Abstract
Protein folding and assembly into macromolecule complexes within the living cell are complex processes requiring intimate coordination. The biogenesis of complex iron sulfur molybdoenzymes (CISM) requires use of a system specific chaperone – a redox enzyme maturation protein (REMP) – to help mediate final folding and assembly. The CISM dimethyl sulfoxide (DMSO) reductase is a bacterial oxidoreductase that utilizes DMSO as a final electron acceptor for anaerobic respiration. The REMP DmsD strongly interacts with DMSO reductase to facilitate folding, cofactor-insertion, subunit assembly and targeting of the multi-subunit enzyme prior to membrane translocation and final assembly and maturation into a bioenergetic catalytic unit. In this article, we discuss the biogenesis of DMSO reductase as an example of the participant network for bacterial CISM maturation pathways.
Introduction
It was proposed in the early 1970s that the information required for a correctly folded protein is contained inclusively within its primary sequence (1). This concept was later challenged in the mid-1980s with the observation that in vivo a heat-shock protein assisted as a ‘chaperone’ in protein folding (2). Non-productive folding routes, such as those leading to aggregation, degradation, or abortive maturation of inactive apoenzymes, must be prevented in order to ensure correct subcellular targeting of cofactor-containing enzymes. Thus, the ‘protein-folding problem’ presents a myriad of interrelated questions: What controls the protein folding process in vivo in addition to the peptide chain searching the thermodynamic landscape for a conformational energetic minimum? How are folding and targeting coordinated with cofactor insertion? How are proteins targeted to their correct subcellular locations? What controls the assembly of large multimeric complexes?
Bacteria are resilient organisms that can subsist in diverse environments. In addition to oxygen, bacteria can exploit a myriad of substrates as terminal electron acceptors for respiration in anaerobic environments. Respiratory redox enzymes catalyze these oxidation/reduction reactions by transferring electrons from a donor to an acceptor molecule, most often operating at the cytoplasmic membrane by forming a redox loop between periplasmic and cytoplasmic enzymes connected by the quinone pool (3), (4). The facultative anaerobic model organism, Escherichia coli, has a variety of characterized anaerobic electron acceptors that include nitrate (NO3−), nitrite (NO2−), fumarate (C4H4O4), trimethylamine N-oxide [(CH3)3NO, TMAO] and dimethyl sulfoxide [(CH3)2SO, DMSO] (4).
The twin-arginine translocase (Tat) system is the purported protein-conducting channel used by enzymes that exhibit cytoplasmic cofactor insertion and subsequent membrane localization (5), (6), (7). Proteins that utilize the Tat pathway are characterized by two defining features: (i) they contain a consensus S/T-R-R-x-F-L-K twin arginine (RR) motif in their N-terminal leader peptide sequence (RR-leader); and (ii) they are usually translocated across the cytoplasmic membrane as an active and folded holoenzyme (8), (9), (10). The Tat translocon consists of the TatABC subunits and the current model identifies TatA as the homo-oligomeric pore subunit, whereas TatB and TatC act in substrate recognition and delivery (11), (12).
Respiratory oxidoreductases that contain a molybdenum-bis(pyranopterin guanine dinucleotide) (Mo-bisPGD) catalytic cofactor are grouped under the molybdoenzyme superfamily (13), (14), (15), (16), (17). In many instances, the catalytic subunit also coordinates an iron-sulfur cluster ([Fe-S]) near the catalytic site, an electron transfer subunit that typically contains typically four [Fe-S] clusters of various Fe:S ratios, and a membrane anchor subunit required to connect the [Fe-S] cluster electrical conduit to the quinone pool (16), (18) which may also contain hemes. These enzymes are now described as a complex iron-sulfur molybdoenzyme (CISM) to distinguish from those iron-sulfur-lacking molybdoenzymes (19). The architecture, function, biogenesis, and maturation of CISMs have been a major topic of recent study (16), (17), (20).
When oxygen-sensing regulatory proteins in the cell (e.g. Fnr) detect a lack of oxygen the biogenesis of a CISM begins (21). In many instances, the induction of CISM maturation too requires the presence of a substrate (e.g. in Rhodobacter capsulatus, the substrate DMSO is induced 150 fold upon commencing DMSO reductase biogenesis) (22). Maturation of a CISM into a functional holoenzyme requires numerous stages that involve initial translational ribosome integrated folding, cofactor insertion and coordination, subsequent folding and assembly with other subunits that may have had similarly complex folding pathway. The many steps comprising the cytoplasmic biogenesis processes are highly complex and must be intricately coordinated by numerous assistant proteins to produce a functional CISM. Of special interest to CISM maturation are system specific chaperones that help mediate the complete final folding and assembly (23), (24). Such chaperones were termed redox enzyme maturation proteins (REMPs) and are essential for the proper assembly of CISMs albeit absent in the final assembled holoenzyme (25), (26), (27), (28).
The importance of REMPs in CISM maturation is becoming accepted (17), (29). A key mechanistic finding is that REMPs do not chaperone in isolation. Rather, they ‘escort’ their CISM substrates though the entire maturation process as implied by their interactome (30) (Figure 1). Accordingly, many potential roles for REMPs have been proposed, including functioning as: (i) foldases to ensure correct secondary and tertiary structure (25), (31); (ii) unfoldases to correct folding mistakes (25); (iii) avoidance chaperones to prevent incorrect membrane targeting during folding and assembly (32); (iv) cofactor-assembly chaperones to maintain apoenzymes in a cofactor-binding competent conformation (30), (31); (v) cofactor-binding proteins, which bind the cofactor prior to its transfer to the apoenzyme (30), (33); (vi) targeting proteins directing substrates to specific cellular locations (34), (35); (vii) escort chaperones to promote transmembrane transport of enzyme complexes (30), (31), (34); (viii) proofreading chaperones to suppress transport until essential prior steps in the assembly process are complete (33), (36), (37); and (ix) protease protection chaperones to prevent degradation during assembly (38). Yet despite these proposed roles, a detailed path of actions for REMPs has yet to be defined.

Interactome for DmsD.
The interaction web for DmsD derived from pair wise protein-protein interaction experiments (30), (51). The proteins identified to interact with DmsD through various biochemical studies are listed below.
| Protein | Class/group | Reference |
| TatB | Translocase | Papish et al. (52); Kostecki et al. (51) |
| TatC | Translocase | Papish et al. (52); Kostecki et al. (51) |
| GroEL | General molecular chaperone | Li et al. (30) |
| DnaK | General molecular chaperone | Li et al. (30) |
| DnaJa | General molecular chaperone | Li et al. (30) |
| GrpEa | General molecular chaperone | Li et al. (30) |
| TufA/Ef-Tu | Ribosome-associated | Li et al. (30) |
| TF | Ribosome-associated | Li et al. (30) |
| MoeA | MobisPGD biosynthesis | Li et al. (30) |
| MoeB | MobisPGD biosynthesis | Li et al. (30) |
| MogA | MobisPGD biosynthesis | Li et al. (30) |
| MobB | MobisPGD biosynthesis | Li et al. (30) |
TufA-Ef-Tu translation elongation factor, TF trigger factor.
aThese co-chaperones work concurrently with DnaK in the DnaK-DnaJ-GrpE chaperone assembly.
DMSO reductase is an archetypical heterotrimer CISM enzyme comprised of three modular subunits, DmsABC (15), (39). DmsA, the RR-leader containing subunit of the enzyme, functions as the catalytic subunit that coordinates the MobisPGD catalytic cofactor in addition to one [4Fe-4S] (40), (41). DmsB serves as the electron conduit subunit coordinating four [4Fe-4S] clusters through conserved Cys residues, and is essential for anchoring to the cytoplasmic membrane via the integral membrane protein DmsC (42). Importantly, insertion of DmsC into the cytoplasmic membrane in the absence of DmsAB is lethal or suppresses growth (43), (44). DmsC is the subunit responsible for subunit anchoring to the membrane and is involved with binding and oxidation (44), (45).
An important component of DMSO reductase is the REMP dimethyl sulfoxide protein D, DmsD (17). Twin-arginine leader peptide binding is a canonical feature of REMPs, and as such, binding of DmsD to the DmsA RR-leader peptide is an absolute prerequisite for DMSO reductase biogenesis (26), (27). It is clear that DmsD binds the DmsA RR-leader at 1:1 stoichiometry and that the entire hydrophobic region is required for binding (46) although specificity is more difficult to define (47). Phylogenetic analyses found DmsD to be related to system-specific chaperones of other MobisPGD-containing oxidoreductases that included TorD for TMAO reductase and NarJ for cytoplasmic nitrate reductase A, both of which are CISM enzymes related to DMSO reductase (19), (25). Regardless of whether the RR-leaders from each substrate have a final teleological purpose of targeting to the Tat system, the structure of the signal peptides must be fine-tuned for a singular REMPs specificity and unique CISM maturation pathway. Here we discuss the proposed schematic network (Figure 2) of E. coli DmsD in the participant pathway for DMSO reductase biogenesis.
![Figure 2: Maturation pathway model for DMSO reductase.The cartooned pathway shows stages (10–15). The nascent chain exiting the ribosome and the RR-leader interacts with DnaK and trigger factor (TF) (1 and 2). The REMP chaperone DmsD (2), in addition to DnaJ and GrpE (3), join the interactome. DmsD interaction with GroEL (4) allows insertion of the molybdopterin cofactor through contact with MoeB (5). After DmsA is fully folded (6), the 4[4Fe-4S] center containing DmsB must fold and interact with DmsA (7 and 8). With DmsAB assembled, it is targeted to the Tat translocon (9–11) and translocated (12–14). Finally, DmsAB docks to the membrane anchor subunit DmsC to complete the accomplishment of securing a fully functional respiratory enzyme (15).](/document/doi/10.1515/bmc-2017-0011/asset/graphic/j_bmc-2017-0011_fig_002.jpg)
Maturation pathway model for DMSO reductase.
The cartooned pathway shows stages (10–15). The nascent chain exiting the ribosome and the RR-leader interacts with DnaK and trigger factor (TF) (1 and 2). The REMP chaperone DmsD (2), in addition to DnaJ and GrpE (3), join the interactome. DmsD interaction with GroEL (4) allows insertion of the molybdopterin cofactor through contact with MoeB (5). After DmsA is fully folded (6), the 4[4Fe-4S] center containing DmsB must fold and interact with DmsA (7 and 8). With DmsAB assembled, it is targeted to the Tat translocon (9–11) and translocated (12–14). Finally, DmsAB docks to the membrane anchor subunit DmsC to complete the accomplishment of securing a fully functional respiratory enzyme (15).
Stages of DMSO reductase biogenesis
As a result of research on the system specific chaperones over the past 15 years allows the opportunity to conceptualize an assembly pathway of CISM maturation. This pathway is illustrated in Figure 2.
Stages 1 and 2: the ribosome, trigger factor (TF) and DnaK
A key event in CISM maturation involves binding of the RR-leader peptide as the nascent polypeptide exits the ribosome. TF and DnaK are promiscuous chaperones that assist in protein folding and have a combined substrate pool of over 1000 cellular proteins (48). Studies have demonstrated that TF and DnaK interact with Tat system substrates, and both chaperones have an apparently similar substrate specificity: a stretch of hydrophobic amino acid residues flanked on either end by positively charges residues (38), (49), (50). Indeed, DmsD has some form of association with the general chaperones, trigger factor (TF) and DnaK, which implicates DmsD as a central connector for all processes required for DMSO reductase maturation (30). We have established an interaction web with differential affinities for the DmsA RR-leader in the following order: DmsD>TF>>DnaK (29), suggesting that an initial chaperone cascade from ribosome to TF to DnaK could be involved in DMSO reductase maturation with each chaperone having overlapping and distinct roles in folding modulation thereby promoting productive folding and avoiding proteolytic degradation (51), (52).
Stage 3: DnaJ and GrpE
Many of the DnaK functions appear to rely on the presence of the co-chaperone partners, DnaJ and GrpE. The DnaK-DnaJ-GrpE assembly is a well-studied chaperone cascade machine in E. coli initially identified for their roles in heat shock rescue (53). Our interactome data suggests that there is a DmsD-DnaJ interaction; however, it is unclear whether this interaction is direct or detected due to the presence of DnaK or another effector molecular (30).
Stage 4: the DmsD-GroEL interaction
A large number of targets in the interaction proteome of DmsD have been identified that include general molecular chaperones, ribosomal components, and cofactor biosynthesis proteins. Figure 1 demonstrates the possible processes of these non-substrate proteins that were identified in the DmsD interactome from several studies (30), (51), (52). These interactions implicate DmsD as a hub of the DMSO reductase maturation pathway for connecting the nascent DmsA polypeptide to proteins that would assist in active holoenzyme maturation. Interaction of DmsD with the general chaperone, GroEL, is an example of one such downstream connection (30), (54). As GroEL provides a protective cavity for newly synthesized or misfolded polypeptides to fold or refold (55), it is plausible to hypothesize that DmsD having received DmsA from TF, directs it towards GroEL which then further assists in the folding of DmsA (56). Further folding of DmsA outside of the GroEL cavity would allow for subsequent insertion of [4Fe-4S] and MobisPGD during maturation.
Stage 5: molybdenum cofactor (Moco) insertion
A critical step to procure an enzymatically active DmsABC holoenzyme is the insertion of the MobisPGD cofactor (42), (57). Studies involving another MobisPGD-containing CISM, TMAO reductase, showed that mutants deficient in molybdate-uptake and cofactor biosynthesis were improperly localized within the cell (7). Accordingly, MobisPGD could be considered a chemical chaperone for DMSO reductase folding and stabilization of the tertiary structure of DmsA for cofactor coordination (58), further implicating the secondary importance of MobisPGD in CISM maturation. Two enzymes that catalyze terminal steps in the Moco biosynthesis pathway, MoeB and MoaA, have been demonstrated through both in vivo and in vitro protein-protein interaction methods to interact with DmsD as well as the RR-leader peptide (30). However, complete processing of the RR-leader peptide is not dependent on cofactor availability, as moeB chromosomal mutation does not prevent maturation and membrane insertion of the apomolybdoenzyme (59). Although the highly-labile and complex MobisPGD cofactor is an essential component for DMSO reductase activity, understanding its insertion in the maturation pathway poses as a challenge to understanding CISM folding. Currently, it is postulated that MoeB and MoaA are involved in the final cofactor hand-off during CISM maturation (60), (61), (62), (63).
Stages 6 to 8: DmsA and DmsB interaction
The importance of [Fe-S] clusters is recognized in all kingdoms of life because they can bind electron-rich enzymatic substances, accept or donate single electrons, and can also stabilize protein conformations (64). The biogenesis of [Fe-S] clusters in E. coli is a highly conserved and complex process mediated by several dedicated coordinated systems (65), (66). The proposed roles of the assembly components include: iron donors (67), (68), intermediate [Fe-S] cluster scaffold proteins (69), (70), (71), carrier proteins (72), and putative [Fe-S] cluster repair proteins (73). The biogenesis system may have evolved to protect and enshroud [Fe-S] clusters during the majority of the biogenesis and transfer process, as the [Fe-S] species are notoriously vulnerable to oxidative stress (74), (75). Recently, the biosynthesis and distribution of [Fe-S] clusters has been linked to the maturation of molybdoenzymes and in particular to the synthesis of molybdenum cofactor (76), (77). While the four [4Fe-4S] clusters of DmsB are essential to receive and conduit electrons from the cytoplasmic quinone pool, defining the exact molecular mechanism of [Fe-S] cluster assembly in DmsB insertion is still a challenge due to the spontaneous chemical nature of the species. Yet it is considered that the fully [Fe-S] holo DmsB needs to fold with cofactors prior to interaction with holo DmsA (16).
It is suggested that RR-leader bearing oxido-reductases with more than one subunit are targeted to the translocase via a ‘piggy-back’ mechanism, as all identified multi-subunit oxidoreductases contain a RR-leader in only one of the subunits (78). In this way the non RR-leader containing subunit folds and interacts with the RR-leader containing subunit in order to catch a ride across the cell membrane. While recent studies have shown that the loss of several general chaperone genes in vivo leads to loss or retardation of respiratory growth on DMSO, the final assembly of the DmsAB catalytically active holoenzyme is an important step in the maturation pathway that remains to be investigated. However, it is known that most of the CISM enzymes will fold a catalytically active form of the enzyme prior to interaction and subsequent translocation across the cell membrane (28), (79).
Stages 9 to 11: targeting to the Tat translocon
The Tat translocon in E. coli is comprised of Tat A, B, and C subunits of which multiple copies of each are involved in the translocon (80), (81). Tat-targeted respiratory enzymes appear to have RR-leaders in only one subunit of the holoenzyme (8) that differs from general secretory pathway (Sec) signal peptides (82). Specifically, the RR-leader peptide is recognized at the membrane by a receptor complex formed from copies of TatB and TatC (80), (81). It has been established by both in vivo and in vitro experiments that DmsD interacts with both TatB and TatC (34), (51), (52). It is likely that the associated REMP docks at TatBC to donate the substrate to the TatBC receptor, with unfolded proteins rejected by the TatBC receptor complex through an undefined quality control mechanism (83). In addition to leader recognition, membrane targeting has been suggested to also involve association with membrane phospholipids by the RR-leader (35), (84), (85).
The events surrounding Tat targeting likely include substrate handover in conjunction with DmsD dissociation. Undoubtedly, DmsA can interact with the Tat translocon independent of DmsD, further supporting the existence of a handover or dissociation step (30). Moreover, DmsD association with the cytoplasmic membrane occurs only in the presence of TatB or TatC (51), (52). Although a transient tripartite interaction is too complicated to analyze by most biochemical techniques, an interaction between DmsD, TatB/C, and the RR-leader suggests that an intermediate complex involving DmsA(B)/DmsD/TatB(C) is highly probable.
Stages 12 to 14: crossing of the cytoplasmic membrane
The most remarkable feature of the Tat system is that its substrates are usually fully-folded prior to membrane translocation (10). Substrate protein recognition by the TatBC complex subsequently results in the recruitment and oligomerization of TatA protomers to form the active translocon pore using energy from the proton motive force (86), (87). As the purported protein-conducting channel, the TatA multi-protomer pore unit must be large enough to permit passage of large protein substrates possessing secondary, tertiary, and in many cases quaternary structure with sizes ranging from ~10 kDa to ~140 kDa (88), (89), (90). Following translocation initiation from the TatBC recognition unit, substrates are translocated through the TatA multimer pore across the cytoplasmic membrane utilizing the electrochemical gradient (86), (88), (90), (91), (92), (93). Several studies have found that the substrate protein interacts with TatA and TatBC separately, suggesting that the translocon is not fully assembled until receiving a substrate (88), (89), (90), (94). Currently, it is proposed that both TatA and TatB use the same binding site on TatC, and since TatB occupies this site in the resting TatBC complex, TatA must displace TatB form the site at some stage in the translocation cycle (95). Despite the various data available on Tat complex formation, piecing together the translocation pathway from various studies still proves to be a daunting task. To elucidate the mechanism of Tat transport, it is critical to better define how the multiple Tat components are arranged within the translocation complex.
Following successful translocation, the RR-leader is assumed to be cleaved off by leader peptidease I, LepB (96). However, mutagenesis to the LepB recognition and cleavage site does not appear to totally inhibit RR-leader processing of DmsA (97). To date no study has shown LepB interacting with TatABC and studies suggest that the RR-leader containing substrate is released from the Tat translocon prior to cleavage (98). This study suggests that the uncleaved RR-leader could anchor the protein to the membrane until the leader is cleaved.
Stage 15: anchoring to DmsC
The final stage of DMSO reductase assembly involved attachment of DmsAB to the DmsC membrane anchor subunit. As DmsC is a polytopic integral membrane protein with eight transmembrane segments, it is likely translocated and inserted into the membrane by one of the two known pathways in bacteria as opposed to spontaneous insertion (99), (100). Truncation studies have demonstrated that the C-terminus of DmsB is indispensable for anchoring to DmsC, suggesting that the direct docking of DmsAB to DmsC is likely solely through interactions with DmsB (42) in agreement with the overall structural architecture of CISM systems. While much is known on the mechanism of electron transfer between the DmsABC subunits, exactly how the final complex is assembled remains to be elucidated.
Molecular regulation of the system
Proteostasis
Protein quality control ensures that proteins are correctly folded and functional at the right place and time and is essential in cellular life (101). This regulation is mediated by chaperone and protease systems, in accordance with cellular clearance mechanisms (102). The most intensively studied folding chaperones, such as the GroEL-GroES and DnaK systems, facilitate substrate protein folding through ATP- and cofactor-driven conformational changes that convert the chaperones from a state of high substrate affinity to low substrate affinity, thereby allowing substrate proteins to be released (103). An emerging functional feature of ATP binding chaperones is their highly dynamic behavior, and it seems that their massive domain movement is only weakly coupled to their nucleotide states. Rather, these ATPases are in a continual state of rapid flux (104). A mechanism for GroEL-assisted folding of large proteins involves their binding to the open (trans) ring, and subsequent folding in the bulk solution outside of the cavity (55). If folding does not succeed, the cycle then repeats after the substrate is released by hydrolysis of ATP in the GroEL cis ring. Additionally, ‘adaptor’ proteins may be utilized to connect GroEL with protein substrates. Such an interaction was recently discovered between DmsD and GroEL, in that DmsD was shown to function as a connector to carry the DmsA substrate preprotein to GroEL for assisted folding (54). Although the numerous known folding chaperones have a fairly broad range of substrates, each chaperone family has a distinct mode of ATP binding (104). Currently, a major aim in the field is obtaining high-resolution structures of chaperone complexes acting on misfolded or unfolded proteins to better define the energetic regulation of protein quality control.
Additionally, degradation of protein in E. coli is generally executed by ATP-dependent proteases (105). The proteases form large multisubunit machines with an internal proteolytic chamber accessible to unfolded proteins only. The denatured polypeptides are translocated into the proteolytic chamber in an ATP-dependent manner (106). It was demonstrated that TorD (the REMP for the CISM TMAO reductase) binding to the core of apoTorA prevents proteolytic attack of the Lon protease and also the proteolysis of the N-terminal extremity by an additional, still unknown protease (50), (107). Although the region of TorD involved in the recognition of apoTorA is well defined, the region and energetic mechanism responsible for the protection of the apoTorA signal sequence is still under debate (49), (108).
Guanosine 5′-triphosphate (GTP)
Molybdoenzymes are present in all domains of life and they catalyze critical steps in carbon, sulfur, and nitrogen metabolism. A large diversity of organic and inorganic substrates are utilized for catalyzing oxidation-reduction reactions, and as such, molybdoenzymes play a crucial role in fundamental biogeochemical cycles (109), (110).
The biosynthesis of the molybdenum cofactor (Moco) is a highly conserved and complex pathway (60), (111). For instance, in E. coli, nine proteins with known function are directly involved in Moco biosynthesis (35), (112). The biosynthesis of Moco begins from guanosine 5-triphosphate (GTP) and is catalyzed by the two proteins MoaA and MoaC in bacteria (76), the former containing two [4Fe-4S] clusters (113). The different final forms of Moco are inserted into molybdoenzymes that are categorized based on the ligands at the molybdenum atom (14). The proteins of the DMSO reductase family in bacteria contain the bisPGD form of the cofactor (14), and the synthesis of which has been shown to occur in a two-step reaction that requires hydrolysis of Mg-GTP (114). As bisPGD is not stable in its free form, it is immediately bound by Moco-binding chaperones, which insert the cofactor specifically into its target enzymes. The system specific chaperone of Nitrate reductase A, NarJ, has been shown to interact with Moco biosynthesic machinery and to facilitate final complex assembly prior to Tat translocation (48), (115), (116). Moreover, TorD, induces a conformational change of apoTorA to allow competency for Moco insertion, as well as interaction with the MobA protein involved in bisMGD formation (49). The maturation steps associated with the Moco biosynthesis involve a myriad of interactions with enzymes, chaperones and cofactors alike.
GTP binding and hydrolysis are hypothesized to govern the activity of CISM maturation during the final stages of protein folding (33). It has been suggested that GTP binding by REMP proteins themselves may play a role in the REMP/substrate maturation process (33). Accordingly, the domain-swapper dimer of TorD, showed an increase in GTP affinity following TorA ligand binding (27), (117), (118), and it was suggested to in fact be binding to the mature molybdopterin-guanine dinucleotide (MGD) form of Moco as a step in the cofactor insertion event (49). Recent investigation has illustrated strong cooperativity in the DmsD binding to both GTP and the RR-leader, suggesting that GTP binding at one site on DmsD alters affinity at additional binding sites (other protomers of a multimeric state of DmsD). Oddly, similar dissociation constants for DmsD release of the RR-leader were determined with all guanine nucleotides (GTP, GDP, and GMP), implying that hydrolysis is not involved and that the recognition is directed via the guanosine moiety rather than attached phosphate groups (119). Overall, GTP is crucial factor in CISM maturation both as a component of the molybdopterin cofactor as well as potentially regulating the REMP’s protein-protein interactions.
Protein translocation energetics
Two transport mechanisms have evolved to facilitate passage of proteins across the cytoplasmic membrane. In bacteria, secretory proteins are translocated either via the general secretion (Sec) pathway (120), (121) or the Tat pathway (8). The fundamental functional difference between the two pathways is that the Sec system is involved in the secretion and insertion of unfolded proteins, while that Tat system is implicated in the secretion of folded and/or cofactor containing proteins (122), (123). It has been proposed that the Tat system is utilized only when cytoplasmic folding of the protein substrate therefore omitted the use of the Sec system (124). There may be specific folding-related motives for a Tat system preference, such as labile cofactors in the example of CISM (125), (126). If cytoplasmic folding is not required for a Tat substrate, the question arises why the Tat system is still preferred (122). Although the Sec and Tat pathways translocate proteins by distinctive mechanisms, many common fundamentals can be recognized.
The bacterial Sec translocase is composed of a membrane embedded protein conducting channel (PCC) that consists of three integral membrane proteins, SecY, SecE, and SecG, and a peripheral associated ATPase, SecA, which functions as a molecular motor to drive membrane translocation (120), (123), (127). SecA associates peripherally to the PCC, where it accepts secretory proteins from chaperones to subsequently thread the unfolded protein through the narrow PCC transmembrane channel in an ATP dependent fashion. Secretory proteins can be targeted to the Sec translocase by two different mechanisms, the co-translational and the post-translational targeting (128), both of which occurring through a step-wise process wherein a catalytic turnover of the SecA ATPase is couple to the translocation of the unfolded protein. Comparatively, as described earlier, the Tat translocase consists of three membrane integrated subunits, TatA, TatB, and TatC, which together form a receptor (TatBC) and a protein conducting channel (TatA) for protein translocation (122).
A hypothesized possible role of the REMP chaperones was for Sec system avoidance. It was noted that Tat dependent proteins could default to the Sec translocase in a ∆tatABC mutant, albeit not leading to a folded functional protein. Following this premise, it was observed that SecA ATPase activity in vitro was activated in the presence of a RR-leader containing substrate, but inhibited in the presence of its cognate REMP (unpublished results).
Protein translocation requires the input of an energy source. In general, the two main sources are chemical energy (ATP, GTP) and electrochemical energy (PMF) (129), (130). Whereas GTP is the main driving force during Sec co-translational protein translocation, ATP is the main source of energy utilized during the Sec post-translational route. Quantitative estimates of nucleotide requirements for ATP-driven Sec translocation indicated that about 500 molecules of ATP are required per translocated polypeptide of 25 kDa in length (131). Interestingly while ATP is essential for initiation of Sec mediated translocation, depending on the substrate, translocation may be further stimulated by the PMF (128), (131). Although the PMF cannot initiate translocation, PMF-driven translocation is highly efficient in the absence of SecA, and it is suggested that the PMF acts as a driving force possibly involving vectorial proton movements (129), (130).
In contrast, Tat transport is triggered and driven by the PMF. Specifically, transport is driven by the PMF of which the pH is the dominant component (132), and there is no requirement for ATP (133), (134). Only a small ΔpH is required for Tat transport in vitro (132). Measurement with bacterial Tat systems in vivo indicate that nearly 80000 protons pass the membrane per translocated protein (132), which render Tat mediated translocation energetically more costly than the Sec system. Furthermore, TatA oligomerizes and associates with the TatBC complex only in the presence of a ΔpH, which highlights the absolute ΔpH requirement for successful protein translocation (88). Recent studies suggest that DmsA RR-leader release from DmsD is reduced below pH 6.0 (135), a range that corresponds to a transition in DmsD folding forms (136). Collectively, these results suggest that a ΔpH is imperative for not only successful protein translocation, but H+ may also serve as a molecular regulator in protein maturation prior to Tat system targeting.
Moreover, an important mechanistic question is whether protonation and deprotonation events are required for the conformational changes during Tat transport. In other PMF dependent motors, such as the ATP synthase, there are acidic residues located within the membrane that couple proton transfer to molecular movements and conformational changes. To date, no such acidic residues have been identified for TatA, TatB, or TatC (137), which raises the question as to the direct coupling of proton flux to Tat translocation. While numerous aspects of Sec-dependent protein translocation have been functionally reconstituted with purified translocase components, it still remains a significant challenge to address the many remaining mechanistic questions regarding Tat translocation.
Expert opinion
The process by which nascent polypeptides fold into functional proteins in vivo may be simple or complex. Dissecting the biochemical dance involved for a given biomolecular complex is typically poorly understood and/or studied, presenting researchers with a myriad of interrelated questions that fall under the umbrella of the ‘protein-folding problem’. Our research has demonstrated that multimeric bacterial respiratory protein complexes frequently require protein chaperones for functional assembly. The working hypothesis for the maturation of CISMs is that a system-specific chaperone, is necessary to pilot its cognate substrate through the numerous folding, cofactor loading and protein-protein interaction targeting steps.
The past few years have seen the development of novel quantitative genetic interaction (GI) screens aimed to elucidate the relationship between gene function and higher-level protein complexes. The group of M. Babu has applied this approach to E. coli (138), (139), and their findings support our proposed DMSO reductase maturation pathway. Specifically, the work suggests that protein homeostasis groups together the roles of chaperoning, protein folding, and proteases. Thus, in our maturation cascade, stages 1 through 4 may be intertwined. Additionally, Babu et al. identified a new gene responsible for molybdopterin biosynthesis, which was then linked with genes involved with the [Fe-S] cluster assembly. This information implies that the assembly of the molybdopterin and the [Fe-S] cluster of DmsA may be coupled, and that the coordination of stages 5 through 7 may be more intimate than our pathway suggests.
Bettering our understanding of in vivo folding, targeting, and assembly mechanisms of large multi-subunit CISMs will have a huge biological impact. It will provide insights into how large multi-subunit respiratory enzymes assemble in all organisms. Considerable effort has been expended to define how these complexes function, but there is paucity of information available on how they come to be. A comprehensive model of how respiratory enzyme subunits are guided from the ribosome, to their final subcellular destinations, sheds light on the temporal regulation of critical maturation steps, and on how multiple subunits are brought together to form active enzyme complexes.
Outlook
Modeling bacterial chaperones and their cognate nascent polypeptides as tractable archetypal systems address a fundamental biological question that applies to the bioenergetic systems of organisms across all domains in the tree of life. Investigation of REMP interactomes illustrates participation in a complex cascade of interactions, and the study of which will provide insights into the fundamental area of protein-protein interactions. This is particularly relevant in cases of transient, but highly specific, interactions that are necessary to coordinate complex multi-step enzyme maturation pathways.
Moreover, approximately one third of all proteins contain metal ions. Of these, transition metal irons, such as iron and molybdenum, are assembled into complex redox-active cofactors. Mo-containing enzymes exist in almost all organisms and catalyze key redox reactions that are critical to the global carbon, nitrogen, and sulfur cycles. As well, a subset of bacterial membrane-bound Mo-enzymes play a critical bioenergetic role, as their catalytic turnover contributes directly to the transmembrane proton gradient across the cytoplasmic membrane (20). The maturation and targeting of Mo-enzymes frequently depends on system-specific chaperones, most notably the REMPs and multi-cofactor containing multimeric enzymes. As such, investigation into REMP chaperones and the respiratory enzyme maturation processes in bacteria have the potential to reveal novel antimicrobial targets, and could provide broad-spectrum antibiotics against bacterial pathogens. Therefore, it is of utmost importance to better understand the exquisite coordination of cofactor insertion, translocation, multimeric subunit association, and membrane insertion in bacterial Mo-enzymes.
Highlights
Many proteins require assistance of system specific chaperones to achieve their native fold, including correct cofactor insertion and subcellular localization.
The twin-arginine translocase system transports full-folded proteins across the bacterial cytoplasmic membrane, and proteins that utilize the Tat pathway contain a twin-arginine motif leader peptide sequence at the N-terminus.
Anaerobic respiratory redox enzymes exploit the use of alternative terminal electron acceptors to subsist in anoxic environments.
Respiratory oxidoreductases that contain a molybdenum-bis(pyranopterin guanine dinucleotide) (Mo-bisPGD) catalytic cofactor are characterized as molybdoenzymes, and are further classified as complex iron its sulfur molybdoenzymes (CISMs) when iron-sulfur groups are also included in the final enzyme fold.
DMSO reductase is a CISM that utilizes DMSO as a terminal electron acceptor, and requires the use of various general chaperones as well as its system specific chaperone DmsD, for cytoplasmic folding and assembly prior to Tat translocation.
A large number of targets in the interaction proteome of DmsD have been identified, and recent studies has illustrated strong cooperativity in binding of DmsD to both GTP and the RR-leader.
Next steps are to better understand exactly how the DmsAB enzyme is recognized by TatBC, the translocation across the membrane by TatA and how the final DMSO reductase complex (DmsABC) is assembled following Tat translocation.
Acknowledgments
This research was supported by a Canadian Institutes of Health Research operating grant (MOP-49422) to R.J.T. The authors wish to thank the many colleagues with whom we have discussed fundamental aspects related to the investigation of this topic. We also thank C. Yu specifically for graphics help to generate Figure 2.
References
1. Fisher WR, Taniuchi T, Anfinsen B. On the role of heme in the formation of the structure of cytochrome c. J Biol Chem 1973: 248; 3188–95.10.1016/S0021-9258(19)44026-XSearch in Google Scholar
2. Morimoto RI. Regulation of the heat-shock transcriptional response: cross talk between a family of heat-shock factors, molecular chaperones, and negative regulators. Genes Dev 1998; 12: 3788–96.10.1101/gad.12.24.3788Search in Google Scholar
3. Bueno E, Mesa S, Bedmar EJ, Richardson DJ, Delgado MJ. Bacterial adaptation of respiration from oxic to microoxic and anoxic conditions: redox control. Antioxid Redox Signal 2012; 16: 819–52.10.1089/ars.2011.4051Search in Google Scholar
4. Richardson DJ. Bacterial respiration: a flexible process for a changing environment. Microbiology 2000; 146: 551–71.10.1099/00221287-146-3-551Search in Google Scholar
5. Weiner JH, Bilous PT, Shaw GM, Lubitz SP, Frost L, Thomas GH, Cole JA, Turner RJ. A novel and ubiquitous system for membrane targeting and secretion of cofactor-containing proteins. Cell 1998; 93: 93–101.10.1016/S0092-8674(00)81149-6Search in Google Scholar
6. Sargent F, Bogsch EG, Stanley NR, Wexler M, Robinson C, Berks BC, et al. Overlapping functions of components of a bacterial Sec-independent protein export pathway. EMBO J 1998; 17: 3640–50.10.1093/emboj/17.13.3640Search in Google Scholar
7. Santini CL, Ize B, Chanal A, Müller M, Giordano G, Wu LF. A novel Sec-independent periplasmic protein translocation pathway in Escherichia coli. EMBO J 1998; 17: 101–12.10.1093/emboj/17.1.101Search in Google Scholar
8. Berks BC. A common export pathway for proteins binding complex redox cofactors? Mol Microbiol 1996; 22: 393–404.10.1046/j.1365-2958.1996.00114.xSearch in Google Scholar
9. Wexler M, Bogsch EG, Klösgen RB, Palmer T, Robinson C, Berks BC. Targeting signals for a bacterial Sec-independent export system direct plant thylakoid import by the ΔpH pathway. FEBS Lett 1998; 431: 339–42.10.1016/S0014-5793(98)00790-XSearch in Google Scholar
10. DeLisa MP, Tullman D, Georgiou G. Folding quality control in the export of proteins by the bacterial twin-arginine translocation pathway. Proc Natl Acad Sci USA 2003; 100: 6115–20.10.1073/pnas.0937838100Search in Google Scholar PubMed PubMed Central
11. Palmer T, Berks BC. The twin-arginine translocation (Tat) protein export pathway. Nat Rev Microbiol 2012; 10: 483–96.10.1038/nrmicro2814Search in Google Scholar
12. Robinson C, Matos CFRO, Beck D, Ren C, Lawrence J, Vasisht N, Mendel S. Transport and proofreading of proteins by the twin-arginine translocation (Tat) system in bacteria. Biochim Biophys Acta Biomembr 2011; 1808: 876–84.10.1016/j.bbamem.2010.11.023Search in Google Scholar
13. McCrindle SL, Kappler U, McEwan AG. Microbial dimethylsulfoxide and trimethylamine-N-oxide respiration. Adv Microb Physiol 2005; 50: 147–98.10.1016/S0065-2911(05)50004-3Search in Google Scholar
14. Hille R. The molybdenum oxotransferases and related enzymes. Dalton Trans 2013; 42: 3029–42.10.1039/c2dt32376aSearch in Google Scholar PubMed
15. Rothery RA, Stein B, Solomonson M, Kirk ML, Weiner JH. Pyranopterin conformation defines the function of molybdenum and tungsten enzymes. Proc Natl Acad Sci USA 2012; 109: 2–7.10.1073/pnas.1200671109Search in Google Scholar PubMed PubMed Central
16. Iobbi-Nivol C, Leimkühler S. Biochimica et biophysica acta molybdenum enzymes, their maturation and molybdenum cofactor biosynthesis in Escherichia coli. Biochim Biophys Aacta Bioenerg 2013; 1827: 1086–101.10.1016/j.bbabio.2012.11.007Search in Google Scholar PubMed
17. Magalon A, Fedor JG, Walburger A, Weiner JH. Molybdenum enzymes in bacteria and their maturation. Coord Chem Rev 2011; 255: 1159–78.10.1016/j.ccr.2010.12.031Search in Google Scholar
18. Cheng VWT, Johnson A, Rothery RA, Weiner JH. S13.5 An alternative site for proton entry from the cytoplasm to the quinone binding site in the Escherichia coli succinate dehydrogenase. Biochim Biophys Acta Bioenerg 2008; 1777(Supp 0): S89.10.1016/j.bbabio.2008.05.349Search in Google Scholar
19. Rothery RA, Workun GJ, Weiner JH. The prokaryotic complex iron-sulfur molybdoenzyme family. Biochim Biophys Acta 2008; 1778: 1897–929.10.1016/j.bbamem.2007.09.002Search in Google Scholar PubMed
20. Grimaldi S, Schoepp-Cothenet B, Ceccaldi P, Guigliarelli B, Magalon A. The prokaryotic Mo/W-bisPGD enzymes family: A catalytic workhorse in bioenergetic. Biochim Biophys Acta Bioenerg 2013; 1827: 1048–85.10.1016/j.bbabio.2013.01.011Search in Google Scholar PubMed
21. Crack JC, Green J, Hutchings MI, Thomson AJ, Le Brun NE. Bacterial iron-sulfur regulatory proteins as biological sensor-switches. Antioxid Redox Signal 2012; 17: 1215–31.10.1089/ars.2012.4511Search in Google Scholar PubMed PubMed Central
22. Kappler U, Huston WM, McEwan AG. Control of dimethylsulfoxide reductase expression in Rhodobacter capsulatus: The role of carbon metabolites and the response regulators DorR and RegA. Microbiology 2002; 148: 605–14.10.1099/00221287-148-2-605Search in Google Scholar PubMed
23. Chan CS, Chang L, Rommens KL, Turner RJ. Differential interactions between tat-specific redox enzyme peptides and their chaperones. J Bacteriol 2009; 191: 2091–101.10.1128/JB.00949-08Search in Google Scholar PubMed PubMed Central
24. Ilbert M, Méjean V, Iobbi-Nivol C. Functional and structural analysis of members of the TorD family, a large chaperone family dedicated to molybdoproteins. Microbiology 2004; 150: 935–43.10.1099/mic.0.26909-0Search in Google Scholar PubMed
25. Turner RJ, Papish AL, Sargent F. Sequence analysis of bacterial redox enzyme maturation proteins (REMPs). Can J Microbiol 2004; 50: 225–38.10.1139/w03-117Search in Google Scholar PubMed
26. Oresnik IJ, Ladner CL, Turner RJ. Identification of a twin-arginine leader-binding protein. Mol Microbiol 2001; 40: 323–31.10.1046/j.1365-2958.2001.02391.xSearch in Google Scholar PubMed
27. Pommier J, Méjean V, Giordano G, Iobbi-Nivol C. TorD, a cytoplasmic chaperone that interacts with the unfolded trimethylamine N-oxide reductase enzyme (TorA) in Escherichia coli. J Biol Chem 1998; 273: 16615–20.10.1074/jbc.273.26.16615Search in Google Scholar PubMed
28. Blasco F, Pommier J, Augier V, Chippaux M, Giordano G. Involvement of the narJ or narW gene product in the formation of active nitrate reductase in Escherichia coli. Mol Microbiol 1992; 6: 221–30.10.1111/j.1365-2958.1992.tb02003.xSearch in Google Scholar PubMed
29. Chan CS, Bay DC, Leach TGH, Winstone TML, Kuzniatsova L, Tran VA, Turner RJ. “Come into the fold”: a comparative analysis of bacterial redox enzyme maturation protein members of the NarJ subfamily. Biochim Biophys Acta 2014; 1838: 2971–84.10.1016/j.bbamem.2014.08.020Search in Google Scholar PubMed
30. Li H, Chang L, Howell JM, Turner RJ. DmsD, a Tat system specific chaperone, interacts with other general chaperones and proteins involved in the molybdenum cofactor biosynthesis. Biochim Biophys Acta 2010; 1804: 1301–9.10.1016/j.bbapap.2010.01.022Search in Google Scholar PubMed PubMed Central
31. Shanmugham A, Bakayan A, Völler P, Grosveld J, Lill H, Bollen YJM. The hydrophobic core of twin-arginine signal sequences orchestrates specific binding to Tat-pathway related chaperones. PLoS One 2012; 7: e34159.10.1371/journal.pone.0034159Search in Google Scholar PubMed PubMed Central
32. Porcelli I, De Leeuw E, Wallis R, Van den Brink-van der Laan E, De Kruijff B, Wallace BA, Palmer T, Berks BC. Characterization and membrane assembly of the TatA component of the Escherichia coli twin-arginine protein transport system. Biochemistry 2002; 41: 13690–7.10.1021/bi026142iSearch in Google Scholar PubMed
33. Hatzixanthis K, Clarke TA, Oubrie A, Richardson DJ, Turner RJ, Sargent F. Signal peptide-chaperone interactions on the twin-arginine protein transport pathway. Proc Natl Acad Sci USA 2005; 102: 8460–5.10.1073/pnas.0500737102Search in Google Scholar PubMed PubMed Central
34. Kuzniatsova L, Winstone TML, Turner RJ. Identification of protein-protein interactions between the TatB and TatC subunits of the twin-arginine translocase system and respiratory enzyme specific chaperones. Biochim Biophys Acta Biomembr 2016; 1858: 767–75.10.1016/j.bbamem.2016.01.025Search in Google Scholar PubMed
35. Shanmugham A, Sang HWWF, Bollen YJM, Lill H. Membrane binding of twin arginine preproteins as an early step in translocation. Biochemistry 2006; 45: 2243–9.10.1021/bi052188aSearch in Google Scholar PubMed
36. Maillard J, Spronk CAEM, Buchanan G, Lyall V, Richardson DJ, Palmer T, Vuister GW, Sargent F. Structural diversity in twin-arginine signal peptide-binding proteins. Proc Natl Acad Sci USA 2007; 104: 15641–6.10.1073/pnas.0703967104Search in Google Scholar PubMed PubMed Central
37. Chan CS, Chang L, Rommens KL, Turner RJ. Differential Interactions between Tat-specific redox enzyme peptides and their chaperones. J Bacteriol 2009; 191: 2091–101.10.1128/JB.00949-08Search in Google Scholar PubMed PubMed Central
38. Pérez-Rodríguez R, Fisher AC, Perlmutter JD, Hicks MG, Chanal A, Santini CL, Wu LF, Palmer T, DeLisa MP. An essential role for the DnaK molecular chaperone in stabilizing over-expressed substrate proteins of the bacterial twin-arginine translocation pathway. J Mol Biol 2007; 367: 715–30.10.1016/j.jmb.2007.01.027Search in Google Scholar PubMed
39. Bilous PT, Weiner JH. Molecular cloning and expression of the Escherichia coli dimethyl sulfoxide reductase operon. J Bacteriol 1988; 170: 1511–8.10.1128/jb.170.4.1511-1518.1988Search in Google Scholar PubMed PubMed Central
40. Heffron K, Léger C, Rothery RA, Weiner JH, Armstrong FA. Determination of an optimal potential window for catalysis by E. coli dimethyl sulfoxide reductase and hypothesis on the role of Mo(V) in the reaction pathway. Biochemistry 2001; 40: 3117–26.10.1021/bi002452uSearch in Google Scholar PubMed
41. Tang H, Rothery RA, Voss JE, Weiner JH. Correct assembly of iron-sulfur cluster FS0 into Escherichia coli dimethyl sulfoxide reductase (DmsABC) is a prerequisite for molybdenum cofactor insertion. J Biol Chem 2011; 286: 15147–54.10.1074/jbc.M110.213306Search in Google Scholar PubMed PubMed Central
42. Sambasivarao D, Scraba DG, Trieber C, Weiner JH. Organization of dimethyl sulfoxide reductase in the plasma membrane of Escherichia coli. J Bacteriol 1990; 172: 5938–48.10.1128/jb.172.10.5938-5948.1990Search in Google Scholar PubMed PubMed Central
43. Geijer P, Weiner JH. Glutamate 87 is important for menaquinol binding in DmsC of the DMSO reductase (DmsABC) from Escherichia coli. Biochim Biophys Acta Biomembr 2004; 1660: 66–74.10.1016/j.bbamem.2003.10.016Search in Google Scholar PubMed
44. Sambasivarao D, Dawson HA, Zhang G, Shaw G, Hu J, Weiner JH. Investigation of Escherichia coli dimethyl sulfoxide reductase assembly and processing in strains defective for the sec-independent protein translocation system membrane targeting and translocation. J Biol Chem 2001; 276: 20167–74.10.1074/jbc.M010369200Search in Google Scholar PubMed
45. Turner RJ, Busaan JL, Lee JH, Michalak M, Weiner JH. Expression and epitope tagging of the membrane anchor subunit (DmsC) of Escherichia coli dimethyl sulfoxide reductase. Protein Eng 1997; 10: 285–90.10.1093/protein/10.3.285Search in Google Scholar PubMed
46. Winstone TL, Workentine ML, Sarfo KJ, Binding AJ, Haslam BD, Turner RJ. Physical nature of signal peptide binding to DmsD. Arch Biochem Biophys 2006; 455: 89–97.10.1016/j.abb.2006.08.009Search in Google Scholar PubMed
47. Winstone TML, Tran VA, Turner RJ. The hydrophobic region of the DmsA twin-arginine leader peptide determines specificity with chaperone DmsD. Biochemistry 2013; 52: 7532–41.10.1021/bi4009374Search in Google Scholar PubMed PubMed Central
48. Vergnes A, Gouffi-Belhabich K, Blasco F, Giordano G, Magalon A. Involvement of the molybdenum cofactor biosynthetic machinery in the maturation of the Escherichia coli nitrate reductase A. J Biol Chem 2004; 279: 41398–403.10.1074/jbc.M407087200Search in Google Scholar PubMed
49. Genest O, Neumann M, Seduk F, Stöcklein W, Méjean V, Leimkühler S, Iobbi-Nivol C. Dedicated metallochaperone connects apoenzyme and molybdenum cofactor biosynthesis components. J Biol Chem 2008; 283: 21433–40.10.1074/jbc.M802954200Search in Google Scholar PubMed
50. Genest O, Seduk F, Ilbert M, Méjean V, Iobbi-Nivol C. Signal peptide protection by specific chaperone. Biochem Biophys Res Commun 2006; 339: 991–5.10.1016/j.bbrc.2005.11.107Search in Google Scholar PubMed
51. Kostecki JS, Li H, Turner RJ, DeLisa MP. Visualizing interactions along the Escherichia coli twin-arginine translocation pathway using protein fragment complementation. PLoS One 2010; 5: 21–3.10.1371/journal.pone.0009225Search in Google Scholar PubMed PubMed Central
52. Papish AL, Ladner CL, Turner RJ. The twin-arginine leader-binding protein, DmsD, interacts with the TatB and TatC subunits of the Escherichia coli twin-arginine translocase. J Biol Chem 2003; 278: 32501–6.10.1074/jbc.M301076200Search in Google Scholar PubMed
53. Genevaux P, Georgopoulos C, Kelley WL. The Hsp70 chaperone machines of Escherichia coli: a paradigm for the repartition of chaperone functions. Mol Microbiol 2007; 66: 840–57.10.1111/j.1365-2958.2007.05961.xSearch in Google Scholar PubMed
54. Chan CS, Song X, Qazi SJS, Setiaputra D, Yip CK, Chao TC, Turner RJ. Unusual pairing between assistants: interaction of the twin-arginine system-specific chaperone DmsD with the chaperonin GroEL. Biochem Biophys Res Commun 2015; 456: 841–6.10.1016/j.bbrc.2014.12.046Search in Google Scholar PubMed
55. Marchenkov VV, Semisotnov GV. GroEL-assisted protein folding: does it occur within the chaperonin inner cavity? Int J Mol Sci 2009; 10: 2066–83.10.3390/ijms10052066Search in Google Scholar PubMed PubMed Central
56. Houry WA, Frishman D, Eckerskorn C, Lottspeich F, Hartl FU. Identification of in vivo substrates of the chaperonin GroEL. Nature 1999; 402: 147–54.10.1038/45977Search in Google Scholar PubMed
57. Solomon PS, Shaw AL, Lane I, Hanson GR, Palmer T, McEwan AG. Characterization of a molybdenum cofactor biosynthetic gene cluster in Rhodobacter capsulatus which is specific for the biogenesis of dimethylsulfoxide reductase. Microbiology 1999; 145: 1421–9.10.1099/13500872-145-6-1421Search in Google Scholar PubMed
58. Tang H, Rothery RA, Weiner JH. A variant conferring cofactor-dependent assembly of Escherichia coli dimethylsulfoxide reductase. Biochim Biophys Acta Bioenerg 2013; 1827: 730–7.10.1016/j.bbabio.2013.02.009Search in Google Scholar PubMed
59. Sambasivarao D, Turner RJ, Bilous PT, Rothery RA, Shaw G, Weiner JH. Differential effects of a molybdopterin synthase sulfurylase (moeB) mutation on Escherichia coli molybdoenzyme maturation. Biochem Cell Biol 2001; 80: 435–43.10.1139/o02-131Search in Google Scholar PubMed
60. Leimkühler S, Wuebbens MM, Rajagopalan KV. The history of the discovery of the molybdenum cofactor and novel aspects of its biosynthesis in bacteria. Coord Chem Rev 2011; 255: 1129–44.10.1016/j.ccr.2010.12.003Search in Google Scholar PubMed PubMed Central
61. Leimkühler S, Rajagopalan KV. In vitro incorporation of nascent molybdenum cofactor into human sulfite oxidase. J Biol Chem 2001; 276: 1837–44.10.1074/jbc.M007304200Search in Google Scholar PubMed
62. Leimkühler S, Wuebbens MM, Rajagopalan KV. Characterization of Escherichia coli MoeB and its involvement in the activation of molybdopterin synthase for the biosynthesis of the molybdenum cofactor. J Biol Chem 2001; 276: 34695–701.10.1074/jbc.M102787200Search in Google Scholar PubMed
63. Rajagopalan KV. Biosynthesis and processing of the molybdenum cofactors. Biochem Soc Trans 1997; 3: 757–61.10.1042/bst0250757Search in Google Scholar PubMed
64. Pandey A, Pain J, Ghosh AK, Dancis A, Pain D. Fe-S cluster biogenesis in isolated mammalian mitochondria: coordinated use of persulfide sulfur and iron and requirements for GTP, NADH, and ATP. J Biol Chem 2015; 290: 640–57.10.1074/jbc.M114.610402Search in Google Scholar PubMed PubMed Central
65. Lee JH, Yeo WS, Roe JH. Induction of the sufA operon encoding Fe-S assembly proteins by superoxide generators and hydrogen peroxide: Involvement of OxyR, IHF and an unidentified oxidant-responsive factor. Mol Microbiol 2004 ;51: 1745–55.10.1111/j.1365-2958.2003.03946.xSearch in Google Scholar PubMed
66. Jang S, Imlay JA. Hydrogen peroxide inactivates the Escherichia coli Isc iron-sulphur assembly system, and OxyR induces the Suf system to compensate. Mol Microbiol 2010; 78: 1448–67.10.1111/j.1365-2958.2010.07418.xSearch in Google Scholar PubMed PubMed Central
67. Adinolfi S, Iannuzzi C, Prischi F, Pastore C, Iametti S, Martin SR, Bonomi F, Pastore A. Bacterial frataxin CyaY is the gatekeeper of iron-sulfur cluster formation catalyzed by IscS. Nat Struct Mol Biol 2009; 16: 390–6.10.1038/nsmb.1579Search in Google Scholar PubMed
68. Velayudhan J, Castor M, Richardson A, Main-hester KL, Fang FC. The role of ferritins in the physiology of Salmonella enterica sv . Typhimurium: a unique role for ferritin B in iron-sulphur cluster repair and virulence. Mol Microbiol 2007; 63: 1495–507.10.1111/j.1365-2958.2007.05600.xSearch in Google Scholar PubMed
69. Angelini S, Gerez C, Choudens SO De, Sanakis Y, Fontecave M, Barras F, Py B. NfuA, a new factor required for maturing Fe/S proteins in Escherichia coli under oxidative stress and iron starvation conditions. J Biol Chem 2008; 283: 14084–91.10.1074/jbc.M709405200Search in Google Scholar PubMed
70. Yeung N, Gold B, Liu NL, Prathapam R, Sterling HJ, Willams ER, Butland, G. The E. coli monothiol glutaredoxin GrxD forms homodimeric and heterodimeric FeS cluster containing complexes. Biochemistry 2011; 50: 8957–69.10.1021/bi2008883Search in Google Scholar PubMed PubMed Central
71. Boyd JM, Sondelski JL, Downs DM. Bacterial apbC protein has two biochemical activities that are required for in vivo function. J Biol Chem 2009; 284: 110–8.10.1074/jbc.M807003200Search in Google Scholar
72. Vinella D, Loiseau L, Choudens SO De, Fontecave M, Barras F. In vivo [ Fe-S ] cluster acquisition by IscR and NsrR, two stress regulators in Escherichia coli 2013; 87: 493–508.10.1111/mmi.12135Search in Google Scholar
73. Justino MC, Almeida CC, Teixeira M, Saraiva LM. Escherichia coli Di-iron YtfE protein is necessary for the repair of stress-damaged iron-sulfur clusters. J Biol Chem 2007; 282: 10352–9.10.1074/jbc.M610656200Search in Google Scholar
74. Bitto E, Bingman CA, Bittova L, Kondrashov DA, Bannen RM, Fox BG, Markley JL, Phillips GN Jr. Structure of human J-type co-chaperone HscB reveals a tetracysteine metal-binding domain. J Biol Chem 2008; 283: 30184–92.10.1074/jbc.M804746200Search in Google Scholar
75. Uhrigshardt H, Singh A, Kovtunovych G, Ghosh M, Rouault TA. Characterization of the human HSC20, an unusual DnaJ type III protein, involved in iron-sulfur cluster biogenesis. Hum Mol Genet 2010; 19: 3816–34.10.1093/hmg/ddq301Search in Google Scholar
76. Wuebbens MM, Rajagopalan KV. Structural characterization of a molybdopterin precursor. J Biol Chem 1993; 268: 13493–8.10.1016/S0021-9258(19)38676-4Search in Google Scholar
77. Hanzelmann C, Weimann I, Feller J, Žak Z. Synthesis and characterization of the new copper indium phosphate Cu/n 8/n In/n 8/n P/n 4/n O/n 30. Zeitschr Anorg Allg Chem 2014; 640: 213–8.10.1002/zaac.201300251Search in Google Scholar
78. Mallee JJ, Salvatore CA, LeBourdelles B, Oliver KR, Longmore J, Koblan KS, Kane SA. Receptor activity-modifying protein 1 determines the species selectivity of non-peptide CGRP receptor antagonists. J Biol Chem 2002; 277: 14294–8.10.1074/jbc.M109661200Search in Google Scholar PubMed
79. Blasco F, Dos Santos JP, Magalon A, Frixon C, Guigliarelli B, Santini CL, Giordano G. NarJ is a specific chaperone required for molybdenum cofactor assembly in nitrate reductase A of Escherichia coli. Mol Microbiol 1998; 28: 435–47.10.1046/j.1365-2958.1998.00795.xSearch in Google Scholar PubMed
80. Mori H, Cline K. A twin arginine signal peptide and the pH gradient trigger reversible assembly of the thylakoid ∆pH/Tat translocase. J Cell Biol 2002; 157: 205–10.10.1083/jcb.200202048Search in Google Scholar PubMed PubMed Central
81. Chen S, Buchanan G, Greene NP, Lea SM, Palmer T, Tarry MJ, Schäfer E, Saibil HR, Berks BC. Structural analysis of substrate binding by the TatBC component of the twin-arginine protein transport system 2009; 106: 13284–9.10.1073/pnas.0901566106Search in Google Scholar
82. Chaddock a M, Mant a, Karnauchov I, Brink S, Herrmann RG, Klösgen RB, Robinson C. A new type of signal peptide: central role of a twin-arginine motif in transfer signals for the delta pH-dependent thylakoidal protein translocase. EMBO J 1995; 14: 2715–22.10.1002/j.1460-2075.1995.tb07272.xSearch in Google Scholar
83. Panahandeh S, Maurer C, Moser M, DeLisa MP, Müller M. Following the path of a twin-arginine precursor along the TatABC translocase of Escherichia coli. J Biol Chem 2008; 283: 33267–75.10.1074/jbc.M804225200Search in Google Scholar
84. Brehmer T, Kerth A, Graubner W, Malesevic M, Hou B, Brüser T, Blume A. Negatively charged phospholipids trigger the interaction of a bacterial tat substrate precursor protein with lipid monolayers. Langmuir 2012; 28: 3534–41.10.1021/la204473tSearch in Google Scholar
85. Bageshwar UK, Whitaker N, Liang FC, Musser SM. Interconvertibility of lipid- and translocon-bound forms of the bacterial Tat precursor pre-Sufl. Mol Microbiol 2009; 74: 209–26.10.1111/j.1365-2958.2009.06862.xSearch in Google Scholar
86. Alcock F, Baker MAB, Greene NP, Palmer T, Wallace MI, Berks BC. Live cell imaging shows reversible assembly of the TatA component of the twin-arginine protein transport system. Proc Natl Acad Sci USA 2013; 110: E3650–9.10.1073/pnas.1306738110Search in Google Scholar
87. Dabney-Smith C, Mori H, Cline K. Oligomers of Tha4 organize at the thylakoid Tat translocase during protein transport. J Biol Chem 2006; 281: 5476–83.10.1074/jbc.M512453200Search in Google Scholar
88. De Leeuw E, Granjon T, Porcelli I, Alami M, Carr SB, Müller M, Sargent F, Palmer T, Berks BC. Oligomeric properties and signal peptide binding by Escherichia coli Tat protein transport complexes. J Mol Biol 2002; 322: 1135–46.10.1016/S0022-2836(02)00820-3Search in Google Scholar
89. Richter S, Brüser T. Targeting of unfolded PhoA to the TAT translocon of Escherichia coli. J Biol Chem 2005; 280: 42723–30.10.1074/jbc.M509570200Search in Google Scholar PubMed
90. Fröbel J, Rose P, Müller M. Early contacts between substrate proteins and TatA translocase component in twin-arginine translocation. J Biol Chem 2011; 286: 43679–89.10.1074/jbc.M111.292565Search in Google Scholar PubMed PubMed Central
91. Gohlke U, Pullan L, McDevitt CA, Porcelli I, de Leeuw E, Palmer T, Saibil HR, Berks BC. The TatA component of the twin-arginine protein transport system forms channel complexes of variable diameter. Proc Natl Acad Sci USA 2005; 102: 10482–6.10.1073/pnas.0503558102Search in Google Scholar PubMed PubMed Central
92. Bageshwar UK, Musser SM. Two electrical potential-dependent steps are required for transport by the Escherichia coli Tat machinery. J Cell Biol 2007; 179: 87–99.10.1083/jcb.200702082Search in Google Scholar PubMed PubMed Central
93. Gérard F, Cline K. Efficient twin arginine translocation (Tat) pathway transport of a precursor protein covalently anchored to its initial cpTatC binding site. J Biol Chem 2006; 281: 6130–5.10.1074/jbc.M512733200Search in Google Scholar PubMed
94. McDevitt CA, Buchanan G, Sargent F, Palmer T, Berks BC. Subunit composition and in vivo substrate-binding characteristics of Escherichia coli Tat protein complexes expressed at native levels. FEBS J 2006; 273: 5656–68.10.1111/j.1742-4658.2006.05554.xSearch in Google Scholar PubMed
95. Alcock F, Stansfeld PJ, Basit H, Habersetzer J, Baker MAB, Palmer T, Wallace MI, Berks BC. Assembling the Tat protein translocase. eLife 2016; 5: e20718.10.7554/eLife.20718Search in Google Scholar PubMed PubMed Central
96. Lüke I, Handford JI, Palmer T, Sargent F. Proteolytic processing of Escherichia coli twin-arginine signal peptides by LepB. Arch Microbiol 2009; 191: 919–25.10.1007/s00203-009-0516-5Search in Google Scholar PubMed
97. Sambasivarao D, Turner RJ, Simala-Grant JL, Shaw G, Hu J, Weiner JH. Multiple roles for the twin arginine leader sequence of dimethyl sulfoxide reductase of Escherichia coli. J Biol Chem 2000; 275: 22526–31.10.1074/jbc.M909289199Search in Google Scholar PubMed
98. Ren C, Patel R, Robinson C. Exclusively membrane-inserted state of an uncleavable Tat precursor protein suggests lateral transfer into the bilayer from the translocon 2013; 280: 3354–64.10.1111/febs.12327Search in Google Scholar PubMed
99. Engelman DM, Steitz TA. The spontaneous insertion of proteins into and across membranes: the helical hairpin hypothesis. Cell 1981; 23: 411–22.10.1142/9789811215865_0012Search in Google Scholar
100. Kudva R, Denks K, Kuhn P, Vogt A, Müller M, Koch HG. Protein translocation across the inner membrane of Gram-negative bacteria: the Sec and Tat dependent protein transport pathways. Res Microbiol 2013; 164: 505–34.10.1016/j.resmic.2013.03.016Search in Google Scholar PubMed
101. Gidalevitz T, Prahlad V, Morimoto RI. The stress of protein misfolding: from single cells to multicellular organisms. Cold Spring Harb Perspect Biol 2011; 3: a009704.10.1101/cshperspect.a009704Search in Google Scholar PubMed PubMed Central
102. Taylor RC, Dillin A. Aging as an event of proteostasis collapse. Cold Spring Harb Perspect Biol 2011; 3: 1–17.10.1101/cshperspect.a004440Search in Google Scholar PubMed PubMed Central
103. Saibil H. Chaperone machines for protein folding, unfolding and disaggregation. Nat Rev Mol Cell Biol 2013; 14: 630–42.10.1038/nrm3658Search in Google Scholar PubMed PubMed Central
104. Stull F, Koldewey P, Humes JR, Radford SE, Bardwell JCA. Substrate protein folds while it is bound to the ATP-independent chaperone Spy. Nat Struct Mol Biol 2015; 23: 53–8.10.1038/nsmb.3133Search in Google Scholar PubMed PubMed Central
105. Inobe T, Matouschek A. Protein targeting to ATP-dependent proteases. Curr Opin Struct Biol 2008; 18: 43–51.10.1016/j.sbi.2007.12.014Search in Google Scholar PubMed PubMed Central
106. Van Melderen L, Aertsen A. Regulation and quality control by Lon-dependent proteolysis. Res Microbiol 2009; 160: 645–51.10.1016/j.resmic.2009.08.021Search in Google Scholar PubMed
107. Redelberger D, Genest O, Arabet D, Méjean V, Ilbert M, Iobbi-nivol C. Quality control of a molybdoenzyme by the Lon protease. FEBS Lett 2013; 587: 3935–42.10.1016/j.febslet.2013.10.045Search in Google Scholar PubMed
108. Genest O, Méjean V, Iobbi-Nivol C. Multiple roles of TorD-like chaperones in the biogenesis of molybdoenzymes. FEMS Microbiol Lett 2009; 297: 1–9.10.1111/j.1574-6968.2009.01660.xSearch in Google Scholar PubMed
109. Zhang Y, Gladyshev VN. Molybdoproteomes and evolution of molybdenum utilization. J Mol Biol 2008; 379: 881–99.10.1016/j.jmb.2008.03.051Search in Google Scholar PubMed PubMed Central
110. Carpenter LJ, Archer SD, Beale R. Ocean-atmosphere trace gas exchange. R Soc Chem 2012; 41: 6473–506.10.1039/c2cs35121hSearch in Google Scholar PubMed
111. Rajagopalan S, Meng X-P, Ramasamy S, Harrison DG, Gallis ZS. Reactive oxygen species produced by macrophage-derived foam cells regulate the activity of vascular matrix metalloproteinases in vitro. J Clin Invest 1996; 98: 2572–9.10.1172/JCI119076Search in Google Scholar
112. Iobbi-Nivol C, Leimkühler S. Molybdenum enzymes, their maturation and molybdenum cofactor biosynthesis in Escherichia coli. Biochim Biophys Acta 2013; 1827: 1086–101.10.1016/j.bbabio.2012.11.007Search in Google Scholar
113. Hänzelmann P, Schindelin H. Crystal structure of the S-adenosylmethionine-dependent enzyme MoaA and its implications for molybdenum cofactor deficiency in humans. Proc Natl Acad Sci USA 2004; 101: 12870–5.10.1073/pnas.0404624101Search in Google Scholar
114. Reschke S, Sigfridsson KGV, Kaufmann P, Leidel N, Horn S, Gast K, Schulzke C, Haumann M, Leimkühler S. Identification of a bis-molybdopterin intermediate in molybdenum cofactor biosynthesis in Escherichia coli. J Biol Chem 2013; 288: 29736–45.10.1074/jbc.M113.497453Search in Google Scholar
115. Vergnes A, Pommier J, Toci R, Blasco F, Giordano G, Magalon A. NarJ chaperone binds on two distinct sites of the aponitrate reductase of Escherichia coli to coordinate molybdenum cofactor insertion and assembly. J Biol Chem 2006; 281: 2170–6.10.1074/jbc.M505902200Search in Google Scholar
116. Lanciano P, Vergnes A, Grimaldi S, Guigliarelli B, Magalon A. Biogenesis of a respiratory complex is orchestrated by a single accessory protein. J Biol Chem 2007; 282: 17468–74.10.1074/jbc.M700994200Search in Google Scholar
117. Tranier S, Iobbi-Nivol C, Birck C, Ilbert M, Mortier-Barrière I, Méjean V, Samama J-P. A novel protein fold and extreme domain swapping in the dimeric TorD chaperone from Shewanella massilia. Structure 2003; 11: 165–74.10.1016/S0969-2126(03)00008-XSearch in Google Scholar
118. Hatzixanthis K, Clarke TA, Oubrie A, Richardson DJ, Turner RJ, Sargent F. Signal peptide-chaperone interactions on the twin-arginine protein transport pathway. Proc Natl Acad Sci USA 2005; 102: 8460–5.10.1073/pnas.0500737102Search in Google Scholar PubMed PubMed Central
119. Cherak SJ, Turner RJ. Influence of GTP on system specific chaperone – Twin arginine signal peptide interaction. Biochem Biophys Res Commun 2015; 465: 753–7.10.1016/j.bbrc.2015.08.079Search in Google Scholar PubMed
120. Wickner W, Driessen a J, Hartl FU. The enzymology of protein translocation across the Escherichia coli plasma membrane. Annu Rev Biochem 1991; 60: 101–24.10.1146/annurev.bi.60.070191.000533Search in Google Scholar PubMed
121. Driessen AJM, Manting EH, van der Does C. The structural basis of protein targeting and translocation in bacteria. Nat Struct Biol 2001; 8: 492–8.10.1038/88549Search in Google Scholar PubMed
122. Robinson C, Bolhuis A. Tat-dependent protein targeting in prokaryotes and chloroplasts. Biochim Biophys Acta Mol Cell Res 2004; 1694: 135–47.10.1016/j.bbamcr.2004.03.010Search in Google Scholar PubMed
123. De Keyzer J, Van Der Does C, Driessen AJM. The bacterial translocase: A dynamic protein channel complex. Cell Mol Life Sci 2003; 60: 2034–52.10.1007/s00018-003-3006-ySearch in Google Scholar PubMed
124. Rose RW, Brüser T, Kissinger JC, Pohlschröder M. Adaptation of protein secretion to extremely high-salt conditions by extensive use of the twin-arginine translocation pathway 2002; 45: 943–50.10.1046/j.1365-2958.2002.03090.xSearch in Google Scholar PubMed
125. Lindenstrauß U, Brüser T. Conservation and variation between Rhodobacter capsulatus and Escherichia coli Tat systems. J Bacteriol 2006; 188: 7807–14.10.1128/JB.01139-06Search in Google Scholar PubMed PubMed Central
126. Widdick DA, Dilks K, Chandra G, Bottrill A, Naldrett M, Pohlschro M, Palmer T. The twin-arginine translocation pathway is a major route of protein export in Streptomyces coelicolor. Proc Natl Acad Sci USA 2006; 103: 17927–32.10.1073/pnas.0607025103Search in Google Scholar PubMed PubMed Central
127. Osborne AR, Rapoport TA, van den Berg B. Protein translocation by the Sec61/Secy channel. Annu Rev Cell Dev Biol 2005; 21: 529–50.10.1146/annurev.cellbio.21.012704.133214Search in Google Scholar PubMed
128. Van Der Wolk JPW, De Wit JG, Driessen AJM. The catalytic cycle of the Escherichia coli SecA ATPase comprises two distinct preprotein translocation events. EMBO J 1997; 16: 7297–304.10.1093/emboj/16.24.7297Search in Google Scholar PubMed PubMed Central
129. Driessen AJM. Precursor protein translocation by the Escherichia coli translocase is directed by the protonmotive force. EMBO J 1992; 11: 847–53.10.1002/j.1460-2075.1992.tb05122.xSearch in Google Scholar PubMed PubMed Central
130. Driessen AJM, Wickner W. Proton transfer is rate-limiting for translocation of precursor proteins by the Escherichia coli translocase. Proc Natl Acad Sci USA 1991; 88: 2471–5.10.1073/pnas.88.6.2471Search in Google Scholar PubMed PubMed Central
131. Schiebel E, Driessen AJM, Hartl F-U, Wickner W. Delta muH+ and ATP function at different steps of the catalytic cycle of preprotein translocase. Cell 1991; 64: 927–39.10.1016/0092-8674(91)90317-RSearch in Google Scholar
132. Klösgen RB, Brock IW, Herrmann RG, Robinson C. Proton gradient-driven import of the 16 kDa oxygen-evolving complex protein as the full precursor protein by isolated thylakoids. Plant Mol Biol 1992; 18: 1031–4.10.1007/BF00019226Search in Google Scholar
133. Cline K, Ettingerf WF, Thegll M. Protein-specific energy requirements for protein transport across or into thylakoid membranes 1992; 267: 2688–96.10.1016/S0021-9258(18)45935-2Search in Google Scholar
134. Robinson C, Klösgen RB, Herrmann RG, Shackleton JB. Protein translocation across the thylakoid membrane – a tale of two mechanisms 1993; 325: 67–9.10.1016/0014-5793(93)81415-VSearch in Google Scholar
135. Cherak SJ, Turner RJ. Exploring GTP control of bacterial respiratory system specific chaperone interaction with its substrate proteins twin-arginine signal peptide. FASEB J 2016; 30: 810.1.Search in Google Scholar
136. Sarfo KJ, Winstone TL, Papish AL, Howell JM, Kadir H, Vogel HJ, Turner RJ. Folding forms of Escherichia coli DmsD, a twin-arginine leader binding protein. Biochem Biophys Res Commun 2004; 315: 397–403.10.1016/j.bbrc.2004.01.070Search in Google Scholar PubMed
137. Fincher V, Dabney-Smith C, Cline K. Functional assembly of thylakoid ∆pH-dependent/Tat protein transport pathway components in vitro. Eur J Biochem 2003; 270: 4930–41.10.1046/j.1432-1033.2003.03894.xSearch in Google Scholar PubMed
138. Babu M, Arnold R, Bundalovic-Torma C, Gagarinova A, Wong KS, Kumar A, Stewart G, Samanfar B, Aoki H, Wagih O, Vlasblom J, Phanse S, Lad K, Yeou Hsiung Yu A, Graham C, Jin K, Brown E, Golshani A, Kim P, Moreno-Hagelsieb G, Greenblatt J, Houry WA, Parkinson J, Emili A. Quantitative genome-wide genetic interaction screens reveal global epistatic relationships of protein complexes in Escherichia coli. PLoS Genet 2014; 10: e1004120.10.1371/journal.pgen.1004120Search in Google Scholar PubMed PubMed Central
139. Gagarinova A, Stewart G, Samanfar B, Phanse S, White CA, Aoki H, Deineko V, Beloglazova N, Yakunin AF, Golshani A, Brown ED, Babu M, Emili A. Systematic genetic screens reveal the dynamic global functional organization of the bacterial translation machinery. Cell Rep 2016; 17: 904–16.10.1016/j.celrep.2016.09.040Search in Google Scholar PubMed
©2017 Walter de Gruyter GmbH, Berlin/Boston
Articles in the same Issue
- Frontmatter
- Reviews
- Breaking in and busting out: cell-penetrating peptides and the endosomal escape problem
- Protein kinase C-α and the regulation of diverse cell responses
- Assembly pathway of a bacterial complex iron sulfur molybdoenzyme
- Biotechnology of siderophores in high-impact scientific fields
- Mechanisms of tail resorption during anuran metamorphosis
- Breastfeeding and the gut-brain axis: is there a role for melatonin?
- Short Conceptual Overview
- Role of TRAF3 in neurological and cardiovascular diseases: an overview of recent studies
Articles in the same Issue
- Frontmatter
- Reviews
- Breaking in and busting out: cell-penetrating peptides and the endosomal escape problem
- Protein kinase C-α and the regulation of diverse cell responses
- Assembly pathway of a bacterial complex iron sulfur molybdoenzyme
- Biotechnology of siderophores in high-impact scientific fields
- Mechanisms of tail resorption during anuran metamorphosis
- Breastfeeding and the gut-brain axis: is there a role for melatonin?
- Short Conceptual Overview
- Role of TRAF3 in neurological and cardiovascular diseases: an overview of recent studies