3.D.1 The H+ or Na+-translocating NADH Dehydrogenase (NDH) Family

NADH:ubiquinone oxidoreductases type I (NDH-Is) of bacteria and of eukaryotic mitochondria and chloroplasts couple electron transfer to the electrogenic transport of protons (Brandt, 2006) or Na+ (Gemperli, et al., 2007). The subunit, NuoL, is related to Na+/ H+antiporters of 2.A.63.1.1 (PhaA and PhaD). NuoL has been shown to transport Na+ and K+ independently of other Nuo subunits (Gemperli et al., 2007). These protein complexes are multi-subunit complexes with 13 or 14 recognized subunits in the E. coli or P. denitrificans complex and about 30-50 distinct subunits in the complexes of eukaryotic mitochondria.  The structures and mechanisms of respiratory chain complex I have been reviewed (Sazanov 2015).  There is a probable link between redox changes at heme a and modulation of intramolecular proton-transfer rates (Vilhjálmsdóttir et al. 2015).  Respiratory complexes I, III and IV form a complex, and it's archetecture has been studied (Gu et al. 2016). The mechanism of electron flow-coupled proton pumping may involve the semi-quinone free radical (Ohnishi et al. 2018). The mammalian complex I pumps protons within water wires using tightly linked conformational and electrostatic coupling principles (Röpke et al. 2020). The binding pocket for quinones in mitochondrial respiratory complex I is also the site of binding of several chemical inhibitors of complex I activity (Murai 2020). Damage to cerebral mitochondria, particularly opening of the mitochondrial permeability transition pore (MPTP), provides a key mechanism of ischemic brain injury; therefore, modulation of MPTP may be a target for a neuroprotective strategy in ischemic brain pathologies. Brain tissue necrosis can be prevented by the biguanides, metformin and phenformin, as well as Complex I inhibitors rotenone and amobarbital and by the MPTP inhibitor cyclosporine A (Skemiene et al. 2020). The structure of the Escherichia coli respiratory complex I, reconstituted into lipid nanodiscs, revealed an uncoupled conformation (Kolata and Efremov 2021). Transcriptomic and metabolomic analyses revealed that the potent antifungal pyrylium salt inhibits mitochondrial complex I in Candida albicans (Lv et al. 2023).  Quinone catalysis modulates proton transfer reactions in the membrane domain of respiratory complex I (Kim et al. 2023). 

Complex I contains a ubiquinone binding pocket at the interface of the 49-kDa and PSST subunits. Close to iron-sulfur cluster N2, the proposed immediate electron donor for ubiquinone, a highly conserved tyrosine constitutes a critical element of the quinone reduction site. A possible quinone exchange path leads from cluster N2 to the N-terminal beta-sheet of the 49-kDa subunit (Tocilescu et al., 2010). All 45 subunits of the bovine NDHI have been sequenced (Cardol et al., 2004; Gabaldon et al., 2005). Each complex contains noncovalently bound FMN, coenzyme Q and several iron-sulfur centers. The bacterial NDHs have 8-9 iron-sulfur centers. Electrons pass from NADH to FMN, to the various iron-sulfur centers, and finally to ubiquinone or another quinone. NDHI, a complex I in mitochondria, forms a 'supercomplex' with complex III (3.D.3) and complex IV (3.D.4) (Schafer et al., 2007). The mechanistic details are discussed by Hirst (2010). The interaction of mitochondrial complex I with complex III is discussed by Dudkina et al. (2009). The x-ray structures of the functional modules of the 40 subunit Yarrowia lipolytica Complex I have been examined, and the basis for conformational coupling has been presented (Hunte et al., 2010).  Moreover, a molecular mechanism of the proton pump has been proposed (Verkhovskaya and Bloch 2013). The cardiac mitochondria respirasome is adapted for the β-oxidation of fatty acids (Panov 2024).

The positions of all iron-sulfur clusters relative to the membrane arm were determined in the complete enzyme complex. The ubiquinone reduction site resides close to 30 Å above the membrane domain. The arrangement of functional modules suggests conformational coupling of redox chemistry with proton pumping and essentially excludes direct mechanisms. (Hunte et al., 2010) suggest that a ~60 Å long helical transmission-element is critical for transducing conformational energy to proton-pumping elements in the distal module of the membrane arm.  The x-ray structures of various complexes have been solved, and a coupling mechanism involving long range conformational changes has been proposed (Sazanov et al. 2013).  Thjere are several assembly proteins for complex 1, and NDUFAF6 (333 aas; Q3300K2) plays important roles in complex I assembly and mitochondrial disease (Sung et al. 2024).

The mammalian enzyme complexes contain several water soluble peripheral membrane proteins which are anchored to the integral membrane constituents. The seven most hydrophobic proteins of the complex are encoded within mammalian or fungal mitochondrial genomes while the remainder are nuclearly encoded (Brandt, 2006; Gray et al., 2004). All thirteen of the E. coli proteins, which comprise NADH dehydrogenase I and are encoded within the nuo operon, are homologous to mitochondrial complex I subunits. Fearnley and Walker (1992) have provided multiple alignments and hydropathy plots for many of the NDH complex subunits, but phylogenetic analyses are not included. NDH-I complexes contrast with NDH-II complexes that do not transport H+ or Na+. Nelfinavir, and anticancer drug, induces lipid bilayer stress in cellular organelles that disrupts mitochondrial respiration and transmembrane protein transport via metabolic rewiring and activation of the unfolded protein response (Besse et al. 2021).

Efremov et al. (2010) have determined the structure of the E. coli NDH-I transmembrane structure at 3.9 A resolution. The antiporter-like subunits NuoL/M/N each contains 14 conserved transmembrane (TM) helices. Two of them are discontinuous, but subunit NuoL contains a 110-A long amphipathic α-helix, spanning the entire length of the domain. The structure of the entire complex I from Thermus thermophilus at 4.5 A resolution reveals an L-shaped assembly with 63 TM helices. The architecture of the complex suggests that the conformational changes at the interface of the two main domains drives the long amphipathic α-helix of NuoL in a piston-like motion, tilting nearby discontinuous TM helices, resulting in proton translocation (Efremov et al., 2010). This work has been reviewed and evaluated (Efremov and Sazanov, 2011b). Terminal respiratory oxidases allow microorganisms to adapt to low-oxygen conditions, survive in chemically aggressive environments and acquire antibiotic resistance (Siletsky and Borisov 2021). Three-dimensional structures with atomic resolution of members of all major groups of terminal respiratory oxidases, heme-copper oxidases, and bd-type cytochromes, have been obtained. These groups of enzymes have different origins, but all of them are united by a catalytic reaction of four-electron reduction in oxygen into water which proceeds without the formation and release of potentially dangerous ROS from active sites. The review by Siletsky and Borisov 2021 analyzes structural and functional studies of oxygen reduction intermediates in the active sites of terminal respiratory oxidases, the features of catalytic cycles, and the properties of the active sites of these enzymes.

The NDH family includes complexes in bacteria, archaea and eukaryotic organelles (mitochondria and chloroplasts) that may use different electron donors. Thus, the bacterial and mitochondrial complexes function as NADH dehydrogenases, but some of the archaeal complexes function as F420H2 dehydrogenases (see TC# 3.D.9). The electron donor(s) for the cyanobacterial and plastid complexes is (are) not yet known. Despite the potentially different electron input sites, eleven polypeptide chains constitute the structural framework for proton translocation and quinone binding. At least six of these subunits are also present in a family of membrane-bound multisubunit [NiFe] hydrogenases. One of these hydrogenases (from Paracoccus furiosus) has been shown to couple electron transfer to H+ translocation (Friedrich and Scheide, 2000; Sapra et al., 2003). In E. coli these hydrogenases are hydrogenases 3 and 4 of the formate hydrogen lyase system. They also include the CO-induced hydrogenase of Rhodospirillum rubrum and the Ech hydrogenase of Methanosarcina barkerei. The archaeal hydrogenase transfers reducing equivalents generated by the oxidation of the low potential electron donor ferridoxin to protons, yielding H2, coupled to the generation of a pmf (Sapra et al., 2003). Several archaea have two such enzymes of unknown physiological function.

Campylobacter jejuni encodes 12 of the 14 subunits that make up the respiratory NADH ubiquinone oxidoreductase (complex I). The two nuo genes not present in C. jejuni encode the NADH dehydrogenase, and in their place in the operon are the novel genes designated as Cj1575c (CAL25672) and Cj1574c (CAL35671; 24% identical and 41% similar (e-5) to ORF1 of plasmid pIP501 (3.A.7.14.1)). Mutants were generated in which each of the 12 nuo genes (homologues to known complex I subunits) were disrupted or deleted (Weerakoon and Olson, 2008). Each of the nuo mutants do not grow in amino acid-based media unless supplemented with an alternative respiratory substrate such as formate. Unlike the nuo genes, Cj1574c is an essential gene and could not be disrupted unless an intact copy of the gene was provided at an unrelated site on the chromosome. A nuo deletion mutant can efficiently respire formate but is deficient in α-ketoglutarate respiratory activity when compared to wild type cells. In C. jejuni, α-ketoglutarate respiration is mediated by 2-oxoglutarate:acceptor oxidoreductase. Mutagenesis of this enzyme abolished α-ketoglutarate-dependent O2 uptake and failed to reduce the electron transport chain. The electron acceptor for this 2-oxoglutarate oxidoreductase was determined to be flavodoxin, which was also found to be an essential protein in C. jejuni. A model was presented by Weerakoon and Olson (2008) in which CJ1574 mediates electron flow into the respiratory transport chain from reduced flavodoxin and through complex I to ubiquinone.

Oxidation of NADH by submitochondrial particles (SMPs) from the yeast Yarrowia lipolytica is coupled to protonophore-resistant Na+ uptake, indicating that a redox-driven, primary Na+ pump is operative in the inner mitochondrial membrane. A respiratory NADH dehydrogenase couples NADH-dependent reduction of ubiquinone to Na+ translocation. NADH-driven Na+ transport was sensitive rotenone, a specific inhibitor of complex I (Lin et al., 2008).

E. coli complex I (NADH dehydrogenase) is capable of proton translocation in the same direction to the established Δψ, showing that in the tested conditions, the coupling ion is H+ (Batista and Pereira 2011). Na+ transport in the opposite direction was observed, and although Na+ was not necessary for the catalytic or proton transport activities, its presence increased the latter. H+ was translocated by the P. denitrificans complex I, but in this case, H+ transport was not influenced by Na+, and Na+ transport was not observed. Possibly, the E. coli complex I has two energy coupling sites (one Naindependent and the other Na+ dependent), as observed for the Rhodothermus marinus complex I, whereas the coupling mechanism of the P. denitrificans enzyme is completely Na+ independent. It is also possible that another transporter catalyzes the uptake of Na+.  Complex I energy transduction by proton pumping may not be exclusive to the R. marinus enzyme. The Na+/H+ antiport activity seems not to be a general property of complex I (Batista and Pereira 2011). 

The NADH-driven respiratory Na+ pump in Klebsiella pneumoniae was first reported by Peter Dimroth and co-workers (Steffen and Steuber 2013). The 3D structures of complex I from different organisms support the idea that the mechanism of cation transport by complex I involves conformational changes of the membrane-bound NuoL, NuoM and NuoN subunits. In vitro methods to follow Na+ transport were compared with in vivo approaches to test whether complex I, or its individual NuoL, NuoM or NuoN subunits, extrude Na+ from the cytoplasm to the periplasm of bacterial host cells. A truncated NuoL subunit of the E. coli complex which comprises amino acids 1-369 exhibits Na+ transport activity in vitro. This observation, together with an analysis of putative cation channels in NuoL, suggests that there exists in NuoL at least one continuous pathway for cations lined by amino acid residues from transmembrane segments 3, 4, 5, 7 and 8. Finally, Na+ transport by mitochondrial complex I has been discussed with respect to its possible role in the cycling of Na+ across the inner mitochondrial membrane (Steffen and Steuber 2013).

Three of the conserved, membrane-bound subunits in NADH dehydrogenase are related to each other, and to Mrp sodium-proton antiporters. Structural analysis of two prokaryotic complexes I revealed that the three subunits each contain fourteen transmembrane helices that overlay in structural alignments: the translocation of three protons may be coordinated by a lateral helix connecting them together (Efremov et al., 2010). The architecture of respiratory complex I. Nature 465, 441-447). Birrel & Hirst (2010) showed that in higher metazoans the threefold symmetry is broken by the loss of three helices from subunit ND2.

M. mazei and M. barkeri belong to the group of aceticlastic methanogens and converts acetate into the potent greenhouse gases CO2 and CH4. The aceticlastic respiratory chain involved in methane formation comprises the three transmembrane proteins Ech hydrogenase, F(420) nonreducing hydrogenase and heterodisulfide reductase. All three contribute to the proton motive force. ATP synthesis was observed in a cytoplasm-free vesicular system that was dependent on the oxidation of reduced ferredoxin and the formation of molecular hydrogen (Welte et al., 2010). ATP formation was not observed in a deletion mutant strain. Proton protonophores led to complete inhibition of ATP formation without inhibiting hydrogen production, whereas sodium ion ionophores did not affect ATP formation. Welte et al. (2010) concluded that the Ech hydrogenase complex acts as a primary proton pump in a ferredoxin-dependent electron transport system.

Efremov and Sazanov (2011) have reported the crystal structure of the membrane domain of complex I at 3.0 Å resolution. It includes six subunits, NuoL, NuoM, NuoN, NuoA, NuoJ and NuoK, with 55 transmembrane helices. The fold of the homologous antiporter-like subunits L, M and N is novel, with two inverted structural repeats of five transmembrane helices arranged, unusually, face-to-back. Each repeat includes a discontinuous transmembrane helix and forms half of a channel across the membrane. A network of conserved polar residues connects the two half-channels, completing the proton translocation pathway. Unexpectedly, lysines rather than carboxylate residues act as the main elements of the proton pump in these subunits. The fourth probable proton-translocation channel is at the interface of subunits N, K, J and A. The structure indicates that proton translocation in complex I, uniquely, involves coordinated conformational changes in six symmetrical structural elements.

The architecture of the hydrophobic region of complex I shows multiple proton transporters that are mechanically interlinked. Transduction of conformational changes to drive the transmembrane transporters linked by a 'connecting rod' during the reduction of ubiquinone (Q) can account for two or three of the four protons pumped per NADH oxidized. The remaining proton must be pumped by direct coupling at the Q-binding site. Treberg et al. (2011) proposed direct and indirect coupling mechanisms to account for the pumping of the four protons.

The first crystal structure of the entire, intact complex I (from Thermus thermophilus) at 3.3 A resolution has been reported (Baradaran et al. 2013). The structure of the 536-kDa complex comprises 16 different subunits, with a total of 64 transmembrane helices and 9 iron-sulphur clusters. The core fold of subunit Nqo8 (ND1 in humans) is, unexpectedly, similar to a half-channel of the antiporter-like subunits. Small subunits nearby form a linked second half-channel, which completes the fourth proton-translocation pathway (present in addition to the channels in three antiporter-like subunits). The quinone-binding site is unusually long, narrow and enclosed. The quinone headgroup binds at the deep end of this chamber, near iron-sulphur cluster N2. The chamber is linked to the fourth channel by a 'funnel' of charged residues. The link continues over the entire membrane domain as a flexible central axis of charged and polar residues, and probably has a leading role in the propagation of conformational changes, aided by coupling elements. The structure suggests that a unique, out-of-the-membrane quinone-reaction chamber enables the redox energy to drive concerted long-range conformational changes in the four antiporter-like domains, resulting in translocation of four protons per cycle (Baradaran et al., 2013). 

The x-ray structure of mitochondrial complex I has been solved at a resolution of 3.6 to 3.9 angstroms, describing in detail the central subunits that execute the bioenergetic function (Zickermann et al. 2015). A continuous axis of basic and acidic residues running centrally through the membrane arm connects the ubiquinone reduction site in the hydrophilic arm to four putative proton-pumping units. The binding position for a substrate analogous inhibitor and blockage of the predicted ubiquinone binding site provide a model for the 'deactive' form of the enzyme. The proposed transition into the active form is based on a concerted structural rearrangement at the ubiquinone reduction site, providing support for a two-state stabilization-change mechanism of proton pumping (Zickermann et al. 2015).  The largest membrane-embedded subunits, ND5, ND4, and ND2, share a similar core of 14 TMSs with two repeats of five TMHs (A, TMH4–8; B, TMH9–13) in inverted topology. Each repeat features a discontinuous helix (TMH7a/b, TMH12a/b). Such helices are hallmarks of ion-translocating membrane proteins (Zickermann et al. 2015). Indeed, ND5, ND4, and ND2 are homologous to the Mrp Na+/H+ antiporter family, which suggests a role in proton pumping (Zickermann et al. 2015). ND5 has a C-terminal extension with a lateral helix, >60 Å long, lining ND5, ND4, and ND2 on the concave side of the arm close to the matrix side. The C terminus of ND5 is anchored to ND2 via a V-shaped arrangement of TMSs 16 and 17. The lateral helix was previously identified in the low-resolution analysis of mitochondrial complex I from Y. lipolytica, and is also present in the bacterial and bovine complexes in which its C terminus is anchored by one TMS only.  

Berrisford et al. 2016; have solved the crystal structures of the hydrophilic domain of complex I from Thermus thermophilus, the membrane domain from Escherichia coli and more recently of the intact, entire complex I from T. thermophilus (536 kDa, 16 subunits, 9 iron- sulphur clusters, 64 transmembrane helices). The 95 Å long electron transfer pathway through the enzyme proceeds from the primary electron acceptor flavin mononucleotide through seven conserved Fe-S clusters to the unusual elongated quinone-binding site at the interface with the membrane domain. Four putative proton translocation channels are found in the membrane domain, all linked by the central flexible axis containing charged residues. The redox energy of electron transfer is coupled to proton translocation by the as yet undefined mechanism proposed to involve long-range conformational changes. 

Mitochondrial electron transport chain complexes are organized into supercomplexes responsible for carrying out cellular respiration. Letts et al. 2016 presented the architectures of mammalian (ovine) supercomplexes determined by cryo-electron microscopy. The former group identified two distinct arrangements of supercomplex CICIII2CIV (the respirasome)-a major 'tight' form and a minor 'loose' form (resolved at the resolution of 5.8 A and 6.7 A, respectively), which may represent different stages in supercomplex assembly or disassembly.They also determined an architecture of supercomplex CICIII2 at 7.8 A resolution. All observed density could be attributed to the known 80 subunits of the individual complexes, including 132 transmembrane helices. The individual complexes form tight interactions that vary between the architectures, with complex IV subunit COX7a switching contact from complex III to complex I. The arrangement of active sites within the supercomplex may help control reactive oxygen species production (Letts et al. 2016). Gu et al. 2016 reported the 5.4 Å structure of the 1.7 MDa respirasome from pig heart.  The CIII dimer as well as CIV bind to the same side of the L-shaped CI with their TMSs aligned to form a transmembrane disk. Complexes I, III and IV, comprising the respirasome, have been solved by cryo-EM, showing the architecture of the respirasome with near-atomic detail. ATP synthase occurs as dimers in the inner membrane, which by its curvature is responsible for the folding of the membrane into cristae. This allows for a huge increase in available surface area that makes mitochondria the efficient energy organelles of the eukaryotic cell (Sousa et al. 2018).

Fiedorczuk et al. 2016 identified All 14 conserved core subunits and 31 mitochondrial supernumerary subunits within the L-shaped molecule. The hydrophilic matrix arm includes flavin mononucleotide and 8 iron-sulfur clusters involved in electron transfer, and the membrane arm contains 78 transmembrane helices, mostly contributed by antiporter-like subunits involved in proton translocation. Supernumerary subunits form an interlinked, stabilizing shell around the conserved core. Tightly bound lipids (including cardiolipins) further stabilize interactions between the hydrophobic subunits. Subunits with possible regulatory roles contain additional cofactors, NADPH and two phosphopantetheine molecules, which are probably involved in inter-subunit interactions. The two different conformations of the complex, may be related to the conformationally driven coupling mechanism and to the active-deactive transition of the enzyme.

In Prevotella bryantii, the Na+-translocating NADH:quinone oxidoreductase (NQR) and quinol:fumarate reductase (QFR) interact using menaquinone as electron carrier, catalyzing NADH:fumarate oxidoreduction (Hau et al. 2023). P. bryantii NQR establishes a sodium-motive force, whereas P. bryantii QFR does not contribute to membrane energization. Hau et al. 2023 presented 3D structural models of NQR and QFR from P. bryantii to predict cofactor-binding sites, electron transfer routes and interaction with substrates. Molecular docking revealed the proposed mode of menaquinone binding to the quinone site of subunit NqrB of P. bryantii NQR. A comparison of the 3D model of P. bryantii QFR with experimentally determined structures suggests alternative pathways for transmembrane proton transport in this type of QFR. These findings are relevant for NADH-dependent succinate formation in anaerobic bacteria which operate both NQR and QFR.

 

The generalized transport reactions catalyzed by NDH family members are:

NADH + ubiquinone + 4H+ (in) → NAD+ + ubiquinol + 4H+ (out)



This family belongs to the Multidrug Resistance Protein, Na+ Transporting Mrp (Mrp) Superfamily.

 

References:

Abdrakhmanova, A., V. Zickermann, M. Bostina, M. Radermacher, H. Schägger, S. Kerscher, and U. Brandt. (2004). Subunit composition of mitochondrial complex I from the yeast Yarrowia lipolytica. Biochim. Biophys. Acta. 1658: 148-156.

Bagramyan, K., A. Vassilian, N. Mnatsakanyan, and A. Trchounian. (2001). Participation of hyf-encoded hydrogenase 4 in molecular hydrogen release coupled with proton-potassium exchange in Escherichia coli. Membr Cell Biol 14: 749-763.

Bagramyan, K., N. Mnatsakanyan, A. Poladian, A. Vassilian, and A. Trchounian. (2002). The roles of hydrogenases 3 and 4, and the F0F1-ATPase, in H2 production by Escherichia coli at alkaline and acidic pH. FEBS Lett. 516: 172-178.

Baradaran, R., J.M. Berrisford, G.S. Minhas, and L.A. Sazanov. (2013). Crystal structure of the entire respiratory complex I. Nature 494: 443-448.

Barquera, B., P. Hellwig, W. Zhou, J.E. Morgan, C.C. Häse, K.K. Gosink, M. Nilges, P.J. Bruesehoff, A. Roth, C.R.D. Lancaster, and R.B. Gennis. (2002). Purification and characterization of the recombinant Na+-translocating NADH:quinone oxidoreductase from Vibrio cholerae. Biochemistry 41: 3781-3789.

Batista, A.P. and M.M. Pereira. (2011). Sodium influence on energy transduction by complexes I from Escherichia coli and Paracoccus denitrificans. Biochim. Biophys. Acta. 1807: 286-292.

Berrisford, J.M., R. Baradaran, and L.A. Sazanov. (2016). Structure of bacterial respiratory complex I. Biochim. Biophys. Acta. 1857: 892-901.

Besse, L., A. Besse, S.C. Stolze, A. Sobh, E.A. Zaal, A.J. van der Ham, M. Ruiz, S. Phuyal, L. Büchler, M. Sathianathan, B.I. Florea, J. Borén, M. Ståhlman, J. Huber, A. Bolomsky, H. Ludwig, J.T. Hannich, A. Loguinov, B. Everts, C.R. Berkers, M. Pilon, H. Farhan, C.D. Vulpe, H.S. Overkleeft, and C. Driessen. (2021). Treatment with HIV-Protease Inhibitor Nelfinavir Identifies Membrane Lipid Composition and Fluidity as a Therapeutic Target in Advanced Multiple Myeloma. Cancer Res 81: 4581-4593.

Birrell, J.A. and J. Hirst. (2010). Truncation of subunit ND2 disrupts the threefold symmetry of the antiporter-like subunits in complex I from higher metazoans. FEBS Lett. 584: 4247-4252.

Brandt, U. (1997). Proton-translocation by membrane-bound NADH:ubiquinone-oxidoreductase (complex I) through redox-gated ligand conduction. Biochim Biophys Acta 1318: 79-91.

Brandt, U. (2006). Energy converting NADH:quinone oxidoreductase (complex I). Annu. Rev. Biochem. 75:69-92.

Caporali, L., L. Iommarini, C. La Morgia, A. Olivieri, A. Achilli, A. Maresca, M.L. Valentino, M. Capristo, F. Tagliavini, V. Del Dotto, C. Zanna, R. Liguori, P. Barboni, M. Carbonelli, V. Cocetta, M. Montopoli, A. Martinuzzi, G. Cenacchi, G. De Michele, F. Testa, A. Nesti, F. Simonelli, A.M. Porcelli, A. Torroni, and V. Carelli. (2018). Peculiar combinations of individually non-pathogenic missense mitochondrial DNA variants cause low penetrance Leber''s hereditary optic neuropathy. PLoS Genet 14: e1007210. [Epub: Ahead of Print]

Cardol, P., F. Vanrobaeys, B. Devreese, J. Van Beeumen, R.F. Matagne, and C. Remacle. (2004). Higher plant-like subunit composition of mitochondrial complex I from Chlamydomonas reinhardtii: 31 conserved components among eukaryotes. Biochim. Biophys. Acta. 1658:212-224.

Dudkina NV., Kouril R., Peters K., Braun HP. and Boekema EJ. (2010). Structure and function of mitochondrial supercomplexes. Biochim Biophys Acta. 1797(6-7):664-70.

Efremov, R.G. and L.A. Sazanov. (2011). Respiratory complex I: 'steam engine' of the cell? Curr. Opin. Struct. Biol. 21: 532-540.

Efremov, R.G. and L.A. Sazanov. (2011). Structure of the membrane domain of respiratory complex I. Nature 476: 414-420.

Efremov, R.G., R. Baradaran, and L.A. Sazanov. (2010). The architecture of respiratory complex I. Nature 465: 441-445.

Fearnley, I.M. and J.E. Walker. (1992). Conservation of sequences of subunits of mitochondrial complex I and their relationships with other proteins. Biochim. Biophys. Acta 1140: 105-134.

Fiedorczuk, K., J.A. Letts, G. Degliesposti, K. Kaszuba, M. Skehel, and L.A. Sazanov. (2016). Atomic structure of the entire mammalian mitochondrial complex I. Nature 538: 406-410.

Forzi, L., J. Koch, A.M. Guss, C.G. Radosevich, W.W. Metcalf, and R. Hedderich. (2005). Assignment of the [4Fe-4S] clusters of Ech hydrogenase from Methanosarcina barkeri to individual subunits via the characterization of site-directed mutants. FEBS J. 272: 4741-4753.

Fox, J.D., R.L. Kerby, G.P. Roberts, and P.W. Ludden. (1996b). Characterization of the CO-induced, CO-tolerant hydrogenase from Rhodospirillum rubrum and the gene encoding the large subunit of the enzyme. J. Bacteriol. 178: 1515-1524.

Fox, J.D., Y. He, D. Shelver, G.P. Roberts, and P.W. Ludden. (1996a). Characterization of the region encoding the CO-induced hydrogenase of Rhodospirillum rubrum. J. Bacteriol. 178: 6200-6208.

Friedrich, T. (1998). The NADH:ubiquinone oxidoreductase (complex I) from Escherichia coli. Biochim. Biophys. Acta 1364: 134-146.

Friedrich, T. and D. Scheide. (2000). The respiratory complex I of bacteria, archaea and eukarya and its module common with membrane-bound multisubunit hydrogenases. FEBS Lett. 479: 1-5.

Friedrich, T., B. Brors, P. Hellwig, L. Kintscher, T. Rasmussen, D. Scheide, U. Schulte, W. Mäntele, and H. Weiss. (2000). Characterization of two novel redox groups in the respiratory NADH: ubiquinone oxidoreductase (complex I). Biochim. Biophys. Acta 1459: 305-309.

Gabaldón, T., D. Rainey, and M.A. Huynen. (2005). Tracing the evolution of a large protein complex in the eukaryotes, NADH:ubiquinone oxidoreductase (Complex I). J. Mol. Biol. 348: 857-870.

Gemperli, A.C., C. Schaffitzel, C. Jakob, and J. Steuber. (2007). Transport of Na+ and K (+) by an antiporter-related subunit from the Escherichia coli NADH dehydrogenase I produced in Saccharomyces cerevisiae. Arch. Microbiol. 188: 509-521.

Gemperli, A.C., P. Dimroth, and J. Steuber. (2003). Sodium ion cycling mediates energy coupling between complex I and ATP synthase. Proc. Natl. Acad. Sci. USA 100: 839-844.

Gennis, R.B. and V. Stewart. (1996). Respiration. In: F.C. Neidardt et al. (eds.), Escherichia coli and Salmonella. Cellular and Molecular Biology, Vol. 1, 2nd Ed. Washington, DC: ASM Press, pp. 217-261.

Gray, M.W., B.F. Lang, and G. Burger. (2004). Mitochondria of protists. Annu. Rev. Genet. 38:477-524.

Grigorieff, N. (1999). Structure of the respiratory NADH:ubiquinone oxidoreductase (complex I). Curr. Opin. Struc. Biol. 9: 476-483.

Gu, J., M. Wu, R. Guo, K. Yan, J. Lei, N. Gao, and M. Yang. (2016). The architecture of the mammalian respirasome. Nature 537: 639-643.

Gupta, C., U. Khaniya, C.K. Chan, F. Dehez, M. Shekhar, M.R. Gunner, L. Sazanov, C. Chipot, and A. Singharoy. (2020). Charge Transfer and Chemo-Mechanical Coupling in Respiratory Complex I. J. Am. Chem. Soc. 142: 9220-9230.

Hau, J.L., L. Schleicher, S. Herdan, J. Simon, J. Seifert, G. Fritz, and J. Steuber. (2023). Functionality of the Na-translocating NADH:quinone oxidoreductase and quinol:fumarate reductase from Prevotella bryantii inferred from homology modeling. Arch. Microbiol. 206: 32.

Hedderich, R. and L. Forzi. (2005). Energy-converting [NiFe] hydrogenases: more than just H2 activation. J. Mol. Microbiol. Biotechnol. 10: 92-104.

Hirst, J. (2003). The dichotomy of complex I: a sodium ion pump or a proton pump. Proc. Natl. Acad. Sci. USA 100: 773-775.

Hirst, J. (2010). Towards the molecular mechanism of respiratory complex I. Biochem. J. 425: 327-339.

Hunte, C., V. Zickermann, and U. Brandt. (2010). Functional modules and structural basis of conformational coupling in mitochondrial complex I. Science 329: 448-451.

Ishikawa, N., A. Takabayashi, S. Ishida, Y. Hano, T. Endo, and F. Sato. (2008). NDF6: a thylakoid protein specific to terrestrial plants is essential for activity of chloroplastic NAD(P)H dehydrogenase in Arabidopsis. Plant Cell Physiol. 49: 1066-1073.

Kim, H., P. Saura, M.C. Pöverlein, A.P. Gamiz-Hernandez, and V.R.I. Kaila. (2023). Quinone Catalysis Modulates Proton Transfer Reactions in the Membrane Domain of Respiratory Complex I. J. Am. Chem. Soc. 145: 17075-17086.

Kolata, P. and R.G. Efremov. (2021). Structure of respiratory complex I reconstituted into lipid nanodiscs reveals an uncoupled conformation. Elife 10:.

Krebs, W., J. Steuber, A.C. Gemperli, and P. Dimroth. (1999). Na+ translocation by the NADH: ubiquinone oxidoreductase (complex I) from Klebsiella pneumoniae. Mol. Microbiol. 33: 590-598.

Künkel, A., J.A. Vorholt, R.K. Thauer, and R. Hedderich. (1998). An Escherichia coli hydrogenase-3-type hydrogenase in methanogenic archaea. Eur. J. Biochem. 252: 467-476.

Letts, J.A., K. Fiedorczuk, and L.A. Sazanov. (2016). The architecture of respiratory supercomplexes. Nature 537: 644-648.

Lin, P.C., A. Puhar, and J. Steuber. (2008). NADH oxidation drives respiratory Na+ transport in mitochondria from Yarrowia lipolytica. Arch. Microbiol. 190: 471-480.

Lu, H. and X. Cao. (2008). GRIM-19 is essential for maintenance of mitochondrial membrane potential. Mol. Biol. Cell 19: 1893-1902.

Lv, Q., L. Yan, J. Wang, J. Feng, L. Gao, L. Qiu, W. Chao, Y.L. Qin, and Y. Jiang. (2023). Combined Transcriptome and Metabolome Analysis Reveals That the Potent Antifungal Pyrylium Salt Inhibits Mitochondrial Complex I in Candida albicans. Microbiol Spectr e0320922. [Epub: Ahead of Print]

Masuya, T., S. Uno, M. Murai, and H. Miyoshi. (2021). Pinpoint Dual Chemical Cross-Linking Explores the Structural Dynamics of the Ubiquinone Reaction Site in Mitochondrial Complex I. Biochemistry 60: 813-824.

McDowall, J.S., B.J. Murphy, M. Haumann, T. Palmer, F.A. Armstrong, and F. Sargent. (2014). Bacterial formate hydrogenlyase complex. Proc. Natl. Acad. Sci. USA 111: E3948-3956.

Murai, M. (2020). Exploring the binding pocket of quinone/inhibitors in mitochondrial respiratory complex I by chemical biology approaches. Biosci. Biotechnol. Biochem. 84: 1322-1331.

Nakamaru-Ogiso, E., M.C. Kao, H. Chen, S.C. Sinha, T. Yagi, and T. Ohnishi. (2010). The membrane subunit NuoL(ND5) is involved in the indirect proton pumping mechanism of Escherichia coli complex I. J. Biol. Chem. 285: 39070-39078.

Ohnishi, T., S.T. Ohnishi, and J.C. Salerno. (2018). Five decades of research on mitochondrial NADH-quinone oxidoreductase (complex I). Biol Chem. [Epub: Ahead of Print]

Padavannil, A., A. Murari, S.K. Rhooms, E. Owusu-Ansah, and J.A. Letts. (2023). Resting mitochondrial complex I from adopts a helix-locked state. Elife 12:.

Pandya, B.U., N.A. Takyi, A.R. Vosoughi, E.A. Margolin, and J.A. Micieli. (2024). Novel Mutations in the ND5 Gene Associated With Leber Hereditary Optic Neuropathy. J Neuroophthalmol 44: e227-e229.

Panov, A.V. (2024). The Structure of the Cardiac Mitochondria Respirasome Is Adapted for the β-Oxidation of Fatty Acids. Int J Mol Sci 25:.

Peng, L., H. Yamamoto, and T. Shikanai. (2011). Structure and biogenesis of the chloroplast NAD(P)H dehydrogenase complex. Biochim. Biophys. Acta. 1807: 945-953.

Roca, F.J., L.J. Whitworth, H.A. Prag, M.P. Murphy, and L. Ramakrishnan. (2022). Tumor necrosis factor induces pathogenic mitochondrial ROS in tuberculosis through reverse electron transport. Science 376: eabh2841.

Röpke, M., P. Saura, D. Riepl, M.C. Pöverlein, and V.R.I. Kaila. (2020). Functional Water Wires Catalyze Long-Range Proton Pumping in the Mammalian Respiratory Complex I. J. Am. Chem. Soc. 142: 21758-21766.

Sapra, R., K. Bagramyan, and M.W.W. Adams. (2003). A simple energy-conserving system: proton reduction coupled to proton translocation. Proc. Natl. Acad. Sci. USA 100: 7545-7550.

Sato, M., P.K. Sinha, J. Torres-Bacete, A. Matsuno-Yagi, and T. Yagi. (2013). Energy transducing roles of antiporter-like subunits in Escherichia coli NDH-1 with main focus on subunit NuoN (ND2). J. Biol. Chem. 288: 24705-24716.

Sazanov, L.A. (2015). A giant molecular proton pump: structure and mechanism of respiratory complex I. Nat Rev Mol. Cell Biol. 16: 375-388.

Sazanov, L.A., R. Baradaran, R.G. Efremov, J.M. Berrisford, and G. Minhas. (2013). A long road towards the structure of respiratory complex I, a giant molecular proton pump. Biochem Soc Trans 41: 1265-1271.

Schafer, E., N.A. Dencher, J. Vonck, and D.N. Parcej. (2007). Three-dimensional structure of the respiratory chain supercomplex I1III2IV1 from bovine heart mitochondria. Biochemistry. 46(44):12579-12585.

Schuller, J.M., J.A. Birrell, H. Tanaka, T. Konuma, H. Wulfhorst, N. Cox, S.K. Schuller, J. Thiemann, W. Lubitz, P. Sétif, T. Ikegami, B.D. Engel, G. Kurisu, and M.M. Nowaczyk. (2019). Structural adaptations of photosynthetic complex I enable ferredoxin-dependent electron transfer. Science 363: 257-260.

Siletsky, S.A. and V.B. Borisov. (2021). Proton Pumping and Non-Pumping Terminal Respiratory Oxidases: Active Sites Intermediates of These Molecular Machines and Their Derivatives. Int J Mol Sci 22:.

Skemiene, K., E. Rekuviene, A. Jekabsone, P. Cizas, R. Morkuniene, and V. Borutaite. (2020). Comparison of Effects of Metformin, Phenformin, and Inhibitors of Mitochondrial Complex I on Mitochondrial Permeability Transition and Ischemic Brain Injury. Biomolecules 10:.

Soboh, B., D. Linder, and R. Hedderich. (2002). Purification and catalytic properties of a CO-oxidizing:H2-evolving enzyme complex from Carboxydothermus hydrogenoformans. Eur. J. Biochem. 269: 5712-5721.

Soboh, B., D. Linder, and R. Hedderich. (2004). A multisubunit membrane-bound [NiFe] hydrogenase and an NADH-dependent Fe-only hydrogenase in the fermenting bacterium Thermoanaerobacter tengcongensis. Microbiology 150: 2451-2463.

Sousa, J.S., E. D''Imprima, and J. Vonck. (2018). Mitochondrial Respiratory Chain Complexes. Subcell Biochem 87: 167-227.

Steffen, W. and J. Steuber. (2013). Cation transport by the respiratory NADH:quinone oxidoreductase (complex I): facts and hypotheses. Biochem Soc Trans 41: 1280-1287.

Steuber, J., C. Schmid, M. Rufibach, and P. Dimroth. (2000). Na+ translocation by complex I (NADH: quinone oxidoreductase) of Escherichia coli. Mol. Microbiol. 35: 428-434.

Stroud, D.A., E.E. Surgenor, L.E. Formosa, B. Reljic, A.E. Frazier, M.G. Dibley, L.D. Osellame, T. Stait, T.H. Beilharz, D.R. Thorburn, A. Salim, and M.T. Ryan. (2016). Accessory subunits are integral for assembly and function of human mitochondrial complex I. Nature. [Epub: Ahead of Print]

Sung, A.Y., R.M. Guerra, L.H. Steenberge, C.L. Alston, K. Murayama, Y. Okazaki, M. Shimura, H. Prokisch, D. Ghezzi, A. Torraco, R. Carrozzo, A. Rötig, R.W. Taylor, J.L. Keck, and D.J. Pagliarini. (2024). Systematic analysis of NDUFAF6 in complex I assembly and mitochondrial disease. Nat Metab 6: 1128-1142.

Tocilescu MA., Zickermann V., Zwicker K. and Brandt U. (2010). Quinone binding and reduction by respiratory complex I. Biochim Biophys Acta. 1797(12):1883-90.

Treberg, J.R., C.L. Quinlan, and M.D. Brand. (2011). Evidence for two sites of superoxide production by mitochondrial NADH-ubiquinone oxidoreductase (complex I). J. Biol. Chem. 286: 27103-27110.

Verkhovskaya, M. and D.A. Bloch. (2013). Energy-converting respiratory Complex I: on the way to the molecular mechanism of the proton pump. Int J Biochem. Cell Biol. 45: 491-511.

Vilhjálmsdóttir, J., A.L. Johansson, and P. Brzezinski. (2015). Structural Changes and Proton Transfer in Cytochrome c Oxidase. Sci Rep 5: 12047.

Vinothkumar, K.R., J. Zhu, and J. Hirst. (2014). Architecture of mammalian respiratory complex I. Nature 515: 80-84.

Walker, J.E. (1992). The NADH:ubiquinone oxidoreductase (complex I) of respiratory chains. Quart. Rev. Biophys. 25: 253-324.

Weerakoon, D.R., and J.W. Olson. (2008). The Campylobacter jejuni NADH:ubiquinone oxidoreductase (complex I) utilizes flavodoxin rather than NADH. J. Bacteriol. 190: 915-925.

Weidner, U., S. Geier, A. Ptock, T. Friedrich, H. Leif, and H. Weiss. (1993). The gene locus of the proton-translocating NADH:ubiquinone oxidoreductase in Escherichia coli. J. Mol. Biol. 233: 109-122.

Welte C. and Deppenmeier U. (2014). Bioenergetics and anaerobic respiratory chains of aceticlastic methanogens. Biochim Biophys Acta. 1837(7):1130-47.

Welte, C., C. Krätzer, and U. Deppenmeier. (2010). Involvement of Ech hydrogenase in energy conservation of Methanosarcina mazei. FEBS J. 277: 3396-3403.

Yagi, T., T. Yano, and A. Matsuno-Yagi. (1993). Characteristics of the energy-transducing NADH-quinone oxidoreductase of Paracoccus denitrificans as revealed by biochemical, biophysical and molecular biological approaches. J. Bioener. Biomembr. 25: 339-345.

Yagi, T., T. Yano, S. Di Bernardo, and A. Matsuno-Yagi. (1998). Procaryotic complex I (NDH-1), an overview. Biochim. Biophys. Acta 1364: 125-133.

Zickermann, V., C. Wirth, H. Nasiri, K. Siegmund, H. Schwalbe, C. Hunte, and U. Brandt. (2015). Structural biology. Mechanistic insight from the crystal structure of mitochondrial complex I. Science 347: 44-49.

Examples:

TC#NameOrganismal TypeExample
3.D.1.1.1

NADH dehydrogenase I, NuoA-N.  NuoL probably comprises part of the proton pathway (Nakamaru-Ogiso et al. 2010).  NuoL (ND5), NuoM (ND4) and NuoN (ND2) are all homologous to proton:sodium antiporters and may all play reoles in pumping protons using a similar mechanism (Sato et al. 2013).

Bacteria

NDH of E. coli NuoA-N

 
Examples:

TC#NameOrganismal TypeExample
3.D.1.2.1NADH dehydrogenase I

Bacteria

NDH of Paracoccus denitrificans Nqo1-14

 
Examples:

TC#NameOrganismal TypeExample
3.D.1.3.1

NADH Dehydrogenase, NDH (Baradaran et al. 2013).  The x-ray structures of various complexes have been solved, and a coupling mechanism involving long range conformational changes has been proposed (Sazanov et al. 2013). The complex includes 16 subunits with nine iron-sulfur clusters, reduced by electrons from NADH. Employing the latest crystal structure of T. thermophilus complex I, Gupta et al. 2020 used microsecond-scale molecular dynamics simulations to study the chemo-mechanical coupling between redox changes of the iron-sulfur clusters and conformational transitions across complex I. The simulations revealed the molecular design principles linking redox reactions to quinone turnover and proton translocation in complex I. Using a zebrafish model of TB infection, Roca et al. 2022 found that tumor necrosis factor (TNF) induces reverse electron transport (RET) in mitochondrial complex I. This in turn drives the production of mitochondrial reactive oxygen species (mROS), causing macrophage necrosis. The complex I inhibitor metformin could be repurposed to inhibit TNF-induced mROS and necrosis in infected zebrafish and human macrophages, suggesting that this common antidiabetes drug may also be a useful adjunct therapy for TB (Roca et al. 2022).

Bacteria

NDH of Thermus thermophilus HB-8 Nqo1-15

 
Examples:

TC#NameOrganismal TypeExample
3.D.1.4.1Proton-translocating hydrogenase, H2ase

Archaea

H2ase of Pyrococcus furiosus MbhA-N (PF1423-PF1436)

 
3.D.1.4.2

[Ni2+-4Fe-4S] H+ translocating, quinone-independent ferredoxin:H+ oxidoreductase, EchA-F (Hedderich and Forzi, 2005; Künkel et al., 1998; Welte and Deppenmeier 2013) The Ech hydrogenases of M. mazei and M. barkeri have been characterized and were shown to pump protons (Welte et al., 2010).  The EchC and EcnF subunits include [Fe4S4] centers, and that in the EchC subunit exhibits a pH dependency to suggest that it plays a role in proton pumping (Forzi et al. 2005).  EchE has the NiFe center that converts 2H+ to H2.  The transmembrane subuits are EchA and EchB (Welte and Deppenmeier 2013).

Archaea

EchABCDEF of Methanosarcina barkeri
EchA (639 aas; 16-18 TMSs; CAA76117)
EchB (285 aas; 7 TMSs; CAA76118)
EchC (156 aas; 0 TMSs; CAA76119)
EchD (113 aas; 0 or 1 TMSs; CAA76120)
EchE (358 aas; 4 TMSs; CAA76121)
EchF (122 aas; 0 TMSs; CAA76122)

 
3.D.1.4.3

Carbon monoxide-induced, H+ translocating, quinone-independent, polyferredoxin (CooF):H+ oxidoreductase, H2ase (Fox et al., 1996a,b; Hedderich and Forzi, 2005) [oxidation of CO to CO2 by CO dehydrogenase results in transfer of electrons to polyferredoxin which reduces H2ase] (Soboh et al., 2002, 2004).

Bacteria

CooMKLXUH of Rhodospirillum rubrum
CooM (1265 aas; 32 TMSs; AAC45116)
CooK (323 aas; 7 TMSs; AAC45117)
CooL (142 aas; 4 TMSs; AAC45118)
CooX (166 aas; 4 TMSs; AAC45119)
CooU (172 aas; 0-4 TMSs; AAC45120)
CooH (361 aas; 2-4 TMSs; AAC45121)

 
3.D.1.4.4

Energy conserving probable carbon monoxide-inducible hydrogenase, CooMKLXUH (Martins et al. 2016).

CooMKLXUH of Desulfovibrio vulgaris

 
Examples:

TC#NameOrganismal TypeExample
3.D.1.5.1

Proton-translocating NADH dehydrogenase I

Bacteria

NDH of Geobacter sulfurreducens
NuoA (142 aas) NP_954485
NuoB (170 aas) NP_951399
NuoC (162 aas) NP_951400
NuoD (390 aas) NP_951401
NuoE (173 aas) NP_951402
NuoF (591 aas) NP_951403
NuoG (677 aas) NP_954479
NuoH (329 aas) NP_954476
NuoI (176 aas) NP_954474
NuoJ (163 aas) NP_954473
NuoK (102 aas) NP_954472
NuoL (624 aas) NP_954471
NuoM (496 aas) NP_954470
NuoN (466 aas) NP_954469

 
Examples:

TC#NameOrganismal TypeExample
3.D.1.6.1

The vertebrate H+-translocating NADH dehydrogenase (NDH) complex (45 subunits) (Cardol et al., 2004).  The 3-d structure of the 44 subunit complex (14 core subunits present in bacteria, plus 30 additional subunits) with a molecular weight of 1 MDa, has been solved at 5 Å resolution by single particle electron cryo-microscopy (Vinothkumar et al. 2014).The core subunits contain eight iron-sulphur clusters and 60 transmembrane helices. The structures of many of the supernumerary subunits were determined or modeled. The structure provided insight into the roles of the supernumerary subunits in regulation, assembly and homeostasis.  One such subunit, GRIM-19 or NDUFA13, (Q9P0J0 of the human homologue) is essential for membrane potential formation and NADH assembly (Lu and Cao 2008).  Stroud et al. 2016 showed that 25 of the 31 accessory subunits in the 45 subunit human NADH dehydrogenase complex are required for assembly of a functional complex, and 1 subunit is essential for cell viability. Quantitative proteomic analysis revealed that loss of each subunit affects the stability of other subunits residing in the same structural module (Stroud et al. 2016). Leber's hereditary optic neuropathy (LHON) in humans is associated with combinations of individually non-pathogenic missense mitochondrial DNA (mtDNA) variants, affecting the MT-ND4, MT-ND4L and MT-ND6 subunit genes of Complex I (Caporali et al. 2018). It has been proposed that the quinone reaction cavity is indeed from the matrix-side region covered by the ND3 TMH1-2 loop (Masuya et al. 2021). Resting mitochondrial complex I from Drosophila melanogaster adopts a helix-locked state (Padavannil et al. 2023).  Mutations in the ND5 gene are associated with Leber Hereditary Optic Neuropathy (Pandya et al. 2024).

Animals

NDH of Bos taurus (45 subunits)
PSST (20) (NuoB) (P42026)
TYKY (23) (NuoI) (P42028)
24 (NuoE) (P04394)
30 (NuoC) (P23709)
49 (NuoD) (P17694)
51 (NuoF) (P25708)
75 (NuoG) (P15690)
ND1 (36) (NuoH) (P03887)
ND2 (39) (NuoN) (P03892)
ND3 (13) (NuoA) (P03898)
ND4 (52) (NuoM) (P03910)
ND4L (11) (NuoK) (P03902)
ND5 (67) (NuoL) (P03920)
ND6 (19) (NuoJ) (P03924)
MWFE (7.5) (Q02377)
SDAP (8) (P52505)
B8 (Q02370)
B12 (Q02365)
B13 (P23935)
13a (P23934)
B14 (Q02366)
ESSS (14.5) (Q8HXG5)
PFFD (15) (Q02379)
B16.6 (Q95KV7)
B17.2 (O97725)
B18 (Q02368)
AQDQ (18) (Q02375)
PGIV (19) (P42029)
B22 (Q02369)
PDSW (22) (Q9DCS9)
39 (P34943)
B14.7 (Q8HXG6)
B15 (P48305)
ASHI (19) (Q02372)
KFYI (6)6 (Q02376)
MNLL (7) (Q02378)
AGGG (8) (Q02374)
B95,6 (Q02371)
MLRQ (9) (Q01321)
106 (P25712)
B14.5a (Q05752)
B14.5b (Q02827)
SGDH (16) (Q02380)
B17 (Q02367)
42 (P34942)

 
3.D.1.6.2

The fungal H+ translocating NADH dehydrogenase (NDH) complex (38 subunits; 35 included here) (Cardol et al., 2004).  The high resolution (3.6 - 3.9 Å) structure of the mitochondrial Yarrowia lipolytica enzyme, showing all central subunits that execute the bioenergetic functions, has been solved (see discussion for TC# 3.A.1 and Zickermann et al. 2015). The subunit inventory of mitochondrial complex I from the obligate aerobic yeast Yarrowia lipolytica involved a total of 37 subunits (Abdrakhmanova et al. 2004).

Fungi

NDH of Neurospora crassa (35 subunits):
19.3 (NuoB) (O47950)
21.3c (NuoI) (Q12644)
24 (NuoE) (P40915)
31 (NuoC) (P23710)
49 (NuoD) (P22142)
51 (NuoF) (P24917)
78 (NuoG) (P24918)
ND1 (42) (NuoH) (P08774)
ND2 (66) (NuoN) (Q35140)
ND3 (NuoA) (Q35141)
ND4 (NuoM) (P15582)
ND4L (10) (NuoK) (P05509)
ND5 (80) (NuoL) (P05510)
ND6 (NuoJ) (P15959)
9.8 (Q6MFL8)
9.6 (P11943)
10.5 (Q07842)
10.63 (XP_331394)
29.9 (P24919)
18.43(Q7RWU3)
14.8 (P42114)
11.73 (XP_324110)
11.53 (Q9P6T9)
13.53 (P7S1I2)
13.43 (Q8X0V6)
89.73 (Q7RZ09)
21 (P25711)
20.8 (P21976)
183 (Q875B9)
12.3 (Q03015)
40 (P25284)
21.3B (P25710)
73 (XP_322246)
20.13 (XP_332152)
20.9 (Q02854)

 
3.D.1.6.3The higher plant H+ translocating NADH dehydrogenase (NDH) complex (41 subunits; 23 included here) (Cardol et al., 2004)PlantsNDH of Arabidopsis thaliana (32 subunits)
24 (NuoB) (Q42577)
25.5 (NuoI) (Q9FX83)
28.3 (NuoE) (O22769)
ND9 (22.6) (NuoC) (Q95748)
ND7 (44.6) (NuoD) (P93306)
53.5 (NuoF) (Q9FNN5)
81.5 (NuoG) (Q9FGI6)
ND1 (36) (NuoH) (NP_085565)
ND2 (55) (NuoN) (NP_085584)
ND3 (14) (NuoA) (NP_085553)
ND4 (55) (NuoM) (NP_085518)
ND4L (11) (NuoK) (NP_051111)
ND5 (74) (NuoL) (NP_085478)
ND6 (23.5) (NuoJ) (NP_085495)
7.5 (Q84MD9)
14 (O80800)
10.8 (Q8FIJ2)
7 (AK059007)
19.2 (Q9FLX7)
12.2 (Q9M9M6)
15 (Q9LHI0)
133 (Q8L3S7)
14 (Q9LZI6)
16.1 (Q8RWA7)
18 (Q9M9M9)
12 (Q8GXZ6)
17.1 (Q9FJW4)
12 (Q8LGE7)
13.6 (Q945M1)
12.5 (Q84W12)
44 (Q9SK66)
 
3.D.1.6.4The green algal H+ translocating NADH dehydrogenase (NDH) complex (42 subunits; 33 included here) (Cardol et al., 2004)AlgaeNDH of Chlamydomonas reinhardtii (33 subunits)
18 (NuoB) (Q6V9B0)
23 (NuoI) (Q6V9B1)
27 (NuoE) (Q6V9B3)
25 (NuoC) (Q6V5O7)
43 (NuoD) (Q6V9A8)
50 (NuoF) (Q6V9B2)
75 (NuoG) (Q6UKY6)
ND1 (31.6) (NuoH) (P11658)
ND2 (42.4) (NuoN) (P08740)
14 (NuoA) (Q6V502)
ND4 (48.7) (NuoM) (P20113)
24.2 (NuoK) (Q84K56)
ND5 (59) (NuoL) (P08739)
ND6 (17.7) (NuoJ) (P10329)
7.53 (Q6Q1V4)
143 (Q6UKY4)
11 (Q6V9A9)
6.53 (Q6Q1V8)
18 (Q6UKY3)
133 (Q6UP30)
14 (Q6UKY9)
17 (Q6Q1W0)
11 (Q6TH88)
16 (Q6UP32)
18 (Q6UP31)
12 (Q6UKY7)
19 (Q6UP29)
12.93 (Q6V504)
13.93 (Q6UKY8)
17 (Q6V505)
38 (Q6V506)
23 (Q6QAY4)
13 (Q9UP28)
 
Examples:

TC#NameOrganismal TypeExample
3.D.1.7.1The NADH: ubiquinone oxidoreductase homologue using flavodoxin rather than NADH as electron donor (Weerakoon and Olson, 2008), Bacteria

Flavodoxin: ubiquinone oxidoreductase of Campylobacter jejuni
NuoA Q0P849
NuoB Q0P850
NuoC Q0P851
NuoD Q9PM99
NuoG Q0P855
NuoH Q9PMA3
NuoI Q0P857
NuoJ Q0P858
NuoK Q0P859
NuoL Q9PMA7
NuoM Q0P861
NuoN Q0P862
Cj1575c (NuoE) Q0P853
Cj1574c (NuoF) Q0P854

 
Examples:

TC#NameOrganismal TypeExample
3.D.1.8.1

The chloroplast NDH-1 complex with form subcomplexes, subcomplex: M (membrane). L (lumen), A and B (Peng et al., 2011).  The NDF6 (PNSB6) protein is essential for activity (Ishikawa et al. 2008).

Plants

NDH complex of Arabidopsis thaliana chloroplasts
NdhA (Q37165) = NuoH in E. coli
NdhB (P0CC32) = NuoN in E. coli
NdhC (P56751) = NuoA in E. coli
NdhD (P26288) = NuoM in E. coli
NdhE (P26289) = NuoK in E. coli
NdhF (P56752) = NuoL in E. coli
NdhG (Q95695) = NuoJ in E. coli
NdhH (P56753) = NuoD in E. coli
NdhI (P56755) = NuoI in E. coli
NdhJ (P56754) =  NuoC in E. coli
NdhK (P56756) = NuoB in E. coli
NdhL (Q9CAC5)
NdhM (Q2V2S7)
NdhN (B3H5R4)
NdhO (Q9S829)
NDF1 (Q9S9N6)
NDF2 (C0Z2H4)
NDF4 (Q9LU21)
NDF6 (B3H6Z4)
NDH18 (Q9FG89)
PQL (Q9SGH4)
PPL2 (O80634)
CYP20-2 (F4K2G0)
FKBP16-2 (F4JW56)
PQL (Q9XI73) 

 

 
3.D.1.8.2

The photosynthetic/respiratory NAD(P)H-quinone oxidoreductase subunits A - Q and S. NdhP and NdhQ are peptides of 42 and 39 aas, identified by cryoEM at 3.3 Å resolution (Schuller et al. 2019). NDH-1 (NdhA) shuttles electrons from reduced ferridoxin, via FMN and iron-sulfur (Fe-S) centers, to quinones in the respiratory and photosynthetic chains. The immediate electron acceptor for the enzyme in this species is believed to be plastoquinone. The system couples the redox reaction to proton translocation, and thus conserves the redox energy in a proton gradient (Schuller et al. 2019). Ferridoxin directly mediates electron transfer between photosystem I and complex I. Ferridoxin efficiently binds to complex I with subunit NdhS being the key component in this process (Schuller et al. 2019).

NAD(P)H-quinone oxidoreductase of Thermosynechococcus elongatus (strain BP-1)
NdhA, 372 aas, Q8DL32
NdhB, 515 aas, Q8DMR6
NdhC, 132 aas, Q8DJ02
NdhD1, 526 aas, Q8DKY0
NdhE, 101 aas, Q8DL29
NdhF1, 656 aas, Q8DKX9
NdhG, 200 aas, Q8DL30
NdhH, 394 aas, Q8DJD9
NdhI, 196 aas, Q8DL31
NdhJ, 168 aas, Q8DJ01
NdhK, 237 aas, Q8DKZ4
NdhL, 76 aas, Q8DKZ3
NdhM, 111 aas, Q8DLN5
NdhN, 150 aas, Q8DJU2
NdhO, 70 aas, Q8DMU4
NdhS, 110 aas, Q8DL61

 
Examples:

TC#NameOrganismal TypeExample
3.D.1.9.1

Hydrogenase 4, Hyf. Catalyzes H2 production and H+/K+ exchange (Bagramyan et al., 2001; 2002).

Bacteria

Hydrogenase 4 (HyfABCDEFGHI) of E. coli
HyfA (P23481)
HyfB (P23482)
HyfC (P77858)
HyfD (P77416)
HyfE (P0AEW1)
HyfF (P77437)
HyfG (P77329)
HyfH (P77423)
HyfI (P77668)

 
3.D.1.9.2

Hydrogenase 3, Hyc. Catalyzes H2 production coupled to H+ export (Bagramyan et al., 2002). The HycC and HycD proteins span the membrane multiple times; FdhF is a molybdenum-dependent formate dehydrogenase while HycE (Hyd-3) is a NiFe hydrogenase (McDowall et al. 2014).

Bacteria

Hydrogenase 3, HycBCDEF of E. coli 
HycB (like HyfA) (P0AAK3)
HycC (like HyfB) (P16429) 
HycD (like HyfC) (P16430)
HycE (like HyfD) (P16431)
HycF (like HyfE) (P16432)
HycG (like HyfF) (P16433)
FdhF (FHL) (P07658)

 
3.D.1.9.3

Putative hydrogenase 4, HyfA-I. 

Spirochaetes

HyfA-I (H missing) of Leptospira interrogans 
HyfA (99aas) (Q8EXV0)
HyfB (573aas) (Q8EYE3)
HyfC (295aas) (Q8EYE2)
HyfD (519aas) (Q8F7D6)
HyfE (206aas) (Q8EYE1)
HyfF (402aas) (Q8EYD0)
HyfG (466aas) (Q8EYD9)
HyfI (273aas) (Q8EYD8)