Skip to main content

ORIGINAL RESEARCH article

Front. Chem., 13 April 2021
Sec. Theoretical and Computational Chemistry
Volume 9 - 2021 | https://doi.org/10.3389/fchem.2021.663706

The Ion-Translocating NrfD-Like Subunit of Energy-Transducing Membrane Complexes

Filipa Calisto1,2 Manuela M. Pereira1,2*
  • 1Instituto de Tecnologia Química e Biológica-António Xavier, Universidade Nova de Lisboa, Oeiras, Portugal
  • 2BioISI–Biosystems & Integrative Sciences Institute, Faculdade de Ciências, Universdade de Lisboa, Lisboa, Portugal

Several energy-transducing microbial enzymes have their peripheral subunits connected to the membrane through an integral membrane protein, that interacts with quinones but does not have redox cofactors, the so-called NrfD-like subunit. The periplasmic nitrite reductase (NrfABCD) was the first complex recognized to have a membrane subunit with these characteristics and consequently provided the family's name: NrfD. Sequence analyses indicate that NrfD homologs are present in many diverse enzymes, such as polysulfide reductase (PsrABC), respiratory alternative complex III (ACIII), dimethyl sulfoxide (DMSO) reductase (DmsABC), tetrathionate reductase (TtrABC), sulfur reductase complex (SreABC), sulfite dehydrogenase (SoeABC), quinone reductase complex (QrcABCD), nine-heme cytochrome complex (NhcABCD), group-2 [NiFe] hydrogenase (Hyd-2), dissimilatory sulfite-reductase complex (DsrMKJOP), arsenate reductase (ArrC) and multiheme cytochrome c sulfite reductase (MccACD). The molecular structure of ACIII subunit C (ActC) and Psr subunit C (PsrC), NrfD-like subunits, revealed the existence of ion-conducting pathways. We performed thorough primary structural analyses and built structural models of the NrfD-like subunits. We observed that all these subunits are constituted by two structural repeats composed of four-helix bundles, possibly harboring ion-conducting pathways and containing a quinone/quinol binding site. NrfD-like subunits may be the ion-pumping module of several enzymes. Our data impact on the discussion of functional implications of the NrfD-like subunit-containing complexes, namely in their ability to transduce energy.

Introduction

All living organisms need energy to fuel life processes. External energy sources, light or chemical compounds, are converted to biologically usable forms of energy, such as adenosine triphosphate (ATP) or electrochemical gradients. According to Peter Mitchell's chemiosmotic hypothesis, the transmembrane difference of the electrochemical potential (Δμ~) can be established by energy-transducing membrane protein complexes that couple the energy released by light or chemical reactions (Gibbs energy change, ΔG) to the translocation of charges (electrons or ions) across the membrane (Mitchell, 1961). The energy stored in the form of the electrochemical potential can drive different energy-requiring reactions of the cells, such as synthesis of cellular components, solute transport or motility.

Energy transducing membrane complexes are usually composed of catalytic subunits and transmembrane proteins that perform translocation of charges, electrons or cations, across the membrane (Calisto et al., 2021). The most common membrane charge-translocating subunits so far observed in energy transduction complexes are the di-heme cytochrome b-like subunits and the so called NrfD-like subunits. The di-heme cytochrome b-like subunits are involved in the transport of electrons, whereas the NrfD-like subunits translocate ions. In this way this type of subunits is devoid of redox cofactors but contain ion-conducting pathways.

The NrfD-like subunits are present in many and diverse membrane complexes, widespread in Bacteria and Archaea, that can take part in oxygen, nitrogen, sulfur, arsenate or hydrogen metabolism (Figure 1, Table 1) (Rothery et al., 2008; Refojo et al., 2010, 2019; Marreiros et al., 2016). These subunits thus compose the NrfD family, which was named after the characterization of the periplasmic nitrite reductase (NrfABCD) complex, the first complex recognized to have a NrfD-like subunit (Simon, 2002; Rothery et al., 2008) (Figure 1, NrfABCD). Structures of representatives of the NrfD family were first obtained for the PsrC subunit from the Thermus thermophilus polysulfide reductase (PsrABC) complex (Figures 1, 2, PsrABC) and later for the ActC and ActF subunits from the respiratory alternative complex III (ACIII) from Rhodothermus marinus (Figure 2, ACIII) and Flavobacterium johnsoniae (Sousa et al., 2018; Sun et al., 2018) and for the photosynthetic ACIII from Roseiflexus castenholzii (Shi et al., 2020) (Figure 1, ACIII). The structural data showed these membrane subunits have 8 common transmembrane helixes, organized in two four-helix bundles (TMHs 1–4 and TMHs 5–8) related by a 180° rotation around an axis perpendicular to the membrane (Sousa et al., 2018) (Figure 3), and contain one quinone/quinol-binding site close to the periphery of the membrane at the side at which the peripheral subunits are bound to. ActF does not contain any quinone/quinol binding site and it is the only NrfD-like subunit present in a complex in which another subunit of this type (which contains a quinone/quinol-binding site) is present. In addition, for all these subunits, putative ion-conducting pathways were proposed (Calisto et al., 2021). NrfD-like proteins also function as the link of the peripheral subunits to the membrane (Figure 1).

FIGURE 1
www.frontiersin.org

Figure 1. Composition and diversity of NrfD-like subunit-containing complexes. Schematic representation of periplasmic nitrite reductase (NrfABCD), polysulfide reductase (PsrABC), respiratory alternative complex III (ACIII), DMSO reductase (DmsABC), group-2 [Ni-Fe] hydrogenase (Hyd-2), arsenate reductase (ArrABC), tetrathionate reductase (TtrABC), quinone reductase complex (QrcABCD), sulfur reductase complex (SreABC), nine-heme cytochrome complex (NhcABCD), dissimilatory sulfite-reductase complex (DsrMKJOP), multiheme cytochrome c sulfite reductase (MccACD), formate:quinone oxidoreductase (FqoABC), sulfite dehydrogenase (SoeABC). P-side, positive-side of the membrane; N-side, negative-side of the membrane. The NrfD-like subunits (NrfD, PsrC, ActC, ActF, DmsC, HybB, ArrC, ttrC, QrcD, SreC, NhcC, DsrP, MccD, fqoC, and SoeC) are colored, the two-color tones indicate the two structural repeats, each composed of a four-helix bundle (TMHs 1–4 and TMHs 5–8) and related by a 180° rotation around an axis perpendicular to the membrane. The circle, inside the colored subunits, indicates the presence of a quinone/quinol-binding site. The red arrows represent the pathway for electron transfer. The color code is used in all the following figures and in Tables 1, 2.

TABLE 1
www.frontiersin.org

Table 1. NrfD-like subunit-containing complexes.

FIGURE 2
www.frontiersin.org

Figure 2. Structures of NrfD-like subunit-containing complexes. Overall structure of PsrABC from Thermus thermophilus (left, PDB: 2VPZ) and Alternative Complex III from Rhodothermus marinus (right, PDB: 6FOK, ActH was not included for clarity). PsrC (yellow), ActC (pink) and ActF (blue) are NrfD-like subunits. The two complexes are oriented in relation to each other by the superimposition of the PsrC and ActC subunits. Iron-sulfur clusters represented by orange/yellow spheres; Cytochrome c-type hemes represented by red sticks; Mo-bisPGD cofactor represented by blue/red sticks with molybdenum shown as a blue sphere. Orange dots stand for the quinone-binding site. P-side, positive-side of the membrane; N-side, negative-side of the membrane.

FIGURE 3
www.frontiersin.org

Figure 3. Fold of homologous membrane subunits ActC and ActF. ActC (pink) and ActF (blue) seen from the membrane (left) and from the top (right). The N- and C-terminal helices (TMHs A-B), which form a helix dimer, are colored in gray. The four-helix bundles are shown in light pink/blue (TMHs 1–4) and dark pink/blue (TMHs 5–8).

Besides Escherichia coli NrfABCD (Hussain et al., 1994; Clarke et al., 2007), ACIII and PsrABC complexes, the enzymes containing a NrfD-like subunits are dimethyl sulfoxide (DMSO) reductase (DmsABC), tetrathionate reductase (TtrABC), sulfur reductase complex (SreABC), sulfite dehydrogenase (SoeABC) and arsenate reductase (ArrABC) complexes, members of the so called complex iron-sulfur molybdoenzyme (CISM) family, which have a similar subunit composition to that of PsrABC complex (Figure 1) (Rothery and Weiner, 1993; Weiner et al., 1993; Berks, 1996; Hensel et al., 1999; Hinsley and Berks, 2002; Laska et al., 2003; Geijer and Weiner, 2004; Guiral et al., 2005; Duval et al., 2008; Dahl et al., 2013; Grimaldi et al., 2013; Tang et al., 2013; Steinmetz et al., 2014; Kurth et al., 2015; Boughanemi et al., 2020). The subunits and prosthetic groups of these enzymes are specified in Table 1. DmsABC is a quinol:DMSO oxidoreductase with its peripheral DmsAB subunits facing the P-side of the membrane (Rothery and Weiner, 1993; Berks, 1996). TtrABC, expressed under anaerobic conditions in Proteobacteria (Marreiros et al., 2016), catalyzes the endergonic reduction of tetrathionate to thiosulfate with the concomitant oxidation of quinol to quinone. As DmsAB, TtrAB subunits are facing the P-side of the membrane (Hensel et al., 1999; Hinsley and Berks, 2002). SreABC, characterized in Acidianus ambivalens, was proposed to have quinol:sulfur oxidoreductase activity and SreAB were suggested to be located at the P-side of the membrane (Laska et al., 2003). SoeABC is a sulfite:quinone oxidoreductase complex described in Aquifex aeolicus and Allochromatium vinosum (Dahl et al., 2013; Boughanemi et al., 2020). SoeAB subunits were hypothesized to be located at the negative-side (N-side) of the membrane. ArrABC, involved in arsenate respiration, seems to have the peripheral ArrAB subunits at the P-side of the membrane (Stolz et al., 2006; Duval et al., 2008).

TABLE 2
www.frontiersin.org

Table 2. Amino acid residues conservation in NrfD-like subunit-containing complexes.

The multiheme cytochrome c sulfite reductase complex (MccACD), biochemically characterized from W. succinogenes and Shewanella oneidensis (Kern et al., 2011; Shirodkar et al., 2011) (Figure 1, MccACD, Table 1), the quinone reductase complex (QrcABCD), homologous to ACIII and present in sulfate-reducing bacteria, are other enzymes with NrfD-like subunits. Qrc catalyzes the transfer of electrons from a type-1 tetraheme cytochrome c3 (TpIc3) to quinone (Figure 1, Qrc) and so the peripheral subunits are facing the P-side of the membrane (Pereira et al., 2011; Venceslau et al., 2011). Group-2 [NiFe] hydrogenase (Hyd-2), produced under anaerobic conditions, is a hydrogen:quinone oxidoreductase complex with the peripheral subunits facing the P-side of the membrane (Laurinavichene et al., 2002; Sargent, 2016; Beaton et al., 2018) (Figure 1, Hyd-2, Table 1). The nine-heme cytochrome complex (NhcABCD), a TpIc3:quinone oxidoreductase, is a member of a family of transmembrane complexes with multiheme cytochrome c subunit, described in Desulfovibrio species (Matias et al., 1999, 2005; Saraiva et al., 2001; Bento et al., 2003; Pereira et al., 2011) (Figure 1, Nhc, Table 1). The dissimilatory sulfite-reductase complex (DsrMKJOP) is a transmembrane complex with peripheral subunits at both sides of the membrane (Grein et al., 2010). DsrJ and DsrO subunits are anchored to the P-side of the membrane (Pires et al., 2006). DsrM is a di-heme cytochrome b membrane subunit (Pires et al., 2006). DsrK is facing the N-side of the membrane (Mander et al., 2002; Pires et al., 2006). DsrK is able to reduce the small protein DsrC, while DsrJ was suggested to interact with thiosulfate (Bamford et al., 2002; Denkmann et al., 2012; Venceslau et al., 2014) (Figure 1, DsrMKJOP, Table 1). Since DsrP and DsrM are membrane subunits with hypothetical quinone-binding sites, a quinone cycling of reduction and oxidation between DsrMK and DsrJOP was proposed to take place (Grein et al., 2013).

In this work we thoroughly analyze the primary structures of the members of the NrfD family and predicted the respective tertiary structures, the data provided allowed a deep and broad discussion on the presence of ion-conducting pathways and quinone/quinol-binding sites, which impacts on the function of the several complexes, namely in their ability to transduce energy.

Results and Discussion

Taxonomic Distribution

NrfD-like subunits are the anchor protein of several modular complexes, which take part in a vast array of metabolisms, such as the oxygen, nitrogen, sulfur, arsenate and hydrogen cycles (Rothery et al., 2008; Refojo et al., 2010, 2019; Marreiros et al., 2016) (Figure 1). In this work, we performed sequence alignments and taxonomic profiling to investigate the distribution of the NrfD-like subunit-containing complexes in microbial species. We gathered 4,545 NrfD-like amino acid sequences present in the genomes of 1,822 distinct species (96% Bacteria, 4% Archaea), with an average of 2.5 NrfD-like subunits per organism. In order to reduce the size of the obtained dataset, the 4,545 amino acid sequences were clustered according to their identity (50% identity) and 551 representative sequences were aligned, allowing to generate the correspondent Neighbor-Joining (NJ) dendrogram. From the obtained dendrogram we were able to identify several branches belonging to different groups of NrfD-like subunit-containing complexes (Figure 4).

FIGURE 4
www.frontiersin.org

Figure 4. Neighbor-Joining dendrogram of the NrfD-like subunits. The dendrogram was obtained using 551 representative sequences of NrfD-like subunit sequences. Branches are colored according to the color code indicated in the Figure 1. The external ring of the dendrogram indicates the four major identified groups discussed in the text (groups A to D).

We observed that close to the root of the dendrogram, the sequences of NrfD-like subunits separated into four major groups (Figure 4, dendrogram groups A–D). The amino acid sequences of SreC (21 sequences) (Figure 4, group A) and ArrC (9 sequences) (Figure 4, group B) seem to be less related with the others NrfD-like subunits and constitute two distinct groups. The SreABC complex was biochemically characterized from the sulfur-dependent archaeon A. ambivalens (Laska et al., 2003) and genes coding for SreC subunit were only identified in archaea, in 21 species (27% of Crenarchaeota species) (Figure 5, SreC). ArrAB complex was described as a periplasmic complex, which is associated with the transmembrane ArrC subunit only in few microorganisms (Duval et al., 2008). In agreement, we identified nine genes coding for ArrC subunit, distributed in Gammaproteobacteria (four sequences), Betaproteobacteria (four sequences) and Euryarchaeota (one sequence) species (Figure 5, ArrC).

FIGURE 5
www.frontiersin.org

Figure 5. Distribution of NrfD-like subunits in Prokaryotes. The diagram indicates the percentage of the species within each phylum containing at least one gene encoding a NrfD-like subunit. The color code is indicated in the figure.

The third group (Figure 4, group C) englobes the amino acid sequences of ActC (521 sequences), ActF (521 sequences), DsrP (88 sequences) and DmsC (42 sequences) subunits. Intriguingly, we identified two branches of DsrP amino acid sequences, which we called DsrP1 and DsrP2 (Figure 4). The amino acid sequence of DsrP subunit, biochemically characterized from Desulfovibrio vulgaris DsrMKJOP complex, was found in the DsrP1 branch (66 amino acid sequences of sulfate-reducing bacteria). Besides their presence in Deltaproteobacteria (46 species, 48%), genes coding for DsrP1 subunit were identified in species from Thermodesulfobacteria (6 species, 86%), Chlorobi (6 species, 35%) and Firmicutes-Clostridia (6 species, 3%) phyla (Figure 5, DsrP1). Genes coding for DsrP2 subunit were observed in 36 species of sulfur-oxidizing bacteria and from Archaeoglobus genus (23 species, 3% of Gammaproteobacteria; 10 species, 1% of Alphaproteobacteria; 3 species of 1% of Euryarchaeota species) (Figure 5, DsrP2). The amino acid sequences of ActC and ActF subunits (521 sequences) were naturally separated into two branches (Figure 4, ActC and ActF). Genes coding for ActC and ActF subunits were identified in species from Gammaproteobacteria-Others (11 species, 1%), Betaproteobacteria (10 species, 2%), Deltaproteobacteria (35 species, 37%), Alphaproteobacteria (46 species, 6%), Oligoflexia (nine species, 75%), Verrucomicrobia (10 species, 67%), Spirochaetes (14 species, 16%), Acidobacteria (12 species, 86%), Gemmatimonadetes (three species, 100%), Planctomycetes (24 species, 75%), Bacteroidetes (290 species, 76%), Chlorobi (two species, 12%), Chloroflexi (six species, 16%), Deinococcus-Thermus (11 species, 33%) and Terrabacteria (three species, 100%) (Figure 5, ActC/F). Group C of the dendrogram also includes 42 amino acid sequences identified as DmsC subunit (Figure 4, group DmsC), namely the DmsC amino acid sequence from Halobacterium sp. strain NRC-1 (Müller and DasSarma, 2005). Genes coding for DmsC present in group C are all observed in 24 archaeal species from Euryarchaeota (10%) (Figure 5, DmsC).

The fourth group (Figure 4, group D) is composed of amino acid sequences of PsrC (48 sequences), DmsC (793 sequences), NrfD (411 sequences), TtrC (292 sequences), NhcC (81 sequences), HybB (320 sequences), QrcD (35 sequences), SoeC (239 sequences) and MccD (79 sequences) subunits. As observed before (Duarte et al., 2018), the amino acid sequences of PsrC subunit forms two separate branches, one represented by PsrC from W. succinogenes and the other pictured by PsrC from T. thermophilus. The amino acid sequence of PsrC subunit from T. thermophilus PsrABC complex, clusters within branch PsrC T with other 34 amino acid sequences (Figure 4, PsrC T) and PsrC from W. succinogenes PsrABC complex is present in branch PsrC W (Figure 4, PsrC W; 33 sequences). Genes coding for PsrC T subunits were found in 34 species (36%) from Deltaproteobacteria and one species (7%) from Acidobacteria, while genes coding for PsrC W subunits were observed in 26 species (13%) from Epsilonproteobacteria (Figure 5, PsrC T and PsrC W).

Genes coding for bacterial DmsC and NrfD subunits were identified in 80% (444 species) and 61% (383 species) of species from Gammaproteobacteria, respectively (Figure 5, DmsC and NrfD). Amino acid sequences of TtrC subunits were expected to be produced in 291 species, distributed in Proteobacteria (219 species, 35% Gammaproteobacteria; 47 species, 10% Betaproteobacteria; eight species, 4% Epsilonproteobacteria; two species, 2% Deltaproteobacteria; 12 species, 2% Alphaproteobacteria), Aquificae (three species, 21%) and Armatimonadetes (one species, 50%) phyla (Figure 5, TtrC). NhcC subunit coding genes were found in 64 bacteria species (36 from Deltaproteobacteria, 38%; 16 from Firmicutes-Clostridia, 9%; two from Acidobacteria, 14%; 2 from Bacteroidetes; one from Chlorobi, 6%; 1 from Nitrospirae, 11%; three from Thermodesulfobacteria, 43%) (Figure 5, NhcC). Genes coding for HybB subunit were observed in 314 species, 277 from Gammaproteobacteria (50%), 7 from Betaproteobacteria (2%), 13 from Deltaproteobacteria (4%), 11 from Alphaproteobacteria (2%), 2 from Acidobacteria (14%), 1 from Aquificae (7%) and three from Deferribacteres (60%) (Figure 5, HybB). The presence of genes coding for QrcD subunit was restricted to 35 species (36%) from Deltaproteobacteria (Figure 5, QrcD). Genes coding for SoeC subunit were present in 236 species from Proteobacteria (45 species from Gammaproteobacteria, 5%; 83 species from Betaproteobacteria, 18%; 100 species from Alphaproteobacteria, 14%), Firmicutes-Clostridia (three species, 1%) and Aquificae (five species, 36%) phyla (Figure 5, SoeC). While genes coding for MccD subunit were identified in 51 species (35 from Gammaproteobacteria, 4%; 15 from Epsilonproteobacteria, 7%) (Figure 5, MccD).

In group D of the dendrogram (Figure 4, FqoC), we were able to identify 110 amino acid sequences present in an uncharacterized complex, that we tentatively assigned as formate:quinone oxidoreductase complex (Figure 1, FqoABC). Our assignment is based on the observations that the complex is possibly composed of three subunits: two peripheral FqoAB subunits homologous to FdnGH subunits from E. coli formate dehydrogenase complex (Fdn-N) and a FqoC subunit, a NrfD-like subunit with eight predicted TMHs. Genes coding for FqoC were identified in 110 species present in three different dendrogram groups: FqoC1 includes genes coding for FqoC subunit from Actinobacteria (59 species, 7%), Chlorobi (15 species, 88%) and Crenarchaeota (9 species, 12%) phyla; FqoC2 gathers the genes coding for FqoC from Chloroflexi phylum (15 species, 48%); and FqoC3 includes genes coding for FqoC subunit from Deltaproteobacteria phylum (12 species, 13%) (Figure 5, FqoC).

Homology Models

The NrfD-like subunits are transmembrane proteins, with 8 to 10 TMHs, that interact with quinones but do not have redox cofactors. The structures of T. thermophilus PsrC and R. marinus ActC and ActF revealed details of the overall architecture of these NrfD-like subunits: the common 8 TMHs are organized in two four-helix bundles (TMHs 1–4 and TMHs 5–8), which form two structural repeats related by a 180° rotation around an axis perpendicular to the membrane layer plane (Jormakka et al., 2008; Sousa et al., 2018). Two additional TMHs (TMHs A-B), present in ActC and ActF, cross each other at an angle of ~45° at the periphery of each subunit (Sousa et al., 2018) (Figure 3). ActC and ActF subunits have their N- and C-terminal located at N-side of the membrane, while PsrC have both termini located at the P-side of the membrane. T. thermophilus PsrC and R. marinus ActC subunits contain one quinol-binding site at the P-side of the membrane of the first four-helix bundle (Jormakka et al., 2008; Sousa et al., 2018) (Figure 2).

We performed structural homology models of the NrfD-like subunits from W. succinogenes PsrABC, E. coli NrfABCD, E. coli DmsABC, E. coli Hyd-2, Salmonella enterica TtrABC, A. ambivalens SreABC, A. aeolicus SoeABC, D. vulgaris QrcABCD, D. vulgaris DsrMKJOP1, A. vinosum DsrMKJOP2, Desulfovibrio desulfuricans NhcABCD, Alkalilimnicola ehrlichii ArrABC, W. succinogenes MccACD and Chlorobaculum tepidum FqoABC complexes. The homology models were calculated in Phyre2 without imposing any template. In all cases the randomly selected template was the structure from R. marinus ActC subunit. The homology models have confidence scores higher than 90%. The final models presented the common 8 TMHs organized in a similar arrangement as those of ActC, ActF and PsrC subunits. As in the case of T. thermophilus PsrC subunit, we predicted only 8 TMHs for W. succinogenes PsrC, E. coli DmsC, E. coli NrfD, A. aeolicus SoeC, W. succinogenes MccD and C. tepidum FqoC subunits. S. enterica TtrC subunit was predicted to have 9 TMHs, with the N-terminal at the P-side and the C-terminal at the N-side of the membrane, the extra TMH is equivalent to TMH B from ActC. The other NrfD-like subunits (E. coli HybB, A. ehrlichii ArrC, D. vulgaris DsrP1, A. vinosum DsrP2, D. vulgaris QrcD, A. ambivalens SreC and D. desulfuricans NhcC) have 10 predicted TMHs, the two extra TMHs are equivalent to TMH A and B of ActC and ActF subunits. The models predict for all NrfD-like subunits the presence of the structural repeats, composed of the two four-helix bundles, harboring putative ion-conducting pathways and the existence of a quinone/quinol-binding site (Figure 1, Table 2).

Quinone-Binding Site

The crystal structures of T. thermophilus PsrC co-crystallized with either menaquinone-7 or ubiquinone-1 showed that the quinone-binding site is located in the first four-helix bundle of PsrC (TMHs 1-4) on the P-side of the membrane, in close proximity to the [4Fe-4S]2+/1+ cluster of PsrB subunit (Jormakka et al., 2008). Highly conserved amino acid charged residues, present in this region, were suggested as essential for quinone-binding and coordination: His21PsrCT, Asn18PsrCT and Tyr130PsrCT in T. thermophilus PsrC and His139ActC, Asp169ActC and Asp253ActC in R. marinus ActC (Jormakka et al., 2008; Sousa et al., 2018) (Figure 6, ActC and PsrC T).

FIGURE 6
www.frontiersin.org

Figure 6. Quinone-binding site. Zoomed views of structural models of NrfD-like subunits showing the respective putative quinone/quinol-binding sites, located close to the positive-side (P-side) of the membrane. The SoeC subunit is expected to be in an inverted orientation toward the membrane in relation to the other NrfD-like subunits and thus its quinone/quinol binding site is located close to the negative-side (N-side). The amino acid residues composing the different quinone/quinol-binding sites are indicated and depicted as sticks. Please note that all models are oriented with the P-side of the membrane at the top of each panel, except for SoeC that is oriented with the N-side of the membrane at the top of the respective panel.

To further investigate structurally relevant elements and/or amino acid residues in the quinone-binding site, we analyzed the obtained structural homology models of the NrfD-like subunits and performed sequence alignment to identify the conserved residues involved in the binding and coordination of quinone/quinol molecule.

We identified for all NrfD-like subunits, with the exception of ActF subunit, the only NrfD-like that has been shown not to interact with quinones/quinols and part of a complex in which another NrfD-like subunit is present, the presence of quinone/quinol-binding site in the same spatial position of those observed for T. thermophilus PsrC and R. marinus ActC. These are all located in the first four-helix bundle (TMHs 1–4), close to the P-side of the membrane. In the case of SoeC the quinone/quinol-binding site is also present in the first four-helix bundle (TMHs 1–4) at the same special position of sites present in the other NrfD-like subunits, however it is expecting to be facing the N-site of membrane. This is because the catalytic subunit of SoeABC complex is predicted to be oriented toward the N-side of the membrane (Dahl et al., 2013; Boughanemi et al., 2020) (Figure 1) and thus SoeC would be expected to have an inverted orientation inside the membrane plane comparing to the other NrfD-like proteins.

We observed, within the predicted quinone/quinol-binding site of all NrfD-like subunits analyzed, the presence of 5, on average, amino acid residues (histidine, arginine, aspartate, glutamate, serine, tyrosine, threonine and asparagine) that may be involved in quinone/quinol-binding and stabilization in each protein (Figure 6, Table 2).

Although histidine, serine, glutamate, aspartate and tyrosine residues were described as being involved in the coordination of quinones in several quinone-interacting complexes, there seems to be no common pattern for quinone-binding (Abramson et al., 2000; Fisher and Rich, 2000; Zhang et al., 2002; Iwaki et al., 2003; Horsefield et al., 2006; Kleinschroth et al., 2008). Nevertheless, we noticed a conserved serine residue in all NrfD-like subunits, close to the entry of the quinone/quinol pocket (Ser164ActC, Ser87PsrC_T, Ser94PsrC_W, Ser129HybB, Ser82DmsC, Ser126ArrC, Ser124DsrP1, Ser128DsrP2, Ser138QrcD, Ser128SreC, Ser92NrfD, Ser93TtrC, Ser137NhcC, Ser82SoeC, Ser93MccD and Ser81FqoC). We hypothesized that this serine residue is involved in quinone coordination, since it is only absent in R. marinus ActF subunit, the only NrfD-like that has been shown not to interact with quinone (Sousa et al., 2018). Mutation of Ser94PsrC_W residue in W. succinogenes PsrC caused partially inhibition of polysulfide respiration (Dietrich and Klimmek, 2002). We also identified in TMH 2, in the same spatial position in all NrfD-like subunits (again with the exception of ActF), the presence of a glutamate/aspartate/histidine/serine residue (Glu67PsrC_T, Glu107ArrC, Asp76PsrC_W, Asp109HybB, Asp106DsrP1, Asp110DsrP2, Asp120QrcD, Asp75TtrC, Asp119NhcC, Asp75MccD, Asp65FqoC, His139ActC, His65DmsC, His65SoeC, His74NrfD and Ser110SreC) that seems to be required for protonation/deprotonation of the quinone/quinol molecule (Table 2). Mutational studies support the relevance of this position. Replacement of Asp76PsrC_W and His65DmsC residues, in W. succinogenes PsrC and E. coli DmsC, respectively, blocked quinol oxidation (Rothery and Weiner, 1996; Zhao and Weiner, 1998; Geijer and Weiner, 2004) and lead to full inhibition of polysulfide respiration (Dietrich and Klimmek, 2002). In T. thermophilus PsrC structure, Tyr130PsrC_T and His21PsrC_T residues were proposed to act as ligands to menaquinone and pentachlorophenol, a quinone inhibitor (Jormakka et al., 2008), and, in fact, mutation of Tyr23PsrC_W residue, the correspondent residue in W. succinogenes PsrC, also inhibit polysulfide respiration (Dietrich and Klimmek, 2002). Additional mutational studies, conducted in E. coli DmsABC (Rothery and Weiner, 1996; Zhao and Weiner, 1998), W. succinogenes PsrABC (Dietrich and Klimmek, 2002) and E. coli Hyd-2 (Lubek et al., 2019), also showed the importance of Glu87DmsC, Tyr159PsrC_W, Thr160PsrC_W and Asp58HybB residues for quinone/quinol coordination (Rothery and Weiner, 1996; Zhao and Weiner, 1998; Dietrich and Klimmek, 2002; Geijer and Weiner, 2004; Lubek et al., 2019).

Our models indicate that all NrfD-like proteins (except for ActF) are able to interact with quinones/quinols (Figures 1, 6).

Ion Translocation Pathways

NrfD-like subunit-containing complexes were hypothesized to be capable of ion-translocation across the membrane (Calisto et al., 2021) and, in fact, proton-conducting pathways have been identified in the structures of the PsrC, ActC and ActF subunits (Jormakka et al., 2008; Sousa et al., 2018; Sun et al., 2018; Shi et al., 2020). Proton-conducting pathways are formed by amino acid residues with side chains that can establish hydrogen bonds, constituting a hydrogen bond network. This allows proton transfer by a Grotthuss-type mechanism, which involves successive breaking and concomitant formation of hydrogen bonds (de Grotthuss, 1806; Cukierman, 2006).

In R. marinus ActC and ActF subunits, the putative ion-conducting pathway was suggested to be formed by two half-channels: a N-side half-channel in the first four-helix bundle (TMH 1–4) and a P-side half-channel in the second four-helix bundle (TMH 5–8) (Sousa et al., 2018). ActC and ActF ion-conducting pathways are composed of conserved amino acid residues within all ACIII complexes (Sousa et al., 2018) (Figure 7, ActC; Table 2).

FIGURE 7
www.frontiersin.org

Figure 7. Ion-translocation pathway. Structural models of NrfD-like subunits with amino acid residues putatively involved in ion-translocation. The top left side panel contains a general schematic representation of NrfD-like subunits. Each protein is composed by eight common TMH that form two four-helix bundle (TMHs 1–4, light gray [lighter colors in the following panels) and TMHs 5–8, dark gray (darker colors in the following panels)] related by a 180° rotation around an axis perpendicular to the membrane. The light blue line schematizes the ion-conducting pathway and the red dot indicates the presence of the arginine residue (Arg395ActC), possibly acting as the gate of the pathway. The orange circle points to the quinone/quinol-binding site, which is in close proximity to the [4Fe-4S]+2/+1 cluster ([3Fe-4S]1+/0 in ActC) (cube) present at the peripheral subunit. The other panels contain the structural models obtained for the 8 common TMHs, the amino acid residues, putatively involved ion-pathways are indicated and depicted as sticks along the light blue line and are highlighted in black (conserved residues in respective sequence alignments), gray (amino acid residues present in the model able to conduct protons, but not conserved in respective sequence alignments) and red (equivalent to Arg395ActC). Please note that all models are oriented with the P-side of the membrane at the top of each panel, except for SoeC that is oriented with the N-side of the membrane at the top of the respective panel (see more in text and in legend of Figure 6).

We were able to identify, in all NrfD-like subunits, amino acid residues that may constitute a N-side half-channel in the first four-helix bundle (TMH 1–4) and a P-side half-channel in the second four-helix bundle (TMH 5–8). The amino acid residues that compose these putative ion-conducting pathways are conserved within each subunit: ActC, ActF, HybB, ArrC, DsrP, QrcD, SreC, NrfD and NhcC subunits (Figure 7, Table 2). However, in PsrC, DmsC, TtrC, SoeC, MccD and FqoC subunits, the two half-channels were not so easy to define, as we observed less conserved residues.

In T. thermophilus PsrC structure, the putative ion-conducting pathway was suggested to be formed by a N-side half-channel in the second four-helix bundle (TMH 5–8) and a P-side half-channel in the first four-helix bundle (TMH 1–4) (Jormakka et al., 2008). Of the proposed residues in the N-side half-channel at TMH 5-8 (Glu224PsrC_T, Thr220PsrC_T, Arg177PsrC_T) (Jormakka et al., 2008), only Arg177PsrC_T residue is conserved among other PsrC T subunits. Mutation of Asp218PsrC_W, the equivalent to Arg177PsrC_T in W. succinogenes PsrC, and Ser192PsrC_W residues, both located at TMH 5-8 in the N-side half-channel, resulted in inhibition or reduction of polysulfide respiration, respectively (Jormakka et al., 2008). Nevertheless, we identified conserved amino acid residues (Figure 7, PsrC T; Table 2) that could form two half-channels, resembling those identified in ActC and ActF subunits. Our identification is supported by substitution studies of amino acid residues located at TMH 1–4 on our proposed N-side half-channel (Tyr106PsrC_W and Glu146PsrC_W) and located at TMH 5-8 on the P-side half-channel (Glu225PsrC_W, Ser185PsrC_W, Ser188PsrC_W and Tyr310PsrC_W) of W. succinogenes PsrC model, resulted in strains with a compromised polysulfide respiration (Dietrich and Klimmek, 2002). These data suggest those residues may be important for proton translocation, since PsrABC catalyzes an endergonic reaction dependent of electrochemical potential (Dietrich and Klimmek, 2002; Calisto et al., 2021).

The presence of an ion-conducting pathway in HybB is also supported by the observations that replacement of Arg89HybB, Tyr99HybB, Glu148HybB and His184HybB in E. coli HybB (N-side half-channel TMH 5–8) significantly decreased hydrogen oxidation (Lubek et al., 2019). Addition of a protonophore increases hydrogen oxidation in these mutated strains (Lubek et al., 2019), showing that ion translocation and catalytic activity are coupled.

Previously for QrcD, only the N-side half-channel, present in the first four-helix bundle (TMH 1–4) as observed for ActC and ActF subunits, was proposed to be present (Duarte et al., 2018). This N-side half-channel was hypothesized to translocate protons from the N-side of the membrane to the quinone-binding site, for quinone reduction (Duarte et al., 2018). However, we identified conserved amino acid residues that could be part of a P-side half-channel in the second four-helix bundle (TMH 5–8) (Figure 7, QrcD; Table 2).

Although the ion-conducting pathways present in the different NrfD-like subunits are not composed by the same amino acid residues, they are localized at the same spatial position. Noticeably, we identified a conserved arginine residue (Arg395ActC, Arg239PsrC_T, Arg305PsrC_W, Arg329HybB, Arg271DmsC, Arg324ArrC, Arg358QrcD, Arg306NrfD, Arg280TtrC, Arg345NhcC, Arg297SoeC, Arg306MccD, Arg293FqoC) located in middle of the membrane (in TMH 8) in a position that coincides with that at which the two proton half-channels converge (Figure 7). In Desulfovibrio vulgaris DsrP this is occupied by a lysine (Lys329DsrP1). We hypothesized that the residue at this position may perform a gate keeping role for proton translocation across the membrane. In fact, water molecules around Arg239PsrC_T were observed in the structure of PsrC from T. thermophilus (Jormakka et al., 2008) and mutation of Arg305PsrC_W resulted in inhibition of polysulfide respiration (Dietrich and Klimmek, 2002). Although, we hypothesized a relevant function for this arginine (or Lys329DsrP1) residue, it is not conserved in ActF, DsrP2 and SreC. In these subunits the gating role may be performed by a conserved serine residue (Ser238ActF, Ser204DsrP2 and Ser235SreC), which is located in middle of the membrane in TMH 5, also coinciding with the convergence of the two half-channels (Figure 7).

The members of the NrfD family are transmembrane proteins (8 to 10 TMHs) characterized by the presence of structural repeats, composed of two four-helix bundles, harboring ion-translocation pathways and a quinone/quinol-binding site.

The quinone/quinol-binding site of NrfD-like subunits is located at the P-side of the membrane in the first four-helix bundle (TMHs 1–4), always in vicinity of the peripheral iron-sulfur subunit. The peripheral subunits of SoeABC complex are hypothesized to be located at the N-side (Dahl et al., 2013; Boughanemi et al., 2020) and thus SoeC would be expected to be in an inverted orientation in the membrane when comparing to the other NrfD-like proteins. In this way, its quinone-binding site would be present on the N-side of the membrane. We identified, in TMH 2 close to the entry of the quinone/quinol pocket, a serine residue (Ser164ActC) that appears to be important for interaction with quinone/quinol molecule, since this serine residue is strictly conserved in all NrfD-like subunits that interact with quinone/quinol.

Our structural models reinforce the possible existence of ion-translocation pathways in all NrfD-like subunits and the contribution of the NrfD-like subunit-containing complexes to energy transduction. NrfD-like subunit-containing complexes may perform energy transduction by an indirect-coupling mechanism and may generate or consume electrochemical potential (Calisto et al., 2021). W. succinogenes PsrABC activity was shown to be dependent on electrochemical potential (Dietrich and Klimmek, 2002), while QrcABCD activity was shown to be coupled to the formation of electrochemical potential (Duarte et al., 2018). The ion-translocation pathways, observed in NrfD-like subunits, are formed by amino acid residues that may establish hydrogen bonds, allowing proton translocation by a Grotthuss-type mechanism (de Grotthuss, 1806; Cukierman, 2006). A semiconserved arginine amino acid residue, located in the middle of the membrane at the intersection of the two ion half-channels, was here suggested to play an important role as gate keeper in ion-translocation. In ActF, DsrP2 and SreC the gating role may be performed by a conserved serine.

The data here presented indicate that NrfD-like subunits are possibly the ion translocating modules of different enzymes, involved in a vast array of metabolisms, such as the oxygen, nitrogen, sulfur, arsenate and hydrogen cycles. All NrfD-like subunits containing complexes may be thus energy transducing membrane machines that contribute to energy conservation in vast range of organisms and under multiple growth conditions.

Materials and Methods

The selection of the complete NrfD-like subunits dataset was performed using protein BLAST (pBLAST) analysis tool (default parameters) running at KEGG's (Kyoto Encyclopedia of Genes and Genomes) database, which only contains data on fully sequenced genomes (Ogata et al., 1999; Kanehisa et al., 2006, 2008). The information used in this study was that available by October 2020.

Protein search was performed using the amino acid sequences of ActC and ActF from R. marinus (rmr:rmar_0223, rmr:rmar_0226), PsrC from T. thermophilus (tth:tt_c0153) and W. succinogenes (wsu:ws0118), NrfD (eco:b4073), DmsC (eco:b0896), and HybB (eco:b2995) from E. coli, TtrC from S. enterica (sty:sty1737), SreC from A. ambivalens (aamb:d1866_07880), SoeC from A. aeolicus (aae:aq_1231), QrcD (dvu:dvu0692) and DsrP (dvu:dvu1286) from D. vulgaris, NhcC from D. desulfuricans (dds:ddes_2040), ArrC from A. ehrlichii (aeh:mlg_0214) and MccD from W. succinogenes (wsu:ws0382). Results with e-value <0.01 were accepted for further analysis (ca. 4,545 protein sequences).

The 4,545 amino acid sequences were clustered according to their identity using the CD-HIT tool (50% identity) (Huang et al., 2010). The resulting 551 representative sequences were aligned using PROMALS3D (standard parameters, using R. marinus ActC as structural template, PDB: 6F04) (Pei et al., 2008). The dendrogram was constructed using RAxML toll and the Neighbor-Joining (NJ) method at the CIPRES gateway portal (Miller et al., 2010). The obtained dendrogram was visualized and manipulated in Dendroscope (Huson et al., 2007).

To generate the structural models of NrfD-like subunits we used the amino acid sequence from W. succinogenes PsrC (wsu:ws0118), E. coli NrfD (eco:b4073), E. coli DmsC (eco:b0896), E. coli HybB (eco:b2995), S. enterica TtrC (sty:sty1737), A. ambivalens SreC (aamb:d1866_07880), A. aeolicus SoeC (aae:aq_1231), D. vulgaris QrcD (dvu:dvu0692), D. vulgaris DsrP1 (dvu:dvu1286), Allochromatium vinosum DsrP2 (alv:alvin_1262), D. desulfuricans NhcC (dds:ddes_2040), A. ehrlichii ArrC (aeh:mlg_0214), W. succinogenes MccD (wsu:ws0382) and C. tepidum FqoC (cte:ct0494) subunits. We developed the structural models using Phyre2 for predicting protein structure by homology modeling under intensive mode, and ActC R. marinus (PDB: 6F0K2) structure was selected as template (Kelley et al., 2015; Sousa et al., 2018). Protein structural model visualization and figure construction were performed using Chimera (Pettersen et al., 2004).

Quinone/quinol-binding sites and ion-translocation pathways were identified by sequence conservation analysis, after declustering the amino acid sequences from dendrogram branches, selected as containing ActC, ActF, PsrC, NrfD, DmsC, HybB, TtrC, SreC, SoeC, QrcD, DsrP, NhcC, ArrC, MccD and FqoC sequences. Sequence alignments were performed using PROMALS3D for each dendrogram group and visualized using JalView software (Waterhouse et al., 2009).

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors, without undue reservation.

Author Contributions

FC and MP conceived the study, analyzed the results, and wrote the manuscript. All authors contributed to the article and approved the submitted version.

Funding

The work was funded by Fundação para a Ciência e a Tecnologia (PTDC/BIA-BQM/28827/2017 and PTDC/BIA-BQM/30528/2017). The project was further supported by UIDB/04046/2020 and UIDP/04046/2020 Centre grants from FCT, Portugal (to BioISI), by LISBOA-01-0145-FEDER-007660 cofunded by FEDER through COMPETE2020-POCI and by Fundacão para a Ciência e a Tecnologia and by UIDB/04612/2020 and UIDP/04612/2020 research unit grants from FCT (to Mostmicro).

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

Abramson, J., Riistama, S., Larsson, G., Jasaitis, A., Svensson-Ek, M., Laakkonen, L., et al. (2000). The structure of the ubiquinol oxidase from Escherichia coli and its ubiquinone binding site. Nat. Struct. Biol. 7, 910–917. doi: 10.1038/82824

PubMed Abstract | CrossRef Full Text | Google Scholar

Bamford, V. A., Bruno, S., Rasmussen, T., Appia-Ayme, C., Cheesman, M. R., Berks, B. C., et al. (2002). Structural basis for the oxidation of thiosulfate by a sulfur cycle enzyme. EMBO J. 21, 5599–5610. doi: 10.1093/emboj/cdf566

PubMed Abstract | CrossRef Full Text | Google Scholar

Beaton, S. E., Evans, R. M., Finney, A. J., Lamont, C. M., Armstrong, F. A., Sargent, F., et al. (2018). The structure of hydrogenase-2 from Escherichia coli: implications for H2-driven proton pumping. Biochem. J. 475, 1353–1370. doi: 10.1042/BCJ20180053

PubMed Abstract | CrossRef Full Text | Google Scholar

Bento, I., Teixeira, V. H., Baptista, A. M., Soares, C. M., Matias, P. M., and Carrondo, M. A. (2003). Redox-Bohr and other cooperativity effects in the nine-heme cytochrome c from Desulfovibrio desulfuricans ATCC 27774: crystallographic and modeling studies. J. Biol. Chem. 278, 36455–36469. doi: 10.1074/jbc.M301745200

PubMed Abstract | CrossRef Full Text | Google Scholar

Berks, B. C. (1996). A common export pathway for proteins binding complex redox cofactors? Mol. Microbiol. 22, 393–404. doi: 10.1046/j.1365-2958.1996.00114.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Boughanemi, S., Infossi, P., Giudici-Orticoni, M.-T., Schoepp-Cothenet, B., and Guiral, M. (2020). Sulfite oxidation by the quinone-reducing molybdenum sulfite dehydrogenase SoeABC from the bacterium aquifex aeolicus. Biochim. Biophys. Acta Bioenerg. 1861:148279. doi: 10.1016/j.bbabio.2020.148279

PubMed Abstract | CrossRef Full Text | Google Scholar

Calisto, F., Sousa, F. M., Sena, F. V., Refojo, P. N., and Pereira, M. M. (2021). Mechanisms of energy transduction by charge translocating membrane proteins. Chem. Rev. 121, 1804–1844. doi: 10.1021/acs.chemrev.0c00830

PubMed Abstract | CrossRef Full Text | Google Scholar

Clarke, T. A., Cole, J. A., Richardson, D. J., and Hemmings, A. M. (2007). The crystal structure of the pentahaem c-type cytochrome NrfB and characterization of its solution-state interaction with the pentahaem nitrite reductase NrfA. Biochem. J. 406, 19–30. doi: 10.1042/BJ20070321

PubMed Abstract | CrossRef Full Text | Google Scholar

Cukierman, S. (2006). Et tu, grotthuss! and other unfinished stories. Biochim. Biophys. Acta Bioenerg. 1757, 876–885. doi: 10.1016/j.bbabio.2005.12.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Dahl, C., Franz, B., Hensen, D., Kesselheim, A., and Zigann, R. (2013). Sulfite oxidation in the purple sulfur bacterium allochromatium vinosum: identification of SoeABC as a major player and relevance of SoxYZ in the process. Microbiology 159, 2626–2638. doi: 10.1099/mic.0.071019-0

PubMed Abstract | CrossRef Full Text | Google Scholar

de Grotthuss, C. J. T. (1806). On the decomposition of water and the species held in solution by the aid of galvanic electricity. Annu. Chem. 58, 54–74.

Google Scholar

Denkmann, K., Grein, F., Zigann, R., Siemen, A., Bergmann, J., van Helmont, S., et al. (2012). Thiosulfate dehydrogenase: a widespread unusual acidophilic c-type cytochrome. Environ. Microbiol. 14, 2673–2688. doi: 10.1111/j.1462-2920.2012.02820.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Dietrich, W., and Klimmek, O. (2002). The function of methyl-menaquinone-6 and polysulfide reductase membrane anchor (PsrC) in polysulfide respiration of Wolinella succinogenes. Eur. J. Biochem. 269, 1086–1095. doi: 10.1046/j.0014-2956.2001.02662.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Duarte, A. G., Catarino, T., White, G. F., Lousa, D., Neukirchen, S., Soares, C. M., et al. (2018). An electrogenic redox loop in sulfate reduction reveals a likely widespread mechanism of energy conservation. Nat. Commun. 9:5448. doi: 10.1038/s41467-018-07839-x

CrossRef Full Text | Google Scholar

Duval, S., Ducluzeau, A.-L., Nitschke, W., and Schoepp-Cothenet, B. (2008). Enzyme phylogenies as markers for the oxidation state of the environment: the case of respiratory arsenate reductase and related enzymes. BMC Evol. Biol. 8:206. doi: 10.1186/1471-2148-8-206

PubMed Abstract | CrossRef Full Text | Google Scholar

Fisher, N., and Rich, P. R. (2000). A motif for quinone binding sites in respiratory and photosynthetic systems. J. Mol. Biol. 296, 1153–1162. doi: 10.1006/jmbi.2000.3509

PubMed Abstract | CrossRef Full Text | Google Scholar

Geijer, P., and Weiner, J. H. (2004). Glutamate 87 is important for menaquinol binding in DmsC of the DMSO reductase (DmsABC) from Escherichia coli. Biochim. Biophys. Acta Biomembr. 1660, 66–74. doi: 10.1016/j.bbamem.2003.10.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Grein, F., Pereira, I. A. C., and Dahl, C. (2010). Biochemical characterization of individual components of the allochromatium vinosum DsrMKJOP transmembrane complex aids understanding of complex function in vivo. J. Bacteriol. 192, 6369–6377. doi: 10.1128/JB.00849-10

PubMed Abstract | CrossRef Full Text | Google Scholar

Grein, F., Ramos, A. R., Venceslau, S. S., and Pereira, I. A. C. (2013). Unifying concepts in anaerobic respiration: insights from dissimilatory sulfur metabolism. Biochim. Biophys. Acta Bioenerg. 1827, 145–160. doi: 10.1016/j.bbabio.2012.09.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Grimaldi, S., Schoepp-Cothenet, B., Ceccaldi, P., Guigliarelli, B., and Magalon, A. (2013). The prokaryotic Mo/W-bisPGD enzymes family: a catalytic workhorse in bioenergetic. Biochim. Biophys. Acta Bioenerg. 1827, 1048–1085. doi: 10.1016/j.bbabio.2013.01.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Guiral, M., Tron, P., Aubert, C., Gloter, A., Iobbi-Nivol, C., and Giudici-Orticoni, M. T. (2005). A membrane-bound multienzyme, hydrogen-oxidizing, and sulfur-reducing complex from the hyperthermophilic bacterium Aquifex aeolicus. J. Biol. Chem. 280, 42004–42015. doi: 10.1074/jbc.M508034200

PubMed Abstract | CrossRef Full Text | Google Scholar

Hensel, M., Hinsley, A. P., Nikolaus, T., Sawers, G., and Berks, B. C. (1999). The genetic basis of tetrathionate respiration in Salmonella typhimurium. Mol. Microbiol. 32, 275–287. doi: 10.1046/j.1365-2958.1999.01345.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Hinsley, A. P., and Berks, B. C. (2002). Specificity of respiratory pathways involved in the reduction of sulfur compounds by Salmonella enterica. Microbiology 148, 3631–3638. doi: 10.1099/00221287-148-11-3631

PubMed Abstract | CrossRef Full Text | Google Scholar

Horsefield, R., Yankovskaya, V., Sexton, G., Whittingham, W., Shiomi, K., Omura, S., et al. (2006). Structural and computational analysis of the quinone-binding site of complex II (succinate-ubiquinone oxidoreductase). J. Biol. Chem. 281, 7309–7316. doi: 10.1074/jbc.M508173200

PubMed Abstract | CrossRef Full Text | Google Scholar

Huang, Y., Niu, B., Gao, Y., Fu, L., and Li, W. (2010). CD-HIT suite: a web server for clustering and comparing biological sequences. Bioinformatics 26, 680–682. doi: 10.1093/bioinformatics/btq003

PubMed Abstract | CrossRef Full Text | Google Scholar

Huson, D. H., Richter, D. C., Rausch, C., Dezulian, T., Franz, M., and Rupp, R. (2007). Dendroscope: an interactive viewer for large phylogenetic trees. BMC Bioinformatics 8:460. doi: 10.1186/1471-2105-8-460

PubMed Abstract | CrossRef Full Text | Google Scholar

Hussain, H., Grove, J., Griffiths, L., Busby, S., and Cole, J. (1994). A seven-gene operon essential for formate-dependent nitrite reduction to ammonia by enteric bacteria. Mol. Microbiol. 12, 153–163. doi: 10.1111/j.1365-2958.1994.tb01004.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Iwaki, M., Giotta, L., Akinsiku, A. O., Schägger, H., Fisher, N., Breton, J., et al. (2003). Redox-Induced Transitions in Bovine Cytochrome bc 1 Complex Studied by Perfusion-Induced ATR-FTIR Spectroscopy †. Biochemistry 42, 11109–11119. doi: 10.1021/bi0343020

CrossRef Full Text | Google Scholar

Jormakka, M., Yokoyama, K., Yano, T., Tamakoshi, M., Akimoto, S., Shimamura, T., et al. (2008). Molecular mechanism of energy conservation in polysulfide respiration. Nat. Struct. Mol. Biol. 15, 730–737. doi: 10.1038/nsmb.1434

PubMed Abstract | CrossRef Full Text | Google Scholar

Kanehisa, M., Araki, M., Goto, S., Hattori, M., Hirakawa, M., Itoh, M., et al. (2008). KEGG for linking genomes to life and the environment. Nucleic Acids Res. 36, D480–D484. doi: 10.1093/nar/gkm882

PubMed Abstract | CrossRef Full Text | Google Scholar

Kanehisa, M., Goto, S., Hattori, M., Aoki-Kinoshita, K. F., Itoh, M., Kawashima, S., et al. (2006). From genomics to chemical genomics: new developments in KEGG. Nucleic Acids Res. 34, D354–D357. doi: 10.1093/nar/gkj102

PubMed Abstract | CrossRef Full Text | Google Scholar

Kelley, L. A., Mezulis, S., Yates, C. M., Wass, M. N., and Sternberg, M. J. E. (2015). The Phyre2 web portal for protein modeling, prediction and analysis. Nat. Protoc. 10, 845–858. doi: 10.1038/nprot.2015.053

PubMed Abstract | CrossRef Full Text | Google Scholar

Kern, M., Klotz, M. G., and Simon, J. (2011). The Wolinella succinogenes mcc gene cluster encodes an unconventional respiratory sulphite reduction system. Mol. Microbiol. 82, 1515–1530. doi: 10.1111/j.1365-2958.2011.07906.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Kleinschroth, T., Anderka, O., Ritter, M., Stocker, A., Link, T. A., Ludwig, B., et al. (2008). Characterization of mutations in crucial residues around the Qo binding site of the cytochrome bc1 complex from Paracoccus denitrificans. FEBS J. 275, 4773–4785. doi: 10.1111/j.1742-4658.2008.06611.x

CrossRef Full Text | Google Scholar

Kurth, J. M., Dahl, C., and Butt, J. N. (2015). Catalytic protein film electrochemistry provides a direct measure of the tetrathionate/thiosulfate reduction potential. J. Am. Chem. Soc. 137, 13232–13235. doi: 10.1021/jacs.5b08291

PubMed Abstract | CrossRef Full Text | Google Scholar

Laska, S., Lottspeich, F., and Kletzin, A. (2003). Membrane-bound hydrogenase and sulfur reductase of the hyperthermophilic and acidophilic archaeon Acidianus ambivalens. Microbiology 149, 2357–2371. doi: 10.1099/mic.0.26455-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Laurinavichene, T. V., Zorin, N. A., and Tsygankov, A. A. (2002). Effect of redox potential on activity of hydrogenase 1 and hydrogenase 2 in Escherichia coli. Arch. Microbiol. 178, 437–442. doi: 10.1007/s00203-002-0471-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Lubek, D., Simon, A. H., and Pinske, C. (2019). Amino acid variants of the HybB membrane subunit of Escherichia coli [NiFe]-hydrogenase-2 support a role in proton transfer. FEBS Lett. 593, 2194–2203. doi: 10.1002/1873-3468.13514

CrossRef Full Text | Google Scholar

Mander, G. J., Duin, E. C., Linder, D., Stetter, K. O., and Hedderich, R. (2002). Purification and characterization of a membrane-bound enzyme complex from the sulfate-reducing archaeon Archaeoglobus fulgidus related to heterodisulfide reductase from methanogenic archaea. Eur. J. Biochem. 269, 1895–1904. doi: 10.1046/j.1432-1033.2002.02839.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Marreiros, B. C., Calisto, F., Castro, P. J., Duarte, A. M., Sena, F. V., Silva, A. F., et al. (2016). Exploring membrane respiratory chains. Biochim. Biophys. Acta Bioenerg. 1857, 1039–1067. doi: 10.1016/j.bbabio.2016.03.028

CrossRef Full Text | Google Scholar

Matias, P. M., Coelho, R., Pereira, I. A. C., Coelho, A. V., Thompson, A. W., Sieker, L. C., et al. (1999). The primary and three-dimensional structures of a nine-haem cytochrome c from Desulfovibrio desulfuricans ATCC 27774 reveal a new member of the Hmc family. Structure 7, 119–130. doi: 10.1016/S0969-2126(99)80019-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Matias, P. M., Pereira, I. A. C., Soares, C. M., and Carrondo, M. A. (2005). Sulphate respiration from hydrogen in Desulfovibrio bacteria: a structural biology overview. Prog. Biophys. Mol. Biol. 89, 292–329. doi: 10.1016/j.pbiomolbio.2004.11.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, M. A., Pfeiffer, W., and Schwartz, T. (2010). “Creating the CIPRES science gateway for inference of large phylogenetic trees,” in 2010 Gateway Computing Environments Workshop (GCE) (New Orleans, LA: IEEE), 1–8. doi: 10.1109/GCE.2010.5676129

CrossRef Full Text | Google Scholar

Mitchell, P. (1961). Coupling of phosphorylation to electron and hydrogen transfer by a chemi-osmotic type of mechanism. Nature 191, 144–148. doi: 10.1038/191144a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Müller, J. A., and DasSarma, S. (2005). Genomic analysis of anaerobic respiration in the archaeon Halobacterium sp. strain NRC-1: dimethyl sulfoxide and trimethylamine N-oxide as terminal electron acceptors. J. Bacteriol. 187, 1659–1667. doi: 10.1128/JB.187.5.1659-1667.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Ogata, H., Goto, S., Sato, K., Fujibuchi, W., Bono, H., and Kanehisa, M. (1999). KEGG: kyoto encyclopedia of genes and genomes. Nucleic Acids Res. 27, 29–34. doi: 10.1093/nar/27.1.29

PubMed Abstract | CrossRef Full Text | Google Scholar

Pei, J., Kim, B.-H., and Grishin, N. V. (2008). PROMALS3D: a tool for multiple protein sequence and structure alignments. Nucleic Acids Res. 36, 2295–2300. doi: 10.1093/nar/gkn072

PubMed Abstract | CrossRef Full Text | Google Scholar

Pereira, I. A. C., Ramos, A. R., Grein, F., Marques, M. C., Silva, S. M., and Venceslau, S. S. (2011). A comparative genomic analysis of energy metabolism in sulfate reducing bacteria and archaea. Front. Microbiol. 2:69. doi: 10.3389/fmicb.2011.00069

CrossRef Full Text | Google Scholar

Pettersen, E. F., Goddard, T. D., Huang, C. C., Couch, G. S., Greenblatt, D. M., Meng, E. C., et al. (2004). UCSF Chimera–a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612. doi: 10.1002/jcc.20084

PubMed Abstract | CrossRef Full Text | Google Scholar

Pires, R. H., Venceslau, S. S., Morais, F., Teixeira, M., Xavier, A. V., and Pereira, I. A. C. (2006). Characterization of the Desulfovibrio desulfuricans ATCC 27774 DsrMKJOP complex - a membrane-bound redox complex involved in the sulfate respiratory pathway. Biochemistry 45, 249–262. doi: 10.1021/bi0515265

PubMed Abstract | CrossRef Full Text | Google Scholar

Refojo, P. N., Sena, F. V., Calisto, F., Sousa, F. M., and Pereira, M. M. (2019). The plethora of membrane respiratory chains in the phyla of life. Adv. Microb. Physiol. 74, 331–414. doi: 10.1016/bs.ampbs.2019.03.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Refojo, P. N., Sousa, F. L., Teixeira, M., and Pereira, M. M. (2010). The alternative complex III: a different architecture using known building modules. Biochim. Biophys. Acta Bioenerg. 1797, 1869–1876. doi: 10.1016/j.bbabio.2010.04.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Rothery, R. A., and Weiner, J. H. (1993). Topological characterization of Escherichia coli DMSO reductase by Electron paramagnetic resonance spectroscopy of an engineered [3Fe-4S] cluster. Biochemistry 32, 5855–5861. doi: 10.1021/bi00073a019

CrossRef Full Text | Google Scholar

Rothery, R. A., and Weiner, J. H. (1996). Interaction of an engineered [3Fe-4S] cluster with a menaquinol binding site of Escherichia coli DMSO reductase. Biochemistry 35, 3247–3257. doi: 10.1021/bi951584y

PubMed Abstract | CrossRef Full Text | Google Scholar

Rothery, R. A., Workun, G. J., and Weiner, J. H. (2008). The prokaryotic complex iron-sulfur molybdoenzyme family. Biochim. Biophys. Acta Biomembr. 1778, 1897–1929. doi: 10.1016/j.bbamem.2007.09.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Saraiva, L. M., da Costa, P. N., Conte, C., Xavier, A. V., and LeGall, J. (2001). In the facultative sulphate/nitrate reducer Desulfovibrio desulfuricans ATCC 27774, the nine-haem cytochrome c is part of a membrane-bound redox complex mainly expressed in sulphate-grown cells. Biochim. Biophys. Acta 1520, 63–70. doi: 10.1016/S0167-4781(01)00250-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Sargent, F. (2016). The model [NiFe]-hydrogenases of Escherichia coli. Adv. Microb. Physiol. 68, 433–507. doi: 10.1016/bs.ampbs.2016.02.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Shi, Y., Xin, Y., Wang, C., Blankenship, R. E., Sun, F., and Xu, X. (2020). Cryo-EM structures of the air-oxidized and dithionite-reduced photosynthetic alternative complex III from Roseiflexus castenholzii. Sci. Adv. 6:eaba2739. doi: 10.1126/sciadv.aba2739

PubMed Abstract | CrossRef Full Text | Google Scholar

Shirodkar, S., Reed, S., Romine, M., and Saffarini, D. (2011). The octahaem SirA catalyses dissimilatory sulfite reduction in Shewanella oneidensis MR-1. Environ. Microbiol. 13, 108–115. doi: 10.1111/j.1462-2920.2010.02313.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Simon, J. (2002). Enzymology and bioenergetics of respiratory nitrite ammonification. FEMS Microbiol. Rev. 26, 285–309. doi: 10.1111/j.1574-6976.2002.tb00616.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Sousa, J. S., Calisto, F., Langer, J. D., Mills, D. J., Refojo, P. N., Teixeira, M., et al. (2018). Structural basis for energy transduction by respiratory alternative complex III. Nat. Commun. 9:1728. doi: 10.1038/s41467-018-04141-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Steinmetz, P. A., Degreif-Dünnwald, P., and Unden, G. (2014). The aerobic and anaerobic respiratory chain of Escherichia coli and Salmonella enterica: enzymes and energetics. EcoSal Plus 6. doi: 10.1128/ecosalplus.ESP-0005-2013

PubMed Abstract | CrossRef Full Text | Google Scholar

Stolz, J. F., Basu, P., Santini, J. M., and Oremland, R. S. (2006). Arsenic and selenium in microbial metabolism. Annu. Rev. Microbiol. 60, 107–130. doi: 10.1146/annurev.micro.60.080805.142053

PubMed Abstract | CrossRef Full Text | Google Scholar

Sun, C., Benlekbir, S., Venkatakrishnan, P., Wang, Y., Hong, S., Hosler, J., et al. (2018). Structure of the alternative complex III in a supercomplex with cytochrome oxidase. Nature 557, 123–126. doi: 10.1038/s41586-018-0061-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Tang, H., Rothery, R. A., and Weiner, J. H. (2013). A variant conferring cofactor-dependent assembly of Escherichia coli dimethylsulfoxide reductase. Biochim. Biophys. Acta Bioenerg. 1827, 730–737. doi: 10.1016/j.bbabio.2013.02.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Venceslau, S. S., Matos, D., and Pereira, I. A. C. (2011). EPR characterization of the new Qrc complex from sulfate reducing bacteria and its ability to form a supercomplex with hydrogenase and TpIc3. FEBS Lett. 585, 2177–2181. doi: 10.1016/j.febslet.2011.05.054

PubMed Abstract | CrossRef Full Text | Google Scholar

Venceslau, S. S., Stockdreher, Y., Dahl, C., and Pereira, I. A. C. (2014). The “bacterial heterodisulfide” DsrC is a key protein in dissimilatory sulfur metabolism. Biochim. Biophys. Acta Bioenerg. 1837, 1148–1164. doi: 10.1016/j.bbabio.2014.03.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Waterhouse, A. M., Procter, J. B., Martin, D. M. A., Clamp, M., and Barton, G. J. (2009). Jalview version 2-a multiple sequence alignment editor and analysis workbench. Bioinformatics. 25, 1189–1191. doi: 10.1093/bioinformatics/btp033

PubMed Abstract | CrossRef Full Text | Google Scholar

Weiner, J. H., Shaw, G., Turner, R. J., and Trieber, C. A. (1993). The topology of the anchor subunit of dimethyl sulfoxide reductase of Escherichia coli. J. Biol. Chem. 268, 3238–3244. doi: 10.1016/S0021-9258(18)53684-X

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, J., Oettmeier, W., Gennis, R. B., and Hellwig, P. (2002). FTIR spectroscopic evidence for the involvement of an acidic residue in quinone binding in cytochrome bd from Escherichia coli. Biochemistry 41, 4612–4617. doi: 10.1021/bi011784b

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, Z., and Weiner, J. H. (1998). Interaction of 2-n-heptyl-4-hydroxyquinoline-N-oxide with dimethyl sulfoxide reductase of Escherichia coli. J. Biol. Chem. 273, 20758–20763. doi: 10.1074/jbc.273.33.20758

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: NrfD-like, membrane protein, ion translocation, quinine/quinol binding site, CISM family, energy transduction

Citation: Calisto F and Pereira MM (2021) The Ion-Translocating NrfD-Like Subunit of Energy-Transducing Membrane Complexes. Front. Chem. 9:663706. doi: 10.3389/fchem.2021.663706

Received: 03 February 2021; Accepted: 19 March 2021;
Published: 13 April 2021.

Edited by:

Petra Hellwig, Université de Strasbourg, France

Reviewed by:

Blanca Barquera, Rensselaer Polytechnic Institute, United States
Giovanni La Penna, National Research Council (CNR), Italy

Copyright © 2021 Calisto and Pereira. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Manuela M. Pereira, mmpereira@fc.ul.pt

Download