Fact-checked by Grok 2 weeks ago

Electron transport chain

The electron transport chain (), also known as the respiratory chain, is a series of membrane-bound protein complexes and mobile carriers that transfer high-energy electrons derived from NADH and FADH₂ to molecular oxygen (O₂), the terminal , while simultaneously pumping protons across the to establish an essential for ATP production via . This process occurs primarily in the mitochondria of eukaryotic cells and represents the final stage of aerobic , efficiently converting the energy from nutrient breakdown into usable ATP, with approximately 30–32 molecules of ATP generated per glucose molecule oxidized. In prokaryotes, a similar ETC operates in the plasma membrane, and analogous systems exist in chloroplasts for . The ETC comprises four large protein complexes (I–IV), along with the lipid-soluble carrier ubiquinone (coenzyme Q) and the water-soluble , all embedded in or associated with the . Complex I () initiates the chain by accepting electrons from NADH (produced in the and ), passing them through iron-sulfur clusters and (FMN) to reduce ubiquinone while translocating four protons per two electrons. Complex II (), the only complex not pumping protons directly, feeds electrons from FADH₂ (generated in the ) into ubiquinone without contributing to the initial proton gradient. Electrons then move via reduced ubiquinone to Complex III (cytochrome bc₁ complex), which uses the Q-cycle mechanism to transfer them to while pumping four additional protons. Finally, Complex IV () accepts electrons from , reduces O₂ to water, and pumps two more protons, completing the chain and preventing electron leakage that could generate harmful (ROS). Beyond ATP synthesis, the ETC plays critical roles in cellular , including ROS production as a (primarily at Complexes I and III), which serves as signaling molecules but can contribute to , aging, and diseases like Parkinson's when dysregulated. Recent structural studies, including cryo-electron microscopy of supercomplexes (e.g., I₁III₂IV₁), have revealed how these assemblies enhance efficiency through substrate channeling and stabilization by lipids, underscoring the ETC's evolutionary optimization for energy conservation. Disruptions in ETC function, such as mutations in complex subunits, are implicated in mitochondrial disorders, highlighting its indispensability for aerobic life.

Introduction

Definition and biological role

The electron transport chain (ETC) is a series of multi-protein complexes embedded in cellular membranes that catalyze sequential reactions, transferring electrons from donor molecules to acceptor molecules while coupling this process to proton translocation across the membrane. In eukaryotic cells, the ETC resides in the ; in prokaryotes, it is located in the plasma membrane; and in photosynthetic organisms, it occurs in the membrane of chloroplasts. This conserved mechanism underpins in and in , enabling efficient energy conversion across diverse life forms. The primary biological role of the ETC is to generate a proton motive force—an () across the membrane—through proton pumping driven by exergonic electron transfers, which powers ATP synthesis via . In aerobic respiration, electrons derived from NADH and FADH₂ (produced in , the , and fatty acid oxidation) flow through the chain to oxygen as the terminal acceptor, yielding and facilitating the production of approximately 30–32 ATP molecules per glucose molecule oxidized. The overall reaction for the mitochondrial ETC can be summarized as:
\ce{NADH + 1/2 O2 + H+ -> NAD+ + H2O}
(with associated proton pumping omitted for simplicity). In photosynthesis, electrons originate from (split by light energy) and are transferred via to NADP⁺, producing NADPH and oxygen, while the proton gradient drives ATP formation essential for carbon fixation.
This universal process is vital for cellular energy homeostasis, as disruptions in the ETC can impair ATP production and lead to metabolic disorders, underscoring its evolutionary conservation from bacteria to higher eukaryotes.

Historical background

The discovery of the electron transport chain (ETC) emerged from early 20th-century studies on cellular respiration. In the 1920s, Otto Warburg investigated oxygen consumption in living cells using manometric techniques, revealing that respiration involves an iron-containing enzyme that facilitates atmospheric oxygen reduction, for which he received the 1931 Nobel Prize in Physiology or Medicine. Concurrently, David Keilin identified cytochromes as key respiratory pigments in 1925, observing their characteristic absorption bands in yeast, insects, and vertebrates, establishing them as ubiquitous components of aerobic respiration. In the 1940s and , advances in mitochondrial isolation enabled fractionation and component identification. and colleagues developed methods to disrupt mitochondria into functional assemblies, isolating the "cyclophorase" system in 1951 that integrated the with . Britton Chance applied rapid-mixing to elucidate cytochrome kinetics, demonstrating sequential reactions in the respiratory chain during the . The role of quinones was clarified with the 1957 discovery of ubiquinone (coenzyme Q) by Frederick L. Crane, , and associates, who isolated it from beef heart mitochondria as a lipid-soluble electron essential for NADH and succinate oxidation. Efraim Racker's group isolated the F1 portion of from mitochondria in 1960, reconstituting it into vesicles to confirm its role in ATP synthesis. Peter Mitchell proposed the chemiosmotic hypothesis in 1961, positing that ETC-driven proton translocation across the generates an powering ATP synthesis, a concept validated over the next decade and awarded the 1978 . The 1970s and 1980s saw structural insights through and early ; for instance, spectroscopic studies in the 1970s and 1980s mapped electron pathways in Complex I (NADH:ubiquinone ). Recent decades have benefited from cryo-electron microscopy (cryo-EM) for high-resolution structures. Judy Hirst's laboratory determined the near-atomic structure of mammalian Complex I in 2016, revealing its L-shaped architecture and proton-pumping mechanism. In the 2020s, single-molecule techniques have provided dynamic views; for example, tracking of bacterial outer-membrane in 2021 illuminated protein diffusion and rates in respiratory chains. These milestones continue to refine our understanding of assembly and function.

General principles

Redox carriers and electron transfer

The electron transport chain (ETC) relies on a series of carriers that facilitate the transfer of electrons from high-energy donors like NADH to the terminal acceptor oxygen. These carriers are prosthetic groups embedded within membrane-bound protein complexes or mobile within the . Key types include flavins such as (FMN) and (FAD), which accept two electrons and two protons from substrates like NADH or succinate, forming semiquinone intermediates before passing electrons singly. Iron-sulfur (Fe-S) clusters, consisting of 2Fe-2S or 4Fe-4S configurations, serve as one-electron carriers with variable redox potentials tuned to sequential steps in the chain. groups in (b, c, a types) coordinate iron atoms that cycle between Fe³⁺ and Fe²⁺ states, enabling one-electron transfers, while centers in the terminal oxidase handle electron delivery to O₂. Quinones, particularly ubiquinone (coenzyme Q), act as lipid-soluble mobile carriers that shuttle two electrons and protons between complexes, cycling between oxidized ubiquinone and reduced forms. Electron transfer in the ETC proceeds sequentially through these carriers in a downhill gradient, ensuring exergonic . NADH donates two electrons to FMN in Complex I via a one-electron semiquinone step, followed by transfers to Fe-S clusters and then to , which accepts two electrons to form . diffuses to the next complex, where it donates electrons—one via an Fe-S cluster to a , and the other through a bifurcated path involving additional hemes. , another mobile one-electron carrier, relays electrons to centers and hemes in the terminal , reducing O₂ to H₂O. This process involves both one- and two-electron steps, with carriers insulated by protein scaffolds to direct and prevent leakage, often via quantum tunneling across short distances up to 2 . The overall gradient spans from the low potential of NADH/NAD⁺ (E°' ≈ -320 ) to the high potential of O₂/H₂O (E°' ≈ +820 ), driving irreversible electron . The difference (ΔE) between and donor dictates the feasibility and yield of each transfer, calculated as ΔE = E°'_acceptor - E°'_donor. For the full NADH to O₂ span, ΔE ≈ 1.14 V, making the highly favorable. The associated standard change is given by ΔG°' = -nFΔE, where n is the number of electrons transferred (typically 2 for NADH), and F is the (96.485 kJ/mol/V). This yields ΔG°' ≈ -220 kJ/mol for NADH oxidation, releasing in discrete steps to avoid dissipation as . In mitochondrial systems, this couples to proton translocation, with approximately 10 protons pumped across the per NADH oxidized—4 at Complex I, 4 at Complex III, and 2 at Complex IV—establishing the proton motive force. A specialized amplifying proton translocation occurs in Complex III via the Q-cycle, a two-step process of quinone oxidation and reduction. In the first step, at the Qo donates one to the Rieske Fe-S (leading to ) and the other to a low-potential , releasing two protons to the . The second reduces another at the Qi , taking up two protons from the matrix to form semiquinol, which in a subsequent fully reduces to . This bifurcation effectively translocates 4 H⁺ per 2 transferred (beyond scalar protons), doubling the proton/ compared to linear transfer. The Q-cycle thus enhances bioenergetic efficiency without additional energy input.

Proton translocation and chemiosmosis

Proton translocation in the electron transport chain occurs through two primary mechanisms: vectorial transport driven by conformational changes in membrane-bound complexes and redox-linked scalar proton release or uptake associated with chemistry. In complexes such as Complex I, induces long-range conformational changes that propagate across the protein, facilitating the active pumping of protons from to the via dedicated proton channels. This vectorial mechanism ensures that proton movement is tightly coupled to the reactions without direct scalar release. In contrast, redox-linked translocation, exemplified by the Q-cycle in Complex III, involves the scalar protons liberated or consumed during the reduction and oxidation of at the quinone-binding sites, contributing to net proton displacement across the membrane. The primary sites of proton pumping in the mitochondrial electron transport chain are Complexes I, III, and IV, where from NADH ultimately results in the translocation of 10 protons per two electrons (H⁺/2e⁻) into the . This collective pumping establishes an that stores the energy derived from electron flow. Complex II does not contribute to proton pumping, as its role is limited to without associated translocation. The chemiosmotic theory, proposed by Peter Mitchell, posits that the proton gradient generated by these translocation events forms a proton motive force (Δp) comprising a (Δψ) and a gradient (ΔpH), which drives ATP synthesis through indirect osmotic coupling rather than direct chemical intermediates. This delocalized mechanism allows the gradient to power processes across the membrane without requiring physical linkage between electron transport and phosphorylation sites. The proton motive force is quantitatively expressed as: \Delta p = \Delta \psi - 2.303 \frac{RT}{F} \Delta \mathrm{pH} where Δp and Δψ are in millivolts, R is the gas constant, T is temperature in Kelvin, F is the Faraday constant, and ΔpH is the difference between internal and external pH (positive when the matrix is more alkaline). In mitochondria, the membrane potential component (Δψ) typically ranges from 150 to 200 mV, negative inside relative to the intermembrane space, reflecting the high energetic barrier protons must overcome to re-enter the matrix. In prokaryotes, this gradient also maintains cytoplasmic pH homeostasis by counteracting environmental acidity through proton influx.

Mitochondrial electron transport chain

Complex I: NADH:ubiquinone oxidoreductase

Complex I, also known as NADH:ubiquinone oxidoreductase, serves as the entry point for electrons from NADH into the mitochondrial electron transport chain, linking catabolic pathways like the tricarboxylic acid cycle to . It catalyzes the transfer of two electrons from NADH to ubiquinone while translocating protons across the , thereby contributing to the essential for ATP synthesis. The enzyme exhibits an L-shaped architecture, with a hydrophilic peripheral arm projecting into the mitochondrial matrix and a hydrophobic membrane arm integrated into the lipid bilayer. In mammalian mitochondria, Complex I consists of 45 subunits totaling approximately 1 MDa, including 14 conserved core subunits critical for electron transfer and proton pumping—seven of which (ND1–ND6 and ND4L) are encoded by mitochondrial DNA, while the remainder are nuclear-encoded. The peripheral arm, primarily composed of nuclear-encoded subunits, accommodates the NADH oxidation site, a flavin mononucleotide (FMN) cofactor, and eight iron-sulfur (Fe-S) clusters arranged in a linear chain. The membrane arm, dominated by mitochondrially encoded subunits, harbors the ubiquinone (Q) binding pocket and four putative proton channels formed by antiporter-like structures in ND2, ND4, and ND5. High-resolution cryo-electron microscopy structures have elucidated these features, revealing intricate subunit interfaces and cofactor positions. Complex I oxidizes NADH to NAD⁺ in the matrix, reducing ubiquinone to (QH₂) at the domain and pumping four protons (4 H⁺/2 e⁻) from the matrix to the . This process establishes a proton motive force without net consumption of matrix protons beyond the stoichiometry. The balanced equation is: \text{NADH} + \text{[Q](/page/Q)} + 5\text{H}^+_\text{matrix} \to \text{NAD}^+ + \text{QH}_2 + 4\text{H}^+_\text{intermembrane} Electrons flow sequentially from NADH to FMN, through the eight Fe-S clusters (seven participating in transfer, one structural), and finally to , with midpoint potentials increasing from approximately −320 mV (NADH/NAD⁺) to +90 mV (/QH₂) to drive the exergonic transfer. Proton pumping is coupled to Q reduction via a mechanism involving conformational rearrangements that propagate as an electrostatic "wave" across the membrane arm. Reduction of Q induces loop movements in the Q-binding cavity, tilting transmembrane helices in ND1 and rotating the ND6 subunit, which in turn facilitates proton release through coordinated water networks and charged residues in the antiporter subunits. This indirect coupling avoids short-circuiting the redox reaction while ensuring efficient energy transduction. Beyond the core subunits, approximately 31 accessory proteins provide structural stability, facilitate biogenesis via dedicated factors, and modulate activity under physiological . Mutations in Complex I genes, particularly in core subunits like NDUFS1 or mtDNA-encoded ND genes, disrupt or function and are implicated in mitochondrial diseases, including neurodegenerative conditions such as .

Complex II: Succinate:ubiquinone oxidoreductase

Complex II, also known as or , is a heterotetrameric complex embedded in the , consisting of four nuclear-encoded subunits: SDHA, SDHB, SDHC, and SDHD. The SDHA subunit, a , binds (FAD) at its and catalyzes the oxidation of succinate to fumarate, while SDHB serves as the iron-sulfur protein subunit containing three iron-sulfur clusters ([2Fe-2S], [4Fe-4S], and [3Fe-4S]) that facilitate . The membrane-anchoring subunits SDHC and SDHD form a heterodimer that harbors the ubiquinone-binding site at their interface, but unlike other respiratory complexes, Complex II lacks proton-translocating channels or membrane-spanning helices dedicated to pumping. In its primary function, Complex II links the tricarboxylic acid (TCA) cycle to the electron transport chain by oxidizing succinate to fumarate in the TCA cycle while reducing ubiquinone (Q) to ubiquinol (QH2) in the respiratory chain, serving as a dual-role enzyme without net proton translocation across the membrane.01789-1/fulltext) The overall reaction is: \text{Succinate} + \text{Q} \rightarrow \text{Fumarate} + \text{QH}_2 This process bypasses Complex I, delivering electrons directly to the ubiquinone pool and subsequently to Complex III, resulting in a lower ATP yield of approximately 1.5 ATP per pair of electrons compared to 2.5 ATP from NADH oxidation via Complex I, due to fewer protons pumped downstream. The redox potentials are tuned for efficient electron transfer, with the succinate/fumarate couple at E°' = +30 mV and the ubiquinone/ubiquinol couple at approximately +90 mV, ensuring thermodynamically favorable reduction of Q. A unique aspect of Complex II is its exclusive position as the only cycle enzyme that directly participates in the electron transport chain, integrating carbon with respiration.01789-1/fulltext) Germline mutations in the SDH subunit genes (SDHA, SDHB, SDHC, SDHD) are associated with hereditary paraganglioma-pheochromocytoma syndromes, where impaired Complex II activity disrupts succinate and stabilizes hypoxia-inducible factors, promoting tumorigenesis.

Complex III: Ubiquinol:cytochrome c oxidoreductase

Complex III, also known as ubiquinol:cytochrome c oxidoreductase or the cytochrome bc1 complex, is a dimeric embedded in the . Each consists of 11 subunits in mammalian mitochondria, including three core redox-active subunits: (encoded by mtDNA), cytochrome c1, and the Rieske iron-sulfur protein (RISP), along with eight additional nuclear-encoded subunits that provide structural stability. The dimer formation, with each featuring four metal centers—hemes b_H and b_L in , heme c1 in cytochrome c1, and a [2Fe-2S] cluster in RISP—facilitates efficient and is essential for the enzyme's stability and function within respiratory supercomplexes. The primary function of Complex III is to oxidize (QH₂) at the Qo site on the side and transfer electrons to , while simultaneously translocating protons across the to contribute to the proton motive . This process involves bifurcating the two electrons from QH₂: one follows the high-potential chain through the RISP [2Fe-2S] cluster to heme c1 and then to , while the other traverses the low-potential chain via hemes b_L and b_H to reduce ubiquinone (Q) at the Qi site on the matrix side. Through this mechanism, Complex III achieves a proton-to-electron of 4 H⁺ translocated per 2 electrons transferred, amplifying the essential for ATP synthesis. The Q-cycle is the hallmark mechanism of Complex III, enabling proton amplification by coupling quinol oxidation and quinone reduction in a cyclic manner. At the Qo site, the first QH₂ molecule binds and undergoes oxidation: one electron is transferred to the RISP [2Fe-2S] cluster (the rate-limiting step, with a timescale of approximately 770 µs), reducing it and leading to the formation of a transient, unstable semiquinone anion (SQ⁻, with low steady-state occupancy around 0.01); this electron then passes to heme c1 and , releasing two protons into the . The second electron from the semiquinone reduces heme b_L, which relays it to heme b_H and ultimately to a ubiquinone at the Qi site, forming a semiquinone intermediate. A second QH₂ oxidation at Qo repeats the process, providing the second electron to fully reduce the Qi semiquinone to QH₂, taking up two protons from the matrix. This bifurcation and recycling ensure that two molecules are reduced per net QH₂ oxidized, without net consumption at Qi. The net reaction catalyzed by Complex III, encompassing one full Q-cycle, is: \mathrm{QH_2 + 2\ cyt\ c^{3+} + 2\ H^+_{matrix} \to Q + 2\ cyt\ c^{2+} + 4\ H^+_{intermembrane}} This equation reflects the overall electron transfer from ubiquinol to two molecules of oxidized cytochrome c, with enhanced proton release due to the Q-cycle. Unique aspects of Complex III include its two distinct quinone-binding sites: the Qo site, which accommodates stigmatellin binding to inhibit quinol oxidation near the RISP cluster, and the Qi site, targeted by antimycin to block electron transfer from heme b_H. The dimeric architecture not only supports intra-monomer electron transfer (with edge-to-edge distances of about 7-10 Å between hemes) but is also crucial for assembly into supercomplexes with Complexes I and IV, which stabilize the respiratory chain and optimize electron flux while minimizing reactive oxygen species production. The mobility of the RISP extrinsic domain further enables conformational changes essential for Qo site access and efficient bifurcation during the Q-cycle.

Complex IV: Cytochrome c oxidase

Complex IV, also known as , is the terminal of the mitochondrial electron transport chain in eukaryotic cells, consisting of 13 to 14 subunits in mammalian forms, with a catalytic core formed by three mitochondrially encoded subunits (MT-CO1, MT-CO2, and MT-CO3) and the rest as nuclear-encoded supernumerary subunits that stabilize the structure and regulate activity. The MT-CO1 subunit houses heme a and the binuclear center (heme a3 coordinated to CuB), while MT-CO2 contains the CuA binuclear center, and assembly requires specific chaperones for insertion of these metal cofactors into the membrane-embedded complex. These supernumerary subunits, such as COX4 and COX5, form a that positions the for efficient electron transfer from cytochrome c and proton translocation. The primary function of is to catalyze the four-electron reduction of molecular oxygen to two molecules of , utilizing electrons from four reduced cytochrome c molecules, while contributing to the proton motive force by translocating protons across the . This reaction consumes four protons from for scalar chemistry and pumps an additional four protons to the per oxygen reduced, though the pumped proton per two electrons (2-4 H+/2e-) has been debated, with experimental evidence supporting a total of four vectorial protons per O2 alongside the scalar protons. The overall balanced equation is: $4 \ \ce{cyt \ c^{2+}} + \ce{O2} + 8 \ce{H+_{matrix}} \rightarrow 4 \ \ce{cyt \ c^{3+}} + 2 \ce{H2O} + 4 \ce{H+_{intermembrane}} This process ensures complete reduction of O2 without releasing harmful intermediates and links exergonic electron transfer to energy conservation via the proton gradient. The catalytic mechanism involves sequential electron transfer from reduced cytochrome c docked at the extramembranous domain of subunit MT-CO2 to the CuA center, which relays electrons through heme a in MT-CO1 to the binuclear center (heme a3-CuB), where O2 binds to ferrous heme a3 and undergoes four-electron reduction with concomitant cleavage of the O-O bond to form water. Proton loading sites and channels, including the D-pathway in MT-CO1, facilitate both chemical proton uptake for water formation and pumped proton ejection, with the redox-driven conformational changes at the binuclear center ensuring vectorial transport without back-leakage. This tightly coupled electron-proton transfer prevents reactive oxygen species formation during O2 activation. Cytochrome c oxidase activity is allosterically regulated by the ATP/ADP ratio, where ATP binds to nuclear-encoded subunits like COX3 to inhibit and when cellular energy is high, while relieves this inhibition to accelerate turnover. The enzyme is reversibly inhibited by (NO) at nanomolar concentrations, which competitively binds the binuclear center with O2 to modulate mitochondrial in response to signaling, and by (CO), which similarly targets the reduced heme a3 to suppress activity. In brown adipose tissue, supports non-shivering by maintaining high electron flux rates that, coupled with uncoupling protein 1, dissipate the proton gradient as heat rather than ATP synthesis.

Mobile electron carriers: Ubiquinone and cytochrome c

Ubiquinone, also known as coenzyme Q or , is a lipid-soluble embedded in the that serves as a mobile carrier in the electron transport chain (). It consists of a redox-active ring attached to a polyisoprenoid tail, which in humans comprises 10 units (CoQ10), conferring hydrophobicity and enabling within the membrane lipid bilayer. Ubiquinone cycles between its oxidized form (Q) and reduced form (QH2, ) as it accepts two electrons and two protons from upstream complexes, such as Complex I or II, and donates them to Complex III. This two-electron transfer is governed by the reaction: \text{Q} + 2\text{H}^+ + 2\text{e}^- \rightleftharpoons \text{QH}_2 with a standard (E°') of +90 mV at 7. The ubiquinone pool in the inner maintains a high local concentration, approximately 10-fold greater than that of the respiratory complexes, ensuring rapid and efficient without significant limitations during steady-state . This large pool behaves as a homogeneous reservoir, allowing electrons from multiple dehydrogenases to converge and distribute to downstream acceptors. In , the isoprenoid tail is typically shorter, with 9 units (CoQ9), reflecting species-specific variations in biosynthesis that do not substantially alter its function. Beyond electron shuttling between the Q-binding sites of Complexes I, II, and III, reduced (QH2) exhibits properties by donating electrons to neutralize , thereby protecting lipids from peroxidation. Cytochrome c is a small, water-soluble protein that functions as a one-electron carrier in the mitochondrial ETC, diffusing freely within the to shuttle electrons from Complex III to Complex IV. It has a molecular weight of approximately 12 kDa and contains a covalently bound heme c group, where the iron atom cycles between Fe3+ (oxidized, ferricytochrome c) and Fe2+ (reduced, ferrocytochrome c) states, with a midpoint of about +250 mV. The protein's compact structure, featuring lysine-rich surfaces, facilitates electrostatic interactions with the membrane-associated complexes and anionic phospholipids like , optimizing docking and rates. Maintaining a concentration of around 10 μM in the , cytochrome c ensures kinetic efficiency in electron delivery, supporting maximal respiratory flux under physiological conditions. In pathological contexts, such as , cytochrome c is released from the into the , where it promotes activation and .

Integration with oxidative phosphorylation

ATP synthase and proton motive force

The , also known as Complex V of the mitochondrial electron transport chain, is a rotary embedded in the that harnesses the proton motive force (PMF) to synthesize ATP from and inorganic phosphate (Pi). The PMF, consisting of the (Δψ) and gradient (ΔpH) across the membrane, provides the energy for this process, with protons flowing back into through the enzyme driving rotation and catalysis. Structurally, ATP synthase comprises two main domains: the membrane-embedded F0 portion, which includes a proton-translocating c-ring channel formed by 8–15 c-subunits (8 in mammalian mitochondria), and the peripheral F1 portion in , featuring a catalytic hexameric head of three α and three β subunits arranged as (αβ)3 around a central γ rotor stalk. The mechanism of ATP synthesis, elucidated by Paul Boyer's binding change model and confirmed by structural studies from John Walker, involves proton-driven rotation of the c-ring coupled to conformational changes in the F1 head. Protons enter the F0 channel via an a-subunit half-channel from the , bind to aspartate or glutamate residues on c-subunits, and drive stepwise rotation of the c-ring (45° per proton in mammalian mitochondria, or 8 protons per full 360° turn). This rotation turns the γ stalk, which induces sequential conformational shifts in the β-subunits: from open (O, nucleotide release), to loose (L, /Pi binding), to tight (T, ATP formation without energy input), releasing one ATP per 120° rotation and thus three ATP per full turn. The overall reaction is + Pi + n H⁺_out → ATP + n H⁺_in, where n reflects the proton . Stoichiometry links PMF utilization to ATP yield: in mammalian mitochondria, a c-ring of 8 subunits translocates 8 protons per full rotation for 3 ATP synthesized at the (mechanistic stoichiometry of ~2.67 H⁺/ATP). Accounting for additional protons required for ADP import and Pi uptake (~1 H⁺ each via translocators), the effective is ~4 H⁺/ATP overall. Combined with ~10 H⁺ pumped per NADH oxidized by the upstream complexes, this results in a P/O ratio of ~2.5 ATP per NADH (or ~1.5 per FADH₂, with 6 H⁺ pumped via Complexes ). This efficiency ensures that the PMF, generated by electron transport, directly powers without net ATP hydrolysis under physiological conditions. Regulation of ATP synthase maintains cellular energy homeostasis and prevents wasteful ATP hydrolysis during low PMF states. The inhibitory factor 1 (IF1) binds to the F1 β-subunits in a pH- and Δp-dependent manner, stabilizing the enzyme in an inactive conformation to block hydrolysis while allowing synthesis when PMF is high. Oligomycin, a natural antibiotic, specifically inhibits proton translocation through the F0 c-ring by binding the a/c interface, halting both synthesis and hydrolysis. The enzyme's activity is thus finely tuned to the PMF magnitude, slowing rotation at low gradients. F-type ATP synthases exhibit remarkable evolutionary conservation across domains of life, with bacterial homologs like that in Escherichia coli (also featuring a 10-c-subunit ring) operating analogously to drive ATP synthesis using PMF from diverse respiratory chains. This conservation underscores their origin as ancient proton pumps adapted for energy conservation in membranes.

Reverse electron transport and ROS generation

Reverse electron transport (RET) refers to the backward flow of electrons within the mitochondrial electron transport chain, where reducing equivalents from (QH₂) are transferred to Complex I, reducing NAD⁺ to NADH against the typical gradient. This process is driven by a high proton motive force (), which provides the energy to overcome the endergonic nature of the reaction, particularly when the ubiquinone pool is highly reduced and NAD⁺ levels are elevated. The key reaction can be represented as: \ce{QH2 + NAD+ -> Q + NADH + H+} This thermodynamically unfavorable transfer (negative ΔE) is favored under conditions of high membrane potential and a reduced CoQ/CoQH₂ ratio, such as during succinate accumulation via Complex II. RET prominently occurs in pathophysiological states like ischemia-reperfusion injury, where succinate buildup from reversed Complex II activity fuels electron donation to the CoQ pool, promoting backward flow to Complex I. In such scenarios, RET accounts for a substantial portion of reactive oxygen species (ROS) production, with superoxide (O₂⁻) generated primarily at the flavin mononucleotide (FMN) site of Complex I during reverse electron entry. Additional ROS sites include the FMN site in forward mode and the Q₀ site of Complex III, where semiquinone intermediates leak electrons to molecular oxygen; overall, approximately 1-2% of electrons in the chain divert to O₂ under physiological conditions, escalating during RET. Physiologically, RET-mediated ROS signaling supports adaptive responses, such as hypoxic sensing in cells and immune activation in macrophages, where controlled bursts modulate pathways like HIF-1α stabilization. However, excessive RET-driven ROS contributes to oxidative damage in pathologies, including ischemia-reperfusion, by peroxidizing lipids and proteins. Antioxidants like (SOD) mitigate this by converting to , while targeted agents such as mitoQ accumulate in mitochondria to scavenge RET-derived ROS. Recent studies from the highlight RET's role in cancer metabolism, particularly in brain cancer stem cells of , where elevated RET at Complex I sustains , generates ROS for proliferation signaling, and maintains NAD⁺/NADH balance via activation, rendering these cells vulnerable to RET inhibitors like CPTP.

Prokaryotic electron transport chains

Diversity of electron donors and dehydrogenases

In prokaryotic electron transport chains (ETCs), the entry of electrons occurs through a diverse array of dehydrogenases that oxidize various substrates, enabling adaptation to different environmental conditions and energy demands. Unlike the more conserved mitochondrial systems, bacterial ETCs feature modular components that allow flexible assembly of respiratory modules, facilitating efficient under aerobic, microaerobic, or growth. This modularity supports the integration of multiple electron donors into the quinone pool, enhancing metabolic versatility in diverse bacterial lineages. NADH serves as a primary in many , oxidized by two main types of NADH:quinone oxidoreductases. NDH-1, a complex I homolog, is a proton-pumping that translocates approximately 4 H⁺ per 2 electrons transferred from NADH to , contributing to the proton motive for ATP ; it is prevalent in species like . In contrast, NDH-2 is a simpler, non-pumping single-subunit that transfers electrons from NADH to without generating a proton gradient, often serving as a backup or alternative pathway in organisms such as and . Beyond NADH, alternative dehydrogenases oxidize a range of organic and inorganic substrates to feed electrons into the ETC. For instance, catalyzes the oxidation of to CO₂, transferring electrons to , as exemplified by the reaction: \text{HCOO}^- + \text{Q} \rightarrow \text{CO}_2 + \text{QH}_2 This process is crucial for in E. coli and other enteric bacteria, where FDH enables utilization as an energy source. Similarly, and other substrate-specific enzymes provide additional entry points for electrons from products. Specialized dehydrogenases further expand substrate diversity, including those for gases like H₂ and . Hydrogenases, such as [NiFe]- or [FeFe]-types, oxidize molecular hydrogen (H₂ → 2H⁺ + 2e⁻) and channel electrons into the quinone pool or other carriers, supporting lithoautotrophic growth in like Ralstonia eutropha. dehydrogenase (CODH) oxidizes CO to CO₂, integrating this toxic gas as an energy source in carboxydotrophic prokaryotes such as Oligotropha carboxidovorans, thereby mitigating environmental CO while generating reducing equivalents. In certain marine and pathogenic bacteria, the Na⁺-translocating NADH:quinone reductase (Nqr) couples NADH oxidation to Na⁺ extrusion, pumping 2 Na⁺ per 2 electrons without proton translocation; this is prominent in Vibrio cholerae, where it establishes a sodium motive force for flagellar motility and solute transport under varying oxygen levels. Menaquinol-dependent dehydrogenases, often involved in anaerobic conditions, oxidize substrates using menaquinone as an intermediate, as seen in sulfate-reducing bacteria like Desulfovibrio vulgaris, allowing adaptation to low-potential environments. This diversity of dehydrogenases, through modular assembly, enables bacteria to thrive in microaerobic niches by optimizing electron flow and minimizing reactive oxygen species production.

Variations in quinones, cytochromes, and terminal oxidases

In prokaryotic electron transport chains (ETCs), quinones serve as mobile electron carriers that exhibit significant structural and functional variations to accommodate diverse environmental conditions. Ubiquinone (coenzyme Q), predominant in aerobic respiration, possesses a high midpoint (E°' ≈ +100 mV), facilitating efficient to oxygen under oxic conditions. In contrast, menaquinone () is the primary quinone in anaerobic or microaerobic environments, characterized by a lower (E°' ≈ -74 to -80 mV), which enables coupling to low-potential electron donors and acceptors such as fumarate or . Demethylmenaquinone, a structurally related variant, also features a low (E°' ≈ -36 mV) and participates in anaerobic branches of the ETC, particularly in like , where all three quinones coexist and influence pathway branching by directing electrons to specific downstream complexes based on redox poise and oxygen availability. Cytochrome variations further diversify prokaryotic ETCs, with the bc₁ (analogous to mitochondrial III) oxidizing low-potential quinols like menaquinol and transferring electrons to higher-potential c-type , thereby contributing to proton translocation via the Q-cycle mechanism. Alternative pathways bypass , utilizing bd-type quinol oxidases, which directly oxidize or menaquinol without an intervening and exhibit high affinity for oxygen (K_m ≈ 1-10 nM), allowing under low-oxygen conditions. These bd oxidases, found exclusively in prokaryotes, are particularly vital in for survival during or . Terminal oxidases represent the most variable downstream components, adapting to oxygen levels, altitude, or alternative acceptors. The aa₃-type , resembling eukaryotic complex IV, couples quinol or oxidation to oxygen reduction while pumping protons (2 H⁺/2 e⁻). In contrast, the ba₃-type , prevalent in high-altitude or microaerophilic such as Thermus thermophilus, operates with lower proton-pumping (≈1 H⁺/2 e⁻) but higher oxygen affinity, suited for low-oxygen niches. For in like Pseudomonas stutzeri, nitric oxide reductases serve as terminal enzymes, reducing NO to N₂O as part of the respiratory chain, integrating with upstream quinone-dependent branches. A representative reaction for bd-type oxidases illustrates this adaptability: QH₂ + ½ O₂ → Q + H₂O, accompanied by proton translocation of 2 H⁺/2 e⁻ across the via vectorial quinol oxidation, without active pumping. These variations enable branching in prokaryotic ETCs, as exemplified in E. coli, where cytochrome bo₃ (an aa₃-like quinol ) handles high-oxygen aerobic , while alternative bd oxidases or menaquinone-linked paths activate under low oxygen, optimizing energy yield and survival across gradients.

Examples from aerobic and anaerobic bacteria

In aerobic bacteria, Escherichia coli exemplifies a flexible electron transport chain adapted to varying oxygen levels, featuring two NADH dehydrogenases—NDH-1, a proton-pumping complex I homolog, and NDH-2, a non-proton-pumping —and two terminal quinol oxidases: bo₃, which operates under high oxygen conditions with high proton-pumping efficiency, and bd, which predominates in microaerobic environments for its lower oxygen affinity. This switching mechanism optimizes energy yield while protecting against , with NDH-1 and bo₃ contributing to higher ATP production under normoxia. Similarly, Paracoccus denitrificans possesses a mitochondrial-like respiratory chain, including a canonical complex I (NDH-1), ubiquinone, complex III, , and a terminal aa₃-type oxidase, enabling efficient aerobic respiration with a proton motive force comparable to eukaryotic mitochondria. This configuration supports high ATP yields through , and the bacterium's branched chain allows adaptation to alternative electron acceptors like under oxygen limitation, though aerobic conditions maximize efficiency. Facultative anaerobes like demonstrate branched electron transport with multiple terminal oxidases for environmental versatility, including the quinol oxidase cytochrome aa₃, which pumps protons efficiently under aerobic conditions, and cyanide-resistant cytochrome bd, which sustains in low-oxygen or toxin-exposed settings. The presence of a cytochrome c oxidase branch alongside quinol oxidases allows B. subtilis to balance energy production and stress tolerance, with aa₃ handling bulk oxygen reduction and bd providing a protective, alternative pathway. In strict anaerobes such as species, the electron transport chain couples or oxidation to sulfate reduction via menaquinone (MK-6), involving transmembrane complexes like Qrc (quinone-reducing complex) and Dsr (dissimilatory sulfite reductase) without oxygen involvement, generating a modest proton motive force for ATP synthesis. This menaquinone-dependent pathway supports in sulfate-rich, anoxic environments, though it yields fewer protons translocated per electron compared to aerobic chains. Methanogenic , such as those in the genus , employ a unique electron transport system where reduced coenzyme F₄₂₀ (F₄₂₀H₂) donates electrons to heterodisulfide reductase (), forming a proton-translocating complex that reduces the CoM-S-S-CoB heterodisulfide to thiols during from CO₂ and H₂. This F₄₂₀H₂: pathway generates a proton gradient for but operates without quinones or typical of bacterial respiration, highlighting archaeal divergence. Unique adaptations in prokaryotic chains include the role of cytochrome bd oxidases in respiratory protection, where they act as oxygen scavengers to shield or other O₂-sensitive enzymes in microaerophilic or transitioning , consuming trace oxygen without generating excessive . Recent metagenomic studies from the 2020s have uncovered uncultured prokaryotic diversity, revealing novel electron transport configurations in environmental microbiomes, such as expanded variants in sediment communities that expand known respiratory plasticity. Efficiency varies markedly, with aerobic chains like those in E. coli and P. denitrificans yielding up to 30-38 ATP per glucose equivalent through extensive proton pumping, whereas anaerobic systems in Desulfovibrio and methanogens produce far less—typically 1-5 ATP per reduced substrate—prioritizing and survival in electron acceptor-limited niches over maximal energy harvest.

Photosynthetic electron transport chains

Oxygenic photosynthesis in chloroplasts

In oxygenic photosynthesis, the electron transport chain (ETC) in chloroplasts of , , and converts light energy into by driving non-cyclic electron flow from to NADP⁺, producing oxygen as a byproduct. This process occurs within the membranes and involves two linked by the b₆f complex, forming the Z-scheme. The Z-scheme, first proposed in 1960, illustrates the sequential excitation of electrons at higher redox potentials in (PSII) and (PSI), enabling efficient energy capture despite the thermodynamic barrier between the two systems. The primary components include PSII, PSI, the mobile carrier plastoquinone (PQ), the cytochrome b₆f complex, and plastocyanin (PC). PSII, centered around the reaction core chlorophyll pair P680, contains the oxygen-evolving complex (OEC) with a Mn₄CaO₅ cluster that catalyzes water oxidation to extract electrons. Electrons from the OEC pass through a pheophytin-quinone (Q_A-Q_B) acceptor complex, where Q_B binds PQ to form plastoquinol (PQH₂). The cytochrome b₆f complex, homologous to complex III in respiration, features b-type hemes, a Rieske iron-sulfur protein, and cytochrome f, facilitating electron transfer from PQH₂ to PC via a Q-cycle mechanism. PSI, with its P700 reaction center chlorophyll pair, includes iron-sulfur (Fe-S) clusters (F_X, F_A, F_B) that relay electrons to soluble ferredoxin (Fd), which reduces NADP⁺ to NADPH via ferredoxin-NADP⁺ reductase. Linear electron flow follows the path: H₂O → PSII (P680) → PQ → cytochrome b₆f → PC → PSI (P700) → Fd → NADP⁺. Light absorption at PSII boosts s from P680 to pheophytin, creating P680⁺, which oxidizes the OEC to split (2H₂O → O₂ + 4H⁺ + 4e⁻). These s reduce PQ to PQH₂, which diffuses to cytochrome b₆f. There, the Q-cycle bifurcates s: one to PC via the Rieske center and cytochrome f, and the other across the membrane to reduce another PQ molecule, amplifying proton translocation. PC then delivers s to P700⁺ in PSI, where light re-energizes them to reduce Fd and ultimately NADP⁺. A parallel cyclic electron flow around PSI returns s from Fd to cytochrome b₆f via ferredoxin-dependent pathways, enhancing ATP production without NADPH generation. Proton pumping builds a proton motive force (ΔpH) across the thylakoid membrane, essential for ATP synthesis. In PSII, oxidation releases scalar protons (4H⁺ per O₂) into the . The b₆f Q-cycle translocates 4H⁺ per 2e⁻ (2 from PQH₂ scalar release in the and 2 vectorial from the stromal to side), contributing the majority of the gradient. This ΔpH, along with a minor , drives to produce ATP from and Pᵢ. The overall Z-scheme stoichiometry for linear flow is 2H₂O + 2NADP⁺ + ~8 photons → O₂ + 2NADPH + 2H⁺, balancing NADPH and ATP needs for the , though cyclic flow adjusts the ATP/NADPH ratio. Unique to chloroplast thylakoids, the photosynthetic enables dynamic regulation through and (). involve reversible of II (LHCII) proteins, mediated by the state of PQ; in (oxidized PQ), LHCII associates with PSII, while in (reduced PQ), ~15% of LHCII migrates to via actin-dependent mechanisms, balancing excitation between . dissipates excess light energy as heat in LHCII antennae, primarily through pH-dependent formation and PsbS protein activation in PSII, preventing photodamage under high light. These mechanisms optimize electron flow and protect the chain in varying light conditions. Recent structural studies using cryo-electron microscopy have advanced understanding of the photosynthetic ETC's assembly and regulation. For instance, a 2024 study achieved a 1.7 Å resolution structure of , revealing intricate water channels involved in substrate delivery to the OEC. In 2022, the –NDH supercomplex was resolved at high resolution, elucidating cyclic electron flow mechanisms. Additionally, 2023 cryo-EM structures of LHCII in photo-active and photo-protecting states demonstrated of energy dissipation. These findings highlight the dynamic supramolecular organization that optimizes efficiency under fluctuating light conditions.

Anoxygenic photosynthesis in bacteria

Anoxygenic photosynthesis in involves electron transport chains that utilize light energy to generate a proton motive force for ATP synthesis without producing oxygen, relying instead on alternative electron donors such as reduced compounds. These processes occur in diverse prokaryotes, primarily and green sulfur bacteria, each employing distinct reaction centers and carriers. , such as those in the Rhodospirillaceae family, feature a Type II reaction center with a special pair P870 that absorbs light at approximately 870 nm, initiating to a (Q), followed by the bc1 complex, , and terminal oxidases like aa3 under aerobic conditions. In contrast, green sulfur bacteria utilize a Type I reaction center with P840, where light excitation drives electrons through iron-sulfur (Fe-S) centers to menaquinone, enabling oxidation without . Electron donors in these systems include hydrogen sulfide (H₂S), molecular hydrogen (H₂), and organic compounds, but notably exclude , which limits the process to anoxic environments. For instance, like Chromatium oxidize or H₂S to elemental or via enzymes such as :quinone (SQR), providing electrons to the quinone pool. Green similarly employ H₂S as a primary donor, oxidized by SQR to deposit granules outside the cell. The electron flow typically proceeds from the donor to the reaction , then through quinones or to terminal acceptors, often in a cyclic manner that returns electrons to the or reduces NAD⁺ for carbon fixation, thereby establishing a proton gradient across the for ATP production. A representative in green is $2\mathrm{H_2S} + 2\mathrm{Q} \rightarrow \mathrm{S_0} + 2\mathrm{QH_2}, where quinone accepts electrons from oxidation. Unique structural and functional aspects enhance the efficiency of these chains, including intracytoplasmic membranes in that invaginate from the cytoplasmic membrane to house reaction centers and light-harvesting complexes, optimizing light capture under low-oxygen conditions. These bacteria, exemplified by Rhodobacter capsulatus, demonstrate remarkable versatility, seamlessly switching between photosynthetic electron transport and respiratory modes by sharing components like the bc1 complex and quinones when oxygen becomes available. This homology between bacterial Type I and II reaction centers and eukaryotic I and II underscores their evolutionary conservation, though anoxygenic chains operate with a single .

Regulation, inhibitors, and pathology

Mechanisms of regulation

The electron transport chain () is regulated at multiple levels to fine-tune mitochondrial respiration according to cellular energy demands, preventing excessive (ROS) production and maintaining metabolic . Allosteric mechanisms provide rapid, reversible control by sensing immediate metabolic states, while post-translational modifications allow dynamic adjustments to environmental cues like . Transcriptional adapts the ETC composition over longer timescales in response to or oxygen availability, and structural dynamics of supercomplexes influence electron flux efficiency. In both eukaryotic and prokaryotic systems, these layers integrate with signaling to coordinate overall . Allosteric regulation directly modulates ETC complex activities based on energy status. In mammalian mitochondria, cytochrome c oxidase (Complex IV) is inhibited by ATP binding to its matrix domain under high-energy conditions, reducing electron flux and preserving a low to minimize ROS generation. This ATP-mediated inhibition is particularly pronounced in the phosphorylated, dimeric form of Complex IV, ensuring respiratory control without complete shutdown. Similarly, Complex I activity is allosterically inhibited by a high NADH/NAD⁺ ratio, which reflects substrate accumulation and prevents over-reduction of the chain, thereby linking upstream metabolic states to downstream rates. Post-translational modifications offer precise, covalent control over function, often in response to signaling pathways or stress. of Complex I subunits by (), activated via signaling, alters its assembly, catalytic activity, and supercomplex integration, thereby governing kinetics and oxygen consumption in response to hormonal or adrenergic stimuli. S-glutathionylation, a reversible modification, protects complexes from oxidative damage during ROS elevation; for instance, it targets Complex I and Complex II to attenuate production by blocking electron leakage, while also feeding back to adjust nutrient uptake and mitochondrial . These modifications enable the to adapt swiftly to imbalances without requiring new protein synthesis. Transcriptional control reshapes the ETC proteome to match environmental conditions, particularly oxygen levels. In hypoxic mammalian cells, hypoxia-inducible factor 1 (HIF-1) suppresses and promotes isoform switching in Complex IV, such as upregulating COX4-2 expression while downregulating COX4-1 via LON protease, optimizing respiratory efficiency and reducing oxygen consumption. In , the ArcA/ArcB two-component system senses state to regulate genes; under low oxygen, phosphorylated ArcA activates transcription of cytochrome bd (cydAB) for microaerobic respiration while repressing aerobic cytochrome bo , ensuring adaptive to available acceptors. Supercomplex dynamics further regulate ETC efficiency through assembly and disassembly, stabilizing electron channeling. In mitochondria, factors like COX7A2L (SCAF1) promote the association of Complexes I, III, and IV into respirasomes, enhancing proton pumping and reducing ROS by minimizing diffusible intermediate exposure; disruptions in these factors lead to disassembly and inefficient flux. Assembly of Complex IV itself relies on dedicated factors such as SCO1 and SURF1, which coordinate subunit incorporation and supercomplex integration to maintain structural integrity under varying metabolic loads. Calcium signaling provides an additional layer of , particularly in mitochondria, where MICU1 acts as a gatekeeper for the mitochondrial calcium uniporter (MCU). By sensing cytosolic Ca²⁺ levels, MICU1 modulates Ca²⁺ influx to activate dehydrogenases like and α-ketoglutarate dehydrogenase, thereby stimulating substrate supply and NADH production without overload. In , two-component systems beyond ArcA/B, such as RegB/RegA in photosynthetic species, sense or light cues to regulate ETC components, integrating environmental signals with respiratory adaptation. These mechanisms collectively ensure the ETC responds dynamically to physiological needs across organisms.

Inhibitors and experimental tools

Inhibitors of the are valuable pharmacological and genetic tools that target specific complexes to dissect their functions, measure electron flux, and model mitochondrial dysfunction in settings. These agents bind to distinct sites within the complexes, blocking or proton translocation, and have been instrumental in elucidating the mechanistic details of . For instance, partial inhibition of Complex I by certain compounds can mimic physiological stress and inform therapeutic strategies. For Complex I (NADH:ubiquinone oxidoreductase), binds to the quinone-binding site (Q-site) in the domain, preventing from the iron-sulfur clusters to ubiquinone and halting NADH oxidation. Piericidin A similarly inhibits at the Q-site, competing with ubiquinone for binding and blocking the reduction step.47496-8/fulltext) is extensively used in cellular and animal models of , where chronic exposure induces selective loss by promoting mitochondrial (ROS) production and α-synuclein aggregation. Complex III (cytochrome bc1 complex) inhibitors target the Q-cycle, a bifurcated pathway. Antimycin A binds to the Qi site on the distal side, inhibiting from heme bH to ubiquinone and causing semiquinone accumulation at the Qo site, which facilitates ROS generation. Myxothiazol, in contrast, binds to the Qo site on the proximal side, blocking electron donation from the Rieske iron-sulfur protein to c1 and preventing initial oxidation; this specificity has been crucial for experimentally isolating Q-cycle branches. Complex IV () is inhibited by , which binds tightly to the heme a3 iron in the binuclear center, preventing oxygen binding and reduction to . targets the CuB site adjacent to heme a3, similarly disrupting the by competing with oxygen.45888-5/fulltext) (CO) serves as a spectroscopic tool, reversibly binding to heme a3 to probe the enzyme's state and interactions in low-temperature studies without permanent inactivation. Complex II (succinate:ubiquinone ) is targeted by thenoyltrifluoroacetone (TTFA), which binds to the ubiquinone reduction site in subunit C (Sdha), inhibiting succinate oxidation and flavin-mediated to ubiquinone. For (Complex V), occludes the FO proton channel, blocking proton flow through the c-ring and preventing ATP synthesis while allowing proton leak in some contexts. Genetic tools complement pharmacological inhibitors for ETC studies. (RNAi) knockouts silence nuclear-encoded ETC genes, enabling assessment of complex assembly and compensatory responses in cell lines. (mtDNA)-deficient rho0 cells, generated by treatment, lack mtDNA-encoded subunits and serve as models for ETC impairment, as they exhibit reduced and reliance on . Flux assays, such as those using the XF analyzer, quantify oxygen consumption rate (OCR) to measure ETC activity; inhibitors like and are applied sequentially to isolate contributions from individual complexes via changes in basal, ATP-linked, and maximal . Therapeutically, metformin acts as a partial Complex I inhibitor by binding near the Q-site, modestly reducing NADH oxidation to activate (AMPK) and improve insulin sensitivity in management. In the , proteolysis-targeting chimeras (PROTACs) have emerged as tools for targeted degradation of ETC complexes, such as Complex I subunits, by recruiting E3 ligases for ubiquitin-mediated proteolysis, offering spatiotemporal control in research beyond traditional small-molecule inhibition.

Role in diseases and cellular dysfunction

Dysfunction in the electron transport chain (ETC) is a hallmark of various mitochondrial diseases, primarily arising from in nuclear or that impair respiratory complexes. Leigh syndrome, a severe neurometabolic disorder, often results from defects in Complex I (NADH:ubiquinone oxidoreductase) or Complex IV (cytochrome c oxidase), leading to due to disrupted and energy failure in high-demand tissues like the . These , accounting for about one-third of childhood-onset cases, cause bilateral symmetric lesions in the and , manifesting as , , and seizures. Similarly, (LHON) stems from primary in Complex I genes, such as MT-ND1, MT-ND4, and MT-ND6, resulting in acute or subacute vision loss from degeneration. These point disrupt , reducing ATP production and triggering retinal ganglion cell . In neurodegenerative disorders, ETC impairments contribute to selective neuronal vulnerability. Parkinson's disease features a specific Complex I deficiency in the , where neurons degenerate, leading to motor symptoms like bradykinesia and rigidity. This defect, observed in postmortem tissue, reduces NADH oxidation and elevates (ROS), exacerbating α-synuclein aggregation and neuronal loss. In , Complex IV activity declines in affected brain regions, partly due to amyloid-β peptide binding and inhibition of , which impairs energy metabolism and promotes . This bioenergetic failure correlates with amyloid plaque formation and cognitive decline, as amyloid-β oligomers further suppress mitochondrial . Cancer cells exploit ETC alterations to favor proliferation, exemplified by the effect, where tumors suppress in favor of aerobic , even in oxygen-rich environments, to generate biosynthetic intermediates. Mutations in (Complex II), such as in SDHB or SDHD genes, cause succinate accumulation, mimicking (pseudohypoxia) by stabilizing HIF-1α and driving and in paragangliomas and pheochromocytomas. This oncometabolite inhibits α-ketoglutarate-dependent dioxygenases, leading to epigenetic changes that promote tumorigenesis. ETC dysfunction also drives ROS-mediated pathologies. In aging, somatic mitochondrial DNA mutations accumulate due to chronic ROS exposure from leaky electron transport, impairing respiratory efficiency and accelerating in post-mitotic tissues. This vicious cycle of and oxidative damage contributes to frailty and organ decline without necessarily increasing overall ROS levels. During ischemia-reperfusion , such as in or , reverse electron transport at Complex I generates a burst of ROS upon oxygen restoration, fueled by succinate oxidation, exacerbating tissue damage through and . Emerging therapies in the 2020s target these defects, including EPI-743, a redox-active analog that bypasses CoQ10 deficiencies in disorders by enhancing NADPH-dependent antioxidant defenses and stabilizing mitochondrial function in and related conditions. Gene editing approaches, such as base editors delivered via adeno-associated viruses, have corrected pathogenic mtDNA mutations in Complex I genes in patient-derived cells, restoring activity and offering potential for treating LHON and . These interventions aim to mitigate and neurodegeneration by directly repairing heteroplasmic mutations.

References

  1. [1]
    Biochemistry, Electron Transport Chain - StatPearls - NCBI Bookshelf
    Sep 4, 2023 · The electron transport chain is a series of four protein complexes that couple redox reactions, creating an electrochemical gradient that leads to the creation ...Introduction · Fundamentals · Cellular Level · Molecular Level
  2. [2]
    Electron-Transport Chains and Their Proton Pumps - NCBI - NIH
    The electrons are transferred to these acceptors by a series of electron carriers in the plasma membrane that are comparable to those in mitochondrial ...
  3. [3]
    Photosynthesis - The Cell - NCBI Bookshelf - NIH
    Electrons are transferred sequentially between the two photosystems, with photosystem I acting to generate NADPH and photosystem II acting to generate ATP. The ...
  4. [4]
    Metabolism - PMC - PubMed Central
    This then liberates 30–32 ATP per glucose molecule, a 15–16-fold increase in energy return. These last two pathways, TCA cycle and ETC, require functioning ...
  5. [5]
    Photosynthesis and the Electron Transport Chain - Ask A Biologist
    May 25, 2017 · The electron transport chain is a series of molecules that accept or donate electrons easily. ... Biology; Publisher: Arizona State University ...
  6. [6]
    Regulation and functional role of the electron transport chain ...
    Nov 8, 2021 · This review will focus on the current evidence for the formation of different SCs and will explore how they modulate the ETC organization according to the ...
  7. [7]
    Otto Warburg – Biographical - NobelPrize.org
    For his discovery of the nature and mode of action of the respiratory enzyme, the Nobel Prize has been awarded to him in 1931. This discovery has opened up new ...Missing: 1920s | Show results with:1920s
  8. [8]
    On cytochrome, a respiratory pigment, common to animals, yeast ...
    A respiratory pigment, which he found in muscles and other tissues of representatives of almost all the orders of the animal kingdom.
  9. [9]
    THE CYCLOPHORASE COMPLEX OF ENZYMES - GREEN - 1951
    The cyclophorase complex of enzymes which implement the citric acid cycle is contained within the mitochondrial bodies. 2. Mitochondria behave as fibre-like ...
  10. [10]
    Discovery of ubiquinone (coenzyme Q) and an overview of function
    Details of the discovery of ubiquinone (coenzyme Q) are described in the context of research on mitochondria in the early 1950s.
  11. [11]
    [PDF] Peter Mitchell - Nobel Lecture
    At the beginning of this period, the work of David Keilin (1925, 1929) on the cytochrome system, and work by. Warburg, Wieland and others on the respiratory ...
  12. [12]
    Electron carriers and energy conservation in mitochondrial respiration
    Apr 23, 2019 · The mammalian electron transport chain (ETC) contains the flavins, iron–sulfur clusters (Fe–S), several types of heme, and, in the terminal ...Electron Carriers And Energy... · Introduction · Cofactor Structures And...
  13. [13]
    Mitochondrial electron transport chain: Oxidative phosphorylation ...
    The mitochondrial electron transport chain utilizes a series of electron transfer reactions to generate cellular ATP through oxidative phosphorylation.
  14. [14]
    What is the redox potential of a cell? - Bionumbers book
    The redox potential difference ΔE between the electron donor and acceptor is related to the associated free energy change ΔG of the reaction via ΔG=nFΔE ...
  15. [15]
    A Model of the Proton Translocation Mechanism of Complex I - PMC
    Mar 30, 2011 · These combine direct proton translocation via Q-redox reactions coupled to conformational changes that extend into the hydrophobic domain to ...
  16. [16]
    Mitochondrial electron transport chain: Oxidative phosphorylation ...
    Aug 6, 2020 · In this review, we will provide an overview of the function of the ETC, focusing on oxidative phosphorylation and its relationship to ROS production.
  17. [17]
    Mitochondrial Membrane Potential - an overview - ScienceDirect.com
    Mitochondrial membrane potential (MMP) is the negative potential difference across the inner mitochondrial membrane, typically ranging from −150 to −180 mV.
  18. [18]
    Mitochondrial Respiratory Complex I: Structure, Function and ...
    The matrix arm is related to the small form of NADH: ubiquinone oxidoreductase and consists mostly of the nuclear-encoded subunits, FMN, three iron–sulfur ...
  19. [19]
    High-resolution structure and dynamics of mitochondrial complex I ...
    Nov 12, 2021 · Redox-driven proton translocation by complex I contributes substantially to the proton motive force that drives ATP synthase. Several structures ...
  20. [20]
    Structure of subcomplex Iβ of mammalian respiratory complex I ...
    In mammalian mitochondria, respiratory complex I (NADH:ubiquinone oxidoreductase) (1, 2) contains 45 subunits with a combined mass of 1 MDa (3–6). Fourteen ...<|control11|><|separator|>
  21. [21]
    Real-time electron transfer in respiratory complex I - PNAS
    Complex I couples electron transfer from NADH to ubiquinone to translocation of 2 H+/e− across the membrane (2). It is a true redox-linked proton pump, as is ...
  22. [22]
    The coupling mechanism of mammalian respiratory complex I
    We propose a detailed molecular coupling mechanism of complex I, which is an unexpected combination of conformational changes and electrostatic interactions.
  23. [23]
    From the 'black box' to 'domino effect' mechanism: what have we ...
    Mar 15, 2023 · Since we established that complex I cycles between open and closed conformations during turnover, and we had high-resolution structures (with ...
  24. [24]
    OXPHOS mutations and neurodegeneration | The EMBO Journal
    Mutated structural OXPHOS subunit and assembly factor genes associated with neurodegeneration in humans. Disorder/phenotype, Clinical features, Structural ...Introduction · The Mitochondrial Oxphos... · Oxphos Mutations And...
  25. [25]
    Structure of the human respiratory complex II - PNAS
    Apr 25, 2023 · The SDHA subunit can be divided into four domains (11) termed the FAD-binding domain, capping domain, helical domain, and C-terminal domain (Fig ...
  26. [26]
    Electron-Transfer Pathways in the Heme and Quinone-Binding ...
    Feb 21, 2014 · The first electron initially ends up at the highest potential iron sulfur cluster ([3Fe–4S]), and consistent with the data reported here and our ...<|separator|>
  27. [27]
    Stimulation of Menaquinone-Dependent Electron Transfer in the ...
    −80 mV at pH 7, which is some 110 mV more negative than the Em of the succinate-fumarate couple (ca. +30 mV at pH 7 [9, 14, 37]). Therefore, the succinate: ...
  28. [28]
    Ubiquinone - an overview | ScienceDirect Topics
    ... standard redox potential of the ubiquinone/ubiquinol couple at pH 7 is ~+100 mV. The semiquinone radical is intermediate in most redox reactions of coenzyme Q ...
  29. [29]
    Management of phaeochromocytoma and paraganglioma ... - Nature
    Dec 14, 2023 · Astuti, D. et al. Gene mutations in the succinate dehydrogenase subunit SDHB cause susceptibility to familial pheochromocytoma and to familial ...
  30. [30]
    The Q-cycle reviewed: how well does a monomeric mechanism of ...
    The ubiquinol:cytochrome c oxidoreductase (bc1 complex) family of redox-linked proton pumps form the core of all major electron transfer chains, with an ...
  31. [31]
    The mitochondrial coenzyme Q junction and complex ... - FEBS Press
    Aug 24, 2021 · The cycle starts at the quinol oxidation site (Qo), where the Fe-S cluster of RISP accepts the first electron from a quinol molecule, ...<|separator|>
  32. [32]
    The Q Cycle of Cytochrome bc Complexes: a Structure Perspective
    Structure-function problems discussed recently for the bc1 complex include the role of the monomer and dimer in the electron transport pathway associated with ...Iii. The Q Cycle · Quinone Binding In The Dimer... · V. Role Of The Dimer In...Missing: paper | Show results with:paper
  33. [33]
    Electronic Connection Between the Quinone and Cytochrome c ...
    Jan 1, 2015 · Three of these complexes (complex I, III, and IV) couple the electron transfer with proton translocation across the membrane. Open in Viewer
  34. [34]
    Structure of the intact 14-subunit human cytochrome c oxidase
    Jul 20, 2018 · Therefore, CIV should be a 14-subunit monomer instead of a homodimer with each protomer containing 13 subunits which is described by the ...
  35. [35]
    Biogenesis and Assembly of Eukaryotic Cytochrome c Oxidase ...
    The COX catalytic core is formed by three mitochondrial DNA encoded subunits, Cox1, Cox2 and Cox3, conserved in the bacterial enzyme.
  36. [36]
    Oxygen Activation and Energy Conservation by Cytochrome c Oxidase
    Jan 19, 2018 · A peroxide bridge between Fe and Cu ions in the O2 reduction site of fully oxidized cytochrome c oxidase could suppress the proton pump.Structure of the Active Site · Paths of Electrons, Protons... · Proton Translocation
  37. [37]
    Proton pumping by cytochrome c oxidase – A 40 year anniversary
    Cytochrome c oxidase transduces energy into a proton gradient. The proton pump mechanism involves loading a site, proton transfer, and ejection, creating a ...Missing: O2 | Show results with:O2
  38. [38]
    Electron Transfer Pathways in Cytochrome c Oxidase - PMC - NIH
    The electron transfer pathway in Cytochrome c Oxidase follows the sequence CuA → heme a → heme a3, with arginines 481 and 482 playing a crucial role.
  39. [39]
    Cytochrome c Oxidase at Full Thrust: Regulation and Biological ...
    Feb 22, 2021 · We observed that ATP/ADP controls mitochondrial G3P oxidation through specific and allosteric regulation of COX activity. Additionally, ...
  40. [40]
    Regulation of mitochondrial respiration by nitric oxide inhibition of ...
    Mar 1, 2001 · Nanomolar concentrations of NO immediately, specifically and reversibly inhibit cytochrome oxidase in competition with oxygen.
  41. [41]
    Carbon monoxide signals via inhibition of cytochrome c oxidase and ...
    Jan 30, 2007 · CO acts via inhibition of cytochrome c oxidase leading to the generation of low levels of reactive oxygen species (ROS) that in turn mediate subsequent ...
  42. [42]
  43. [43]
    Coenzyme Q10 | Linus Pauling Institute | Oregon State University
    Mitochondrial ATP synthesis​​ As part of the mitochondrial electron transport chain, coenzyme Q10 accepts electrons from reducing equivalents generated during ...
  44. [44]
    Ubiquinone - an overview | ScienceDirect Topics
    Ubiquinones exist in various chain lengths, with ubiquinone-9 being predominant in rats, while ubiquinone-10 is predominant in humans. 3.3.1 Measurement of ...
  45. [45]
    Coenzyme Q and the Respiratory Chain - PubMed Central - NIH
    As it was mentioned above, CoQ (ubiquinone) is required for the transfer of electrons from NADH- or FAD-dependent enzymes to the respiratory Complex III within ...
  46. [46]
    Redox-State Dynamics of Ubiquinone-10 Imply Cooperative ... - NIH
    First, for RegB, the standard redox potential of the ... The culture redox potential was calculated for pH 7 from the measured electrode potential.
  47. [47]
    Review Mobility and function of Coenzyme Q (ubiquinone) in the ...
    Their results were interpreted to describe the existence of three different pools of CoQ (and cytochrome c): one utilized during steady-state respiration, ...
  48. [48]
    The effect of different ubiquinones on lifespan in Caenorhabditis ...
    The isoprenyl tail length of UQ is species-specific, being ten (UQ10) in humans, nine (UQ9) in rodents and Caenorhabditis elegans (C. elegans), and eight ...
  49. [49]
    The multiple functions of cytochrome c and their regulation in life ...
    Cytc is located in the mitochondrial intermembrane space and functions as a single electron carrier from the bc1 complex to CcO in the final step of the ETC. ...
  50. [50]
    Mechanisms of cytochrome c release from mitochondria - Nature
    May 5, 2006 · Cyt c, a peripheral protein of the mitochondrial inner membrane (IM), functions as an electron shuttle between complex III and complex IV of the ...
  51. [51]
    Structure and Mechanisms of F-Type ATP Synthases | Annual Reviews
    Jun 20, 2019 · Structures of mitochondrial ATP synthase dimers indicate how they shape the inner membrane cristae. The new cryo-EM structures complete our ...
  52. [52]
    Press release: The 1997 Nobel Prize in Chemistry - NobelPrize.org
    In 1960 the American scientist Efraim Racker and co-workers isolated, from mitochondria, the enzyme “F o F 1 ATPase” which we now call ATP synthase. The ...
  53. [53]
    ATP synthase: From sequence to ring size to the P/O ratio - PMC - NIH
    Sep 28, 2010 · The P/O ratio for NADH would be 10/(3.3 + 1) = 10/4.3 ≈ 2.3, thus moving below the range of values directly determined.
  54. [54]
    An overview of ATP synthase, inhibitors, and their toxicity - PMC
    Nov 20, 2023 · IF1, also known as Inhibitory Factor 1, is a protein that naturally inhibits mitochondrial ATP synthase. It comprises 56–87 residues and binds ...
  55. [55]
    Regulation of Mitochondrial Structure and Function by the F1Fo ...
    Jul 2, 2008 · Oligomycin increased NADH fluorescence, consistent with the inhibition of proton influx and inhibition of ATP synthase-dependent respiration, ...
  56. [56]
    Role of Mitochondrial Reverse Electron Transport in ROS Signaling
    Jun 27, 2017 · The redox state of the CoQ pool determines electron leak and superoxide production and can be used by mitochondria as a signaling mechanism.
  57. [57]
    Mitochondrial electron transport chain, ROS generation and ...
    The mammalian mitochondrial electron transport chain (ETC) includes complexes I-IV, as well as the electron transporters ubiquinone and cytochrome c.
  58. [58]
    Regulation of Reverse Electron Transfer at Mitochondrial Complex I ...
    Here we show that reverse electron transfer (RET) through respiratory chain complex I (RC-I) is particularly active in brain cancer stem cells (CSC).
  59. [59]
    Microbial electron transport and energy conservation
    The respiratory chains show a great diversity and variability enclosing 15 primary dehydrogenases to oxidize electron donors and ten terminal reductases (or ...
  60. [60]
    New Insights into Type II NAD(P)H:Quinone Oxidoreductases - PMC
    The main reducing equivalent synthesized by the cell central metabolism is NADH, making it the principal electron donor to respiratory chains. In prokaryotes, ...
  61. [61]
    Purification and Characterization of NDH-2 Protein and Elucidating ...
    May 7, 2019 · NDH-2 is a non-proton pumping NADH dehydrogenase located in the inner membrane of several bacteria like Bacillus subtilis, Escherichia coli, etc.
  62. [62]
    A Bacterial Electron-bifurcating Hydrogenase - PMC - NIH
    Here, we describe an electron-bifurcating hydrogenase that apparently couples the reverse reaction, exergonic downhill flux of two electrons to NAD+ that then ...
  63. [63]
    Structural basis for bacterial energy extraction from atmospheric ...
    Mar 8, 2023 · Huc is a highly efficient oxygen-insensitive enzyme that couples oxidation of atmospheric H 2 to the hydrogenation of the respiratory electron carrier ...
  64. [64]
    Carbon Monoxide and Prokaryotic Energy Metabolism - PMC - NIH
    Mar 20, 2025 · This review discusses the molecular basis of the effects of CO on microbial growth and aerobic respiration supported by different terminal oxidases.
  65. [65]
    Cryo-EM structures of Na+-pumping NADH-ubiquinone ... - Nature
    Jul 26, 2022 · Na+-NQR is the first enzyme in the respiratory chain of many pathogenic bacteria such as Vibrio cholerae, Vibrio alginolyticus, and Haemophilus ...
  66. [66]
    Aerobic respiratory chain of Escherichia coli is not allowed ... - PNAS
    Oct 18, 2011 · NDH-I and. NDH-II transfer electrons to UQ-8 to yield reduced UQ-8. Three quinol:oxygen oxidoreductases, cytochromes bo3, bd-I, bd-II, oxidize ...<|control11|><|separator|>
  67. [67]
    Properties of the two terminal oxidases of Escherichia coli
    Improved production of 2,3‐butanediol and isobutanol by engineering electron transport chain in Escherichia coli. Microbial Biotechnology 2021, 14 (1) , 213-226 ...
  68. [68]
    effects of mutations in components of the aerobic respiratory chain
    E. coli's respiratory chain efficiency depends on the specific enzymes used. The bd-type oxidase is less efficient than the bo-type, and most electron flux ...
  69. [69]
    Modifications of the Aerobic Respiratory Chain of Paracoccus ... - NIH
    Dec 3, 2019 · Paracoccus denitrificans is a strictly respiring bacterium with a core respiratory chain similar to that of mammalian mitochondria.
  70. [70]
    The Paracoccus Denitrificans Electron Transport System
    The electron transport system of Paracoccus denitrificans (then known as Micrococcus denitrificans) ... similar to those found in mitochondria. The identification ...
  71. [71]
    CtaM Is Required for Menaquinol Oxidase aa3 Function in ...
    Jul 12, 2016 · Terminal oxidases of Bacillus subtilis strain 168: one quinol oxidase, cytochrome aa3 or cytochrome bd, is required for aerobic growth. J ...
  72. [72]
    The terminal quinol oxidases of Bacillus subtilis have ... - PubMed
    Cytochrome aa3-600 was found to be the major terminal oxidase in log phase cells. This enzyme was shown to translocate protons across the membrane.Missing: resistant | Show results with:resistant
  73. [73]
    Relationship between cardiolipin metabolism and oxygen ...
    In particular, the electron transport chain in B. subtilis contains two major branches, one quinol oxidase branch and one cytochrome c oxidase one. Three ...
  74. [74]
    New Model for Electron Flow for Sulfate Reduction in Desulfovibrio ...
    Typically, SRB of the genus Desulfovibrio use H2, organic acid substrates, formate, or short-chain alcohols as electron donors for sulfate reduction, a process ...
  75. [75]
    Genetics and Molecular Biology of the Electron Flow for ... - Frontiers
    Progress in the genetic manipulation of the Desulfovibrio strains has provided an opportunity to explore electron flow pathways during sulfate respiration.Missing: chain | Show results with:chain
  76. [76]
    Reduced coenzyme F420: Heterodisulfide oxidoreductase, a - PNAS
    Here we report that the electron transfer from F420H2 to the heterodisulfide as catalyzed by washed vesicles of the meth- anogenic strain Gol gives rise to a A/ ...
  77. [77]
    Reduced coenzyme F420: heterodisulfide oxidoreductase, a proton
    This enzyme system exhibits the phenomenon of coupling and uncoupling and represents a different kind of electron transport chain with the heterodisulfide of 2- ...
  78. [78]
    The cytochrome bd respiratory oxygen reductases - PMC
    Cytochrome bd is a respiratory quinol:O 2 oxidoreductase found in many prokaryotes, including a number of pathogens.
  79. [79]
    Cytochrome bd Displays Significant Quinol Peroxidase Activity
    Jun 9, 2016 · Cytochrome bd has been proposed to confer protection to oxygen-sensitive enzymes and to help protect the cell from nitrosative and ROS stresses7 ...
  80. [80]
  81. [81]
    The genetic basis of energy conservation in the sulfate-reducing ...
    Sulfate-reducing bacteria play major roles in the global carbon and sulfur cycles, but it remains unclear how reducing sulfate yields energy.
  82. [82]
    Electron Bifurcation and Confurcation in Methanogenesis ... - Frontiers
    Jun 19, 2018 · The F420H2 donates electrons to a membrane complex that generates a proton gradient driving ATP synthesis. Introduction. Methane-producing ...
  83. [83]
    Structure of Sr-substituted photosystem II at 2.1 Å resolution ... - PNAS
    Oxygen-evolving complex of photosystem II (PSII) is a tetra-manganese calcium penta-oxygenic cluster (Mn4CaO5) catalyzing light-induced water oxidation ...Missing: Mn4CaO5 | Show results with:Mn4CaO5
  84. [84]
    Molecular basis of plastoquinone reduction in plant cytochrome b 6 f
    Oct 3, 2024 · Cytb6f operates according to a modified 'Q-cycle' that was originally developed for cytochrome bc1 (cytbc1). Cytbc1 is related to cytb6f and ...
  85. [85]
    Structure of the far-red light utilizing photosystem I of Acaryochloris ...
    Apr 20, 2021 · Here, we report the structure of A. marina photosystem I (PSI) reaction center, determined by cryo-electron microscopy at 2.58 Å resolution.
  86. [86]
    Photosynthetic electron partitioning between [FeFe]-hydrogenase ...
    May 23, 2011 · (A) In LEF, light-activated PSII extracts electrons from water and transfers them in a Z-scheme pattern to plastoquinone (PQ), then, through ...
  87. [87]
    The proton to electron stoichiometry of steady-state photosynthesis ...
    The rates of proton pumping through the electron transfer chain and the CFO-CF1 ATP synthase (ATPase) were estimated by measuring the DIRK signals associated ...
  88. [88]
    State transitions in Chlamydomonas reinhardtii strongly modulate ...
    In higher plants, all LHCII is bound to PSII in state 1, whereas in state 2, which can be induced by overexciting PSII, part of LHCII (around 15%) moves to PSI ...
  89. [89]
    A kinetic model of rapidly reversible nonphotochemical quenching
    Oxygen-evolving photosynthetic organisms possess nonphotochemical quenching (NPQ) pathways that protect against photo-induced damage. The majority of NPQ in ...
  90. [90]
    Modeling the electron transport chain of purple non-sulfur bacteria
    Anoxygenic photosynthesis (under anaerobic conditions in the light) and respiration (under aerobic conditions) are the two major strategies of the ...
  91. [91]
    Anoxygenic photosynthesis with emphasis on green sulfur bacteria ...
    Our review explores the ecological significance of anoxygenic photosynthetic bacteria ... photosynthetic Electron transport chain in green sulfur Bacteria.
  92. [92]
    Anoxygenic Photosynthesis in Photolithotrophic Sulfur Bacteria and ...
    May 22, 2021 · In the process of anoxygenic photosynthesis, they can also use molecular hydrogen as an electron donor, and carbon dioxide gas is used in ...
  93. [93]
  94. [94]
    An overview of anoxygenic phototrophic bacteria and their ...
    Anoxygenic phototropic bacteria such as purple bacteria, have a type II (Pheo-Q type) reaction center [52]. In these organisms, when light energy (in photons) ...
  95. [95]
    Leigh Syndrome: A Comprehensive Review of the Disease and ...
    Leigh syndrome (LS) is a severe neurodegenerative condition with an early onset, typically during early childhood or infancy.
  96. [96]
    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 ...
  97. [97]
    Leber Hereditary Optic Neuropathy (LHON) - StatPearls - NCBI - NIH
    These mutations are absent or very rare among normal controls. Except in rare cases of de novo occurrence of a primary LHON mutation, an mtDNA mutation will be ...Continuing Education Activity · Introduction · Etiology · History and Physical
  98. [98]
    Respiratory Chain Complex I Deficiency in Leber Hereditary Optic ...
    Nov 22, 2024 · Although the mutations in complex I of electric chain may cause ... Leber hereditary optic neuropathy primary mitochondrial DNA mutation.
  99. [99]
    Mitochondrial complex I deficiency in Parkinson's disease - PubMed
    These results indicated a specific defect of Complex I activity in the substantia nigra of patients with Parkinson's disease.Missing: seminal paper
  100. [100]
    Mitochondrial Complex I deficiency: guilty in Parkinson's disease
    Apr 23, 2022 · PD is characterized by the loss of dopaminergic neurons in the substantia nigra that is responsible for impaired motor function. Most PD ...Missing: review | Show results with:review
  101. [101]
    Mitochondria, Amyloid β, and Alzheimer's Disease - PMC
    Aβ has been shown to directly inhibit complex IV which would lead to bioenergetic impairment and increased formation of reactive oxygen species. Postmortem ...
  102. [102]
    Cytochrome c oxidase deficiency in neurons decreases both ... - PNAS
    Cytochrome c oxidase deficiency in neurons decreases both oxidative stress and amyloid formation in a mouse model of Alzheimer's disease · Abstract · Results.
  103. [103]
    The Warburg Effect: How Does it Benefit Cancer Cells? - PMC
    It is likely that the Warburg Effect provides an overall benefit that supports a tumor microenvironment conducive to cancer cell proliferation.
  104. [104]
    SDH defective cancers: molecular mechanisms and treatment ... - NIH
    Apr 26, 2025 · In many malignant tumors, the obvious consequence of SDH deficiency is the accumulation of the oncometabolite succinate. Excessive succinate ...
  105. [105]
    Epigenetic and metabolic reprogramming of SDH-deficient ...
    Dec 1, 2020 · SDHx mutations lead to the accumulation of succinate, which acts as an oncometabolite by inhibiting iron(II) and alpha-ketoglutarate-dependent ...
  106. [106]
    Aging: A mitochondrial DNA perspective, critical analysis and an ...
    That is: (1) increased ROS production in aging leads to increased mtDNA mutagenesis; and (2) increased mtDNA mutation loads result in increased mitochondrial ...
  107. [107]
    Somatic mtDNA mutations cause aging phenotypes without ... - PNAS
    Dec 19, 2005 · The profound aging phenotypes generated by accumulation of somatic mtDNA mutations are thus not mediated by dramatically increased ROS ...
  108. [108]
    Reverse Electron Transport at Mitochondrial Complex I in Ischemic ...
    Apr 6, 2023 · In this review, we provide a historical account of the roles of ROS ... Mitochondrial electron transport chain, ROS generation and uncoupling.
  109. [109]
    Initial experience in the treatment of inherited mitochondrial disease ...
    EPI-743 appears to be effective in altering the natural history of mitochondrial disease. Mitochondrial patients showed improvement in quality-of-life.
  110. [110]
    Scientists use gene editing to correct harmful mitochondrial ...
    Jun 24, 2025 · Gene editing technology allows to introduce and correct disease-linked mitochondrial DNA mutations in liver and skin cells.
  111. [111]
    Therapies for Mitochondrial Disease: Past, Present, and Future - PMC
    Jul 25, 2025 · Emerging therapies include dietary intervention, small molecule therapies aimed to restore mitochondrial function, stem cell or liver ...