You are on page 1of 15

BBA - Proteins and Proteomics 1865 (2017) 1455–1469

Contents lists available at ScienceDirect

BBA - Proteins and Proteomics


journal homepage: www.elsevier.com/locate/bbapap

Understanding the response of Desulfovibrio desulfuricans ATCC 27774 to the MARK


electron acceptors nitrate and sulfate - biosynthetic costs modulate substrate
selection
Joana R. Sousaa, Célia M. Silveiraa,b, Pedro Fontesa, Catarina Roma-Rodriguesc,
Alexandra R. Fernandesc, Gonzalez Van Driesschec, Bart Devreesed, Isabel Mouraa,
José J.G. Mouraa, M. Gabriela Almeidaa,e,⁎
a
UCIBIO-REQUIMTE, Departamento de Química, Faculdade de Ciências e Tecnologia, Universidade NOVA de Lisboa, 2829-516 Monte de Caparica, Portugal
b
Instituto de Tecnologia Química e Biológica António Xavier, Universidade NOVA de Lisboa, Av. da República, 2780-157 Oeiras, Portugal
c
UCIBIO, Departamento Ciências da Vida, Faculdade de Ciências e Tecnologia, Universidade NOVA de Lisboa, 2829-516 Monte de Caparica, Portugal
d
Laboratory for Protein Biochemistry and Biomolecular Engineering, Department of Biochemistry and Microbiology, Gent University, 9000 Gent, Belgium
e
Centro de Investigação Interdisciplinar Egas Moniz (CiiEM), Instituto Superior de Ciências da Saúde Egas Moniz, Campus Universitário, Quinta da Granja, 2829-511
Caparica, Portugal

A R T I C L E I N F O A B S T R A C T

Keywords: Sulfate-reducing bacteria (SRB) are a diverse group of anaerobic microorganisms that obtain their energy from
Desulfovibrio desulfuricans ATCC 27774 dissimilatory sulfate reduction. Some SRB species have high respiratory versatility due to the possible use of
Proteomics alternative electron acceptors. A good example is Desulfovibrio desulfuricans ATCC 27774, which grows in the
Respiratory flexibility presence of nitrate (end product: ammonium) with higher rates and yields to those observed in sulfate containing
Nitrate reduction
medium (end product: sulfide). In this work, the mechanisms supporting the respiratory versatility of D. de-
Ammonium assimilation
Energy metabolism
sulfuricans were unraveled through the analysis of the proteome of the bacterium under different experimental
conditions. The most remarkable difference in the two-dimensional gel electrophoresis maps is the high number
of spots exclusively represented in the nitrate medium. Most of the proteins with increase abundance are in-
volved in the energy metabolism and the biosynthesis of amino acids (or proteins), especially those participating
in ammonium assimilation processes. qPCR analysis performed during different stages of the bacterium's growth
showed that the genes involved in nitrate and nitrite reduction (napA and nrfA, respectively) have different
expressions profiles: while napA did not vary significantly, nrfA was highly expressed at a 6 h time point. Nitrite
levels measured along the growth curve revealed a peak at 3 h. Thus, the initial consumption of nitrate and
concomitant production of nitrite must induce nrfA expression. The activation of alternative mechanisms for
energy production, aside several N-assimilation metabolisms and detoxification processes, solves potential
survival problems in adapting to different environments and contributes to higher bacterial growth rates.

1. Introduction of both natural and artificial locations, such as marine sediments, hy-
drothermal vents, oil fields, freshwater sediments and anoxic waste-
Sulfate reducing bacteria (SRB) are a large and diverse group of water treatment plants [1]. There are several examples in the literature
anaerobic microorganisms that obtain their energy through dissim- of SRB species that reduce nitrate as an alternative electron acceptor,
ilatory sulfate reduction (DSR), a pathway of the sulfur cycle where including Desulfobulbus propionicus, Desulforhopalus singaporensis, De-
sulfate is reduced to sulfide as terminal electron acceptor [1–3]. How- sulfobacterium catecholicum, Desulfotomaculum thermobenzoicum, De-
ever, the successive discovery of many other inorganic compounds sulfovibrio oxamicus, Desulfovibrio termitidis, Desulfovibrio furfuralis, De-
(sulfite, thiosulfate, elemental sulfur and even nitrate and nitrite) that sulfovibrio profundus, Desulfovibrio simplex and Desulfovibrio
can also serve as final electron acceptors attested the high flexibility of desulfuricans, thereby involving these bacteria in the biological nitrogen
the energy metabolism of these organisms [1,2,4,5]. This enables them cycle as well [6]. Nitrogen is an essential component in living organ-
to easily adapt to changes in the environment and inhabit a multitude isms and can be found in several oxidation states [7]. The nitrogen


Corresponding author at: UCIBIO, REQUIMTE, Faculdade de Ciências e Tecnologia, Universidade NOVA de Lisboa, 2829-516 Monte de Caparica, Portugal.
E-mail address: mg.almeida@fct.unl.pt (M.G. Almeida).

http://dx.doi.org/10.1016/j.bbapap.2017.07.021
Received 4 April 2017; Received in revised form 12 July 2017; Accepted 21 July 2017
Available online 25 August 2017
1570-9639/ © 2017 Elsevier B.V. All rights reserved.
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

cycle includes six oxidation/reduction reactions wherein the nitrogen 2. Materials and methods
atom passes from the fully oxidized form (nitrate, NO3−, + 5) to its
completely fully reduced form (ammonium, NH4+, −3) [7]. 2.1. Cell growth
The interest on SRB is extended over different research areas. These
microorganisms have been linked to health issues since they can have a D. desulfuricans cells were grown under anoxic conditions at 37 °C in
role in inflammatory bowel diseases [8]. They are involved in corrosion a lactate containing medium supplemented with either nitrate or sulfate
of metal equipment and souring of petroleum reservoirs, which can as electron acceptor substrates. The medium was composed of (in g L− 1
cause serious economic problems in the oil industry [9]. Furthermore, distilled water): KH2PO4, 0.5; NH4Cl, 1.0; CaCl2·2H2O, 0.04 (Merck);
SRB's ability to reduce toxic heavy metals might be valuable for bior- sodium lactate, 6.0 (Sigma); sodium citrate, 0.3 (Ridel-de-Haen); ca-
emediation [10]. The implications in such different areas are surely a saminoacids, 2.0 (Dfico); tryptone, 2.0 (Scharlau); Wolfe's Elixir
consequence of the adaptability of the bacteria to different substrates (1 mL L− 1) and vitamins solution (2 mL L− 1). Please note that tryptone
and strongly compel studies regarding the understanding of its versatile and casaminoacids are digestion products of casein, resulting in a
metabolism. mixture of peptides and amino acids (in the case of casaminoacids),
In this paper, we will focus on the bacterium Desulfovibrio de- which makes their exact composition unknown (semi-defined). The
sulfuricans ATCC 27774 (D. desulfuricans), which grows in medium culture medium was supplemented with Na2SO4, 4.5; FeSO4·7H2O,
containing nitrate with rates and yields ca. 2 times superior to those 0.004 (Merck) and MgSO4·7H2O, 0.060 (Panreac) for sulfate and with
observed in a sulfate rich environment [11]. The changes in cellular NaNO3, 2.4; FeCl2·4H2O, 0.003 (Merck) and MgCl2·6H2O, 0.050
metabolism that enable the switch in oxidizing substrates are not (Panreac) for nitrate and prepared at pH 7.5. The vitamins solution
completely understood. However, it is well known that the nitrate contained (in mg L- 1 distilled water): vit. B2, 0.02; vit. B3, 0.05; vit. B1,
(NapA) and nitrite reductases (NrfA), the two enzymes involved in the 0.06; vit. B5, 0,06; vit. B6, 0.06; vit. C, 0.2 (Merck); vit. B12, 0.05
dissimilatory reduction of nitrate to ammonia (DNRA), are over- (Roche) and vit. H, 0.01 (Fluka). The solution was sterilized with
expressed when nitrate is the sole electron acceptor. In this case, DNRA 0.2 μm filters. The Wolfes's Elixir contained (in g L− 1 distilled water):
is considered a two-step respiratory process where nitrate is the term- nitriloacetic acid, 1.5 (Fluka); MgSO4·7H2O, 3.0 (Panreac); MnSO4·H2O,
inal electron acceptor, initially converted into nitrite, and then reduced 0.5; FeSO4·7H2O, 0.1; CoSO4·7H2O, 0.1; NiCl2·6H2O, 0.1; CuCl2·2H2O,
to ammonia. Interestingly, in the presence of nitrate and sulfate, the 0.1; ZnSO4·7H2O, 0.1; CuSO4·5H2O, 0.01; AlK(SO4)2·12H2O, 0.01;
thermodynamically less-favorable sulfate is preferred [11–13]. Nitrate H3BO3,·0.01; Na2MoO4·2H2O, 0.01; Na2MoO4·2H2O, 0.01,
reduction seems to occur only in the absence of sulfate and sulfide (the Na2SeO3·5H2O, 0.01 (Merck) and NaCl, 1.0 (Ridel-de Haen). The
product of sulfate reduction) [11,14,15]. In fact, opposing to the sulfate medium was purged with argon for 30 min, to remove dissolved oxygen
reduction pathway, which is constitutively expressed when the cells and autoclaved at 121 °C for 20 min. The cells were grown in 1 or 2 L
grow in nitrate, Marietou et al. have shown that nitrate respiration is flasks, specially adapted for anoxic cultures. The nitrate and sulfate cell
regulated at the transcriptional level with the nitrate reductase operon cultures were performed in triplicate. The inoculums (cells grown up to
nap being induced by nitrate but repressed by sulfate [11]. Although the stationary phase) were 10% of the total volume. The growth rate
the nitrate reduction pathway is not completely expressed while sulfate was traced by measuring the optical density (OD) at 600 nm using a
is available, the nrf genes coding for NrfA seem to be expressed con- Shimadzu UV160A spectrophotometer. The specific growth rates (μ)
stitutively in SRB and are present in sulfate grown bacteria [11,15,16]. and culture doubling times (Td) were determined using the slopes of the
In addition to the role in nitrate respiration, NrfA is thought to be in- growth curves (semi logarithmic plots) in the exponential phase. Cells
volved in nitrite and nitric oxide detoxification processes. The latter has were harvested after 16 h of growth by centrifugation (Beckman Avanti
been reported in enteric bacteria that use NrfA to reduce NO [17,18]. J-25) of the culture media at 10000 × g (15 min, 10 °C). The pellet was
Nitrite has been shown to inhibit SRB growth and therefore the pre- then washed in 10 mM phosphate buffer, pH 7.6. Finally, the cells were
sence of NrfA can sustain the cells in environments with millimolar suspended in the same buffer (1 g pellet per 3 mL buffer) and stored at
concentrations of nitrite (e.g. cohabitated by nitrate-reducing and sul- − 80 °C.
fide-oxidizing bacteria) [19,20]. This detoxification effect has been
shown, for example, in Desulfovibrio vulgaris Hildenborough, which is 2.2. Sample preparation
unable to use nitrate as respiratory substrate; however, it can cope with
nitrite through the expression of nrf genes [21,22]. Besides catalyzing The D. desulfuricans cells were homogenized in a Potter-Elvejhem
the six-electron reduction of nitrite to ammonia, some NrfAs were also homogenizer (Multifix) in the presence of a protease inhibitor cocktail
shown to reduce sulfite to sulfide, thus forming the only known link (Complete EDTA - free, Roche) and a mix of endonucleases (DNase I,
between the nitrogen and sulfur biogeochemical cycles. The sulfite re- Roche and RNase A, Sigma; 5 mg/L each). The cell suspensions were
ductase activity in these enzymes is considerably lower than their ni- disrupted in a French-Press (Thermo-FA-080A) at ca. 16000 psi and
trite reductase activity; though, it is sometimes higher than that of centrifuged at 8000 × g for 15 min at 4 °C to remove non-lysed cells and
sulfite reductase enzymes themselves [23,24]. Regarding the peri- debris in a Sigma 3K30 centrifuge. The obtained protein samples were
plasmic nitrate reductase (NapA), which is involved in the reduction of stored at − 80 °C.
nitrate to nitrite, this enzyme participates in several processes, such as
redox balancing, nitrate scavenging, if in residual amounts, and aerobic 2.3. Protein quantification
or anaerobic denitrification [6,25].
To identify the molecular tools involved in the respiration flexibility Protein quantification was performed with the BioRad Protein Assay
of D. desulfuricans cells induced by different electron acceptors (nitrate kit and the 2D Quant kit (GE Healthcare). Bovine serum albumin
or sulfate), we performed a differential proteomic analysis based in (Sigma) was used as protein standard solution.
two-dimensional gel electrophoresis (2DE) combined with mass spec-
trometry (MS) based protein identification. Moreover, we have ana- 2.4. Two-dimensional gel electrophoresis
lyzed the expression of key genes involved in the respiratory chain.
Additionally, we monitored the levels of nitrite present in the extra- Immobiline Drystrip gel (IPG) strips, 18 cm, pH gradient 4–7 (GE
cellular medium and the NrfA and NapA enzyme activities throughout Healthcare) were rehydrated overnight with 340 μL of a 2D rehydration
the bacterial growth, until reaching the stationary phase. solution containing 7 M urea (Merck), 2 M thiourea (Merck), 2% (m/v)
CHAPS (Calbiochem), 0.5% (v/v) IPG buffer (GE Healthcare), 0.002%
(m/v) bromophenol blue (Merck), 0.28% (m/v) DTT (GE Healthcare)

1456
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

and 350 μg of the protein samples. separated using a 2-dimensional NanoAcquity UPLC® system (Waters
The first dimension separation (isoelectric focusing) was performed Corporation) in pseudo 1D mode. For the first dimension (high pH),
in an Ettan IPGphor 3 system (GE Healthcare) at 20 °C with a four step solvent A1 and B1 were composed of 20 mM ammonium formate in
program: step 1–500 V fixed voltage for 4 h, step 2 – voltage gradient water and ACN, respectively. For the second dimension (low pH), sol-
up to 1000 V in 3 h, step 3 – voltage gradient up to 10,000 V in 3 h and vent A2 and B2 were composed of 0.1% formic acid in water and 0.1%
step 4–10,000 V fixed voltage for 3 h, adding up to a total of 61.5 kV formic acid in ACN, respectively. The sample (5 μL) was loaded onto an
ran in 15 h. XbridgeTM BEH130 C18 column (300 μm × 50 mm, 5 μm; Waters) at
After isoelectric focusing the IPG strips were saturated with the 3% solvent B1 at 2 μL/min. Peptides were eluted from the first di-
buffer system required for the second dimension separation in two mension column in 50% of solvent B, and trapped on a Symmetry® C18
consecutive equilibration steps performed at room temperature. The trapping column (180 μm × 20 mm, 5 μm; Waters). The trapped pep-
IPG strips were soaked and rocked for 15 min in SDS equilibration tides were separated on a HSS T3 C18 analytical column
buffer [6 M urea, 50 mM Tris-HCl pH 8.8, 30% (v/v) glycerol (Panreac), (75 μm × 250 mm, 1.8 μm; Waters) at 40 °C at 250 nL/min by in-
2% (m/v) SDS (Sigma) and 0.002% of bromophenol blue] containing i) creasing the acetonitrile concentration from 5 to 50% B2 over 60 min.
10 mg/mL of DTT and ii) 25 mg/mL of iodoacetamide (Merck). For the The eluting peptides were directly ionized using nano-electropray and
second dimension, the proteins were separated by SDS-PAGE in 12.5% analyzed with a SYNAPT G1 HDMS (Waters) using a PicoTip Emitter
acrylamide gels on a vertical unit (EttanDalt Six, GE Healthcare). The from New Objective (uncoated SilicaTipTM 10+/− 1 μm, Woburn).
IPG strips were applied on top of the gel and the system was sealed with The TOF analyzer was externally calibrated with MS/MS fragments of
1% agarose (GE Healthcare) solution prepared in electrophoresis buffer. human [glu1]-fibrinopeptide B (Glu-fib) from m/z 72 to 1285, and the
The SDS-PAGE was carried out at 350 V for ca. 3:30 h after an initial data was corrected post-acquisition using the monoisotopic mass of the
period of 30 min at 80 V to allow the sample to migrate from the IPG doubly charged precursor of Glu-fib (m/z 785.8426) (lock mass cor-
strip to the second dimension gel. The separated proteins were visua- rection). Accurate mass data were collected in a data independent po-
lized by the colloidal Coomassie Blue G-250 (Merck) staining method sitive mode of acquisition (MSE) by alternating between low (5 V) and
[26]. The stained gels were scanned with the ImageScanner II (GE high (ramping from 15 to 35 V) energy scan functions. The selected m/z
Healthcare). range was 125 to 2000 Da. The capillary voltage was set to 3.0 kV, the
Gel image analysis and statistical quantification of relative protein sampling cone voltage was 26 V and the extraction cone voltage on
abundances was performed with Melanie 7.0 software (GeneBio). Three 2.65 V. The source temperature was set on 65 °C.
replicate gels of three independent growths (biological replicates) from The resulting LCMSE data were processed using ProteinLynx Global
each condition (nitrate and sulfate) were used in the analysis and SERVER v3.0 (PLGS, Waters Corporation). In brief, lock mass-corrected
evaluated as a single set of 9 gels. The relative abundance of each spectra (0.250 Da window allowed) were automatically centroided,
protein spot was compared between the nitrate and sulfate growth deisotoped and charge-state reduced to produce a single monoisotopic
conditions: the spots with a relative volume ratio higher than 2 and peak for each peptide and associated fragment ion. The correlation of a
ANOVA p-value bellow than 0.05 were considered differentially ex- precursor and a potential fragment ion was achieved by means of time
pressed. These spots and those observed only in one experimental alignment, in the first instance. The following parameters were used for
condition were selected for protein identification by Matrix-Assisted the data processing in PLGS: the chromatographic peak width, the TOF
Laser Desorption/Ionization - Time-of-Flight (MALDI-TOF) mass spec- resolution and retention time window were determined automatically
trometry. by the software, and the low energy, high energy, and intensity
thresholds, which were set to 250, 100 and 1500 counts respectively. A
2.5. Protein identification database containing Desulfovibrio protein sequences, downloaded from
Uniprot in April 2015, was used. The precursor and fragment ion tol-
Coomassie-stained spots were destained and washed twice with erance were determined automatically. The default protein identifica-
200 mM ammonium bicarbonate in 50% acetonitrile, and subsequently tion criteria used included a maximal protein mass of 250,000 Da, a
dried in a Speedvac Concentrator. The gel was incubated in 50 μL of a detection of minimal 3 fragment ions per peptide, minimal 7 fragment
0.003 μg/μL modified trypsin solution (Promega) in 50 mM ammonium ions per protein and minimal 1 peptide per protein. Variable mod-
bicarbonate, pH 8.6, and incubated at 37 °C overnight. The supernatant ifications (oxidation (M) and deamidation (NQ)) were selected.
was transferred to another tube. Peptides that were retained in the gel Maximally one missed cleavages and a false positive rate of 4% was
spot were extracted using 80% acetonitrile (ACN) containing 0.5% allowed.
trifluoroacetic acid. The pooled tryptic peptide extracts were lyophi-
lized and resuspended in 15 μL of 0.1% formic acid for mass spectro- 2.6. Nitrite quantification
metric analysis.
Tryptic digests from each spot of interest were analyzed using The quantification of nitrite in the growth media of D. desulfuricans
MALDI-TOF MS. Therefore, 1 μL of the peptide mixture was mixed with cells (nitrate or sulfate based) was performed by the Griess colorimetric
an equal volume of matrix solution, i.e. 50 mM α-cyano hydro- assay [27]. The sample (50 μL) was mixed with 0.45 mL of 0.2 M
xycinnamic acid (Sigma) in 50% acetonitrile containing 0.1% formic phosphate buffer, pH 7.6, followed by sequential addition of 0.25 mL of
acid and deposited onto the MALDI target plate. MALDI mass spectro- sulphanilamide (Merck) [1% (m/v)] and 0.25 mL of n-(1-naphtil)
metric analyses were performed on a 4800 Plus MALDI TOF/TOF ethylenediamine (NEDA, Merck) [0.02% (m/v)]. This solution was
Analyzer (Absciex) in MS and MS/MS mode (both in reflectron mode), mixed in a vortex and incubated at room temperature for 10 min, after
both calibrated with a standard calibration mixture provided by the which the absorbance at 540 nm was recorded using a Shimadzu
manufacturer. The MALDI data were processed with MASCOT software UV1800 spectrophotometer.
searching against the non-redundant NCBI protein database
(20150726) selecting for ‘bacteria’, using the public Matrix Science 2.7. Nitrite and nitrate reductases activities
server (www.matrixscience.com). Variable modifications (oxidation
(M) and deamidation (NQ)) were selected. Other parameters were: 1 The nitrite and nitrate reductase activities of the D. desulfuricans
missed cleavage, peptide tolerance 0.35 Da and MS/MS tolerance protein samples were tested in native polyacrylamide gels. Samples
0.25 Da. were incubated in sample buffer [0.0625 M tris-HCl, pH 6.8, 10% (v/v)
Proteins that could not be identified using MALDI-TOF MS were glycerol and 0.01% bromophenol blue] and loaded on native 7.5%
further analyzed using LC-MS. Therefore, peptide mixtures were polyacrylamide gels. Electrophoresis was carried out at 100 V for 1 h

1457
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

after an initial period of 15 min at 80 V. Gels were incubated for 10 min


in three different solutions: 1) control − 100 mM phosphate buffer,
pH 7.6, 0.6 mM methyl viologen (Aldrich), 6 mM sodium dithionite
(Merck); 2) nitrate activity - same solution as control with 20 mM so-
dium nitrate (Merck) as substrate 3) nitrite activity - same solution as
control with 20 mM sodium nitrite (Merck) as substrate. The nitrite (or
nitrate) reductase activity was identified by the appearance of a col-
orless blot in the gels, which indicates that the immobilized enzyme
consumes the reduced methyl viologen (blue colored), the co-substrate
of the catalytic reaction with nitrite (or nitrate). No activity could be
detected when either the purified NrfA or the protein extracts were
incubated in the absence of nitrite or nitrate (i.e. with the control so-
lution).

Fig. 1. Growth curves of D. desulfuricans cells in (◆) nitrate and (■) sulfate based media.
2.8. mRNA Expression levels
Inset: specific growth rates (μ) and doubling times (Td). Error bars represent the standard
deviations from the mean values for three independent cultures.
Cell pellets were resuspended in NZYol (NZYtech) in a proportion of
100 μg pellet/500 μL NZYol. After incubation at room temperature for
15 min, the RNA was extracted according to the manufacturer in- 3.2. Differential protein expression
structions. To ensure total DNA removal, RNA samples were treated for
20 min with DNase I (Fermentas, ThermoFisher Scientific) using the To identify the changes induced by the different respiratory sub-
procedure recommended by the manufacturer. The reaction was strates at the molecular level, 2DE proteomic studies were performed
stopped by adding 2 volumes ethanol 99% (AGA), followed by an on total protein extracts of D. desulfuricans cells grown in sulfate or
overnight incubation and a centrifugation of 15,000 × g, 20 min at 4 °C. nitrate containing media harvested at 16 h (stationary phase). Due the
After removal of the supernatant, the pellet was washed with 1 volume complexity of proteome maps and the high number of proteins visua-
ethanol 75% (v/v), centrifuged at 15000 × g, 15 min at 4 °C and the air- lized in the pH range from 3 to 10 (not shown), we chose to analyze the
dried pellet was resuspended in water treated with diethyl pyr- samples in a narrower interval (4–7) where most proteins are found; at
ocarbonate 0.1 % (v/v) (DEPC, Sigma). cDNA synthesis was performed the same time, gel resolution is increased.
with 500 μg RNA using the NZY M-MuLV First-Strand cDNA synthesis The gels obtained from protein extracts of sulfate and nitrate grown
kit (NZyTech) according to the manufacturer instructions. The expres- cells were compared in terms of the number of spots and their relative
sion of genes involved in nitrate metabolism, cytochrome c nitrite re- intensities (Fig. 2). To obtain reliable results, three gels (i.e. technical
ductase catalytic subunit NrfA (nrfA) and the periplasmic nitrate re- replicates) were analyzed from three different growths cultures of each
ductase (napA), sulfate metabolism, dissimilatory sulfite reductase medium (i.e. biological replicates), summing up to a total of nine gels
alpha subunit (Ddes_0921) and adenylyl-sulfate reductase (apsA), and as per condition. Overall, most protein spots are observed across the range
an internal control, DNA polymerase I (polA) (primer sequences in of pH 5 to 7, with relative molecular masses between 40 and 100 kDa.
Table S1) were evaluated by quantitative real time PCR (qPCR) using Image analysis detects 604 spots in nitrate gels and 519 spots in sulfate
the Hot FirePol Evagreen qPCR Mix Plus (ROX) (Solis BioDyne) per- gels (average values), whereas the overlap between the spots of gels
formed in a Corbett Research Rotor-Gene RG3000 (Qiagen), using the from the same condition is over 98%. Because the total protein extracts
program described in Table S2. The analysis of the relative gene ex- were analyzed, which should encompass all expressed proteins (soluble
pression variation in each sample was performed using the 2− ΔΔCt and membrane) in the cell, we expected a higher number of protein
method [28]. Significant differential expression was considered for spots in the gels. Several reasons may contribute for this. First, it must
samples with 2− ΔΔCt values higher than 1.5 (2− ΔΔCt > 1.5) or lower be considered that the analysis presented here refers only to the pH
than 0.7 (2− ΔΔCt < 0.7). range of 4–7. Second, sampling was restricted to the stationary phase,
which means that proteins expressed in other time windows are not
represented in the proteomes. Third, the inherent limitations of 2DE
3. Results and discussion could also explain the relatively small number of spots detected. For
instance, incomplete protein solubilization or cell lysis may well lead to
3.1. Cell growth the loss of some proteins, namely membrane proteins [29]. Proteins
with a low copy number in the cell will also be underrepresented in the
To evaluate the effects of the oxidizing substrate on cell growth, D. gel maps, both due to the difficulties in extraction/solubilization pro-
desulfuricans was cultured in a lactate containing medium supple- cess and staining detection limits. The staining method used (CCB) is
mented with either nitrate or sulfate salts. As expected, the cell growth adequate for the quantification of spot intensities and their relationship
in nitrate medium is faster than in sulfate medium (Fig. 1) [11]. Ac- with protein concentration, but is less sensitive than, for example,
cordingly, the specific growth rates (μ) and culture doubling times (Td) fluorescent staining, which provides lower detection limits [30].
are ca. 1.5 times lower in the former case (Fig. 1, inset). Next, a comparative analysis on the 2DE gels was conducted to
The maximum OD in the presence of sulfate, observed in the sta- identify the changes in the protein profiles of D. desulfuricans cells
tionary phase, is about 50% of the one obtained in the presence of ni- grown with either sulfate or nitrate. A high number of protein spots
trate. Moreover, the biomass harvested after 16 h of inoculation is al- could not be matched between the gels of the two conditions. These
most doubled in the nitrate containing media. These results have been spots, which we designate “exclusive”, are detected mainly in the gels
attributed to the fact that nitrate reduction is thermodynamically more of nitrate grown cells (81 vs 7 found solely in the gels from sulfate
favorable than sulfate reduction and so the cells grow faster when using obtained cells). Note that, some of these spots may in fact have a very
this electron acceptor. Besides, the final products of sulfate reduction low abundance in the gel in which they were not detected, thus being
(sulfide) might be toxic to the cells and therefore, could inhibit their below the detection limits of the staining method and image analysis
growth [11,16]. software. By using a 2-fold spot volume intensity ratio and p-value
(ANOVA) lower than 0.05 as criteria, we found that 40 additional

1458
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

Fig. 2. 2D gel maps of total protein fractions of D. de-


sulfuricans grown for 16 h in (A) nitrate and (B) sulfate
containing media. Circles indicate protein spots with in-
creased intensity in each condition. The spot numbers
marked with * represent proteins that are exclusively de-
tected in one growth condition. Proteins (350 μg) were se-
parated by 2D gel electrophoresis (12.5% acrylamide) in a
pH gradient of 4–7. Gels were stained with colloidal
Coomassie blue. Gel images were acquired from Melanie 7.0
software.

1459
Table 1
Proteins with changed abundance in the total proteome extracts of D. desulfuricans grown in nitrate or sulfate containing media. *Theoretical values.

Spot nr Gene identifier Protein name and response in nitrate medium Functional categories Molecular mass pI* Fold change
J.R. Sousa et al.

(Da)*

Down-expressed
504 Ddes_0477 Dihydrodipicolinate reductase Amino acid biosynthesis: Aspartate family 38,638 5.68 − 3.68
164 Ddes_2166 Thiazole synthase Biosynthesis of cofactors, prosthetic groups, and carriers: Thiamine 27,537 5.98 − 4.43
10 Ddes_0134 Methyl accepting chemotactic protein Cell motility, Signal transduction mechanisms 87,676 5.21 − 4.50
528 Ddes_1494 Catalase Cellular processes: Detoxification 55,226 6.17 Exclusive
129 Ddes_0602 Peroxiredoxin Cellular processes: Detoxification 20,896 5.33 − 4.08
194 Ddes_0048 Glutaminase Energy metabolism: Amino acids and amines 32,762 5.07 − 2.59
307 Ddes_0173 NAD(P)(+) transhydrogenase (AB-specific) Energy metabolism: Electron transport 41,358 5.95 − 2.08
320 Ddes_0173 NAD(P)(+) transhydrogenase (AB-specific) Energy metabolism: Electron transport 41,332 5.95 − 2.20
340 Ddes_0174 NAD(P) transhydrogenase subunit alpha Energy metabolism: Electron transport 41,331 5.91 − 3.83
523 Ddes_1038 [NiFe] hydrogenase large subunit (Fragment) Energy metabolism: Electron transport 60,026 6.74 Exclusive
529 Ddes_1531 Succinate dehydrogenase or fumarate reductase, flavoprotein subunit Energy metabolism: TCA cycle 67,486 6.07 Exclusive
527 Ddes_1616 3-Oxoacyl-[acyl-carrier-protein] synthase 2 Fatty acid and phospholipid metabolism: Biosynthesis 43,771 5.44 Exclusive
524 Ddes_1751 Hypothetical protein Hypothetical proteins 8825 6.23 Exclusive
120 Ddes_1402 Intracellular protease, PfpI family Protein fate: Degradation of proteins, peptides, and glycopeptides 21,056 5.93 − 2.40
460 Ddes_1785 Elongation factor G Protein synthesis: Translation factors 76,332 5.29 − 2.73
411 Ddes_1338 AMP dependent synthetase and ligase Secondary metabolites biosynthesis, transport and catabolism 61,339 5.06 − 2.05
179 Ddes_1328 Methyltransferase type 12 Secondary metabolites biosynthesis, transport and catabolism 27,581 4.98 − 3.04
68 Ddes_0648 Nickel transporter Transport and binding proteins 30,429 6.54 − 3.14
448 Ddes_1343 TonB-dependent receptor Transport and binding proteins 83,615 4.78 − 2.45

Overexpressed
522 Ddes_1152 Glutamate dehydrogenase Amino acid biosynthesis: Glutamate family 48,913 5.97 Exclusive
275 Ddes_1676 Ornithine carbamoyl transferase Amino acid biosynthesis: Glutamate family 38,392 5.07 2.65
285 Ddes_1676 Ornithine carbamoyltransferase Amino acid biosynthesis: Glutamate family 38,405 5.00 2.48

1460
568 Ddes_1811 1-(5-phosphoribosyl)-5-[(5-phosphoribosylamino)methylideneamino] imidazole-4-carboxamide Amino acid biosynthesis: Histidine family 26,667 4.91 Exclusive
isomerase
564 Ddes_1017 Phosphoserine phosphatase Amino acid biosynthesis: Serine family 22,638 5.23 Exclusive
233 Ddes_1898 Pyridoxal biosynthesis protein; Pyridoxal 5′-phosphate synthase subunit PdxS Biosynthesis of cofactors, prosthetic groups, and carriers: 31,489 5.85 2.81
Pyridoxine
544 Ddes_0164 Cell shape determining protein, MreB/Mrl family Cell envelope: Biosynthesis and degradation of murein sacculus 37,306 5.43 Exclusive
and peptidoglycan
374 Ddes_1108 UDP-N-acetylmuramoyl-L-alanyl-D-glutamate-2,6-diaminopimelate ligase Cell envelope: Biosynthesis and degradation of murein sacculus 52,646 5.49 2.99
and peptidoglycan
534 Ddes_2048 NAD-dependent epimerase/dehydratase Cell wall/membrane/envelope biogenesis 33,603 5.15 Exclusive
575 Ddes_2281 Flagellin Cellular processes: Chemotaxis and motility 41,718 4.84 Exclusive
587 Ddes_2281 Flagellin Cellular processes: Chemotaxis and motility 31,603 5.21 Exclusive
589 Ddes_1028 Flagellin Cellular processes: Chemotaxis and motility 32,110 5.23 Exclusive
396 Ddes_1829 Hydroxylamine reductase Cellular processes: Detoxification 58,499 6.04 2.69
601 Ddes_1829 Hydroxylamine reductase Cellular processes: Detoxification 58,622 5.98 Exclusive
135 Ddes_0538 Superoxide dismutase Cellular processes: Detoxification 22,526 5.43 2.57
494 Ddes_0298 Pyruvate synthase Energy metabolism: Electron transport 128,159 5.32 2.11
267 Ddes_0132 Glyceraldehyde 3-phosphate dehydrogenase Energy metabolism: glycolysis, gluconeogenesis 36,242 5.89 2.67
536 Ddes_0480 Transketolase Energy metabolism: Pentose phosphate pathway 71,527 5.57 Exclusive
314 Ddes_0101 Isocitrate dehydrogenase Energy metabolism: TCA cycle 46,840 5.26 4.28
333 Ddes_0101 Isocitrate dehydrogenase Energy metabolism: TCA cycle 46,840 5.26 3.01
586 Ddes_2012 Lactamase Energy production and conversion 45,380 5.98 Exclusive
554 Ddes_1704 Hypothetical protein Hypothetical proteins 13,641 5.14 Exclusive
420 Ddes_1548 GroEL chaperone Protein fate: Protein folding and stabilization 56,997 4.82 3.00
546 Ddes_2248 GTPase Protein synthesis (ribossome biogenesis) 40,419 4.91 Exclusive
499 Ddes_1785 Translation elongation factor G Protein synthesis: Translation factors 76,437 5.04 2.45
359 Ddes_1628 Elongation factor Tu Protein synthesis: Translation factors 43,293 4.87 3.39
608 Ddes_0010 Aspartyl-tRNA synthetase Protein synthesis: tRNA aminoacylation 70,434 5.60 Exclusive
588 Ddes_2264 Phosphoribosylaminoimidazolecarboxamide formyltransferase Purines, pyrimidines, nucleosides, and nucleotides: Purine 45,565 4.99 Exclusive
BBA - Proteins and Proteomics 1865 (2017) 1455–1469

(continued on next page)


J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

Fold change protein spots display a different relative intensity in the gels from the

Exclusive
Exclusive

Exclusive
two conditions. Among these, 17 have lower abundance and 23 have

2.28

2.05
2.19
higher abundance in the cells grown in nitrate medium. Furthermore, a
careful analysis of the fold change values (Tables 1 and S3) shows that,

5.10
5.89
6.86

5.23
4.89
5.28
in general, they are higher for the proteins increased in sulfate grown
pI*

cells, which is a logical result since the total amount of protein loaded
in the gels, from both samples, sulfate vs nitrate, is the same, but the
number of proteins observed is much lower in the former case. Overall,
Molecular mass

the 128 differentially expressed proteins represent about 25% of the


total spots identified in 2D gels (4–7 pH range), which is a clear in-
41,851
41,054
47,667

38,859

12,404

dication that the respiratory substrates have important consequences in


(Da)*

8793

the cellular composition of D. desulfuricans, considerably affecting its


protein profile. These striking differences and the high number of
protein spots that are present almost exclusively in nitrate respiring
Purines, pyrimidines, nucleosides, and nucleotides: Pyrimidine

Secondary metabolites biosynthesis, transport and catabolism


Secondary metabolites biosynthesis, transport and catabolism

cells reveals that the bacterium requires an additional metabolic effort


to produce the proteins responsible for its adaptation to a nitrate rich
environment. We propose that while growing in sulfate containing
medium, the cells only need to produce the proteins necessary for their
base metabolism. This provides a rationale for the preference that D.
desulfuricans has demonstrated for sulfate reduction in the presence of
both substrates (nitrate and sulfate) [11–13]. To understand the pro-
cesses involved in the adaptation of D. desulfuricans to a nitrate based
culture medium the identity and function of the changed proteins was
ribonucleotide biosynthesis

further assessed in the next section. One should mention that our pre-
liminary work showed that samples collected at 8 and 16 h exhibited
Functional categories

similar proteomic profiles. That is why we opted to analyze cells har-


Unknown function

vested at 16 h (stationary phase), when higher cell yields (particularly


Transcription
Transcription

for sulfate grown cells) could be obtained [31].


metabolism

3.3. Identification and functional categorization of differential proteins

The protein spots showing different intensities (increasing or de-


creasing) and those exclusive found in a single growing condition (ni-
trate vs sulfate) were excised from the 2DE gels for analysis by mass
spectrometry (Tables S3 and S4). The identified proteins were listed
and grouped into functional role categories according to the annotation
provided by TIGRFAMs (Table 1) [32–34].
In general, there is a good agreement between theoretical molecular
mass and pI values and the actual electrophoretic mobility of the pro-
teins in the 2D gels. Discrepancies can be attributed to spots re-
presenting degradation products of larger protein species, protein ag-
gregates that do not disintegrate during electrophoresis and proteins
with extensive post-translational modifications [35,36]. Dissimilar 2D
spots representing the same protein or protein complex are observed for
isocitrate dehydrogenase, ornithine carbamoyl transferase, flagellin,
DNA-directed RNA polymerase subunit omega
Protein name and response in nitrate medium

DNA-directed RNA polymerase subunit alpha

hydroxylamine reductase, NAD(P) transhydrogenase, elongation factor


Pyridine nucleotide-disulfide oxidoreductase

DegT/DnrJ/EryC1/StrS aminotransferase

G and DNA-directed RNA polymerase. The identification of different


forms of a protein species with different expression levels may com-
2-Hydroxyglutaryl-CoA dehydratase

plicate interpretation of proteomic data. This is not the case in our work
since for most spots identified as the same protein the relative abun-
dance values follow the same trend, i.e., they are increased in the same
cell extract. Furthermore, they are usually found in closely related
positions in the 2D gels (same molecular mass and different pIs). This
suggests the spots represent isoforms of the same protein species or the
proteins have undergone post-translational modifications.
The proteins that show significant expression changes belong to
cupin

diverse functional categories, major groups being related with amino


acid or protein biosynthesis and energy metabolism. This is consistent
Gene identifier

with significant metabolism adaptations. No reliable identifications


were obtained for half of all tested spots (Table S3), which can be at-
Ddes_1416

Ddes_0545
Ddes_2174
Ddes_0686
Ddes_2161
Ddes_0206
Table 1 (continued)

tributed to insufficient quality of the MS spectra and/or a low abun-


dance of peptides. In fact, most of the unidentified proteins represent
faint spots, in particular the ones that were detected exclusively in the
Spot nr

gels of nitrate grown cells. In the next sections, we discuss the proteins
582

301
590
302
518
556

and cellular processes affected in D. desulfuricans by the use of different


electron acceptors.

1461
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

Fig. 3. Proposed lactate oxidation/energy transduction


pathway in D. desulfuricans cells grown in nitrate or sulfate
containing media. Orange and dark green boxes indicate
proteins with higher or lower abundance in nitrate grown
cells, respectively. Beige and light green boxes indicate the
proteins involved in nitrate or sulfate reduction, respec-
tively. Gray boxes indicate proteins not identified in this
study. PFOR: pyruvate synthase; Hases: hydrogenases;
APSR: adenylyl-sulfate reductase; SR: sulphite reductase;
[NiFe]Hase:[NiFe] hydrogenase; NrfA: cytochrome c nitrite
reductase, catalytic subunit; NrfH: nitrite reductase, small
subunit; NapA: periplasmic nitrate reductase. (For inter-
pretation of the references to color in this figure legend, the
reader is referred to the web version of this article.)

3.3.1. Energy and carbon metabolism [46]. In addition, we did not identify genes of some TCA enzymes in the
Several proteins involved in central carbon metabolic pathways genome sequence of D. desulfuricans (e.g. citrate synthase and succinyl-
display relative abundance differences. For instance, glyceraldehyde-3- coA synthetase). These evidences indicate that the bacterium does not
phosphate dehydrogenase and transketolase are overexpressed during express a full TCA cycle. Hence, the higher abundance of the TCA en-
growth of D. desulfuricans on nitrate medium. These are key enzymes in zyme isocitrate dehydrogenase, is not directly correlated with an in-
glycolysis and the pentose phosphate pathway, respectively, and pro- creased energy production via TCA cycle. Instead, the up-regulation of a
vide links between the two metabolic pathways. Glyceraldehyde-3- partial TCA cycle (possibly citrate to 2-ketoglutarate) is related with
phosphate dehydrogenase catalyzes the interconversion of 1,3-dipho- biosynthetic pathways. Isocitrate dehydrogenase catalyzes the re-
sphoglycerate and glyceraldehyde-3-phosphate (in glycolysis and glu- versible oxidative decarboxylation of isocitrate to 2-ketoglutarate,
coneogenesis), whereas transketolase is responsible for two separate coupled to the reduction of NAD(P)+ to NAD(P)H; but the reaction is
reactions, in the non-oxidative branch of the pentose phosphate also an important branch-point for amino acid biosynthesis, since 2-
pathway, both yielding glyceraldehyde-3-phosphate [37,38]. The ketoglutarate is used as a carbon skeleton for the production of amino
higher expression of these proteins suggests increased flow through acids from the glutamate family [47], pathways that are also up-regu-
these metabolic pathways. The up-regulation of pyridoxal 5′-phosphate lated in the cells obtained from nitrate culture medium (cf. Section
synthase (cf. Table 1 and Section 3.3.2) indicates that glyceraldehyde-3- 3.3.2). In contrast with isocitrate dehydrogenase, succinate dehy-
phosphate may be directed to the production of vitamin B6, an im- drogenase/fumarate reductase, which also takes part in the TCA cycle
portant cofactor in many enzyme reactions [39,40]. (succinate/fumarate reversible conversion), has lower expression in the
Another protein with increased expression in nitrate grown cells and cells grown in nitrate. The enzyme was previously shown to function as
involved in carbon metabolism is pyruvate synthase, which catalyzes fumarate reductase in D. desulfuricans (strain Essex) enabling fumarate
the decarboxylation of pyruvate to acetyl-CoA. It has been proposed respiration or disproportionation [48,49]. The decrease of fumarate
that this enzyme is very important in the energy metabolism in SRB, reductase expression suggests the reductive branch of the TCA cycle is
since it is involved in energy conservation through substrate level more important for the energy metabolism of D. desulfuricans while
phosphorylation [41,42]. The enzyme is central to the lactate oxidation reducing sulfate.
pathway as pyruvate is a key intermediate of the oxidation of organic Other proteins involved in energy metabolism displaying abundance
substrates (Fig. 3) [43]. Therefore, we can associate the higher abun- changes, also with lower expression in nitrate grown cells, include
dance of pyruvate synthase with the faster cell division rates and con- [NiFe] hydrogenase, NAD(P)(+) transhydrogenase and glutaminase. In
sequent higher energy demand of the cells grown in nitrate medium. fact, high levels of periplasmic [NiFe] hydrogenases were previously
SRB species usually oxidize their organic substrates incompletely to observed in D. desulfuricans strains growing on sulfate and lactate. This
acetate to generate energy (NAD(P)H and ATP production) [44,45]. supports a hydrogen cycling model of energy conservation, according to
Numerous species, including D. vulgaris, display incomplete pentose which cytoplasmatic hydrogenases convert the protons and electrons
phosphate pathway and TCA cycles, which are thought to be used derived from lactate oxidation into hydrogen. This molecule then dif-
mainly for the biosynthesis of amino acids and nucleic acids [34,45]. fuses across the membrane and is reoxidized by periplasmic hydro-
Nevertheless, complete oxidation of acetate to carbon dioxide has been genases, generating a proton gradient that can be used for ATP synth-
observed in some SRB [45]. Certain Desulfobacter species have full TCA esis [34,50,51]. Therefore, [NiFe] hydrogenases are considered key
cycles (slightly different from aerobic bacteria cycles), whereas species players for energy production in SRB growing on sulfate and lactate
that do not possess complete TCA cycles can use the carbon monoxide (Fig. 3). Our proteomic data show that in the absence of sulfate the
dehydrogenase pathway to oxidize acetate [44–46]. Higher production protein is not detected, suggesting that the bacterium favors other
of carbon monoxide dehydrogenase was previously observed in D. de- mechanisms for energy production. The increased flow through the
sulfuricans growing with lactate and sulfate and under molybdate stress pyruvate synthase reaction, for example, points towards ATP synthesis

1462
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

via substrate level phosphorylation. expectedly high concentration of ammonium generated from nitrate
NAD(P) transhydrogenases regulate the NADPH/NADP+ ratio in- respiration (Fig. 4). By taking up a TCA intermediary, glutamate de-
side cells. Under normal conditions, the balance is kept high since hydrogenase is also an important link between the carbon and nitrogen
NADPH is required for the biosynthesis of cell components [52]. The metabolisms. The earlier mentioned up-regulation of isocitrate dehy-
sequence of D. desulfuricans reveals 6 genes for NAD(P) transhy- drogenase can therefore be associated with 2-ketoglutarate derived
drogenase and we have identified three protein spots of similar mole- amino acids biosynthesis. Similarly, higher arginine production may be
cular mass and pI, with lower abundance in nitrate grown cells, which inferred from augmented ornithine carbamoyl transferase levels. This
represent 2 protein species (Ddes_174 and 2 isoforms of Ddes_173). We protein is involved in an intermediary step in the arginine biosynthetic
associate the decrease of NAD(P) transhydrogenase with the higher pathway, where it catalyzes the reversible conversion of carbamoyl
flow through NADPH generating reactions, such as the ones catalyzed phosphate and ornithine (the product of glutamate degradation) into
by glyceraldehyde-3-phosphate and isocitrate dehydrogenases [52,53]. citrulline [57]. Our proteomic data point to the assimilation of am-
In contrast, the reduced flow through energy generating pathways in monium derived from nitrate reduction as a driver of amino acid bio-
sulfate respiring cells may be compensated by a higher abundance of synthesis, thereby contributing to the higher bacterial growth rates and,
NAD(P) transhydrogenases, that increase the levels of NADPH, and thus at the same time, dealing with potentially toxic elevated levels of am-
maintain the cellular redox equilibrium. monium inside the cell (Fig. 4). In addition to the amino acids from the
Glutaminase, catalyzes the deamination of glutamine to glutamate glutamate family, enhanced production of serine and histidine is in-
and ammonium. It has been proposed that the enzyme controls the ferred from the higher abundance of phosphoserine phosphatase and 1-
cellular levels of glutamine, which is a central nitrogen metabolite. (5-phosphoribosyl)-5-[(5-phosphoribosylamino)methylideneamino]
Glutamine is essential to numerous cellular processes, such as the imidazole-4-carboxamide isomerase, respectively (cf. Table 1) [58,59].
synthesis of both purine and pyrimidine nucleotides [54–56]. We as- In contrast, dihydrodipicolinate reductase, a component of lysine bio-
sociate the decreased abundance of glutaminase, in cells obtained from synthetic pathway from aspartate, which is also ammonium assim-
nitrate medium, with the need to maintain adequate levels of in- ilating, has lower levels in the cells obtained from nitrate media. Thus
tracellular glutamine to be used in biosynthetic processes. Additionally, far, we could not find an explanation for this apparently contradictory
because the enzyme reaction releases ammonium, its down-expression result.
may prevent further accumulation of this product, already being gen- Increased biosynthesis of nucleotides in nitrate grown cells is sug-
erated from the respiration of nitrate (Fig. 4). gested by the up-regulation of phosphoribosylaminoimidazolecarbox-
amide formyltransferase. The protein catalyzes the two final reactions
3.3.2. Biosynthesis of amino acids, nucleotides and cofactors of the purine ribonucleotide biosynthetic pathway yielding inosine
A majority of the identified protein spots belonging to functional monophosphate (IMP), a central intermediate in purine metabolism
categories related to biosynthesis of amino acids, nucleotides and sec- [60]. The enzyme substrate, 5-aminoimidazole-4-carboxamide ribonu-
ondary metabolites, display higher relative abundances in D. desulfur- cleotide (AICAR), is a byproduct of histidine biosynthesis also shown to
icans cells grown with nitrate as electron acceptor. The result is in line be augmented [61]. Furthermore, 5-phosphoribosyl diphosphate
with the higher cell yields and overall faster growth observed in nitrate (PRPP), the precursor of purine, pyrimidine and histidine metabolisms
medium. is derived from ribose-5-phosphate, one of the products of the reaction
As shown in Table 1, at least 4 proteins belonging to the biosyn- of the pentose phosphate pathway enzyme transketolase [38,59], which
thetic pathways of different amino acids have increased expression is up-regulated under the same growing conditions (cf. Table 1 and
when nitrate is the respiratory substrate of D. desulfuricans. The in- Section 3.3.1).
creased levels of glutamate dehydrogenase and ornithine carbamoyl Aside from amino acids and nucleotides, our study shows different
transferase are particularly interesting. The first enzyme plays a role in expression levels of proteins involved in cofactor biosynthesis, namely
glutamate synthesis converting 2-ketoglutarate and ammonium into pyridoxal 5′-phosphate synthase (subunit PdxS), augmented in nitrate
glutamate; thus, this reaction is well suited for assimilating the grown cells. The protein is involved in vitamin B6 metabolism; this

Fig. 4. Proposed ammonia assimilation pathways in D. de-


sulfuricans cells grown in nitrate or sulfate containing
media. Orange and dark green boxes indicate proteins with
higher or lower abundance in nitrate grown cells, respec-
tively. Beige boxes indicate the proteins involved in nitrate
reduction. Gray boxes indicate proteins not identified in
this study. IDH: isocitrate dehydrogenase; GDH: glutamate
dehydrogenase. (For interpretation of the references to
color in this figure legend, the reader is referred to the web
version of this article.)

1463
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

cofactor has an important role in several enzyme processes, including Herein we propose that decreased superoxide dismutase levels may be
amino acid and fatty acid metabolism [62,63]. Pyridoxal 5′-phosphate compensated by the other active oxidative stress defenses. Conversely,
synthase contains a second subunit with glutaminase activity, which the protein could play a more significant role in the scavenging of
yields ammonia from glutamine hydrolysis that is incorporated into the oxygen radicals during growth of D. desulfuricans in the presence of
pyridine ring of pyridoxal 5′-phosphate [64]. Consequently, higher nitrate. This activity seems particularly relevant in these conditions in
abundance of pyridoxal 5′-phosphate synthase is consistent with in- order to prevent the formation of the highly reactive peroxynitrite ion.
creased activities of enzymes involved in the biosynthesis of the glu- Our study reveals two protein spots with increased intensity in ni-
tamate family of amino acids and with glyceraldehyde-3-phosphate trate grown cells, which were both identified as hydroxylamine re-
dehydrogenase and transketolase from carbon metabolic pathways. ductase. The protein reduces the highly toxic and mutagenic hydro-
xylamine to ammonium. The reaction product can subsequently be
3.3.3. Transcription and protein synthesis assimilated through amino acid biosynthesis, as discussed in earlier
Higher levels of DNA-directed RNA polymerase, which takes part in sections. Hence, hydroxylamine reductase is likely involved in ni-
the gene transcription process, are observed in the cells obtained from trosative stress response [78]. An increased production of nitrogen re-
nitrate medium. This agrees with the overall up-regulation of protein active species is expected during nitrate respiration. The activation of
biosynthesis observed in that condition. Additionally, several proteins hydroxylamine reductase may indicate an incomplete reduction of ni-
belonging to the functional category of protein synthesis have increased trite to hydroxylamine by NrfA, the enzyme responsible for the final
cellular levels. Obg GTPase and the elongation factors G and Tu are step of the dissimilatory nitrate reduction to ammonium; however,
translational GTPases involved in protein biosynthesis at the ribosome previous work has shown that NrfA enzymes do not release detectable
[65–67]. The latter two have a concerted action during the elongation intermediaries in the course of their catalytic cycle [79]. Still, NrfA is
phase of protein synthesis. The elongation factor Tu's main functions capable of reducing nitric oxide and hydroxylamine, which also sup-
are to bind GTP and aminoacyl-tRNA and deliver it to the ribosome. ports its active role in the protection against nitrogenous stressors [80].
Upon matching to the mRNA, and GTP hydrolysis, the peptide bond is In addition to the detoxification of hydroxylamine, hydroxylamine re-
formed and elongation factor G translocates the mRNA one codon to ductase may be involved in hydrogen peroxide stress defense. Previous
allow for the arrival of a new aminoacyl-tRNA [66,67]. D. desulfuricans studies, with the purified enzyme from D. desulfuricans demonstrated
genome has multiple genes for elongation factors Tu and G. This could that its KM values for hydrogen peroxide are within the range of other
explain why our data also shows one down-expressed elongation factor peroxidases [81].
G protein and suggests different patterns of gene regulation in response
to the available electron acceptor. 3.3.5. Other cellular processes
Cells grown in nitrate containing medium show higher abundance
3.3.4. Stress response and detoxification of flagellin, a main component of the bacterial flagellum. Improved
The stress response proteins peroxiredoxin and catalase are more motility allows the cell to reach available nutrients easily, which sup-
abundant in cells grown in sulfate medium. These proteins are im- ports the faster growth observed in nitrate respiring cells. Similarly,
portant for the protection and survival of cells as they play an important increased levels of a cell shape determining protein from MreB/Mrl
role in antioxidative defense mechanisms from damage induced by family, UDP-N-acetylmuramoyl-L-alanyl-D-glutamate-2,6-diaminopime-
reactive oxygen species (ROS). Catalase catalyzes the reduction of hy- late ligase and NAD-dependent epimerase/dehydratase, all involved in
drogen peroxide to water and oxygen, while peroxiredoxins are able to the formation of the cell wall, is in line with the rapid cell division rates.
eliminate peroxynitrite, organic hydroperoxides as well as hydrogen Moreover, a GroEL chaperone is overexpressed in nitrate cultured
peroxide [68–70]. Some Desulfovibrio species, including desulfuricans cells. The protein is a component of the GroEL/GroES system, which is
strains, can utilize oxygen as respiratory substrate. The SRB oxygen essential for protein folding. Increased abundance of GroEL chaperone
reducing systems are mainly thought to be involved in protective assures the maintenance of optimal conditions for the formation of the
functions and should enable the bacteria to survive in the oxic en- quaternary structure of proteins and the assembly of oligomeric com-
vironments of by lowering the local oxygen concentrations [2,71,72]. plexes [82]. This is consistent with the overall high protein expression
Aside from oxygen reduction, these systems can also generate ROS. in nitrate grown cells. In addition, these molecular chaperones are
Within SRB, D. desulfuricans has a particularly high tolerance to oxygen, important to prevent nonspecific association of polypeptides under
being able to grow under nearly atmospheric oxygen levels. Accord- stress conditions (e.g. end of the exponential growth phase).
ingly, the bacterium expresses not only oxygen reductases but also Interestingly, two transporter proteins are down-expressed in the
proteins that allow the efficient removal of ROS [12]. Under the con- samples obtained from nitrate medium, a TonB-dependent receptor and
ditions employed in this work (anoxic batch cultures), oxygen con- a nickel transporter. TonB-dependent receptors are outer membrane
centrations should be very low. Still, the residual levels may induce the transporters that couple proton gradient with the uptake of iron che-
auto-oxidation of some reduced proteins, such as cytochromes and lates, vitamin B12, carbon substrates and nickel complexes into the
flavoproteins, which can generate ROS [2,73]. More importantly, the periplasmic space [83,84]. As previously reported, the presence of
auto-oxidation of sulfide, the end-product of sulfate reduction, should nickel transport systems is mostly related with the occurrence of nickel-
significantly increase the oxidative stress in the cells in comparison dependent metalloenzymes [85]. In this case, decreased abundance of
with the ones growing in nitrate containing medium [73,74]. There- the TonB-dependent receptor and nickel transporter proteins is in
fore, the increased level of antioxidant proteins is consistent with the agreement with the earlier discussed lower expression of [NiFe] hy-
end of the exponential phase and the accumulation of stress products, drogenase, which is one of the major nickel-binding enzymes in bac-
especially sulfide [71,75]. In contrast with peroxiredoxin and catalase, terial cells [85,86].
superoxide dismutase is down-expressed in the samples from sulfate
cultures. This protein protects cells from free superoxide radicals by 3.4. The terminal enzymes of nitrate reduction
converting them to hydrogen peroxide and oxygen. Previous studies
suggested that superoxide dismutases are constitutively expressed in The enzymes involved in the nitrate reduction pathway in D. de-
some Desulfovibrio species and that activities do not increase sig- sulfuricans, namely the periplasmic NapA and the membrane-bound
nificantly with exposure to oxygen [12,76,77]. The presence of me- NrfA, could not be identified in the proteomic analysis. This is not
chanisms that defend the cells against oxidative stress, even under surprising, considering the intrinsic limitations of 2D electrophoresis to
anoxic conditions, certainly confers a major adaptive advantage to SRB, analyze hydrophobic proteins, such as NrfA. In fact, our previous
facilitating a rapid adjustment to different environmental conditions. findings suggested that NrfA cannot be loaded and/or separated in the

1464
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

Fig. 5. Relative expression of genes involved in nitrate metabolism (periplasmic nitrate Fig. 6. Relative expression of genes involved in nitrate metabolism (periplasmic nitrate
reductase (napA) and cytochrome c nitrite reductase, catalytic subunit (nrfA)) and sulfate reductase (napA) and cytochrome c nitrite reductase, catalytic subunit (nrfA)) at different
metabolism (adenylyl-sulfate reductase (apsA) and dissimilatory sulfite reductase alpha stages of D. desulfuricans growth curve. Data was normalized to the expression of poly-
subunit (dsrA)), when D. desulfuricans is grown for 16 h in nitrate vs sulfate medium. Data merase A (polA) of cells collected after 3 h growth. Error bars represent standard error of
was normalized to the expression of polymerase A (polA) of cells growing in the same at least three independent experiments. The dotted lines represent the cut-off where
condition. Error bars represent standard error of at least three independent experiments. overexpression (2− ΔΔCt > 1.5) or down-expression (2− ΔΔCt < 0.7) was considered.
The dotted lines represent the cut-off where overexpression (2− ΔΔCt > 1.5) or down-
expression (2− ΔΔCt < 0.7) was considered.
nitrate reduction to nitrite by NapA. The gene was expressed in the
early exponential phase, which is consistent with increased energy
2D gels [31]. In the case of NapA, it is possibly excluded from the demands in the cells to support the accelerated growth rates measured.
analysis because its pI (theoretical value 7.97) is outside the range Upon entering the stationary phase, cell division is significantly lower
studied in this work (4–7). Therefore, we have tested NapA and NrfA and the NrfA accumulated earlier was enough to process nitrite and
catalytic activities as well as their gene expression by mRNA quantifi- maintain an adequate flow through energy production pathways.
cation.

3.4.1. Gene expression 3.4.2. In-gel activity


To compare the transcript levels between the nitrate and sulfate In the next step, the nitrate and nitrite reductases activities of the
cultures, we used a normalized quantity (compared to sulfate) that is protein extracts of D. desulfuricans cells were evaluated in native
directly proportional to transcript abundance (Fig. 5). polyacrylamide gels. The colorless activity blot observed in the gel
The analysis revealed that, at the stationary phase, napA is up- containing the nitrate grown cells was considerably larger than the one
regulated in the nitrate grown cells, which agrees with the activation of present in the gel of sulfate cell extracts, both obtained after 16 h of
the nitrate respiration pathway in the presence of that electron ac- growth (Fig. 7A). This indicated a much higher nitrite reducing activity
ceptor, as reported for several SRB species [44]. Surprisingly, the in the former cell samples, which apparently contradicts the nrfA
transcript levels of nrfA, responsible for the second step in bacterial transcript levels measured in that experimental condition, but supports
nitrate reduction, were decreased. In previous works, it was shown that our proposal that after being highly expressed and accumulated during
while napA genes were strongly induced in the presence of nitrate, as the exponential phase, NrfA remains active inside the nitrate grown
also observed in the present study, nrfA may be constitutively expressed cells. Therefore, after being up-regulated, nrfA expression decreases
in D. desulfuricans (nitrite or nitrate only slightly increased its transcript explaining why the abundance of the gene transcript is no longer ele-
levels) [16,87]. The constitutive expression of nrf genes guarantees an vated after 16 h. The results obtained with cells harvested from nitrate
immediate response upon cell exposure to nitrite. Still, this does not medium after 3, 6 and 12 h of culture also back this conclusion; they
explain why nrfA gene was repressed when the cells are grown with reveal increased enzymatic activity, hence higher NrfA concentration in
nitrate as electron acceptor in comparison with the cells respiring sul- the cells, at 6 h and 12 h (Fig. 7B).
fate. Different authors have discussed the proteome/transcriptome re- The residual nitrite reducing activity detected in sulfate respiring
lations in cells. It has been established that although some genes have cells is in line with previous studies where NrfA was shown to be
similar transcript and protein trends, these parameters are not ne- constitutively produced in D. desulfuricans [16]. Unfortunately, the ni-
cessarily correlated [88–91]. To get a better understanding of the reg- trate reducing activity in the cell extracts was below the detection limit
ulation of the nitrate reduction pathway in D. desulfuricans, we eval- and therefore could not be monitored in this assay. As above men-
uated the gene expression levels during different stages of the tioned, this is probably due to the low turnover of NapA in comparison
bacterium's growth in the nitrate containing medium. Particularly, the with NrfA [4].
napA and nrfA levels were reported at 4 different time points (3, 6, 12
and 16 h of growth) as a relative ratio compared to the 3 h of growth 3.4.3. Nitrite levels
(Fig. 6). As a final step to understand the expression of the enzymes re-
The napA gene abundance did not vary significantly, as the tran- sponsible for the dissimilatory nitrate reduction in D. desulfuricans, we
script level ratios were maintained between 1.5 and 0.7 throughout the measured the nitrite levels in the nitrate and sulfate cultures over a 24 h
growth curve (Fig. 6). These results suggest napA transcript levels were period, accompanying the complete cell growth curves (Fig. 8). These
always kept high inside the cells because the gene is induced by the values should enable us to assess the activity of both NrfA and NapA. In
elevated nitrate concentrations in the medium [11,87]. Furthermore, sulfate containing medium, the values of nitrite concentration are in
NapA has a considerably lower turnover when compared to NrfA [4], so general very low. A small increase is observed between 10 and 15 h of
the cells may need to produce the protein continuously to preserve a cell growth for which, thus far, we have not found an explanation.
high nitrate reduction rate. On the other hand, the data showed that Thereafter, during the stationary phase of cell growth, nitrite con-
nrfA was highly expressed around 6 h, during the exponential growth centration stabilizes at a residual value of ca. 35 μM. According to these
phase, and significantly down-regulated at later stages of cell devel- results, the comparatively higher transcript levels of nrfA (at 16 h) in
opment (Fig. 6). nrfA appears to be induced later, possibly after initial the cells from sulfate medium can be explained by the accumulation of

1465
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

Fig. 8. Variation of nitrite concentrations in the culture media during D. desulfuricans


cellular growth. Nitrate (A) and sulfate (B) based media. Error bars represent the standard
Fig. 7. In gel nitrite reductase activities of protein extracts from D. desulfuricans. A) Gels
deviation of mean values of at least three independent measurements.
loaded cells grown for 16 h in medium containing A1) nitrate, A2) sulfate and A3) pur-
ified NrfA. B) Gels loaded with total protein extracts of cells grown in medium containing
nitrate harvested at different time points: B1) 3 h, B2) 6 h; B3) 12 h and B4) purified NrfA. based medium (Fig. 5). This data is in agreement with a previous work
Native polyacrylamide gels (7.5%) were loaded with 16.8 μg of each protein extract and also demonstrating that genes involved in the sulfate reduction
10 μg of NrfA. Gels were incubated for 15 min in staining solution containing 0.6 mM MV,
pathway are constitutively expressed when the cells grow in the pre-
6 mM sodium dithionite and 20 mM sodium nitrite in 100 mM phosphate buffer pH 7.6.
sence of nitrate [11].

nitrite during later stages of cell growth. In addition, the produced NrfA
may contribute to increased sulfite reduction rates [23,24]. Interest- 4. Conclusions
ingly, the variation of nitrite concentration during D. desulfuricans
growth in nitrate medium is non-progressive, i.e., there is an initial As published earlier by others [11], our work clearly shows that D.
peak of nitrite concentration around 3 h (during the lag phase) and desulfuricans is able to grow in nitrate medium with growth rates and
after that nitrite levels decrease to residual values (25 μM) until 10 h of yields higher than those obtained in the presence of its natural electron
growth, when a new progressive increase is detected, similar to the one acceptor, sulfate. To activate the alternative dissimilatory nitrate re-
observed in the sulfate containing medium, that stabilizes around duction pathway, the genes encoding the terminal reductases NapA and
250 μM, after 16 h. These results suggest that in the early stages of NrfA are up-regulated at key points of the cell growth. According to
cellular growth D. desulfuricans must sequentially produce high levels of comparative measurements of the transcripts levels, the nitrate re-
NapA to metabolize the excess of nitrate the cells must deal with upon ductase gene napA, is overexpressed in a nitrate rich environment since
inoculation of the nitrate containing medium. Subsequently, the gen- the lag phase, with no significant variations over the growth curve. In
erated nitrite induces nrfA expression, which leads to a fast decrease in what concerns the nitrite reductase NrfA, its gene expression is induced
nitrite concentration. Nitrite quantification results are also in agree- when considerable amounts of nitrite are already produced (3 h). Ap-
ment with in-gel activity assays, which show a lower abundance of NrfA parently, this enzyme is sufficiently stable and catalytically active over
in the cells harvested at 3 h of growth, when the outburst of nitrite the remaining growth stages, so its expression is silenced thereafter.
concentration is observed. Then, the nrfA gene is up-regulated so that The differentiated set of reactions underlying nitrate reduction (ni-
nitrite may be reduced. This is consistent with the cell requirements at trate → nitrite → ammonium) is thermodynamically more favorable
the exponential phase, when energy demands are higher due to the [11], allowing D. desulfuricans to grow faster. However, the differential
rapid cell division rate. analysis (nitrate vs sulfate) of the total proteomes in the pI range 4–7,
demonstrates that the bacterium's adaptation to the alternative electron
3.5. The terminal enzymes of sulfate reduction acceptor (nitrate) implies the increased production of a large number
(ca. 80) of proteins that are exclusively detected in this condition. It is
Unlike the enzymes involved in nitrate reduction, the transcript possible that these proteins are below the detection limits of the ex-
levels of APS reductase and sulfite reductase genes do not change sig- perimental technique for the samples obtained from sulfate medium,
nificantly between the cell extracts obtained from nitrate and sulfate but they are clearly less significant for D. desulfuricans growth with

1466
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

sulfate as terminal electron acceptor. Overall, these increased proteins nitrate-to-ammonium reducers, Front. Microbiol. 6 (2015) 1–12, http://dx.doi.org/
10.3389/fmicb.2015.01124.
are involved in non-energetic metabolisms, with a special prominence [14] T. Dalsgaard, F. Bak, Nitrate reduction in a sulfate-reducing bacterium, Desulfovibrio
of those related with ammonium assimilation, such as amino acid desulfuricans, isolated from rice paddy soil: sulfide inhibition, kinetics, and reg-
biosynthesis. Besides contributing to the higher bacterial growth rates, ulation, Appl. Environ. Microbiol. 60 (1994) 291–297.
[15] G.J. Mitchell, J.G. Jones, J.A. Cole, Distribution and regulation of nitrate and nitrite
this helps to detoxify the elevated levels of ammonium derived from reduction by Desulfovibrio and Desulfotomaculum species, Arch. Microbiol. 144
nitrate reduction. (1986) 35–40, http://dx.doi.org/10.1007/BF00454953.
Therefore, we propose that in order to adapt to the nitrate medium, [16] H.H. Seitz, H. Cypionka, Chemolithotrophic growth of Desulfovibrio desulfuricans
with hydrogen coupled to ammonification of nitrate or nitrite, Arch. Microbiol. 146
D. desulfuricans activates additional pathways involving amino acid (1986) 63–67, http://dx.doi.org/10.1007/BF00690160.
and/or protein biosynthesis, whereas in the presence of sulfate the [17] P.C. Mills, G. Rowley, S. Spiro, J.C.D. Hinton, D.J. Richardson, A combination of
bacterium expresses predominantly the proteins required for its con- cytochrome c nitrite reductase (NrfA) and flavorubredoxin (NorV) protects
Salmonella enterica serovar Typhimurium against killing by NO in anoxic environ-
stitutive energy metabolism. Such drastic changes in the cellular com-
ments, Microbiology 154 (2008) 1218–1228, http://dx.doi.org/10.1099/mic.0.
position indicate that the metabolism shifting from sulfate to nitrate 2007/014290-0.
respiration has high biosynthetic costs that may well justify the pre- [18] S.R. Poock, E.R. Leach, J.W.B. Moir, J.A. Cole, D.J. Richardson, Respiratory de-
ference of this SRB for sulfate, when in the presence of both electron toxification of nitric oxide by the cytochrome c nitrite reductase of Escherichia coli,
JBC 277 (2002) 23664–23669, http://dx.doi.org/10.1074/jbc.M200731200.
acceptor substrates [11]. [19] S.A. Haveman, E.A. Greene, G. Voordouw, Gene expression analysis of the me-
chanism of inhibition of Desulfovibrio vulgaris Hildenborough by nitrate-reducing,
Transparency document sulfide-oxidizing bacteria, Environ. Microbiol. 7 (2005) 1461–1465, http://dx.doi.
org/10.1111/j.1462-2920.2005.00834.x.
[20] E.A. Greene, C. Hubert, M. Nemati, G.E. Jenneman, G. Voordouw, Nitrite reductase
The http://dx.doi.org/10.1016/j.bbapap.2017.07.021 associated activity of sulphate-reducing bacteria prevents their inhibition by nitrate-reducing,
with this article can be found, in online version. sulphide- oxidizing bacteria, Environ. Microbiol. 5 (2003) 607–617, http://dx.doi.
org/10.1046/j.1462-2920.2003.00446.x.
[21] A.C. Pereira, J. Legall, A.V. Xavier, M. Teixeira, Characterization of a heme c nitrite
Acknowledgements reductase from a non-ammonifying microorganism, Desulfovibrio vulgaris
Hildenborough, Biochim. Biophys. Acta 1481 (2000) 119–130, http://dx.doi.org/
10.1016/S0167-4838(00)00111-4.
The authors acknowledge funding from UCIBIO@REQUIMTE Pest-
[22] S.A. Haveman, E.A. Greene, C.P. Stilwell, J.K. Voordouw, G. Voordouw,
C/EQB/LA0006/2013. The authors thank Ana Teresa Lopes for assis- Physiological and gene expression analysis of inhibition of Desulfovibrio vulgaris
tance with cell cultures. CMS thanks the financial support from Hildenborough by nitrite, J. Bacteriol. 186 (2004) 7944–7950, http://dx.doi.org/
10.1128/JB.186.23.7944.
Fundação para a Ciência e Tecnologia (Fellowship SFRH/BPD/79566/
[23] P. Lukat, M. Rudolf, P. Stach, A. Messerschmidt, P.M.H. Kroneck, J. Simon,
2011). O. Einsle, Binding and reduction of sulfite by cytochrome c nitrite reductase,
Biochemistry 47 (2008) 2080–2086, http://dx.doi.org/10.1021/bi7021415.
Appendix A. Supplementary data [24] I.C. Pereira, I.A. Abreu, A.V. Xavier, J. Legall, M. Teixeira, Nitrite reductase from
Desulfovibrio desulfuricans (ATCC 27774)— a heterooligomer heme protein with
sulfite reductase activity, Biochem. Biophys. Res. Commun. 224 (1996) 611–618,
Supplementary data to this article can be found online at http://dx. http://dx.doi.org/10.1006/bbrc.1996.1074.
doi.org/10.1016/j.bbapap.2017.07.021. [25] J.F. Sparacino-Watkins, Courtney Stolz, P. Basu, Nitrate and periplasmic nitrate
reductases, Chem. Soc. Rev. 43 (2014) 676–706, http://dx.doi.org/10.1039/
c3cs60249d.
References [26] N. Dyballa, S. Metzger, Fast and sensitive colloidal coomassie G-250 staining for
proteins in polyacrylamide gels, J. Vis. Exp. (2009) 2–5, http://dx.doi.org/10.
3791/1431.
[1] G. Muyzer, A.J.M. Stams, The ecology and biotechnology of sulphate-reducing
[27] D.J.D. Nicholas, A. Nason, Determination of nitrate and nitrite, Methods Enzymol.
bacteria, Nat. Rev. Microbiol. 6 (2008) 441–454, http://dx.doi.org/10.1038/
188 (1957) 981–984, http://dx.doi.org/10.1016/S0076-6879(57)03489-8.
nrmicro1892.
[28] K.J. Livak, T.D. Schmittgen, Analysis of relative gene expression data using real-
[2] H. Cypionka, Oxigen respiration by Desulfovibrio species, Annu. Rev. Microbiol. 54
time quantitative PCR and the 2− ΔΔCT method, Methods 25 (2001) 402–408,
(2000) 827–848, http://dx.doi.org/10.1146/annurev.micro.54.1.827.
http://dx.doi.org/10.1006/meth.2001.1262.
[3] I. a C. Pereira, A.R. Ramos, F. Grein, M.C. Marques, S. Marques, D.A. Bryant,
[29] T. Rabilloud, M. Chevallet, S. Luche, C. Lelong, Fully denaturing two-dimensional
T. Pennsylvania, S.M. da Silva, S.S. Venceslau, A comparative genomic analysis of
electrophoresis of membrane proteins: a critical update, Proteomics 8 (2008)
energy metabolism in sulfate reducing bacteria and archaea, Front. Microbiol. 2
3965–3973, http://dx.doi.org/10.1002/pmic.200800043.
(2011) 1–22, http://dx.doi.org/10.3389/fmicb.2011.00069.
[30] P.S. Dudhe, D.M.M. Kshirsagar, A.S. Yerlekar, A review on 2D gel electrophoresis: a
[4] I. Moura, S. Bursakov, C. Costa, J.J.G. Moura, Nitrate and nitrite utilization in
protein identification technique, Int. J. Comput. Sci. Inf. Technol. 5 (2014)
sulfate- reducing bacteria, Anaerobe 3 (1997) 279–290, http://dx.doi.org/10.1006/
856–862.
anae.1997.0093.
[31] C.M. Silveira, Development of Electrochemical Nitrite Biosensors Using Cytochrome
[5] J.R. Postgate, L.L. Campbell, Classification of Desulfovibrio species, the non-
c Nitrite Reductase From Desulfovibrio Desulfuricans ATCC 27774, RUN,
sporulating sulfate-reducing bacteria, Bacteriol. Rev. 30 (1966) 732–738.
Repositório da Universidade Nova de Lisboa, 2011 Retrieved from. (Identifier:
[6] A. López-Cortés, M. Fardeau, G. Fauque, C. Joulian, B. Ollivier, Reclassification of
http://hdl.handle.net/10362/6811 ) (PhD dissertation).
the sulfate- and nitrate-reducing bacterium Desulfovibrio vulgaris subsp. oxamicus as
[32] P.M. Pereira, Q. He, A.V. Xavier, J. Zhou, I.A.C. Pereira, R.O. Louro, Transcriptional
Desulfovibrio oxamicus sp. nov., comb. nov. Int. J. Syst. Evol. Microbiol. 56 (2006)
response of Desulfovibrio vulgaris Hildenborough to oxidative stress mimicking en-
1495–1499, http://dx.doi.org/10.1099/ijs.0.64074-0.
vironmental conditions, Arch. Microbiol. 189 (2008) 451–461, http://dx.doi.org/
[7] P. Cabello, M.D. Roldán, C. Moreno-Vivián, C. Moreno-vivia, M.D. Rolda,
10.1007/s00203-007-0335-5.
C. Moreno-vivia, Nitrate reduction and the nitrogen cycle in archaea, Microbiology
[33] J.D. Peterson, L.A. Umayam, T. Dickinson, E.K. Hickey, O. White, The compre-
150 (2004) 3527–3546, http://dx.doi.org/10.1099/mic.0.27303-0.
hensive microbial resource, Nucleic Acids Res. 29 (2001) 123–125.
[8] W.E.W. Roediger, J. Moore, W. Babidge, Colonic sulfite in pathogenesis and
[34] J.F. Heidelberg, R. Seshadri, S.A. Haveman, C.L. Hemme, I.T. Paulsen, J.F. Kolonay,
treatment of ulcerative colitis, Dig. Dis. Sci. 42 (1997) 1571–1579, http://dx.doi.
J.A. Eisen, N. Ward, B. Methe, L.M. Brinkac, S.C. Daugherty, R.T. Deboy,
org/10.1023/A:1018851723920.
R.J. Dodson, A.S. Durkin, R. Madupu, W.C. Nelson, S.A. Sullivan, D. Fouts,
[9] W.A.W. Hamilton, Bioenergetics of sulphate-reducing bacteria in relation to their
D.H. Haft, J. Selengut, J.D. Peterson, T.M. Davidsen, N. Zafar, L. Zhou, D. Radune,
environmental impact, Biodegradation 9 (1998) 201–212, http://dx.doi.org/10.
G. Dimitrov, M. Hance, K. Tran, H. Khouri, J. Gill, T.R. Utterback, T.V. Feldblyum,
1023/A:1008362304234.
J.D. Wall, G. Voordouw, C.M. Fraser, The genome sequence of the anaerobic, sul-
[10] D.A. Rodionov, I. Dubchak, A. Arkin, E. Alm, M.S. Gelfand, Reconstruction of
fate-reducing bacterium Desulfovibrio vulgaris Hildenborough, Nat. Biotechnol. 22
regulatory and metabolic pathways in metal-reducing δ-proteobacteria, Genome
(2004) 554–559, http://dx.doi.org/10.1038/nbt959.
Biol. 5 (2004) 1–27, http://dx.doi.org/10.1186/gb-2004-5-11-r90.
[35] P. Donoghue, L. Staunton, E. Mullen, G. Manning, K. Ohlendieck, DIGE analysis of
[11] A. Marietou, L. Griffiths, J. Cole, Preferential reduction of the thermodynamically
rat skeletal muscle proteins using nonionic detergent phase extraction of young
less favorable electron acceptor, sulfate, by a nitrate-reducing strain of the sulfate-
adult versus aged gastrocnemius tissue, J. Proteome 73 (2010) 1441–1453, http://
reducing bacterium Desulfovibrio desulfuricans 27774, J. Bacteriol. 191 (2009)
dx.doi.org/10.1016/j.jprot.2010.01.014.
882–889, http://dx.doi.org/10.1128/JB.01171-08.
[36] A. Görg, W. Weiss, M.J. Dunn, Current two-dimensional electrophoresis technology
[12] S.A.L. Lobo, A.M.P. Melo, J.N. Carita, M. Teixeira, L.M. Saraiva, The anaerobe
for proteomics, Proteomics 4 (2004) 3665–3685, http://dx.doi.org/10.1002/pmic.
Desulfovibrio desulfuricans ATCC 27774 grows at nearly atmospheric oxygen levels,
200401031.
FEBS Lett. 581 (2007) 433–436, http://dx.doi.org/10.1016/j.febslet.2006.12.053.
[37] C. Tristan, N. Shahani, T.W. Sedlak, A. Sawa, The diverse functions of GAPDH:
[13] H. Decleyre, K. Heylen, C. Van Colen, A. Willems, Dissimilatory nitrogen reduction
views from different subcellular compartments, Cell. Signal. 23 (2011) 317–323,
in intertidal sediments of a temperate estuary: small scale heterogeneity and novel

1467
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

http://dx.doi.org/10.1016/j.cellsig.2010.08.003. enzymes, Eur. Mol. Biol. Organ. 4 (2003) 850–854, http://dx.doi.org/10.1038/sj.


[38] G. Schenk, R. Layfield, J.M. Candy, R.G. Duggleby, P.F. Nixon, Molecular evolu- embor.embor914.
tionary analysis of the thiamine-diphosphate-dependent enzyme, transketolase, J. [64] M. Strohmeier, T. Raschle, J. Mazurkiewicz, K. Rippe, I. Sinning, T.B. Fitzpatrick,
Mol. Evol. 44 (1997) 552–572, http://dx.doi.org/10.1007/PL00006179. I. Tews, Structure of a bacterial pyridoxal 5′-phosphate synthase complex, PNAS
[39] T.P. Begley, A. Chateerjee, J.W. Hanes, A. Hazra, S.E. Ealick, Cofactor biosynthesis 103 (2006) 19284–19289 (doi:10.1073_pnas.0604950103).
— still yielding fascinating new biological chemistry, Curr. Opin. Chem. Biol. 12 [65] S.J. Sasindran, S. Saikolappan, V.L. Scofield, S. Dhandayuthapani, Biochemical and
(2008) 118–125, http://dx.doi.org/10.1016/j.cbpa.2008.02.006. physiological characterization of the GTP-binding protein Obg of Mycobacterium
[40] S. El Qaidi, J. Yang, J. Zhang, D.W. Metzger, G. Bai, The vitamin B6 biosynthesis tuberculosis, BMC Microbiol. 11 (2011) 1–11, http://dx.doi.org/10.1186/1471-
pathway in Streptococcus pneumoniae is controlled by pyridoxal 5′ -phosphate and 2180-11-43.
the transcription factor PdxR and has an impact on ear infection, J. Bacteriol. 195 [66] A. Kolesnikov, A. Gudkov, Elongation factor G with effector loop from elongation
(2013) 2187–2196, http://dx.doi.org/10.1128/JB.00041-13. factor Tu is inactive in translocation, FEBS 514 (2002) 67–69, http://dx.doi.org/10.
[41] L. Pieulle, V. Magro, E.C. Hatchikian, Isolation and analysis of the gene encoding 1016/S0014-5793(02)02300-1.
the pyruvate-ferredoxin oxidoreductase of Desulfovibrio africanus, production of the [67] A. Savelsbergh, M.V. Rodnina, W. Wintermeyer, Distinct functions of elongation
recombinant enzyme in Escherichia coli, and effect of carboxy-terminal deletions on factor G in ribosome recycling and translocation, RNA 15 (2009) 772–780, http://
its stability, J. Bacteriol. 179 (1997) 5684–5692, http://dx.doi.org/10.1128/jb. dx.doi.org/10.1261/rna.1592509.sequently.
179.18.5684-5692.1997. [68] A.L. Brioukhanov, A.I. Netrusov, R.I.L. Eggen, The catalase and superoxide dis-
[42] W. Buckel, R.K. Thauer, Energy conservation via electron bifurcating ferredoxin mutase genes are transcriptionally up-regulated upon oxidative stress in the strictly
reduction and proton/Na translocating ferredoxin oxidation, Biochim. Biophys. anaerobic archaeon Methanosarcina barkeri, Microbiology 152 (2006) 1671–1677,
Acta 1827 (2013) 94–113, http://dx.doi.org/10.1016/j.bbabio.2012.07.002. http://dx.doi.org/10.1099/mic.0.28542-0.
[43] W. Zhang, D.E. Culley, J.C.M. Scholten, M. Hogan, L. Vitiritti, F.J. Brockman, [69] Z.A. Wood, E. Schroder, J.R. Harris, L.B. Poole, Structure, mechanism and regula-
Global transcriptomic analysis of Desulfovibrio vulgaris on different electron donors, tion of peroxiredoxins, Trends Biochem. Sci. 28 (2003) 32–40.
Antonie Van Leeuwenhoek 89 (2006) 221–237, http://dx.doi.org/10.1007/s10482- [70] M. Trujillo, G. Ferrer-sueta, R. Radi, Kinetic studies on peroxynitrite reduction by
005-9024-z. peroxiredoxins, in: R.B. Thompson, C.A. Fierke (Eds.), Methods Enzymol, Elsevier
[44] L.L. Barton, G.D. Fauque, Biochemistry, physiology and biotecnology of sulfate- Masson SAS, 2008, pp. 173–196, , http://dx.doi.org/10.1016/S0076-6879(08)
reducing bacteria, Adv. Appl. Microbiol. (2009) 41–98, http://dx.doi.org/10.1016/ 01210-X.
S0065-2164(09)01202-7. [71] M. Fournier, C. Aubert, Z. Dermoun, M. Durand, D. Moinier, A. Dolla, Response of
[45] R. Rabus, T. Hansen, F. Widdel, Dissimilatory sulfate-and sulfur-reducing prokar- the anaerobe Desulfovibrio vulgaris Hildenborough to oxidative conditions: proteome
yotes, The Prokaryotes, 2006, pp. 659–768, , http://dx.doi.org/10.1007/0-387- and transcript analysis, Biochimie 88 (2006) 85–94, http://dx.doi.org/10.1016/j.
30742-7_22. biochi.2005.06.012.
[46] R.R. Nair, C.M. Silveira, M.S. Diniz, M.G. Almeida, J.J.G. Moura, M.G. Rivas, [72] M. Fournier, Y. Zhang, J.D. Wildschut, A. Dolla, J.K. Voordouw, D.C. Schriemer,
Changes in metabolic pathways of Desulfovibrio alaskensis G20 cells induced by G. Voordouw, Function of oxygen resistance proteins in the anaerobic, sulfate-re-
molybdate excess, J. Biol. Inorg. Chem. 20 (2014) 311–322, http://dx.doi.org/10. ducing bacterium Desulfovibrio vulgaris Hildenborough, J. Bacteriol. 185 (2003)
1007/s00775-014-1224-4. 71–79, http://dx.doi.org/10.1128/JB.185.1.71.
[47] M. Hodges, V. Flesch, S. Gálvez, E. Bismuth, Higher plant NADP+ − dependent [73] A. Dolla, M. Fournier, Z. Dermoun, Oxygen defense in sulfate-reducing bacteria, J.
isocitrate dehydrogenases, ammonium assimilation and NADPH production, Plant Biotechnol. 126 (2006) 87–100, http://dx.doi.org/10.1016/j.jbiotec.2006.03.041.
Physiol. Biochem. 41 (2003) 577–585, http://dx.doi.org/10.1016/S0981-9428(03) [74] H. Cypionka, F. Widdel, N. Pfennig, Survival of sulfate-reducing bacteria after
00062-7. oxygen stress, and growth in sulfate-free oxygen-sulfide gradients, FEMS Microbiol.
[48] K.L. Keller, B.J. Rapp-giles, E.S. Semkiw, I. Porat, S.D. Brown, J.D. Wall, New model Ecol. 31 (1985) 39–45, http://dx.doi.org/10.1111/j.1574-6968.1985.tb01129.x.
for electron flow for sulfate reduction in Desulfovibrio alaskensis G20, Appl. Environ. [75] S.G. Rhee, H.A. Woo, I.S. Kil, S.H. Bae, Peroxiredoxin functions as a peroxidase and
Microbiol. 80 (2014) 855–868, http://dx.doi.org/10.1128/AEM.02963-13. a regulator and sensor of local peroxides, J. Biol. Chem. 287 (2012) 4403–4410,
[49] T. Zaunmuller, D.J. Kelly, F.O. Glockner, G. Unden, Succinate dehydrogenase http://dx.doi.org/10.1074/jbc.R111.283432.
functioning by a reverse redox loop mechanism and fumarate reductase in sulphate- [76] W.G. Dos Santos, I. Pacheco, M.-Y. Liu, M. Teixeira, A.V. Xavier, J. LeGall,
reducing bacteria, Microbiology 152 (2006) 2443–2453, http://dx.doi.org/10. Purification and characterization of an iron superoxide dismutase and a catalase
1099/mic.0.28849-0. from the sulfate-reducing bacterium Desulfovibrio gigas, J. Bacteriol. 182 (2000)
[50] J.L. Steger, C. Vincent, J.D. Ballard, L.R. Krumholz, Desulfovibrio sp. genes involved 796–804, http://dx.doi.org/10.1128/JB.182.3.796-804.2000.
in the respiration of sulfate during metabolism of hydrogen and lactate, Appl. [77] P. Fareleira, B.S. Santos, C. António, P. Moradas-ferreira, J. Legall, A.V. Xavier,
Environ. Microbiol. 68 (2002) 1932–1937, http://dx.doi.org/10.1128/AEM.68.4. H. Santos, Response of a strict anaerobe to oxygen: survival strategies in
1932. Desulfovibrio gigas, Microbiology 149 (2003) 1513–1522, http://dx.doi.org/10.
[51] J.M. Odom, H.D. Peck, Hydrogen cycling as a general mechanism for energy cou- 1099/mic.0.26155-0.
pling in the sulfate-reducing bacteria, Desulfovibrio sp, FEMS Microbiol. Lett. 12 [78] D. Aragão, S. Macedo, E.P. Mitchell, C.V. Romão, M.Y. Liu, C. Frazão, L.M. Saraiva,
(1981) 47–50, http://dx.doi.org/10.1111/j.1574-6968.1981.tb07609.x. A.V. Xavier, J. Legall, W.M.A.M. van Dongen, W.R. Hangen, M. Teixeira,
[52] S.K. Spaans, R.A. Weusthuis, J. Van Der Oost, S.W.M. Kengen, NADPH-generating M.A. Carrondo, P. Lindley, Reduced hybrid cluster proteins (HCP) from
systems in bacteria and archaea, Front. Microbiol. 6 (2015) 1–27, http://dx.doi. Desulfovibrio desulfuricans ATCC 27774 and Desulfovibrio vulgaris (Hildenborough):
org/10.3389/fmicb.2015.00742. X-ray structures at high resolution using synchrotron radiation, J. Biol. Inorg.
[53] T. Fuhrer, U. Sauer, Different biochemical mechanisms ensure network-wide bal- Chem. 8 (2003) 540–548, http://dx.doi.org/10.1007/s00775-003-0443-x.
ancing of reducing equivalents in microbial metabolism, J. Bacteriol. 191 (2009) [79] O. Einsle, A. Messerschmidt, R. Huber, P.M.H. Kroneck, F. Neese, Mechanism of the
2112–2121, http://dx.doi.org/10.1128/JB.01523-08. six-electron reduction of nitrite to ammonia by cytochrome c nitrite reductase, J.
[54] G. Brown, A. Singer, M. Proudfoot, T. Skarina, Y. Kim, C. Chang, I. Dementieva, Am. Chem. Soc. 124 (2002) 11737–11745, http://dx.doi.org/10.1021/ja0206487.
E. Kuznetsova, C.F. Gonzalez, A. Joachimiak, A. Savchenko, A.F. Yakunin, [80] M. Kern, J. Volz, J. Simon, The oxidative and nitrosative stress defence network of
Functional and structural characterization of four glutaminases from Escherichia coli Wolinella succinogenes: cytochrome c nitrite reductase mediates the stress response
and Bacillus subtilis, Biochemistry 47 (2008) 5724–5735, http://dx.doi.org/10. to nitrite, nitric oxide, hydroxylamine and hydrogen peroxide, Environ. Microbiol.
1021/bi800097h. 13 (2011) 2478–2494, http://dx.doi.org/10.1111/j.1462-2920.2011.02520.x.
[55] J.G. Cory, A.H. Cory, Critical roles of glutamine as nitrogen donors in purine and [81] C.C. Almeida, C.V. Roma, P.F. Lindley, M. Teixeira, L.M. Saraiva, The role of the
pyrimidine nucleotide synthesis: asparaginase treatment in childhood acute lym- hybrid cluster protein in oxidative stress defense, J. Biol. Chem. 281 (2006)
phoblastic leukemia, In Vivo 20 (2006) 587–589. 32445–32450, http://dx.doi.org/10.1074/jbc.M605888200.
[56] N.P. Curthoys, M. Watford, Regulation of glutaminase activity and glutamine me- [82] N.A. Ryabova, V.V. Marchenkov, S.Y. Marchenkova, N.V. Kotova, G.V. Semisotnov,
tabolism, Annu. Rev. Nutr. 15 (1995) 133–159, http://dx.doi.org/10.1146/ Molecular chaperone GroEL/ES: unfolding and refolding processes, Biochemistry 78
annurev.nu.15.070195.001025. (2013) 1405–1414, http://dx.doi.org/10.1134/S0006297913130038.
[57] C.-D. Lu, Pathways and regulation of bacterial arginine metabolism and perspec- [83] N. Noinaj, M. Guillier, T.J. Barnard, S.K. Buchanan, TonB-dependent transporters:
tives for obtaining arginine overproducing strains, Appl. Microbiol. Biotechnol. 70 regulation, structure, and function, Annu. Rev. Microbiol. 64 (2010) 43–60, http://
(2006) 261–272, http://dx.doi.org/10.1007/s00253-005-0308-z. dx.doi.org/10.1146/annurev.micro.112408.134247.
[58] G.J. Germano, K.E. Anderson, Serine biosynthesis in Desulfovibrio desulfuricans, J. [84] A.D. Ferguson, J. Deisenhofer, TonB-dependent receptors — structural perspectives,
Bacteriol. 99 (1969) 893–894. Biochim. Biophys. Acta 1565 (2002) 318–332, http://dx.doi.org/10.1016/S0005-
[59] R.K. Kulis-horn, M. Persicke, J. Kalinowski, Histidine biosynthesis, its regulation 2736(02)00578-3.
and biotechnological application in Corynebacterium glutamicum, Microb. [85] Y. Zhang, D.A. Rodionov, M.S. Gelfand, V.N. Gladyshev, Comparative genomic
Biotechnol. 7 (2014) 5–25, http://dx.doi.org/10.1111/1751-7915.12055. analyses of nickel, cobalt and vitamin B12 utilization, BMC Genomics 10 (2009)
[60] Y. Zhang, M. Morar, S.E. Ealick, Structural biology of the purine biosynthetic 1–26, http://dx.doi.org/10.1186/1471-2164-10-78.
pathway, Cell. Mol. Life Sci. 65 (2008) 3699–3724, http://dx.doi.org/10.1007/ [86] T. Eitinger, M.-A. Mandrand-Berthelot, Nickel transport systems in microorganisms,
s00018-008-8295-8. Arch. Microbiol. 173 (2000) 1–9, http://dx.doi.org/10.1007/s002030050001.
[61] K. Rébora, B. Laloo, B. Daignan-fornier, Revisiting purine-histidine cross-pathway [87] L. Rajeev, A. Chen, A.E. Kazakov, E.G. Luning, G.M. Zane, P.S. Novichkov,
regulation in Saccharomyces cerevisiae: a central role for a small molecule, Genet. J.D. Wall, A. Mukhopadhyay, Regulation of nitrite stress response in Desulfovibrio
Soc. Am. 170 (2005) 61–70, http://dx.doi.org/10.1534/genetics.104.039396. vulgaris Hildenborough, a model sulfate-reducing bacterium, J. Bacteriol. 197
[62] S. Mooney, H. Hellmann, Vitamin B6: killing two birds with one stone, (2015) 3400–3408, http://dx.doi.org/10.1128/JB.00319-15.
Phytochemistry 71 (2010) 495–501, http://dx.doi.org/10.1016/j.phytochem.2009. [88] M.E. Clark, Z. He, A.M. Redding, M.P. Joachimiak, J.D. Keasling, J.Z. Zhou,
12.015. A.P. Arkin, A. Mukhopadhyay, M.W. Fields, Transcriptomic and proteomic analyses
[63] R. Percudani, A. Peracchi, A genomic overview of pyridoxal-phosphatedependent of Desulfovibrio vulgaris biofilms: carbon and energy flow contribute to the distinct

1468
J.R. Sousa et al. BBA - Proteins and Proteomics 1865 (2017) 1455–1469

biofilm growth state, BMC Genomics 13 (2012) 1–17, http://dx.doi.org/10.1186/ S.L. Gomes, A comprehensive genomic, transcriptomic and proteomic analysis of a
1471-2164-13-138. hyperosmotic stress sensitive α-proteobacterium, BMC Microbiol. 15 (2015) 1–15,
[89] A.C. Darby, A.C. Gill, S.D. Armstrong, C.S. Hartley, D. Xia, J.M. Wastling, http://dx.doi.org/10.1186/s12866-015-0404-x.
B.L. Makepeace, Integrated transcriptomic and proteomic analysis of the global [91] T. Maier, M. Güell, L. Serrano, Correlation of mRNA and protein in complex bio-
response of Wolbachia to doxycycline-induced stress, ISME 8 (2014) 925–937, logical samples, FEBS Lett. 583 (2009) 3966–3973, http://dx.doi.org/10.1016/j.
http://dx.doi.org/10.1038/ismej.2013.192. febslet.2009.10.036.
[90] C. Kohler, R.F. Lourenço, J. Bernhardt, D. Albrecht, J. Schüler, M. Hecker,

1469

You might also like