part of the lagoon.
Article
This is the first study investigating the effects of freeze-thaw (FT) and microplastics (MPs) on the distribution of antibiotic resistance genes (ARGs) in soil aggregates (i.e., soil basic constituent and functional unit) via microcosm experiments. The results showed that FT significantly increased the total relative abundance of target ARGs in different aggregates due to the increase in intI1 and ARG host bacteria. However, polyethylene MPs (PE-MPs) hindered the increase in ARG abundance caused by FT. The host bacteria carrying ARGs and intI1 varied with aggregate size, and the highest number of hosts was observed in micro-aggregates (<0.25 mm). FT and MPs altered host bacteria abundance by affecting aggregate physicochemical properties and bacterial community and enhanced multiple antibiotic resistance via vertical gene transfer. Although the dominant factors affecting ARGs varied with aggregate size, intI1 was a co-dominant factor in various-sized aggregates. Furthermore, other than ARGs, FT, PE-MPs, and their integration promoted the proliferation of human pathogenic bacteria in aggregates. These findings suggested that FT and its integration with MPs significantly affected ARG distribution in soil aggregates. They amplified antibiotic resistance environmental risks, contributing to a profound understanding of soil antibiotic resistance in the boreal region.
Preprint
Full-text available
The reductive tricarboxylic acid (rTCA) cycle is touted as a primordial mode of carbon fixation due to its autocatalytic propensity and oxygen intolerance 1,2 . Despite this inferred antiquity, however, the earliest rock record affords scant supporting evidence. In fact, based on the chimeric inheritance of rTCA cycle genes within the Chlorobiaceae ³ , even the utility of the chemical fossil record is now subject to question. While the 1.64-billion-year-old Barney Creek Formation (BCF) contains chemical fossils of the earliest known putative Chlorobiaceae-derived carotenoids, interferences from the accompanying hydrocarbon matrix have hitherto precluded the carbon isotope measurements necessary to establish the physiology of the organisms that produced them. Overcoming this obstacle, here we report a suite of compound-specific carbon isotope measurements identifying a cyanobacterially dominated ecosystem featuring heterotrophic bacteria. Crucially, we demonstrate chlorobactane is 13C-depleted when compared to contemporary equivalents, showing only slight 13C-enrichment over co-existing cyanobacterial carotenoids. These observations demonstrate the late acquisition of autotrophy and the rTCA cycle as predicted via phylogenomic- and molecular-clock-grounded approaches3. We suggest that progressive oxygenation of the Earth System caused an increase in the marine sulfate inventory thereby providing the selective pressure to fuel the Neoproterozoic niche shift toward energy-efficient photoautotrophy within the Chlorobiaceae ⁴ .
Article
Full-text available
We present the latest version of the Molecular Evolutionary Genetics Analysis (MEGA) software, which contains many sophisticated methods and tools for phylogenomics and phylomedicine. In this major upgrade, MEGA has been optimized for use on 64-bit computing systems for analyzing bigger datasets. Researchers can now explore and analyze tens of thousands of sequences in MEGA. The new version also provides an advanced wizard for building timetrees and includes a new functionality to automatically predict gene duplication events in gene family trees. The 64-bit MEGA is made available in two interfaces: graphical and command line. The graphical user interface (GUI) is a native Microsoft Windows application that can also be used on Mac OSX. The command line MEGA is available as native applications for Windows, Linux, and Mac OSX. They are intended for use in high-throughput and scripted analysis. Both versions are available from www.megasoftware.net free of charge.
Article
Full-text available
Here, we report the draft genome sequence of Chlorobium phaeoferrooxidans , a photoferrotrophic member of the genus Chlorobium in the phylum Chlorobi . This genome sequence provides insight into the metabolic capacity that underpins photoferrotrophy within low-light-adapted pelagic Chlorobi .
Article
Full-text available
Nitrogen limitation during the Proterozoic has been inferred from the great expanse of ocean anoxia under low-O2 atmospheres, which could have promoted NO3- reduction to N2 and fixed N loss from the ocean. The deep oceans were Fe rich (ferruginous) during much of this time, yet the dynamics of N cycling under such conditions remain entirely conceptual, as analogue environments are rare today. Here we use incubation experiments to show that a modern ferruginous basin, Kabuno Bay in East Africa, supports high rates of NO3- reduction. Although 60% of this NO3- is reduced to N2 through canonical denitrification, a large fraction (40%) is reduced to NH4+, leading to N retention rather than loss. We also find that NO3- reduction is Fe dependent, demonstrating that such reactions occur in natural ferruginous water columns. Numerical modelling of ferruginous upwelling systems, informed by our results from Kabuno Bay, demonstrates that NO3- reduction to NH4+ could have enhanced biological production, fuelling sulfate reduction and the development of mid-water euxinia overlying ferruginous deep oceans. This NO3- reduction to NH4+ could also have partly offset a negative feedback on biological production that accompanies oxygenation of the surface ocean. Our results indicate that N loss in ferruginous upwelling systems may not have kept pace with global N fixation at marine phosphorous concentrations (0.04-0.13[thinsp][mu]M) indicated by the rock record. We therefore suggest that global marine biological production under ferruginous ocean conditions in the Proterozoic eon may thus have been P not N limited
Article
Full-text available
Iron-rich (ferruginous) ocean chemistry prevailed throughout most of Earth’s early history. Before the evolution and proliferation of oxygenic photosynthesis, biological production in the ferruginous oceans was likely driven by photoferrotrophic bacteria that oxidize ferrous iron {Fe(II)} to harness energy from sunlight, and fix inorganic carbon into biomass. Photoferrotrophs may thus have fuelled Earth’s early biosphere providing energy to drive microbial growth and evolution over billions of years. Yet, photoferrotrophic activity has remained largely elusive on the modern Earth, leaving models for early biological production untested and imperative ecological context for the evolution of life missing. Here, we show that an active community of pelagic photoferrotrophs comprises up to 30% of the total microbial community in illuminated ferruginous waters of Kabuno Bay (KB), East Africa (DR Congo). These photoferrotrophs produce oxidized iron {Fe(III)} and biomass, and support a diverse pelagic microbial community including heterotrophic Fe(III)-reducers, sulfate reducers, fermenters and methanogens. At modest light levels, rates of photoferrotrophy in KB exceed those predicted for early Earth primary production, and are sufficient to generate Earth’s largest sedimentary iron ore deposits. Fe cycling, however, is efficient, and complex microbial community interactions likely regulate Fe(III) and organic matter export from the photic zone.
Book
Anoxygenic Photosynthetic Bacteria is a comprehensive volume describing all aspects of non-oxygen-evolving photosynthetic bacteria. The 62 chapters are organized into themes of: Taxonomy, physiology and ecology; Molecular structure of pigments and cofactors; Membrane and cell wall structure: Antenna structure and function; Reaction center structure and electron/proton pathways; Cyclic electron transfer; Metabolic processes; Genetics; Regulation of gene expression, and applications. The chapters have all been written by leading experts and present in detail the current understanding of these versatile microorganisms. The book is intended for use by advanced undergraduate and graduate students and senior researchers in the areas of microbiology, genetics, biochemistry, biophysics and biotechnology.
Article
The macronutrient phosphorus is thought to limit primary productivity in the oceans on geological timescales. Although there has been a sustained effort to reconstruct the dynamics of the phosphorus cycle over the past 3.5 billion years, it remains uncertain whether phosphorus limitation persisted throughout Earth’s history and therefore whether the phosphorus cycle has consistently modulated biospheric productivity and ocean–atmosphere oxygen levels over time. Here we present a compilation of phosphorus abundances in marine sedimentary rocks spanning the past 3.5 billion years. We find evidence for relatively low authigenic phosphorus burial in shallow marine environments until about 800 to 700 million years ago. Our interpretation of the database leads us to propose that limited marginal phosphorus burial before that time was linked to phosphorus biolimitation, resulting in elemental stoichiometries in primary producers that diverged strongly from the Redfield ratio (the atomic ratio of carbon, nitrogen and phosphorus found in phytoplankton). We place our phosphorus record in a quantitative biogeochemical model framework and find that a combination of enhanced phosphorus scavenging in anoxic, iron-rich oceans and a nutrient-based bistability in atmospheric oxygen levels could have resulted in a stable low-oxygen world. The combination of these factors may explain the protracted oxygenation of Earth’s surface over the last 3.5 billion years of Earth history. However, our analysis also suggests that a fundamental shift in the phosphorus cycle may have occurred during the late Proterozoic eon (between 800 and 635 million years ago), coincident with a previously inferred shift in marine redox states, severe perturbations to Earth’s climate system, and the emergence of animals.
Article
The concept of a last universal common ancestor of all cells (LUCA, or the progenote) is central to the study of early evolution and life's origin, yet information about how and where LUCA lived is lacking. We investigated all clusters and phylogenetic trees for 6.1 million protein coding genes from sequenced prokaryotic genomes in order to reconstruct the microbial ecology of LUCA. Among 286,514 protein clusters, we identified 355 protein families (∼0.1%) that trace to LUCA by phylogenetic criteria. Because these proteins are not universally distributed, they can shed light on LUCA's physiology. Their functions, properties and prosthetic groups depict LUCA as anaerobic, CO2-fixing, H2-dependent with a Wood–Ljungdahl pathway, N2-fixing and thermophilic. LUCA's biochemistry was replete with FeS clusters and radical reaction mechanisms. Its cofactors reveal dependence upon transition metals, flavins, S-adenosyl methionine, coenzyme A, ferredoxin, molybdopterin, corrins and selenium. Its genetic code required nucleoside modifications and S-adenosyl methionine-dependent methylations. The 355 phylogenies identify clostridia and methanogens, whose modern lifestyles resemble that of LUCA, as basal among their respective domains. LUCA inhabited a geochemically active environment rich in H2, CO2 and iron. The data support the theory of an autotrophic origin of life involving the Wood–Ljungdahl pathway in a hydrothermal setting.
Article
The National Center for Biotechnology Information (NCBI) provides a large suite of online resources for biological information and data, including the GenBank® nucleic acid sequence database and the PubMed database of citations and abstracts for published life science journals. Additional NCBI resources focus on literature (PubMed Central (PMC), Bookshelf and PubReader), health (ClinVar, dbGaP, dbMHC, the Genetic Testing Registry, HIV-1/Human Protein Interaction Database and MedGen), genomes (BioProject, Assembly, Genome, BioSample, dbSNP, dbVar, Epigenomics, the Map Viewer, Nucleotide, Probe, RefSeq, Sequence Read Archive, the Taxonomy Browser and the Trace Archive), genes (Gene, Gene Expression Omnibus (GEO), HomoloGene, PopSet and UniGene), proteins (Protein, the Conserved Domain Database (CDD), COBALT, Conserved Domain Architecture Retrieval Tool (CDART), the Molecular Modeling Database (MMDB) and Protein Clusters) and chemicals (Biosystems and the Pub-Chem suite of small molecule databases). The Entrez system provides search and retrieval operations for most of these databases. Augmenting many of the web applications are custom implementations of the BLAST program optimized to search specialized datasets. All of these resources can be accessed through the NCBI home page at www.ncbi.nlm.nih. gov.
Article
Anoxic iron-rich sediment samples that had been stored in the light showed development of brown, rusty patches. Subcultures in defined mineral media with ferrous iron (10 mmol/liter, mostly precipitated as FeCO3) yielded enrichments of anoxygenic phototrophic bacteria which used ferrous iron as the sole electron donor for photosynthesis. Two different types of purple bacteria, represented by strains L7 and SW2, were isolated which oxidized colorless ferrous iron under anoxic conditions in the light to brown ferric iron. Strain L7 had rod-shaped, nonmotile cells (1.3 by 2 to 3 microns) which frequently formed gas vesicles. In addition to ferrous iron, strain L7 used H2 + CO2, acetate, pyruvate, and glucose as substrate for phototrophic growth. Strain SW2 had small rod-shaped, nonmotile cells (0.5 by 1 to 1.5 microns). Besides ferrous iron, strain SW2 utilized H2 + CO2, monocarboxylic acids, glucose, and fructose. Neither strain utilized free sulfide; however, both strains grew on black ferrous sulfide (FeS) which was converted to ferric iron and sulfate. Strains L7 and SW2 grown photoheterotrophically without ferrous iron were purple to brownish red and yellowish brown, respectively; absorption spectra revealed peaks characteristic of bacteriochlorophyll a. The closest phototrophic relatives of strains L7 and SW2 so far examined on the basis of 16S rRNA sequences were species of the genera Chromatium (gamma subclass of proteobacteria) and Rhodobacter (alpha subclass), respectively. In mineral medium, the new isolates formed 7.6 g of cell dry mass per mol of Fe(II) oxidized, which is in good agreement with a photoautotrophic utilization of ferrous iron as electron donor for CO2 fixation. Dependence of ferrous iron oxidation on light and CO2 was also demonstrated in dense cell suspensions. In media containing both ferrous iron and an organic substrate (e.g., acetate, glucose), strain L7 utilized ferrous iron and the organic compound simultaneously; in contrast, strain SW2 started to oxidize ferrous iron only after consumption of the organic electron donor. Ferrous iron oxidation by anoxygenic phototrophs is understandable in terms of energetics. In contrast to the Fe3+/Fe2+ pair (E0 = +0.77 V) existing in acidic solutions, the relevant redox pair at pH 7 in bicarbonate-containing environments, Fe(OH)3 + HCO3-/FeCO3, has an E0' of +0.2 V. Ferrous iron at pH 7 can therefore donate electrons to the photosystem of anoxygenic phototrophs, which in purple bacteria has a midpoint potential around +0.45 V. The existence of ferrous iron-oxidizing anoxygenic phototrophs may offer an explanation for the deposition of early banded-iron formations in an assumed anoxic biosphere in Archean times.