Fact-checked by Grok 2 weeks ago

Respiratory complex I

Respiratory complex I, also known as NADH:ubiquinone oxidoreductase or , is the first and largest enzyme complex in the mitochondrial (), playing a pivotal role in . It catalyzes the oxidation of NADH derived from the tricarboxylic acid cycle, transferring two electrons to ubiquinone (coenzyme Q) to form , while simultaneously pumping four protons from the to the per NADH oxidized. This redox reaction is coupled to proton translocation, establishing a proton motive force across the that drives ATP synthesis via , and it represents the primary entry point for electrons into the respiratory chain. Structurally, complex I adopts an L-shaped architecture, approximately 1 MDa in , comprising a peripheral hydrophilic arm extending into and a membrane-embedded hydrophobic arm. The peripheral arm houses the NADH-binding site, a (FMN) cofactor, and eight iron-sulfur (Fe-S) clusters that facilitate sequential over a distance of about 50 to the quinone-binding site near the . The membrane arm, in contrast, contains proton-pumping channels and is responsible for vectorial proton translocation, with the overall structure divided into four modules: N (NADH oxidation), Q ( reduction), and two proton-pumping modules (P_P and P_D). In mammals, the complex assembles from 45 distinct subunits—14 core subunits conserved across species for catalytic function and 31 accessory (supernumerary) subunits that support assembly, stability, and regulatory roles—encoded by both nuclear and . Beyond energy production, complex I is a major source of (ROS) in mitochondria, particularly under reverse conditions, influencing cellular signaling pathways such as and calcium . Dysfunctions, often arising from in core subunits (e.g., ND1–ND6) or assembly factors, are implicated in a spectrum of human diseases, including mitochondrial encephalomyopathies like , (LHON), and neurodegenerative disorders such as . These pathologies highlight complex I's critical role in cellular metabolism, with therapeutic strategies targeting its assembly and activity under investigation to mitigate and energy deficits.

Overview

Definition and nomenclature

Respiratory complex I, formally known as NADH:ubiquinone (EC 7.1.1.2), is the first enzyme complex in the mitochondrial , where it catalyzes the transfer of electrons from NADH to ubiquinone while coupling this process to proton translocation across the . The complex was first isolated in the early 1960s through fractionation of submitochondrial particles from bovine heart mitochondria, with Hatefi and colleagues purifying it as the DPNH-coenzyme Q reductase and designating it as Complex I within the respiratory chain assemblies. It is commonly synonymous with or Type I NADH dehydrogenase, referring to its multisubunit, proton-pumping nature, in contrast to Type II NADH dehydrogenases, which are single-subunit, non-proton-pumping alternatives found in some organisms and lacking the full complexity of the respiratory chain enzyme. In mammals, respiratory complex I exhibits a molecular weight of approximately 1 and comprises subunits.

Evolutionary conservation

Respiratory complex I traces its origins to the (LUCA), where it likely existed as a simpler with 11 protein subunits capable of NADH oxidation and reduction, as evidenced by phylogenetic analyses of archaeal and bacterial homologs. This ancestral form evolved through modular assembly from [NiFe] hydrogenases and Na+/H+ antiporter complexes, adapting to varying geochemical environments in . Homologs of complex I are widely distributed across the three domains of life: in , it is encoded by the nuo and functions as the primary NADH: oxidoreductase; in , similar enzymes with F420-dependent variants are present in lineages like Crenarchaeota; and in eukaryotes, it resides in mitochondria, reflecting endosymbiotic inheritance from α-proteobacterial ancestors. The core functional unit of complex I consists of 14 highly conserved subunits that span to mammals, encompassing the peripheral arm for (NADH to FMN and Fe-S clusters) and the membrane arm for proton translocation. In , these 14 subunits form the minimal, fully active , sufficient for without accessory proteins. Eukaryotic versions, particularly in mammals, expand to over 44 subunits through the addition of nuclear-encoded accessory proteins that enhance stability, assembly, and regulation, though the catalytic core remains invariant. Evolutionary adaptations have led to variations in complex I across lineages, including its complete loss or reduction in certain parasitic organisms adapted to or low-oxygen environments, such as the kinetoplastid , where its presence is atypical and debated in bloodstream forms. Conversely, in photosynthetic eukaryotes, a homolog known as the NAD(P)H (NDH) complex persists in chloroplasts, derived from cyanobacterial ancestors and functioning in cyclic electron transport around to optimize photoprotection and ATP production. These adaptations highlight complex I's plasticity, balancing energy demands with environmental constraints. Recent advances in cryo-electron microscopy (cryo-EM), particularly since the , have illuminated the structural conservation of complex I, revealing a canonical L-shaped architecture preserved from bacterial minimal forms to elaborate mammalian assemblies. High-resolution structures, such as those of ovine mitochondrial complex I at 3.9 Å (2016), bacterial enzymes like Thermus thermophilus at near-atomic resolution, and more recent porcine structures at ~2.5–3 Å (2024), demonstrate that the L-shape—comprising a matrix-facing peripheral arm and a membrane-embedded arm—underpins electron-proton coupling across species, with only peripheral additions in eukaryotes altering the overall scaffold minimally.

Structure and assembly

Subunit composition and cofactors

Respiratory complex I exhibits a modular subunit composition that varies between prokaryotes and eukaryotes, reflecting differences in complexity and regulatory needs. In bacterial species such as Thermus thermophilus and Escherichia coli, the enzyme assumes a minimalistic form with 14 conserved core subunits—seven in the peripheral (matrix-facing) arm and seven in the membrane arm—sufficient for NADH oxidation, electron transfer to ubiquinone, and proton pumping across the membrane. These core subunits alone enable the full catalytic activity observed in simpler organisms. In mammalian mitochondria, complex I is substantially larger and more intricate, consisting of 45 subunits with a exceeding 1 MDa. Of these, seven are encoded by (MT-ND1, MT-ND2, MT-ND3, MT-ND4, MT-ND4L, MT-ND5, and MT-ND6), primarily integrating into the membrane arm, while the remaining 38 are nuclear-encoded and imported into the . The 14 core subunits, homologous to the bacterial ones, handle the primary and proton translocation functions, with examples including NDUFV1 (51 kDa subunit binding FMN) and NDUFS7 (hosting the terminal Fe-S cluster). The additional 31 accessory subunits, absent in , are eukaryotic innovations that enhance , facilitate stepwise , and enable regulatory control; notable examples include NDUFA1–NDUFA13, which form part of early intermediates and shield the complex from or oxidative damage. Subunits like NDUFS4 stabilize the peripheral arm, NDUFA5 reinforces the Q-module for ubiquinone binding, and NDUFA12 aids in late-stage maturation by promoting cofactor integration. Regulatory accessory subunits, such as NDUFA9 (modulating ubiquinone access) and NDUFA13 (linking complex integrity to apoptotic signaling via STAT3 interaction), allow fine-tuned responses to cellular energy demands. The catalytic prowess of complex I depends on an array of non-protein cofactors embedded within the core subunits. A single (FMN) cofactor per complex is non-covalently bound to the NDUFV1 subunit in the peripheral arm's N-module, serving as the initial electron acceptor from NADH during oxidation. Electrons are subsequently transferred through a chain of eight iron-sulfur (Fe-S) clusters—comprising two binuclear [2Fe-2S] and six tetranuclear [4Fe-4S] centers—spanning approximately 50 Å from FMN to the ubiquinone reduction site; the terminal N2 cluster, a [4Fe-4S] center in NDUFS7, is particularly critical for efficient electron donation to ubiquinone at the membrane interface. Unlike other respiratory complexes, complex I contains no groups, relying solely on FMN and Fe-S clusters for chemistry. Post-translational modifications (PTMs) on complex I subunits provide additional layers of regulation, influencing activity, assembly, and integration into supercomplexes. , mediated by mitochondrial kinases such as or Src family members, has been documented on several core and accessory subunits; for example, the core subunit NDUFS2 (also termed 49 kDa or B14.5a) undergoes in bovine heart mitochondria, potentially at sites near the ubiquinone-binding interface to modulate efficiency or conformational changes. Such modifications, often responsive to cellular signals like levels or calcium flux, underscore the dynamic regulation of complex I beyond its static composition.

Three-dimensional architecture

Respiratory complex I exhibits a characteristic L-shaped , consisting of a hydrophilic peripheral projecting into the and a hydrophobic embedded in the . The peripheral , approximately 20 nm in length, houses the (FMN) and iron-sulfur (Fe-S) clusters essential for , while the , spanning about 15 nm, accommodates the proton translocation channels. The enzyme is organized into four main functional modules: the N-module, located at the tip of the peripheral arm, which binds NADH and contains FMN; the Q-module, positioned at the junction of the two arms, responsible for ubiquinone binding and reduction; and the (proximal) and PD (distal) modules, comprising the membrane arm, which facilitate proton pumping across the membrane. These modules are formed by the conserved core subunits, with the N- and Q-modules in the peripheral arm and the and PD modules in the membrane arm. Recent cryo-electron microscopy (cryo-EM) studies have resolved the at near-atomic below 3 Å, enabling detailed visualization of subunit arrangements and cofactor positions, such as the 2.1 Å structure from Yarrowia lipolytica. Intersubunit interfaces stabilize the modular assembly, with key contacts between the and membrane arms mediated by amphipathic helices and conserved loops, such as those from ND1 interacting with NDUFS2 in the Q-module. Flexible linkers, including loops in subunits like ND1 (between transmembrane helices 5 and 6), connect the modules and permit relative movements while maintaining overall integrity. These interfaces are lined by that further reinforce the structure. While the core architecture is highly conserved across species, mammalian complex I incorporates additional accessory subunits—totaling 45 compared to 14 in minimal bacterial versions—forming a stabilizing shell around the core and extending the peripheral arm slightly. Bacterial structures, such as from or Thermus thermophilus, lack these supernumerary subunits, resulting in a more compact assembly with fewer lipid-binding sites and subtle variations in hydration at proton channels, like in the ND5 subunit.

Biosynthesis and assembly

The biosynthesis of respiratory complex I, also known as NADH:ubiquinone , involves the coordinated integration of subunits encoded by both mitochondrial and genomes. In humans, the enzyme consists of 45 subunits, with seven hydrophobic core subunits (MT-ND1 through MT-ND6 and MT-ND4L) translated from within the , while the remaining 38 hydrophilic and accessory subunits are synthesized in the from DNA and subsequently imported into the mitochondria via complexes in the outer and inner membranes. This dual-genetic origin necessitates precise temporal and spatial regulation to ensure stoichiometric balance and prevent aggregation of unassembled components. Assembly proceeds through a modular, stepwise process in the and , facilitated by at least 18 dedicated chaperone-like assembly factors, including the NDUFAF1 through NDUFAF12 proteins and others such as TIMMDC1 and TMEM70. These factors promote the formation of transient subcomplexes, beginning with the Q-module (comprising nuclear-encoded subunits like NDUFS2 and NDUFS3, responsible for ), followed by integration of the N-module (nuclear-encoded subunits like NDUFV1 and NDUFS1 for NADH oxidation) and the PP and modules (incorporating mitochondrial-encoded subunits like MT-ND5 in the module for proton pumping). Recent studies have identified additional late-stage factors, such as RTN4IP1, which stabilize the ND5-module and enable docking of the ND4-module to complete the arm. Key rate-limiting steps include the biogenesis and insertion of iron-sulfur (Fe-S) clusters, which are essential for across the eight clusters in the enzyme; this process relies on the mitochondrial iron-sulfur cluster (ISC) assembly machinery, with factors like NUBPL ensuring proper maturation of the N- and Q-modules. Mitochondrial translation of the core membrane subunits represents another bottleneck, as delays can stall subcomplex progression. The entire holoenzyme matures within the , where assembly factors dissociate upon completion, allowing incorporation into respiratory supercomplexes with complexes III and IV. Deficiencies in this pathway, often arising from in nuclear-encoded assembly factors such as NDUFAF5 or NDUFAF7, lead to isolated complex I defects characterized by reduced enzyme activity and accumulation of immature subcomplexes, contributing to mitochondrial disorders without affecting other respiratory complexes.

Function

Role in electron transport chain

Respiratory complex I serves as the primary entry point for electrons derived from NADH, which is produced during the tricarboxylic acid (TCA) cycle in the . It catalyzes the oxidation of NADH to NAD⁺, transferring the two electrons along an internal chain of iron-sulfur clusters to the lipid-soluble ubiquinone (coenzyme Q), reducing it to (QH₂). This process initiates the (ETC) in mitochondria, linking catabolic metabolism to . The in complex I is tightly coupled to the translocation of four protons (H⁺) from the to the per two electrons transferred, generating a significant portion of the proton motive force (PMF) across the . This contributes approximately 40% of the total proton flux in the , which drives ATP synthesis by (complex V). Complex I integrates with downstream ETC components—complexes , III, and —where electrons from QH₂ continue through the chain to reduce oxygen to , while the PMF powers ATP . In certain physiological contexts, such as high-intensity exercise in mammalian , alternative metabolic bypasses can enable NADH oxidation independently of complex I, shunting electrons directly to complex II or other pathways to maintain energy production under stress. Overall, the complete oxidation of NADH via complex I yields approximately 2.5 molecules of ATP per NADH, compared to about 1.5 ATP from FADH₂ oxidation entering at complex II, highlighting complex I's central role in maximizing energy efficiency from TCA cycle substrates.

Overall catalytic reaction

The overall catalytic reaction of respiratory complex I (NADH:ubiquinone oxidoreductase) oxidizes NADH to NAD⁺ while reducing ubiquinone (Q) to ubiquinol (QH₂), coupling this redox process to the vectorial translocation of protons from the mitochondrial matrix to the intermembrane space. The net balanced equation is: \text{NADH} + 5\text{H}^+_{\text{matrix}} + \text{Q} \to \text{NAD}^+ + \text{QH}_2 + 4\text{H}^+_{\text{intermembrane}} This reflects the consumption of five protons from the matrix— one from NADH deprotonation, two for QH₂ formation, and two additional equivalents effectively pumped outward—resulting in a net translocation of four protons per reaction cycle. The stoichiometry involves the transfer of two electrons from NADH to Q, fully reducing the quinone to QH₂ without involvement of oxygen as an alternative acceptor under physiological conditions. Respiratory complex I employs a linear chain of eight iron-sulfur (Fe-S) clusters, all of the low-potential, non-Rieske type (with midpoint potentials ranging from approximately -400 mV to -150 mV), which facilitates efficient forward electron flow and minimizes superoxide formation during normoxia by limiting electron dwell times on reactive sites. In contrast to high-potential Fe-S centers like the Rieske protein in complex III, these clusters' negative potentials reduce the thermodynamic favorability of partial reduction of O₂ to superoxide in the forward direction. The reaction is thermodynamically favorable, driven by a substantial span: the NADH/NAD⁺ has a potential (E_m) of -320 mV, while the Q/QH₂ operates at approximately +90 mV, yielding a ΔE of about 410 mV and releasing sufficient (ΔG°' ≈ -79 kJ/mol) to support proton pumping. The (FMN) cofactor, with an E_m near -340 mV, serves as the initial from NADH, bridging to the Fe-S chain. Experimental verification of the reaction kinetics and stoichiometry has relied on stopped-flow spectroscopy, which captures millisecond-scale events by monitoring absorbance changes during NADH oxidation and Q reduction in . These studies confirm the rapid, sequential reduction of FMN and Fe-S clusters, with overall turnover rates supporting the 4 H⁺/2 e⁻ stoichiometry in both bacterial and mammalian enzymes under physiological proton motive force.

Mechanism

Electron transfer pathway

The electron transfer pathway in respiratory complex I initiates with the binding of NADH at the peripheral arm, where it undergoes oxidation by transferring a hydride ion to the non-covalently bound flavin mononucleotide (FMN) cofactor, forming NAD⁺ and FMNH₂. This step occurs rapidly through a proton-coupled hydride transfer mechanism, enabling efficient entry of electrons into the complex. The FMN is embedded within the NDUFV1 subunit and positioned to accept electrons without direct involvement of protein residues in the initial redox event. From FMNH₂, the two electrons are relayed sequentially in one-electron steps through a chain of eight iron-sulfur (Fe-S) clusters spanning over 95 from the FMN to the ubiquinone ()-binding site at the of and membrane arms. The primary linear pathway proceeds as FMN → N3 → N1b → N4 → N5 → N6a → N6b → N2 → , with electron tunneling facilitating transfers across cluster-to-cluster distances of 10–14 . These transfers are enhanced by intervening water molecules and protein residues, such as cysteines, which bridge gaps and increase rates by up to three orders of magnitude in key steps like N5 to N6a. An off-pathway [2Fe-2S] cluster (N1a) near FMN can occasionally participate in reverse transfers under certain conditions but does not contribute to forward flux. The Fe-S clusters comprise two binuclear [2Fe-2S] types (N1a and N1b) and six tetranuclear [4Fe-4S] types (N2, N3, N4, N5, N6a, N6b), coordinated primarily by residues within subunits like NDUFS1, NDUFS7, and NDUFS8. The terminal [4Fe-4S] cluster N2, located in the NDUFS2 subunit approximately 12 from the Q site, acts as a gatekeeper by delivering electrons directly to the ubiquinone headgroup, which binds in a narrow tunnel involving conserved residues like Tyr87 and His38. This positioning ensures controlled access and reduction of Q to semiquinone and then quinol. Redox potentials are tuned progressively higher along the chain to drive forward transfer and minimize reversal, starting at approximately -380 mV for FMN and N3, rising through intermediate clusters (N1b ≈ -250 mV, N4/N5 ≈ -250 to -300 mV), to -150 mV for N2, and culminating at +90 mV for the Q/QH₂ couple. Recent molecular dynamics simulations from the 2020s indicate that the final transfer from N2 to Q involves a coordinated, potentially bifurcated delivery of the two electrons required for full Q reduction, possibly via a stabilized semiquinone intermediate within the Q tunnel to optimize energetics and coupling.

Proton translocation process

Respiratory complex I translocates four protons across the inner mitochondrial membrane for every two electrons transferred from NADH to ubiquinone, establishing a proton motive force essential for ATP synthesis. This process operates independently of a Q-cycle, unlike complexes III and IV, relying instead on direct coupling between quinone reduction and vectorial proton pumping through dedicated membrane channels. The mechanism involves conformational waves initiated at the quinone-binding site (Q-site), which propagate through the peripheral arm to the membrane domain, coordinating proton uptake from the matrix and release to the intermembrane space. Proton translocation occurs via four distinct pumping channels, three of which are formed by the antiporter-like subunits ND2, ND4, and ND5, while the fourth is the primarily involving ND1. Each channel features half-channel architectures: N-side half-channels for proton uptake from , connected by central binding sites, and P-side half-channels for release to the . In ND2, ND4, and ND5, key residues such as conserved glutamates and lysines (e.g., Glu in ND4) facilitate proton conduction along hydrated pathways, with wires enabling Grotthuss-like transfer; the Q-entry path links to the tunnel near ND1 and ND3, while H+ exit paths converge toward ND5 for coordinated ejection. These channels ensure unidirectionality, preventing backflow under the . The energy for uphill proton pumping derives from electrostatic repulsion generated during ubiquinol (QH₂) formation and deprotonation at the Q-site, which triggers a cascade of charged-group rearrangements propagating as an electrostatic wave through the complex. This repulsion displaces ionizable residues and helix bundles in the membrane arm, driving protonation changes in the channels against the proton motive force. Experimental support for this mechanism comes from studies, where introducing disulfide bonds between the ND3 TMH1-2 loop and NDUFS7 subunit decouples conformational changes from proton pumping, abolishing translocation without affecting . Additionally, pH-jump assays combined with simulations demonstrate rapid water chain reorganization in the channels upon redox-state changes, confirming dynamic proton pathways responsive to Q-site events. High-resolution cryo-EM structures further validate the half-channel geometries and key residues, showing hydrated networks essential for conduction.

Conformational dynamics

Respiratory complex I exhibits dynamic conformational changes throughout its , enabling the coupling of from NADH to ubiquinone with proton translocation across the . These dynamics primarily involve transitions between distinct structural states at the (Q)-, captured through high-resolution cryo- (cryo-EM) structures. The enzyme alternates between an open state, which facilitates ubiquinone access and semiquinone release, and a closed state that seals the Q-site to support stable and prevent uncoupled activity. These state transitions occur at the interface between (NADH-oxidizing) arm and the arm, involving coordinated rearrangements of loops and helices in subunits such as NDUFS2, ND1, and ND3. Ubiquinone binding to the Q-site triggers a key conformational rearrangement, including a of the peripheral arm by approximately 10° relative to the arm, which constricts the Q-cavity and positions the for reduction. This motion, observed in bacterial and mammalian structures, tightens interactions around the quinone headgroup, stabilizing the semiquinone and initiating downstream signaling. Cryo-EM intermediates from 2018 onward, including turnover states in ovine and I, reveal open Q-sites with disordered access loops and closed configurations with ordered, solvent-excluded channels, highlighting the precision of these changes during . The conformational dynamics propagate as waves of structural rearrangement from the Q-module (encompassing the FMN and iron-sulfur clusters) to the distal P-module (proton-pumping channels), synchronizing with proton release into four half-channels. These waves, inferred from comparative structural analyses and simulations, ensure mechanical coupling without energy dissipation, with local twisting and tilting motions amplifying the signal across the ~200 span of the arm. Under conditions of low proton motive force (ΔpH), the closed state predominates to inhibit reverse electron flow, blocking futile cycling and generation at the Q-site.

Regulation and modulation

Active/inactive transitions

Respiratory complex I in mammalian mitochondria exists in two interconvertible states: an active (A) form that catalyzes rapid NADH oxidation and a deactive (D) form that is catalytically inactive and predominant in isolated mitochondria lacking substrates. The D-form arises spontaneously during or ischemia when electron turnover ceases, serving as a protective mechanism to halt electron flow and minimize (ROS) production upon reoxygenation. from the D- to A-form requires energization by the proton motive force (ΔpH) across the or exposure to chaotropic agents such as , which disrupt the stable D-state conformation. This transition involves exposure of Cys39 in the ND3 subunit, which becomes solvent-accessible in the D-form, allowing modification by thiol-reactive agents. The physiological role of the A/D transition is to safeguard mitochondria during metabolic stress, such as ischemia, by suppressing ROS generation at complex I, the primary site of superoxide production under reverse electron transfer conditions. During ischemia, the D-form predominates, preventing harmful ROS bursts that could damage cellular components upon reperfusion; reactivation occurs slowly over minutes as oxygen and substrates become available, contrasting sharply with the millisecond timescale of normal catalytic cycles. This temporal mismatch ensures controlled resumption of respiration, reducing oxidative injury in tissues like the heart and brain. Studies in isolated mitochondria and cellular models confirm that enforced maintenance of the A-form exacerbates ischemia-reperfusion damage, underscoring the protective utility of the D-state. Structurally, the D-form features localized disorder in the ubiquinone-binding channel, including a small (approximately 3.4°) of the hydrophilic relative to the arm in the NDUFS1 subunit (the 75-kDa iron-sulfur protein), which displaces key elements like the β1-β2 loop of the 49-kDa subunit and prevents substrate access. Cryo-electron microscopy structures reveal that this and associated unfolding block the Q-site, rendering the enzyme inactive until corrective conformational changes restore order. Recent 2022 investigations have identified thiol oxidation of conserved cysteines in NDUFS2 and adjacent subunits as a key trigger for the D-state, linking signaling to the transition and highlighting its role in hypoxic adaptation across species like C. elegans. These oxidation events propagate structural shifts that stabilize the D-form, reversible by reducing agents or ΔpH-driven . In contrast to mammalian complex I, bacterial NADH:ubiquinone oxidoreductases lack this A/D regulatory switch, remaining constitutively active without the slow deactivation kinetics or protective D-state, a distinction attributed to the absence of certain accessory subunits and the simpler membrane environment in prokaryotes. This eukaryotic-specific mechanism likely evolved to fine-tune in response to fluctuating oxygen levels and metabolic demands in multicellular organisms.

Pharmacological inhibitors

Pharmacological inhibitors of respiratory complex I target various sites within the enzyme, primarily the ubiquinone (Q)-binding channel, to block from NADH to ubiquinone and disrupt proton pumping. These compounds have been instrumental in elucidating the enzyme's mechanism and hold potential for therapeutic applications in cancer and treatments. Rotenone, a natural derived from plant roots, is a prototypical high-affinity inhibitor that binds within the Q-reduction site near the N2 iron-sulfur cluster, involving interactions with conserved residues such as Tyr108 and His59 in the NDUFS2 subunit. This binding sterically hinders ubiquinone access and electron transfer, with an in the low nanomolar range in bovine mitochondrial assays. Piericidin A, an α-pyridone produced by actinomycetes, similarly occupies the Q-binding site at the tunnel's apex, blocking N2 cluster access and ubiquinone reduction, also achieving low nanomolar values; structural studies reveal that may involve a secondary site deeper in the Q-tunnel. ADP-ribose acts as a physiological, reversible competitive inhibitor of NADH oxidation at the FMN-binding site in the enzyme's peripheral arm, with inhibition constants in the micromolar range, potentially modulating complex I activity under conditions. Metformin, an antidiabetic , exerts partial inhibition through binding in the amphipathic Q-channel near the membrane interface, displacing structural elements like the ND5 and preferring the deactive enzyme state, though its potency is weak with values exceeding 10 mM in isolated membranes. Stigmatellin, primarily an of complex III, shows off-target effects on complex I at higher concentrations as a B , noncompetitively blocking activity without direct competition at the Q-site. Cryo-EM structures of bacterial and mammalian complex I have mapped inhibitor binding to the Q-cleft, a narrow tunnel extending from the , informing design of selective agents. In recent developments, IACS-010759 represents a potent, selective small-molecule that binds tightly within the Q-cavity via a "cork-in-bottle" mechanism, achieving subnanomolar potency; it underwent phase I clinical trials for starting in 2016 but development was halted due to concerns as of 2025. These , particularly and piericidin analogs, also show promise in applications by targeting parasite complex I variants.

Reactive oxygen species

Superoxide generation mechanisms

Respiratory complex I generates primarily through the reduction of molecular oxygen by the (FMN) cofactor at site I_F, particularly via its reduced forms, including the FMN semiquinone intermediate, which exhibits a high conducive to electron leakage. A secondary site, I_Q at the quinone-binding channel, also contributes to production, especially during forward when the ubiquinone pool is reduced. This process occurs at the FMN site in the enzyme's hydrophilic arm, where electrons from NADH can divert off the main transfer pathway to react with O₂, forming (O₂⁻•) as the initial product, with subsequent dismutation potentially yielding . Studies on isolated bovine complex I confirm that the terminal iron-sulfur cluster N₂ does not significantly contribute to production, reinforcing the FMN and Q sites as the dominant locations. Superoxide production is modulated by the direction of electron flow within complex I. In the forward mode, driven by NADH oxidation under normoxic conditions, superoxide generation remains low, typically accounting for approximately 1-2% of electrons diverted from the catalytic cycle to O₂. Conversely, during reverse electron transfer (RET)—facilitated by a highly reduced ubiquinone pool and elevated proton-motive force, as seen in hypoxic or ischemic states—superoxide output increases dramatically, up to several-fold higher than in forward mode, due to hyper-reduction of the FMN site. RET-driven superoxide primarily emanates from the fully reduced FMN-hydroquinone (FMNH₂), serving as a key intermediate that reacts with O₂. Kinetic analyses using pulse radiolysis have elucidated the reaction rates at the FMN site, revealing second-order rate constants for the interaction between reduced FMN species and O₂ on the order of 10³ M⁻¹ s⁻¹, which limits superoxide formation to a minor fraction of overall electron flux. These measurements highlight the FMN-hydroquinone as the reactive species, with superoxide production rates in isolated complex I reaching about 0.15% of maximal catalytic turnover under saturating NADH conditions (based on studies in Yarrowia lipolytica). Studies have shown that the mitochondrial matrix NADH/NAD⁺ ratio plays a key role in partitioning between forward and reverse modes. Elevated NADH/NAD⁺ ratios, indicative of reductive stress, enhance FMN and thereby amplify leakage during forward , while favoring RET under high ; this dynamic positions complex I as a sensor responsive to metabolic state.

Physiological and pathological roles

Complex I-derived (ROS) play essential physiological roles at low levels, acting as signaling molecules that facilitate and maintain . In hypoxic conditions, modest ROS production from Complex I stabilizes hypoxia-inducible factor-1α (HIF-1α), promoting adaptive responses such as and metabolic reprogramming to enhance oxygen utilization. This signaling mechanism helps cells sense and respond to environmental stressors, ensuring survival under low-oxygen states. Additionally, low-level Complex I ROS contribute to by modulating defenses and thiol-based signaling pathways, preventing excessive oxidative damage while supporting normal mitochondrial function. Recent studies (as of 2025) have further highlighted Complex I ROS as metabolic sensors in immune cells, where they fine-tune activation and effector functions, such as in macrophages, by integrating nutrient availability with inflammatory signaling during chronic stimulation. In pathological contexts, elevated ROS from Complex I drive cellular damage and dysfunction. High ROS levels oxidize mitochondrial DNA (mtDNA), leading to mutations and impaired respiratory chain assembly, which exacerbates energy deficits in affected cells. This oxidative stress also triggers apoptosis through activation of the intrinsic mitochondrial pathway, involving cytochrome c release and caspase cascades, particularly in response to prolonged metabolic insults. Furthermore, during ischemia-reperfusion injury, burst-like Complex I ROS production via reverse electron transport contributes to tissue damage by promoting inflammation and necrosis in organs like the heart and brain. Modulation of Complex I ROS represents a therapeutic strategy to balance its dual roles, with mitochondria-targeted antioxidants like MitoQ selectively scavenging superoxide at this site to mitigate oxidative harm without fully disrupting electron transport. Evolutionarily, the propensity for Complex I to generate ROS reflects a between high respiratory efficiency and the risk of oxidative leakage, allowing rapid energy production but constraining in high-metabolism organisms.

Pathological implications

Associated mitochondrial diseases

Respiratory complex I deficiency is the most common cause of mitochondrial disease in children, accounting for approximately 30% of cases, and manifests primarily as primary genetic disorders such as Leigh syndrome. Leigh syndrome, the most prevalent presentation, affects about 1 in 40,000 individuals and is linked to complex I defects in roughly 30% of cases, often due to mutations in mitochondrial DNA (mtDNA) genes encoding complex I subunits, particularly the MT-ND genes (e.g., MT-ND1, MT-ND2, MT-ND3, MT-ND4, MT-ND5, MT-ND6). Complex I deficiency accounts for approximately 30% of pediatric mitochondrial diseases, which have an overall prevalence of about 1 in 5,000 births. Clinical features typically include neonatal-onset lactic acidosis, psychomotor regression, hypotonia, seizures, and characteristic bilateral symmetric lesions in the basal ganglia and brainstem visible on neuroimaging. Other associated disorders include (LHON), which primarily causes acute or subacute bilateral vision loss due to from mutations in MT-ND1, MT-ND4, or MT-ND6 (e.g., m.11778G>A in MT-ND4), and may occasionally overlap with Leigh-like features. Some MT-ND mutations, such as those in MT-ND5, can present with MELAS (mitochondrial encephalomyopathy, , and stroke-like episodes) overlap syndromes, featuring recurrent stroke-like episodes alongside and . Isolated complex I-deficient are also reported, characterized by progressive , , and elevated serum lactate, without predominant neurological involvement. Diagnosis of these complex I-related diseases relies on a combination of clinical evaluation, biochemical testing, and genetic analysis. assays measuring NADH:ubiquinone oxidoreductase activity in fibroblasts, muscle, or blood leukocytes confirm reduced complex I function, often below 30% of control values. Muscle may reveal ragged-red fibers, subsarcolemmal accumulation of mitochondria, and (COX) deficiency on histochemistry. Definitive involves next-generation sequencing (NGS) of mtDNA and DNA to identify pathogenic variants, such as heteroplasmic MT-ND with heteroplasmy levels typically exceeding 70-80% in affected tissues. Recent advances in 2024 include preclinical approaches using (AAV) vectors for allotopic expression of MT-ND5 in models caused by mtDNA mutations, demonstrating restored complex I , improved ATP , and extended survival in mice without significant off-target effects. As of 2025, emerging enzyme-based technologies for precise mtDNA in patient-derived cells show promise for correcting complex I mutations, complementing AAV approaches. These strategies relocate the mutant mtDNA to the for cytosolic transcription and mitochondrial import of the protein, offering promise for monogenic mtDNA disorders, though clinical translation remains challenged by and tissue-specific delivery. Respiratory complex I dysfunction plays a central role in () pathogenesis, particularly through its contribution to the selective degeneration of neurons in the . Postmortem analyses of PD brains reveal a 30-40% reduction in complex I activity specifically in this region, which correlates with approximately 50-60% neuronal loss at the symptomatic onset of motor impairments. This deficiency is not uniform across the brain but is most pronounced in vulnerable dopaminergic populations, exacerbating energy deficits and promoting . The toxin , which induces a parkinsonian syndrome, is converted to MPP+ that potently inhibits complex I in a manner akin to , thereby replicating the mitochondrial impairment observed in idiopathic PD and underscoring complex I as a key therapeutic target. In (AD), amyloid-β (Aβ) oligomers inhibit the enzymatic activity of complex I, disrupting electron transport in neuronal mitochondria. This inhibition precedes plaque formation in AD mouse models and contributes to synaptic dysfunction and cognitive decline by reducing ATP production and amplifying . Therapeutic interventions targeting complex I, including supplementation, have demonstrated neuroprotective effects in preclinical AD models by restoring mitochondrial function, reducing Aβ toxicity, and improving memory performance, although large-scale human trials are ongoing to validate these benefits. Aging is associated with a progressive decline in complex I activity, with reported reductions of 20-50% in various tissues, including and muscle, by advanced age, driven by somatic mutations that accumulate over time and lead to impaired assembly and function. This age-related deterioration fosters a of mitochondrial inefficiency, distinct from inherited deficiencies, and heightens susceptibility to neurodegenerative conditions. A 2024 multicenter study stratifying idiopathic patients by complex I deficiency levels identified it as a robust for disease subtypes and progression, with severe deficiencies predicting non-tremor-dominant phenotypes and faster decline, informing potential monitoring strategies in aging populations.

Occurrence in other organisms

Bacterial NADH dehydrogenases

Bacterial NADH dehydrogenases, also known as NDH-1 or type I NADH:quinone oxidoreductases, serve as prokaryotic homologs of mitochondrial respiratory complex I, embedded in the plasma to facilitate . These enzymes are encoded by the conserved nuo , which typically comprises 14 core subunits designated NuoA through NuoN, forming a ~500 kDa L-shaped complex with a peripheral for and a for proton translocation. Prominent examples include the complexes from and Thermus thermophilus, which have been extensively studied as minimal models due to their structural simplicity and genetic tractability. Functionally, bacterial NADH dehydrogenases catalyze the transfer of electrons from NADH to (ubiquinone or menaquinone, depending on the ), coupled to proton pumping across the , generating a proton motive force for ATP synthesis. The proton-to-electron varies but is commonly 4 H⁺ per 2 e⁻, enabling efficient energy coupling in aerobic ; however, some variants exhibit 2–3 H⁺/2 e⁻, reflecting adaptations to environmental conditions. In facultative anaerobes like Rhodobacter sphaeroides, these enzymes also support anaerobic respiration by linking NADH oxidation to alternative electron acceptors such as fumarate, highlighting their versatility in diverse metabolic contexts. Structural studies have provided key mechanistic insights, with the first atomic model of the complete bacterial complex from T. thermophilus determined in 2010 at 4.5 Å resolution using , revealing 63 transmembrane helices in the membrane arm and conserved iron-sulfur clusters in the peripheral arm. Unlike eukaryotic complex I, bacterial versions lack accessory (supernumerary) subunits, resulting in a more streamlined architecture and simpler regulation without the extensive post-translational modifications or assembly factors seen in mitochondria. This minimal design has made bacterial complexes ideal for biophysical and mechanistic investigations, elucidating how long-range conformational changes propagate from the quinone-binding site to drive proton pumping through antiporter-like subunits (NuoL, NuoM, NuoN). Beyond fundamental biology, bacterial NADH dehydrogenases represent promising therapeutic targets, particularly in pathogenic species like Mycobacterium tuberculosis, where complex I is essential for survival under and , and inhibitors disrupting its activity show potential as novel antibiotics against drug-resistant strains.

Chloroplast NDH complexes

The chloroplast NAD(P)H dehydrogenase-like (NDH) complex is a large multi-subunit embedded in the thylakoid membranes of higher chloroplasts, serving as a :plastoquinone reductase that facilitates cyclic electron flow around (PSI). This process recycles electrons from back to , bypassing and thereby enhancing ATP production without net NADPH synthesis, which supports under varying light conditions. The complex's activity is crucial for maintaining balance in the photosynthetic , particularly during high-light stress when linear electron flow may generate excess reducing power. Composed of approximately 29 subunits organized into five subcomplexes, the NDH complex includes 11 plastid-encoded subunits (NdhA through NdhK, homologous to bacterial NDH-1 components) and at least 18 nuclear-encoded subunits unique to plants, such as NDH-M, -N, and -O, which are essential for assembly and function. These nuclear-encoded subunits enable specific adaptations, including interactions with light-harvesting complexes. In contrast to mitochondrial complex I, which pumps four protons per two electrons, the chloroplast NDH complex exhibits a more limited proton-pumping capacity of about two protons per two electrons, prioritizing photoprotection through ΔpH generation for over maximal . This reduced pumping aligns with its role in stromal acidification to regulate electron flow and prevent over-reduction of the pool. The NDH complex performs several key functions beyond cyclic electron transport, including support for CO₂ concentration mechanisms in C₄ plants, where it sustains a high ATP/ADP ratio in bundle sheath cells to enhance carbon fixation efficiency. It also contributes to (ROS) scavenging by oxidizing excess reductants and mitigating photooxidative damage, as evidenced by increased ROS accumulation and photosynthetic impairment in NDH-deficient mutants under high light. Notably, the NDH complex is absent in the plastids of most eukaryotic , with ndh genes lost from all sequenced algal plastid genomes except those of and certain Prasinophyceae, reflecting evolutionary dispensability in non-vascular photosynthetic organisms. Recent cryo-electron microscopy (cryo-EM) structures from 2022 have elucidated the architecture of the NDH complex as part of a hybrid PSI-NDH supercomplex in barley (Hordeum vulgare), revealing how up to two PSI cores associate with one NDH unit via stromal-facing protrusions, stabilized by shared lipids and chlorophylls for efficient electron transfer. These structures highlight plant-specific subunits that facilitate supercomplex formation, underscoring the NDH's integration into the photosynthetic apparatus for coordinated cyclic flow. A more recent 2025 cryo-EM structure of the NDH–PSI–LHCI supercomplex from spinach (Spinacia oleracea) at 3.0–3.3 Å resolution further details the assembly, consisting of 41 protein subunits reinforced by 46 lipids, providing additional insights into subunit interactions and lipid mediation.

Genetics

Encoding genes in humans

In humans, respiratory complex I (NADH:ubiquinone ) is encoded by a combination of mitochondrial and genes, reflecting its dual-genome origin. The seven core subunits derived from (mtDNA) are essential components of the membrane arm, while the 38 nuclear-encoded subunits include both core and accessory proteins that form the peripheral and membrane arms, as well as structural and regulatory elements. These 38 subunits are produced from 37 autosomal genes, as the NDUFAB1 encodes two copies of its acyl carrier protein-like subunit. The mtDNA-encoded genes are MT-ND1, MT-ND2, MT-ND3, MT-ND4, MT-ND4L, MT-ND5, and MT-ND6. These genes reside on the circular, maternally inherited mtDNA molecule and produce proteins synthesized directly within the mitochondrial matrix. Transcription of these genes occurs as part of long polycistronic transcripts from the heavy and light strands of mtDNA, which are processed into mature mRNAs for the 13 mtDNA-encoded respiratory chain proteins, including the seven for complex I. The genome encodes the majority of complex I subunits via 37 autosomal genes, whose products are translated in the and subsequently imported into mitochondria via specific targeting signals. These subunits are organized into functional families based on their roles in , proton pumping, and . For instance, the iron-sulfur (Fe-S) cluster-containing subunits are encoded by NDUFS1 through NDUFS8, which harbor the redox-active Fe-S centers critical for from NADH. Accessory subunits, often involved in and , include those from the NDUFA1 through NDUFA13 family, such as NDUFA1 (MWFE) and NDUFA13 (GRIM-19), which contribute to the matrix-facing arm. Other families encompass NDUFV1 and NDUFV2 (flavin mononucleotide-binding subunits), NDUFB1 through NDUFB11 (membrane arm accessories), NDUFC1 and NDUFC2 (core membrane components), and NDUFAB1 (acyl carrier protein-like subunit, with two copies). This nuclear contribution ensures the complex's ~980 mass and L-shaped architecture. Expression of the nuclear-encoded complex I genes is tightly coordinated by the transcriptional coactivator peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α), which integrates environmental signals to drive . PGC-1α interacts with respiratory factors (NRF1 and NRF2) and estrogen-related receptors (ERRα/γ) to upregulate these genes, ensuring balanced synthesis with mtDNA-encoded subunits. In contrast, mtDNA transcription is regulated independently by mitochondrial transcription factor A (TFAM), though PGC-1α indirectly influences it via NRF1-mediated TFAM expression. The MitoCarta database catalogs over 1,100 human mitochondrial proteins, including all I subunits, with annotations for their subcellular localization, functional domains, and expression patterns based on and GFP-tagging data. This resource facilitates identification and validation of encoding genes in mitochondrial studies.

Mutations and genetic variants

in genes encoding subunits or assembly factors of respiratory I are a leading cause of isolated I deficiency, accounting for approximately 25% of mitochondrial respiratory chain disorders in children. These can occur in either (mtDNA) or nuclear DNA (nDNA), with mtDNA variants exhibiting —the coexistence of and wild-type mtDNA molecules—where manifestation typically requires a mutant load exceeding 70% in affected tissues due to the threshold effect. Common mutation types include missense variants that alter amino acid sequences, often disrupting iron-sulfur (Fe-S) cluster coordination essential for , and frameshift or mutations leading to premature termination and truncated proteins that impair stability. A well-documented mtDNA variant is the m.13513G>A missense mutation in MT-ND5, which substitutes for (p.D393N) in the ND5 subunit, frequently associated with and characterized by variable levels correlating with clinical severity. In nDNA, mutations in NDUFS1, encoding the 75 kDa Fe-S subunit, are prevalent; for instance, the homozygous c.1222C>T (p.R408C) has been identified in an Asian child with Leigh-like , resulting in profoundly reduced I activity in fibroblasts and brain tissue. These examples highlight how specific variants target critical structural domains, such as the Q-module in ND5 or Fe-S binding sites in NDUFS1. The molecular consequences of these mutations often involve defective complex I assembly, leading to accumulation of subcomplexes and decreased holoenzyme formation, as seen in NDUFS1 variants where Fe-S cluster instability reduces NADH oxidation efficiency. Frameshift mutations, such as those introducing premature stop codons, exacerbate this by promoting proteasomal degradation of nascent subunits, further destabilizing the complex. Additionally, mutant complex I generates elevated (ROS) due to electron leakage from disrupted Fe-S centers, contributing to oxidative damage in high-energy tissues like the . Genotype-phenotype correlations reveal that mtDNA mutations like m.13513G>A tend to present with encephalomyopathy at lower heteroplasmy thresholds in neural cells (~60-80%), while severe nDNA biallelic variants, such as in NDUFS1, often cause earlier-onset, rapidly progressive disease with poor prognosis. Recent studies have advanced understanding of assembly defects, with biallelic loss-of-function in NDUFAF2, an chaperone for the N-module attachment during complex I biogenesis, reported in 2024 as causing a severe infantile brainstem-predominant neurodegenerative disorder akin to , underscoring the role of assembly factors in pathogenesis. Experimental approaches, including /Cas9-mediated knockout models of complex I genes like NDUFS4 in induced pluripotent stem cells, have recapitulated assembly failures and ROS overproduction, providing platforms to test therapeutic interventions for these variants. Such mutations link to mitochondrial diseases like , where complex I deficiency manifests as subacute necrotizing encephalomyelopathy.

References

  1. [1]
    Mitochondrial Respiratory Complex I: Structure, Function and ...
    This review provides an updated overview of the structure of complex I, as well as its cellular functions, and discusses the implication of complex I ...
  2. [2]
    Structure and function of mitochondrial complex I - ScienceDirect.com
    Complex I is the largest multi-subunit complex of the respiratory chain. Complex I couples electron transfer from NADH to ubiquinone with transmembrane proton ...
  3. [3]
    Structure of inhibitor-bound mammalian complex I - Nature
    Oct 16, 2020 · Respiratory complex I (NADH:ubiquinone oxidoreductase) captures the free energy from oxidising NADH and reducing ubiquinone to drive protons ...
  4. [4]
    EC 7.1.1.2 - IUBMB Nomenclature
    Comments: The enzyme is a very large complex that participates in electron transfer chains of mitochondria and aerobic bacteria, transferring two electrons ...Missing: definition | Show results with:definition
  5. [5]
    Structure of mammalian respiratory complex I - PMC - PubMed Central
    Mammalian complex I 1 contains 45 subunits, comprising 14 core subunits that house the catalytic machinery and are conserved from bacteria to humans.
  6. [6]
    Studies on the electron transfer system. XL. Preparation ... - PubMed
    Studies on the electron transfer system. XL. Preparation and properties of mitochondrial DPNH-coenzyme Q reductase. J Biol Chem. 1962 May:237:1676-80.Missing: et al. complex paper
  7. [7]
    Structure of the bacterial type II NADH dehydrogenase
    Jan 3, 2014 · These include the proton-pumping type I NADH dehydrogenase (NDH-1, complex I), the non-proton pumping type II NADH dehydrogenase (NDH-2) and the ...Missing: synonyms distinction<|separator|>
  8. [8]
    Structural insight into the type-II mitochondrial NADH dehydrogenases
    Nov 15, 2012 · The single-component type-II NADH dehydrogenases (NDH-2s) serve as alternatives to the multisubunit respiratory complex I (type-I NADH ...
  9. [9]
    Atomic structure of the entire mammalian mitochondrial complex I
    It is the largest protein assembly of the respiratory chain with total mass of 970 kDa. Here we present a nearly complete atomic structure of ovine ...
  10. [10]
    The Evolution of Respiratory Chain Complex I from a Smaller Last ...
    A complex I-homologous enzyme found in some archaea contains an F420 dehydrogenase subunit denoted as FpoF rather than the N-module. In the present study ...
  11. [11]
    Evolution of complex I–like respiratory complexes - PubMed Central
    The modern-day respiratory complex I shares a common ancestor with the membrane-bound hydrogenase (MBH) and membrane-bound sulfane sulfur reductase (MBS).
  12. [12]
    The respiratory complex I of bacteria, archaea and eukarya and its ...
    Aug 11, 2000 · The homologues of the seven hydrophobic subunits (NuoA, H and J–N, Table 1) are mitochondrially encoded in all eukaryotes. The seven ...
  13. [13]
    Respiratory complex I – Mechanistic insights and advances in ...
    Fourteen central subunits are conserved from bacteria to humans and harbor the bioenergetic core functions (Table 1, Fig. 1A). The central subunits are assigned ...Respiratory Complex I... · 2. Complex I Function And... · 5. Ubiquinone Access And...
  14. [14]
    Key role of quinone in the mechanism of respiratory complex I - Nature
    Aug 18, 2020 · Our results suggest that quinone binding and chemistry play a key role in the coupling mechanism of complex I.
  15. [15]
    The Mysterious Multitude: Structural Perspective on the Accessory ...
    CI is composed of 14 core subunits that are conserved across species and an increasing number of accessory subunits from bacteria to mammals. The fact that ...
  16. [16]
    Identification of a functional respiratory complex in chloroplasts ... - NIH
    Identification of a functional respiratory complex in chloroplasts through analysis of tobacco mutants containing disrupted plastid ndh genes. P A Burrows ...
  17. [17]
    High-resolution cryo-EM structures of respiratory complex I
    The arrangement of 14 central subunits in the L-shaped assembly, consisting of a membrane arm and a matrix arm, is conserved from bacteria to man. The ...
  18. [18]
    High-resolution cryo-EM structures of respiratory complex I - Science
    Dec 11, 2019 · We have determined the structure of complex I from the aerobic yeast Yarrowia lipolytica by electron cryo-microscopy at 3.2-Å resolution.
  19. [19]
    Structure of respiratory complex I – An emerging blueprint for the ...
    Mar 19, 2022 · Complex I catalyzes NADH oxidation and quinone reduction to quinol, couples this reaction to pumping of four protons across the membrane, and is ...
  20. [20]
    Cardiac mitochondrial matrix and respiratory complex protein ...
    Another central subunit of the peripheral arm labeled with PhosTag in heart but not in liver (6) is NDUFS2, which lacks redox cofactors but is phosphorylated in ...<|control11|><|separator|>
  21. [21]
  22. [22]
    Conserved in situ arrangement of complex I and III2 in mitochondrial ...
    Mar 8, 2018 · Our results demonstrate that respiratory chain supercomplexes in situ have a conserved core of complex I and III 2 , but otherwise their stoichiometry and ...
  23. [23]
    Structure of bacterial respiratory complex I - ScienceDirect.com
    The L-shaped complex consists of a hydrophilic arm, where electron transfer occurs, and a membrane arm, where proton translocation takes place.
  24. [24]
    High-resolution structure and dynamics of mitochondrial complex I ...
    Nov 12, 2021 · We here present the cryo-EM structure of complex I from Yarrowia lipolytica at 2.1-Å resolution, which reveals the positions of more than 1600 protein-bound ...Results · Materials And Methods · Cryo-Em Structures
  25. [25]
    Architecture of mammalian respiratory complex I - PMC
    Due to its size, L-shaped asymmetry, membrane-bound location, and multi-component structure, mammalian complex I has proved difficult to crystallise, and its ...Missing: studies | Show results with:studies
  26. [26]
  27. [27]
    RTN4IP1 is required for the final stages of mitochondrial complex I ...
    Aug 26, 2025 · We now identify RTN4IP1 as a novel CI assembly factor that completes the late stages of mitochondrial CI assembly.
  28. [28]
    Mammalian Complex I Pumps 4 Protons per 2 Electrons at High and ...
    The correct stoichiometry for complex I is 4H + /2e − in mouse and human cells at high and physiological proton motive force.
  29. [29]
    Complex I is bypassed during high intensity exercise - Nature
    Nov 7, 2019 · A metabolic bypass of mitochondrial complex I is found to increase the ATP synthesis rate per gram of protein compared to full respiration.
  30. [30]
    Mitochondrial complex I inhibition triggers NAD+-independent ... - NIH
    *Based on the assumption that the oxidation of NADH will provide 2.5 ATP, and that the oxidation of FADH2 will provide 1.5 ATP. Fig. 1. Fig. 1. Open in a new ...
  31. [31]
    Stoichiometry of proton translocation by respiratory complex I and its ...
    The stoichiometry of proton translocation is thought to be 4 H + per NADH oxidized (2 e - ). Here we show that a H + /2 e - ratio of 3 appears more likely.
  32. [32]
    The mechanism of superoxide production by NADH:ubiquinone ...
    We conclude that, in isolated complex I, superoxide is produced by the flavin, because it is not produced by any of the FeS clusters. Fig. 6. Open in Viewer EPR ...Missing: forward | Show results with:forward
  33. [33]
    Oxidation of NADH and ROS production by respiratory complex I
    This short review discusses available information on some features of eukaryotic and bacterial complex I as they appear from studies of the reactions (1 and 2).Missing: yield | Show results with:yield
  34. [34]
    Respiratory Complex I: Mechanistic and Structural Insights Provided ...
    The prokaryotic enzyme is simpler and consists of 13−15 subunits with a combined molecular mass of ∼550 kDa ( 2, 5). Analogues of all conserved subunits of ...<|control11|><|separator|>
  35. [35]
    Ubiquinone Binding and Reduction by Complex I—Open Questions ...
    Apr 30, 2021 · Since, in this scenario, there is a redox potential difference between Q in site 1 and Q in the membrane (+90 mV), the release in free ...
  36. [36]
    Real-time electron transfer in respiratory complex I - PNAS
    The reaction of complex I with NADH was stopped in the time domain from 90 μs to 8 ms and analyzed by electron paramagnetic resonance (EPR) spectroscopy at low ...
  37. [37]
    A mechanism to prevent production of reactive oxygen species by ...
    Jun 11, 2019 · 1b). The electron transfer starts with NADH reducing the FMN via hydride transfer18. Subsequently, the electrons are passed on in several one- ...
  38. [38]
    Mitochondrial iron–sulfur clusters: Structure, function, and an ...
    Iron-sulfur (Fe–S) clusters are essential cofactors most commonly known for their role mediating electron transfer within the mitochondrial respiratory chain.
  39. [39]
    Electron tunneling in respiratory complex I - PNAS
    Oct 25, 2010 · Four electron transfer processes of N1b → N4, N4 → N5, N6a → N6b, and N6b → N2 are faster than both the reported estimates of the total transfer ...
  40. [40]
    Ubiquinone Binding and Reduction by Complex I—Open ... - Frontiers
    Apr 29, 2021 · Complex I couples electron transfer from NADH to quinone (Q) to the translocation of protons across the bioenergetic membrane. Note that some ...
  41. [41]
    The coupling mechanism of mammalian respiratory complex I
    Complex I is the largest of the respiratory complexes and, in mammals, is composed of 45 subunits with a total mass of about 1 MDa (Fig. 1A).
  42. [42]
    Symmetry-related proton transfer pathways in respiratory complex I
    The hydration of these channels is sensitive to the protonation state of conserved buried lysine residues, which are in turn coupled to conformational changes ...
  43. [43]
    Cryo-EM structures define ubiquinone-10 binding to mitochondrial ...
    May 19, 2022 · The Q-binding site in complex I is an unusually long and heterogeneous channel. ... However, the Q10 headgroup is ∼100 Å away from cluster N2 ...Missing: gatekeeper | Show results with:gatekeeper
  44. [44]
    Cryo-EM structure of respiratory complex I at work - eLife
    Oct 2, 2018 · Respiratory complex I is a ~1 MDa membrane protein complex with key functions in aerobic energy metabolism (Hirst, 2013; Wirth et al., 2016).
  45. [45]
    Quinone Catalysis Modulates Proton Transfer Reactions in the ...
    Jul 25, 2023 · Here, we study the coupling between the Q catalysis and conformational changes linked to proton translocation in the membrane domain of Complex ...
  46. [46]
    Binding of Natural Inhibitors to Respiratory Complex I - PMC
    Aug 31, 2022 · Rotenone, piericidin A, and annonaceous acetogenins are representatives of complex I inhibitors from biological sources.
  47. [47]
    Current topics on inhibitors of respiratory complex I - ScienceDirect
    Some inhibitors, such as rotenone and piericidin A, have been indispensable molecular tools in mechanistic studies on complex I.
  48. [48]
    A competitive inhibition of the mitochondrial NADH-ubiquinone ...
    In this report we will characterize ADP-ribose (ADPR) as a competitive inhibitor of NADH oxidation. Inability of ADPR to inhibit the reverse reaction is ...
  49. [49]
    Structural Basis of Mammalian Respiratory Complex I Inhibition by ...
    Here we define the inhibitory drug-target interaction(s) of a model biguanide with mammalian respiratory complex I by combining cryo-electron microscopy and ...
  50. [50]
    Complex I and complex III of mitochondria have common inhibitors ...
    Feb 15, 1993 · The stigmatellin analog is more powerful than its parent compound and is noncompetitive with exogenous ubiquinones, rotenone and piericidin.
  51. [51]
    Cork-in-bottle mechanism of inhibitor binding to mammalian complex I
    May 14, 2021 · IACS-010759 is a highly potent and selective small-molecule inhibitor of complex I that was developed by the MD Anderson Cancer Center (10) and ...
  52. [52]
  53. [53]
    High rates of superoxide production in skeletal-muscle mitochondria ...
    Indeed, the frequently quoted value of 1–5% of electrons passing through the chain being diverted to the formation of superoxide is only true under this ...Missing: percentage | Show results with:percentage
  54. [54]
    Control of mitochondrial superoxide production by reverse electron ...
    The generation of mitochondrial superoxide (O2˙̄) by reverse electron transport (RET) at complex I causes oxidative damage in pathologies such as ischemia ...
  55. [55]
    Mitochondrial Management of Reactive Oxygen Species - PMC
    Even so, the rate of reduction of O2 by the reduced FMN of complex I to form O2•− is ∼40 O2•− min−1 [53], corresponding to a second-order rate constant of ∼103 ...
  56. [56]
    [PDF] Superoxide Radical Formation by Pure Complex I (NADH ...
    Complex I forms superoxide at 0.15% of catalytic turnover, with oxygen accepting electrons from FMNH2 or FMN. FMN is proposed as the reductant.
  57. [57]
    Reactive Oxygen Species Production in Cardiac Mitochondria After ...
    Aug 8, 2025 · Reactive oxygen species (ROS) production by isolated complex I is steeply dependent on the NADH/NAD(+) ratio.
  58. [58]
    Mitochondrial reactive oxygen species regulate hypoxic signaling
    During mitochondrial respiration, electrons from NADH and FADH2 are transferred to mitochondrial complex I and complex II respectively. ... HIF-1α during hypoxia ...
  59. [59]
    Physiological roles of mitochondrial reactive oxygen species - PMC
    Oct 26, 2012 · Mitochondrial ROS have been demonstrated to play a role in varied cellular processes, including differentiation, autophagy, metabolic adaptation ...
  60. [60]
    Mitochondrial DNA damage and reactive oxygen species in ...
    Studies done over the last decades have demonstrated that the link between elevated ROS and mtDNA damage is not clear. ... mitochondrial complex I deficiency due ...
  61. [61]
    Mitochondrial Reactive Oxygen Species (ROS) and ROS-Induced ...
    In other cases, apart from their pro-survival signaling role, ROS are apparently involved in a designated physiological function, eliminating unwanted mitotic ...
  62. [62]
    Mitochondrial redox cycling of mitoquinone leads to superoxide ...
    In isolated mitochondria, MitoQ increased complex I-driven mitochondrial ROS production, whereas supplementation with ubiquinone-10 had no effect on ROS ...
  63. [63]
    Reactive oxygen species as universal constraints in life-history ...
    Feb 25, 2009 · Evolutionary theory is firmly grounded on the existence of trade-offs between life-history traits, and recent interest has centred on the ...
  64. [64]
    Mitochondrial complex I ROS production and redox signaling in ...
    Complex I oxidizes nicotinamide adenine dinucleotide (NADH) and transfers electrons to ubiquinone in a reaction coupled with proton pumping. Complex I also ...
  65. [65]
    Complex I deficiency: clinical features, biochemistry and molecular ...
    Complex I deficiency is the most frequent mitochondrial disorder presenting in childhood, accounting for up to 30% of cases.
  66. [66]
    Mitochondrial Diseases: Types, Diagnosis, and New Therapies
    Jun 4, 2025 · According to DelveInsight, Leigh syndrome prevalence is ~1 in 40K, with 80% of cases being the classical form. Higher incidence occurs in ...
  67. [67]
    The genetics of Leigh syndrome and its implications for clinical ...
    Deficiency of complex I is the most commonly identified defect in childhood-onset mitochondrial disease and accounts for approximately a third of all cases of ...
  68. [68]
    Mitochondrial diseases: from molecular mechanisms to therapeutic ...
    Jan 10, 2025 · This review focuses on the physiological mechanisms of mitochondria, the pathogenesis of mitochondrial diseases, and potential diagnostic and therapeutic ...
  69. [69]
    Gene therapy for mitochondrial disorders - Wiley Online Library
    Jan 3, 2024 · In this review, we detail the current state of application of gene therapy to primary mitochondrial disorders (PMDs).<|control11|><|separator|>
  70. [70]
    Randomized, Double-blind, Placebo-Controlled Trial on ...
    May 14, 2007 · Indeed, a 30% to 40% reduction of complex I activity in substantia nigra of PD brains has been found, a defect that does not affect other parts ...
  71. [71]
    Clinical Progression in Parkinson's Disease and the Neurobiology of ...
    Thus, there is a good consistency in the available data to suggest that the motor signs of PD appear when there is about a 30% loss of total SN neurons in ...
  72. [72]
    Toxin Models of Mitochondrial Dysfunction in Parkinson's Disease
    Within DA neurons, the toxins MPP+ and rotenone inhibit mitochondrial complex I (c-I), decreasing endogenous antioxidants and generating oxidative stress from ...
  73. [73]
  74. [74]
    Mitochondrial Complex I and β-Amyloid Peptide Interplay in ...
    The so-called amyloid plaques, i.e., accumulations of β-amyloid (Aβ) protein outside neurons, are clumps that form between neurons and damage surrounding cells, ...
  75. [75]
    CoQ10 and Mitochondrial Dysfunction in Alzheimer's Disease - PMC
    CoQ 10 can be considered promising in restoring the mitochondrial function impaired in AD, or in preventing the onset of mitochondrial dysfunction.
  76. [76]
    Decline in skeletal muscle mitochondrial function with aging ... - PNAS
    MAPR and Citrate Synthase Activity.​​ MAPR per gram of muscle declined with age ≈8% per decade by using substrates that supply electrons primarily to complex I ( ...Materials And Methods · Results · Discussion
  77. [77]
    Mitochondrial complex I deficiency stratifies idiopathic Parkinson's ...
    Apr 29, 2024 · We show that iPD can be stratified according to the severity of neuronal respiratory complex I (CI) deficiency, and identify two emerging disease subtypes.
  78. [78]
    Structure of Escherichia coli respiratory complex I reconstituted into ...
    Jul 26, 2021 · The membrane-embedded arm includes a chain of three antiporter-like subunits, NuoL, NuoM, and NuoN (E. coli nomenclature is used for the ...
  79. [79]
    Different Functions of Phylogenetically Distinct Bacterial Complex I ...
    Complex I activity is required for anaerobic growth of R. sphaeroides. The R. sphaeroides genome has two operons that encode complex I homologues (Fig. 1A). One ...
  80. [80]
    The architecture of respiratory complex I - Nature
    The architecture of respiratory complex I. Download PDF. Article; Published: 27 May 2010. The architecture of respiratory complex I. Rouslan G. Efremov, ...
  81. [81]
    Structure of mycobacterial respiratory complex I - PNAS
    The core Complex I structure consists of 14 subunits, with 31 supernumerary subunits in mammalian mitochondria (28, 34), and total molecular masses ranging from ...Sign Up For Pnas Alerts · Results · Active Mycobacterial Complex...
  82. [82]
    An Src Homology 3 Domain-Like Fold Protein Forms a Ferredoxin ...
    This study indicates that CRR31, a novel NDH subunit, serves in the high-affinity binding of ferredoxin to NDH and proposes that chloroplast NDH is ferredoxin: ...
  83. [83]
    The Significance of Chloroplast NAD(P)H Dehydrogenase Complex ...
    Apr 22, 2021 · The significance of the NDH-mediated cyclic electron transport is discussed from the perspective of plant evolution and physiological ...
  84. [84]
    Chloroplast NDH: A different enzyme with a structure similar to that ...
    Structural evolution of photosynthetic NDH. Chloroplast NDH originated from the cyanobacterial NDH-1 complex. In Synechocystis PCC6803, NDH-1 is involved in the ...
  85. [85]
    Supramolecular assembly of chloroplast NADH dehydrogenase-like ...
    Mar 7, 2022 · In angiosperms, the NDH complex is composed of at least 29 subunits, which can be divided into five subcomplexes (Ifuku et al., 2011; Shikanai, ...
  86. [86]
    Evolution of an assembly factor-based subunit contributed to a novel ...
    Jun 17, 2021 · In contrast, the molecular size of the chloroplast NDH complex has gradually increased during the evolution of land plants. Our results ...
  87. [87]
    The higher plant plastid NAD(P)H dehydrogenase-like complex ...
    We demonstrate that the chloroplast NADPH dehydrogenase complex, a homolog to respiratory Complex I, pumps approximately two protons from the chloroplast stroma ...
  88. [88]
    Chloroplast NADH dehydrogenase‐like complex‐mediated cyclic ...
    Jul 22, 2024 · During the evolutionary transition from C3 to C4 plants, NDH levels increased concurrently with the emergence of the C4 cycle as a CO2 pump (Lyu ...
  89. [89]
    Chloroplastic NAD(P)H Dehydrogenase in Tobacco Leaves ...
    However, there was no direct evidence for involvement of the NDH pathway in ROS scavenging. This work provides direct evidence for the involvement of the ...
  90. [90]
    Plastid ndh genes in plant evolution - ScienceDirect.com
    The ndh genes are absent in all sequenced plastid DNAs of algae except for the Charophyceae and some Prasinophyceae.
  91. [91]
    The nuclear encoded subunits of complex I from bovine heart ...
    The sequences of 45 subunits of complex I have been determined. Seven of them are encoded by mitochondrial DNA, and 38 are products of the nuclear genome.
  92. [92]
    Human CIA30 is involved in the early assembly of mitochondrial ...
    Of the complex I subunits, seven are encoded by mitochondrial DNA (mtDNA) whereas the remainder are encoded by nuclear genes, translated in the cytosol and ...
  93. [93]
    Errα and Gabpa/b specify PGC-1α-dependent oxidative ... - PNAS
    Fig. 2. Proposed model of mechanism of action of PGC-1α. PGC-1α is a highly regulated gene that responds to external stimuli, e.g., reduced in diabetes and ...
  94. [94]
    Mitochondrial DNA heteroplasmy in disease and targeted nuclease ...
    Feb 19, 2020 · This threshold is dependent on the mutation, the cell type, and tissue type, and can vary between 60 and 90% mutant mtDNA for an detectable ...
  95. [95]
    The mitochondrial 13513G > A mutation is most frequent in Leigh ...
    The mitochondrial 13513G > A mutation is most frequent in Leigh syndrome combined with reduced complex I activity, optic atrophy and/or Wolff-Parkinson-White.Missing: 12308A> | Show results with:12308A>
  96. [96]
    The p.M292T NDUFS2 mutation causes complex I-deficient Leigh ...
    NDUFS2 already contains a conserved protein kinase C phosphorylation site (soluble leptin receptor) at the C-terminus of the protein (Loeffen et al., 1998b); ...
  97. [97]
    Natural disease course and genotype-phenotype correlations in ...
    May 30, 2012 · Mutations in one of the nuclear encoded structural or assembly genes of complex I have a dramatic effect on neurodevelopment and overall patient ...<|control11|><|separator|>
  98. [98]
    Lack of mitochondrial complex I assembly factor NDUFAF2 results in ...
    Feb 28, 2024 · To date, only 14 patients with loss-of-function mutations in NDUFAF2 (including our 4 patients) have been reported.Missing: CRISPR | Show results with:CRISPR
  99. [99]
    CRISPR-Cas9 mediated knockout of NDUFS4 in human iPSCs
    Here, we present a novel iPSC model of NDUFS4-related CI deficiency that displays a strong metabolic phenotype in the pluripotent state.