Fact-checked by Grok 2 weeks ago

Oxidative phosphorylation

Oxidative phosphorylation is a central metabolic pathway in aerobic eukaryotic cells that harnesses the energy released from the oxidation of reduced electron carriers, such as NADH and FADH₂, to drive the synthesis of adenosine triphosphate (ATP) through the phosphorylation of ADP, utilizing molecular oxygen as the terminal electron acceptor. This process occurs primarily in the inner mitochondrial membrane, where it is mediated by the electron transport chain (ETC) consisting of four large protein complexes (I–IV) embedded in the lipid bilayer, along with ATP synthase (complex V). Electrons derived from nutrient catabolism flow sequentially through these complexes, releasing energy that pumps protons (H⁺) from the mitochondrial matrix into the intermembrane space, establishing an electrochemical proton gradient known as the proton motive force. ATP synthase then exploits this gradient via chemiosmosis, allowing protons to flow back into the matrix and powering the rotation of its subunits to catalyze ATP formation. As the chief mechanism for ATP production in mitochondria, oxidative phosphorylation accounts for the majority of the cell's energy needs, generating approximately 30–32 ATP molecules per fully oxidized glucose molecule—vastly more efficient than in or the . Most of the usable energy from the breakdown of carbohydrates and fats is captured through this pathway, supporting essential cellular functions in higher animals and while maintaining metabolic . The reduction of oxygen to at complex IV prevents electron buildup and ensures the process's continuity, though dysregulation can lead to formation.

Core Mechanism

Chemiosmosis

Chemiosmosis refers to the process in which the translocation of protons (H⁺ ions) across a , driven by electron transport, establishes an known as the proton motive force, which in turn powers the synthesis of (ATP). This mechanism couples the exergonic flow of electrons through the respiratory chain to the endergonic formation of ATP, without requiring a high-energy chemical intermediate. The chemiosmotic hypothesis was first proposed by British biochemist Peter Mitchell in 1961, positing that the energy released during electron transfer in oxidative phosphorylation is used to actively transport protons across the , creating a transmembrane that drives ATP production via a dedicated . Initially met with significant controversy, as it challenged prevailing chemical coupling models that envisioned direct phosphoryl transfer between respiratory enzymes and , the hypothesis faced skepticism and rejection by many leading biochemists for over a decade. Accumulating experimental evidence from studies on mitochondria, chloroplasts, and bacterial systems, including measurements of proton and their dissipation by uncouplers, ultimately validated the theory, leading to its widespread acceptance and Mitchell's receipt of the in 1978. The proton motive force (Δp), which encapsulates the chemiosmotic potential, is expressed quantitatively by the equation: \Delta p = \Delta \psi - \frac{2.303 RT}{F} \Delta \mathrm{pH} where Δψ represents the electrical membrane potential (in millivolts), ΔpH is the pH gradient across the membrane, R is the gas constant (8.314 J mol⁻¹ K⁻¹), T is the absolute temperature (in kelvin), and F is the Faraday constant (96,485 C mol⁻¹). This force arises primarily from proton pumping during electron transport, where respiratory complexes translocate protons from the mitochondrial matrix (low H⁺ concentration) to the intermembrane space (high H⁺ concentration), generating both a pH differential (alkaline matrix) and a negative-inside membrane potential. ATP synthase, embedded in the inner mitochondrial membrane, serves as a rotary proton channel that facilitates the downhill flow of protons back into the matrix, converting the free energy of the proton motive force into mechanical rotation that catalyzes ATP formation from ADP and inorganic phosphate (Pᵢ). This process exemplifies indirect coupling, as the enzyme harnesses the delocalized electrochemical energy of the gradient rather than a localized chemical bond, ensuring efficient energy transduction in cellular respiration.

Proton Motive Force

The proton motive force (PMF), arising from the chemiosmotic theory, represents the total of protons across the and drives various cellular processes. This force comprises two main components: the electrical (Δψ), which is typically 150–180 mV (negative inside ), and the chemical pH gradient (ΔpH), which is approximately 0.5–1 unit (with more alkaline than the ). The overall PMF (Δp) in energized mitochondria is around 180–220 mV, with Δψ accounting for the majority of the driving force (roughly 70–80%) and ΔpH contributing the remainder (20–30%), though the relative contributions can vary with metabolic state and experimental conditions. Measurement of the PMF components relies on sensitive fluorescent probes that report on the gradients in living cells or isolated mitochondria. For Δψ, lipophilic cationic dyes such as tetramethylrhodamine methyl ester (TMRM) are widely used; these accumulate in the matrix in a potential-dependent manner, with fluorescence intensity calibrated against uncouplers like carbonyl cyanide p-trifluoromethoxyphenylhydrazone (FCCP) to quantify the voltage. For ΔpH, ratiometric pH-sensitive dyes like 2',7'-bis-(2-carboxyethyl)-5-(and-6)-carboxyfluorescein (BCECF) or seminaphthorhodafluor-1 (SNARF-1) are loaded into the matrix and provide pH readings based on emission wavelength shifts, often calibrated with ionophores such as nigericin to equilibrate pH across the membrane. These techniques allow non-invasive assessment but require careful controls for dye distribution, photobleaching, and binding artifacts to ensure accuracy. Maintenance of the PMF depends on the low permeability of the to protons and most ions, preventing passive dissipation of the gradient; this impermeability is conferred by the and selective protein channels. Regulated ion fluxes, such as those mediated by the K⁺/H⁺ exchanger (KHE) or the mitochondrial calcium uniporter (MCU), play a critical role in sustaining the PMF by counteracting osmotic swelling, adjusting matrix volume, and fine-tuning local without compromising the overall proton gradient. Disruptions in these fluxes, such as through pharmacological inhibition, can alter the balance between Δψ and ΔpH components. Collapse of the PMF, often termed uncoupling, occurs when protons re-enter independently of , dissipating the gradient and converting stored energy into rather than . This process is physiologically relevant in , where uncoupling protein 1 () facilitates proton leak to generate for non-shivering . Pathological uncoupling, induced by agents like , similarly leads to release and reduced efficiency of oxidative phosphorylation, highlighting the PMF's role in .

Electron Carriers and Complexes

Mobile Electron Carriers

Mobile electron carriers are essential soluble or lipid-soluble molecules that facilitate between the membrane-bound respiratory complexes in the during oxidative phosphorylation. These carriers operate by diffusing within their respective compartments—either the for hydrophobic molecules or the aqueous —enabling efficient shuttling of electrons without direct contact between distant complexes. In eukaryotic mitochondria, the primary mobile carriers are ubiquinone and , which handle the transfer of electrons derived from NADH and FADH₂ oxidation. Ubiquinone, also known as coenzyme Q (CoQ), is a lipid-soluble benzoquinone derivative that resides within the hydrophobic core of the inner mitochondrial membrane. Its structure consists of a redox-active quinone head group attached to a long polyisoprenoid tail comprising 10 isoprene units in humans, which anchors it in the membrane and allows lateral diffusion. Ubiquinone cycles through three redox states: the fully oxidized form (ubiquinone), the fully reduced form (ubiquinol, carrying two electrons and two protons), and the semiquinone radical intermediate (ubisemiquinone). In its role, ubiquinol accepts two electrons from Complex I (NADH:ubiquinone oxidoreductase) or Complex II (succinate:ubiquinone oxidoreductase), becoming oxidized to ubiquinone, which then diffuses to Complex III (ubiquinol:cytochrome c oxidoreductase) to donate the electrons, contributing to proton translocation across the membrane. Cytochrome c is a small, water-soluble heme-containing protein that operates in the intermembrane space of mitochondria. It features a covalently bound group with an iron center that undergoes reversible one-electron changes between Fe³⁺ (oxidized) and Fe²⁺ (reduced) states, enabling it to shuttle single electrons. receives electrons from reduced at Complex III via the Rieske iron-sulfur protein and , then diffuses to Complex IV (), where it donates the electron to reduce CuA centers, ultimately supporting oxygen to . This one-electron transfer contrasts with ubiquinone's two-electron capacity, allowing precise matching to the requirements of downstream complexes. In prokaryotes, analogous mobile carriers exist, such as various cytochromes c that perform similar one-electron shuttling in respiratory chains, while copper-containing proteins like plastocyanin serve comparable roles in photosynthetic electron transport, highlighting functional parallels to eukaryotic cytochrome c despite structural differences. Diffusion rates and binding affinities are critical for the efficiency of these carriers; for instance, ubiquinone exhibits lateral diffusion in the membrane with coefficients around 10^{-8} cm²/s, and association rate constants with respiratory complexes on the order of 10⁶–10⁸ M⁻¹ s⁻¹, contributing to electron transfer under physiological conditions. Cytochrome c, meanwhile, diffuses primarily in three dimensions within the intermembrane space at rates around 10⁻⁷ cm²/s, with binding affinities to Complex III and IV in the micromolar range (K_d ≈ 1–10 μM under physiological conditions), facilitating frequent collisions and efficient electron handoff. These mobile carriers demonstrate remarkable evolutionary , with ubiquinone and homologs present across eukaryotes and many prokaryotes, underscoring their ancient origins in the development of aerobic from bacterial ancestors. This reflects their fundamental role in maintaining flux and proton motive force generation, with variations primarily in tail length or attachment adapting to diverse membrane environments.

Eukaryotic Respiratory Complexes

The (ETC) in eukaryotic mitochondria comprises four multi-subunit protein complexes (I–IV) embedded in the , which oxidize reducing equivalents derived from nutrient metabolism and reduce molecular oxygen to . These complexes are present in varying abundances that reflect their functional demands, with Complexes I, III, and IV generally more numerous than Complex II. Beyond individual complexes, the respiratory chain organizes into higher-order supercomplexes, often termed respirasomes, which integrate Complexes I, III₂ (as a dimer), and IV in stoichiometries such as I₁III₂IV₁ or I₁III₂IV₂. These assemblies enhance efficiency by providing dedicated pathways for mobile carriers like ubiquinone and , thereby minimizing distances, optimizing substrate channeling, and lowering (ROS) generation compared to fully randomized distributions. The signature inner membrane cardiolipin is critical for stabilizing both individual complexes and supercomplexes, binding to key subunits (e.g., in Complexes I, III, and IV) to promote proper folding, oligomerization, and retention within the membrane lipid environment. deficiency disrupts supercomplex formation and reduces overall respiratory capacity. Electrons enter the from NADH at Complex I or from FADH₂ at , converging on the ubiquinone (Q) pool to form , which donates electrons to Complex III; these are then shuttled via to Complex IV, culminating in O₂ reduction, with concomitant proton extrusion across the membrane at Complexes I, III, and IV to establish the proton motive force.

Organization and Function of Complexes

Complex I: NADH:Ubiquinone Oxidoreductase

Complex I, known as NADH:ubiquinone , is the entry point for electrons from NADH into the mitochondrial and the largest respiratory complex, comprising 45 subunits in mammalian mitochondria with a total of about 1 MDa. Its overall is L-shaped, consisting of a hydrophilic peripheral arm protruding into the and a hydrophobic membrane arm in the . The peripheral arm, formed by 14 core subunits conserved across , contains all the redox-active cofactors, while the membrane arm, with additional supernumerary subunits unique to eukaryotes, facilitates proton translocation. The redox centers include one non-covalently bound (FMN) cofactor and eight iron-sulfur (Fe-S) clusters, which mediate . The FMN is located in the N-terminal domain of the 51 kDa subunit within the peripheral arm, serving as the initial . The Fe-S clusters, including binuclear [2Fe-2S], tetranuclear [4Fe-4S], and one [8Fe-7S] cluster (N2), are distributed across subunits such as the 51 kDa, 24 kDa, 75 kDa, and 49 kDa proteins, forming a linear chain that connects the FMN to the ubiquinone (Q)-binding site. These clusters enable low-potential with minimal energy loss, as their potentials increase progressively along the chain. Electrons enter the pathway when NADH binds to a Rossmann fold in the 51 kDa subunit and transfers a hydride ion (H⁻) to FMN, reducing it to FMNH₂ and releasing a proton into the matrix. The two electrons are then passed singly through the Fe-S clusters—starting from cluster N3 near the FMN, via N4, N5, N6a, N6b, and terminal N2—over a distance of approximately 60 Å to the Q-binding site at the junction of the peripheral and membrane arms. This transfers the two electrons to Q, reducing it to ubiquinol (QH₂), which serves as a mobile electron carrier in the membrane. The entire electron transfer is highly efficient, with no significant side reactions under physiological conditions. Coupled to this , Complex I pumps four protons from to the per two electrons transferred (4 H⁺/2e⁻), contributing substantially to the proton motive force. The mechanism relies on redox-driven conformational changes initiated at the -binding site upon , which propagate through the arm to drive proton translocation via four discrete channels. These channels involve conserved charged residues and molecules in transmembrane helices of subunits ND2, ND4, and ND5, with the process resembling a wave-like propagation rather than a classical Q-cycle; semiquinone intermediates may stabilize the active state but do not directly bifurcate electrons. This coupling ensures without direct chemical linkage between electron and proton paths. High-resolution structural insights into mammalian Complex I were advanced by cryo-electron microscopy, with the bovine structure resolved at 4.3 in 2016, delineating subunit interfaces, cofactor positions, and the Q-site architecture. Earlier, a 3.3 crystal structure of the bacterial homolog from Thermus thermophilus in 2013 provided a template for eukaryotic models, confirming the conserved L-shape and Fe-S chain. Mutations in nuclear-encoded Complex I subunits, such as those in NDUFS2 (encoding the 49 kDa subunit) or NDUFV1 (encoding the 51 kDa subunit), disrupt assembly or redox function, leading to isolated Complex I deficiency and severe mitochondrial disorders like , characterized by neurodegeneration and . For instance, the p.M292T mutation in NDUFS2 abolishes Fe-S cluster N2 binding, severely impairing activity and causing early-onset . Over 100 such mutations have been identified, highlighting Complex I's vulnerability in human disease.

Complex II: Succinate:Ubiquinone Oxidoreductase

Complex II, also known as or , serves as the entry point for electrons derived from the into the , catalyzing the oxidation of succinate to fumarate while reducing ubiquinone to . This enzyme complex is unique among the respiratory complexes as it participates in both the and the , functioning as in the former. Unlike Complexes I, III, and IV, Complex II does not translocate protons across the , contributing electrons to the without directly generating a proton motive force. The structure of Complex II consists of four nuclear-encoded subunits: SDHA (the flavoprotein subunit), SDHB (the iron-sulfur protein subunit), SDHC (cytochrome b small subunit), and SDHD (cytochrome b large subunit). SDHA binds the cofactor and houses the succinate oxidation site, while SDHB coordinates three iron-sulfur clusters—a [2Fe-2S], a [4Fe-4S], and a [3Fe-4S] cluster—that facilitate from FADH₂ (generated during succinate oxidation) to the ubiquinone-binding site at the interface of SDHB, SDHC, and SDHD. A group, coordinated between SDHC and SDHD, is positioned near the [3Fe-4S] cluster and ubiquinone site, where it acts as an electron sink to stabilize the final step to , although its precise role in the forward reaction remains under investigation. The overall architecture includes a hydrophilic domain (SDHA and SDHB) protruding into the and a membrane-embedded domain (SDHC and SDHD) with two transmembrane helices each, anchoring the complex without proton-pumping capability. The reaction catalyzed by Complex II is the two-electron oxidation of succinate to fumarate, coupled to the reduction of ubiquinone (Q) to (QH₂): succinate + Q → fumarate + QH₂. Electrons from succinate reduce to FADH₂ in SDHA, then pass sequentially through the iron-sulfur clusters in SDHB before reaching the and finally reducing ubiquinone at the Q-site. The atomic structure of Complex II was first elucidated from porcine heart mitochondria at 2.4 resolution in , revealing the precise arrangement of prosthetic groups and inhibitor-binding sites, such as those for thenoyltrifluoroacetone and carboxin. This structure confirmed the linear pathway and highlighted the heme's proximity to the Q-site, supporting its role in preventing formation during reverse electron flow. Mutations in SDH genes are associated with human diseases, including hereditary paraganglioma-pheochromocytoma syndromes for SDHB, SDHC, and SDHD variants, which disrupt complex assembly and lead to pseudohypoxic signaling via succinate accumulation. In contrast, SDHA mutations often cause mitochondrial encephalopathies, such as , characterized by severe neurological deficits due to impaired energy metabolism.

Complex III: Ubiquinol:Cytochrome c Oxidoreductase

Complex III, also known as ubiquinol:cytochrome c oxidoreductase or the cytochrome bc1 complex, is a central component of the electron transport chain that catalyzes the transfer of electrons from ubiquinol (QH2) to cytochrome c while translocating protons across the inner mitochondrial membrane. In eukaryotic mitochondria, the complex functions as a symmetric dimer, with each monomer comprising 11 protein subunits, including three core catalytic subunits: cytochrome b, the Rieske iron-sulfur protein, and cytochrome c1. Cytochrome b spans the membrane with eight transmembrane helices and binds two b-type hemes (bL and bH, with low and high midpoint potentials, respectively), while the Rieske protein contains a 2Fe-2S cluster and cytochrome c1 harbors a c-type heme; the remaining subunits include two large core proteins (core 1 and core 2) that resemble mitochondrial processing peptidases and seven smaller supernumerary subunits that stabilize the structure. The overall architecture reveals a monomeric unit with 13 transmembrane helices, forming intermonomer cavities that accommodate ubiquinone binding sites, as determined by the first X-ray crystal structure of the bovine complex at 2.6 Å resolution. The electron transfer in Complex III proceeds via the Q-cycle mechanism, which enables bifurcation of the two electrons from QH2 oxidation, coupling scalar proton release to vectorial translocation and achieving a net translocation of four protons per two electrons transferred to cytochrome c. The cycle begins at the Qo site on the positive (intermembrane space) side of the membrane, where QH2 binds and undergoes oxidation: the high-potential electron path transfers one electron via the Rieske 2Fe-2S cluster to the heme of cytochrome c1 and thence to cytochrome c, while the low-potential electron is passed through heme bL to heme bH; this initial oxidation releases two protons into the intermembrane space and generates a transient, unstable semiquinone anion (Q•−) at the Qo site. In the second half of the cycle, another QH2 molecule binds to the Qo site and repeats the bifurcation: one electron again follows the high-potential path to a second cytochrome c, releasing two more protons to the intermembrane space, while the second low-potential electron traverses bL and bH to the Qi site on the negative (matrix) side, where a ubiquinone (Q) is first reduced to semiquinone (Q•−) by the first such electron (taking up one proton from the matrix) and then fully reduced to QH2 by the second electron (taking up a second proton from the matrix). This bifurcated electron flow in the Q-cycle results in the net oxidation of one QH2 to (with the other QH2 effectively recycled at Qi), transfer of two electrons to two molecules of , and translocation of four protons from the matrix to the per full , thereby contributing to the proton motive force. The semiquinone intermediates, particularly the short-lived Q•− at the Qo site, are stabilized transiently by interactions with residues like His182 of the and the bL but can leak electrons to molecular oxygen, generating as a byproduct and contributing to (ROS) production under conditions of high or partial reduction of the chain. Crystal structures, such as the bovine one at 2.6 Å, have illuminated these sites, showing the Qo site near the convergence of the Rieske head and , and the Qi site deeper within the dimer interface, with semiquinone stability influenced by nearby histidine ligands and quinone headgroup orientation. Blockade at the Qi site, for instance by antimycin, exacerbates Qo semiquinone accumulation and ROS leakage by preventing low-potential electron acceptance, highlighting the mechanistic linkage between cycle progression and oxidative stress potential.

Complex IV: Cytochrome c Oxidase

Complex IV, also known as (CcO), is the terminal enzyme in the mitochondrial , consisting of 13 subunits in bovine mitochondria and up to 18 in some eukaryotic species, with three core subunits (COX1, COX2, and COX3) encoded by and the rest by nuclear DNA. The catalytic core features two a groups—heme a associated with and heme a3 forming part of the binuclear center—and two copper centers: CuA in subunit COX2 for initial electron acceptance and CuB in subunit COX1 coordinated to the binuclear center alongside heme a3. These metal centers enable the four-electron reduction of molecular oxygen to water while contributing to proton translocation across the . The overall reaction catalyzed by Complex IV is: $4 \ cyt \ c^{2+} + O_2 + 8 \ H^+_{matrix} \rightarrow 4 \ cyt \ c^{3+} + 2 \ H_2O + 4 \ H^+_{intermembrane} This process consumes four electrons from reduced cytochrome c, binds one O2 molecule, and utilizes four protons from the matrix to form water, while pumping an additional four protons from the matrix to the intermembrane space, thereby enhancing the proton motive force. The reaction occurs at the binuclear center (BNC) in subunit COX1, where heme a3 and CuB facilitate O2 binding and reduction, preventing harmful intermediates like superoxide. The catalytic cycle of involves sequential transfers and proton movements, progressing through key intermediates at the BNC: the A state (oxygen-bound complex after initial reduction), the P state (peroxy intermediate following O-O bond cleavage), and the F state (ferryl-oxo species with a on a cross-linked residue). These states reflect the stepwise four- reduction of , with proton pumping coupled to delivery, particularly during transitions from P to F and subsequent steps, ensuring vectorial proton without net charge imbalance. The cycle returns to the resting oxidized state (R or E) after full reduction and water release. The atomic structure of bovine heart CcO was first resolved in 1995 at 2.3 resolution, revealing the arrangement of the 13 subunits, molecules, and metal centers, which provided foundational insights into the binuclear center's and subunit interfaces. Subsequent refinements confirmed the dinuclear copper-like nature of and the role of a residue in facilitating . Activity of Complex IV is allosterically regulated by the ATP/ ratio, with ATP binding to a site on subunit COX3 inhibiting at high ATP levels to prevent over-reduction, while relieves this inhibition to match with energy demand.00183-5) This mechanism integrates CcO function with cellular energy status, often modulated by events.

Alternative Oxidases and Reductases

Alternative oxidases represent non-canonical terminal enzymes in the mitochondrial (ETC) of certain eukaryotes, primarily , where they provide an alternative route for from the ubiquinone pool to molecular oxygen without contributing to proton translocation across the . This pathway, known as the alternative oxidase (AOX) route, oxidizes (QH₂) directly to ubiquinone (Q), reducing O₂ to water and dissipating the energy as heat rather than conserving it for ATP synthesis. Unlike the canonical pathway, AOX activity is cyanide-resistant, allowing to continue under conditions that inhibit complex IV. Structurally, plant AOX is an that functions as a homodimer embedded in the , with each featuring a binuclear non-heme iron center at its catalytic site. This di-iron center, coordinated by and residues, facilitates the four-electron of O₂, and the enzyme's activity is regulated by of a bridge and binding of activators like pyruvate. The dimeric form is essential for activity, and the protein belongs to the di-iron superfamily, sharing mechanistic similarities with other oxygen-utilizing enzymes. AOX plays critical roles in stress responses, including the mitigation of (ROS) accumulation by preventing over-reduction of the ubiquinone pool during conditions like attack, , or . In thermogenic tissues, such as the florets of (Symplocarpus renifolius), AOX coexpresses with uncoupling proteins to generate heat for volatilizing attractants and maintaining optimal temperatures around 23°C, even in environments, without ATP production. This heat generation supports reproductive processes by enhancing efficiency. Another key alternative reductase is , a that integrates electrons from β-oxidation and other catabolic pathways into the ETC via the ubiquinone pool. ETF-QO, anchored in the , contains a 4Fe-4S and FAD cofactor, accepting electrons from and transferring them to ubiquinone, thereby linking dehydrogenases to the respiratory chain without direct proton pumping. This enzyme is essential for efficient oxidation of lipids and , preventing metabolic bottlenecks under high substrate loads. Evolutionarily, AOX is absent in mammals and vertebrates but is widely distributed across plants, fungi, and certain protists, such as Naegleria species and trypanosomes, suggesting an ancient origin with lineage-specific losses. Its presence in diverse microbial eukaryotes highlights a conserved role in adaptive respiration, particularly in oxygen-variable environments.

ATP Production

ATP Synthase Structure and Mechanism

, also known as FoF1-ATP synthase or Complex V of the respiratory chain, is a rotary composed of two functionally distinct domains: the membrane-integral Fo sector and the soluble F1 sector. The Fo sector anchors the enzyme in the (or bacterial plasma membrane) and transduces the proton motive force into mechanical rotation, while the F1 sector catalyzes ATP synthesis from and inorganic phosphate. The Fo sector consists primarily of a of c-subunits (c-ring) and the a-subunit, along with additional peripheral elements like the b-subunit dimer that forms part of the . The c-ring, formed by multiple identical c-subunits arranged in a symmetrical oligomeric , typically contains 8 to 15 subunits depending on the , with mammalian mitochondria featuring 8 c-subunits and mitochondria having 10. Each c-subunit spans the with two transmembrane helices and bears a critical proton-binding residue (Asp61 in bacterial homologs) that interacts with protons translocated through the a-subunit channel. The a-subunit, embedded adjacent to the c-ring, contains two half-channels that facilitate proton entry from the and exit to , enabling sequential and of the c-subunits to drive ring rotation. In contrast, the F1 sector protrudes into the and comprises a hexameric head formed by three α-subunits and three β-subunits arranged alternately as an (αβ)3 assembly, along with a central γ-subunit that acts as the , and smaller δ- and ε-subunits. The α- and β-subunits form a globular structure approximately 10 nm in diameter, with the β-subunits housing the three catalytic nucleotide-binding sites at the interfaces with adjacent α-subunits. The γ-subunit extends from the Fo c-ring through the central of the F1 hexamer, coupling from Fo to conformational changes in F1. The elements, including the b-subunit and the OSCP (oligomycin sensitivity-conferring protein) homolog, prevent co-rotation of the F1 head with the . The rotary catalysis mechanism relies on proton flow through Fo, which induces counterclockwise rotation (viewed from the matrix) of the c-ring and attached γ-subunit, typically in 120° steps corresponding to the threefold symmetry of F1. In species with an n-subunit c-ring, a full 360° rotation requires translocation of n protons, with each proton driving an angular step of 360°/n (e.g., ~45° per proton for an 8-subunit ring in mammals). This rotation propagates to F1, where it forces sequential conformational changes in the three β-subunits: from open (O, nucleotide release), to loose (L, and Pi binding), to tight (T, ATP synthesis), cycling through these states with each 120° turn. Central to this process is the binding change mechanism proposed by Paul Boyer, in which ATP synthesis occurs without direct energy input for the chemical bond formation; instead, the energy from proton translocation alters the affinities of the three catalytic sites cooperatively. One site binds substrates loosely, another tightens to form ATP with high affinity (but without release), and the third loosens to release the newly synthesized ATP, with the ~120° rotation per ATP molecule ensuring site interconversion. This mechanism accommodates the observed asymmetry in nucleotide occupancy among the β-subunits. Key insights into the structure were provided by of the bovine mitochondrial F1 sector at 2.8 Å resolution in 1994, revealing the asymmetric (αβ)3 arrangement and distinct conformations of the β-subunits with nucleotides. A higher-resolution view of larger assemblies, including the subcomplex with F1, was achieved at 3.2 Å in 2009, elucidating interactions that maintain rotational asymmetry. More recent cryo-EM structures of intact enzymes have confirmed the c-ring and dynamic interfaces. The enzyme's activity is inhibited by , a that binds within the Fo sector, specifically occluding the proton pathway in the a-subunit-c-ring and preventing rotation by stabilizing the essential carboxylate in a non-protonatable conformation. This blockade halts both ATP synthesis and hydrolysis without affecting isolated F1 activity.

Energetics of ATP Synthesis

The energetics of ATP synthesis in oxidative phosphorylation are quantified by the , defined as the number of ATP molecules produced per atom of oxygen reduced (or per two electrons transferred to oxygen). Experimental measurements in isolated liver mitochondria yield a of approximately 2.5 for NADH-linked substrates and 1.5 for FADH₂-linked substrates such as succinate. These values reflect the overall of coupling electron transport to ATP production under physiological conditions. These ratios arise from the of proton translocation across the and the protons required for ATP and export. Oxidation of NADH transfers two electrons through Complexes I, III, and , pumping a total of 10 protons: 4 from Complex I, 4 from Complex III, and 2 from Complex . In contrast, FADH₂ oxidation via Complex II bypasses Complex I, resulting in 6 protons pumped (4 from III and 2 from ). The mitochondrial F₁F₀-ATP rotates its c-ring (with 8 c-subunits in mammals) to synthesize one ATP per three protons translocated through the F₀ domain, requiring ~2.7 protons per ATP in the matrix; however, exporting ATP to the cytosol via the nucleotide and importing via Pi/H⁺ symport consumes an additional proton equivalent, yielding a total of ~3.7 protons per cytosolic ATP. The ATP yield is thus calculated as: \text{ATP yield} = \frac{\text{H}^+ \text{ pumped}}{\text{H}^+ \text{ per ATP (synthesis + transport)}} For NADH, this gives 10 / 3.7 ≈ 2.7 ATP; for FADH₂, 6 / 3.7 ≈ 1.6 ATP. These theoretical values are consistent with experimental measurements, which often report ~2.5 and ~1.5 due to factors such as proton leaks. Thermodynamically, ATP synthesis under standard biochemical conditions (ΔG°' ≈ 30.5 kJ/mol) is driven by the proton motive force (PMF), which harnesses the energy from proton re-entry. Each proton contributes ~21 kJ/mol across a typical PMF of ~200 mV (calculated as ΔG = F × Δμ_H⁺, where F is the Faraday constant), so three protons provide ~21 kJ/mol to the synthase, enabling an overall efficiency of ~60% when accounting for the full redox span from NADH to O₂ (ΔE°' ≈ 1.14 V, ΔG°' ≈ -220 kJ/mol for 2e⁻). In prokaryotes, variations in stoichiometry enhance yields; without mitochondrial transport costs, the effective H⁺/ATP ratio is closer to 3–4 depending on c-ring size (e.g., c₁₀ in Escherichia coli yields ~3.33 H⁺/ATP, potentially raising P/O to ~3 for NADH), though some bacteria exhibit lower ratios due to alternative complexes or smaller ΔE spans. The process operates near thermodynamic limits but is not 100% efficient due to slippage, where a fraction of electrons traverse the chain without full proton pumping or protons leak back without ATP synthesis, dissipating PMF as heat and allowing metabolic regulation. This intrinsic slip in proton pumps, observed in under high PMF, reduces maximal yields but prevents ROS overproduction and enables in certain tissues.00027-6)

Variations in Prokaryotes

Prokaryotic Electron Transport Chains

In prokaryotes, the (ETC) resides within the , functioning analogously to the or cristae in eukaryotic cells, where it facilitates the transfer of electrons from reduced carriers like NADH and succinate to molecular oxygen while coupling this process to . This integration allows the ETC to contribute directly to cellular without compartmentalization into organelles. The chain's operation generates a (PMF) by translocating protons across the , which powers ATP synthesis and other membrane-associated processes. The conserved core components of prokaryotic ETCs include NDH-1, a proton-pumping NADH: oxidoreductase homologous to eukaryotic Complex I, which oxidizes NADH and transfers electrons to the quinone pool while extruding four protons per two electrons. (SDH), equivalent to Complex II, links the tricarboxylic acid cycle to the by oxidizing succinate to fumarate and reducing ubiquinone without direct proton translocation. Many also feature a cytochrome bc1 complex, akin to Complex III, which oxidizes quinol via a Q-cycle mechanism to translocate additional protons while passing electrons to . Terminal oxidases homologous to Complex IV, such as the aa3-type , complete the chain by reducing oxygen to and pumping protons. harnesses the resulting PMF to synthesize ATP through proton influx. Proton pumping by these complexes occurs outward toward the periplasmic space, establishing a PMF with both electrical (Δψ) and chemical (ΔpH) components across the plasma membrane. This PMF not only drives oxidative phosphorylation but also energizes flagellar rotation for motility and secondary active transport systems in bacteria. In Escherichia coli, for example, the ETC branches after the quinone pool into three terminal oxidases (cytochrome bo₃, bd-I, and bd-II) expressed under varying aerobic conditions, with cytochrome bo₃ serving as the primary high-oxygen quinol oxidase. Cytochrome o and a1 oxidases serve as alternatives to aa3-type enzymes in certain species, enabling flexibility in oxygen affinity and proton pumping efficiency.

Diversity of Prokaryotic Complexes

Prokaryotes exhibit remarkable diversity in their respiratory complexes, allowing to varied environmental conditions such as oxygen availability, , and niches. Unlike the more uniform mitochondrial in eukaryotes, prokaryotic systems often feature branched pathways with alternative terminal acceptors and electron donors, enabling efficient through oxidative phosphorylation in diverse habitats. This variability arises from evolutionary pressures, resulting in specialized complexes that optimize proton motive force generation under specific conditions. Terminal oxidases in prokaryotes include the heme-copper bo₃ oxidase, which operates effectively in high-oxygen environments due to its relatively low oxygen (Km ≈ 0.1–1 μM O₂), facilitating rapid when O₂ is abundant. In contrast, the bd-type oxidase exhibits high oxygen (Km ≈ 0.001–0.01 μM O₂), making it suitable for microaerobic or low-oxygen conditions, where it serves as an oxygen scavenger to protect against while supporting . bo₃ actively pumps protons (approximately 4 H⁺ per 2 electrons), contributing significantly to the proton motive force, whereas bd primarily relies on scalar proton release without substantial vectorial pumping, though it maintains electrogenic activity. These oxidases, often coexisting in bacteria like Escherichia coli, allow seamless switching based on oxygen levels to sustain ATP synthesis. Alternative electron donors further diversify prokaryotic chains, such as the Na⁺-pumping NADH:quinone oxidoreductase (Na⁺-NQR), a complex unique to prokaryotes that translocates Na⁺ ions instead of H⁺ during NADH oxidation to ubiquinone, generating a sodium motive force in marine and pathogenic bacteria like Vibrio cholerae. This mechanism supports oxidative phosphorylation in Na⁺-rich environments, coupling respiration to active Na⁺ extrusion for osmoregulation and motility. Similarly, formate dehydrogenase serves as an electron donor in anaerobic or facultative prokaryotes, oxidizing formate to CO₂ and transferring electrons to menaquinone or the quinone pool, as seen in E. coli during mixed-acid fermentation or nitrate respiration. These donors enable the use of abundant environmental substrates, enhancing metabolic flexibility. In denitrifying prokaryotes, the branches to use as a terminal acceptor under anoxic conditions, with membrane-bound (NarGHI) reducing NO₃⁻ to NO₂⁻ using quinol as the , thereby conserving energy via proton translocation. This pathway, integral to the , replaces O₂ reduction and sustains oxidative phosphorylation in soil and aquatic denitrifiers like Paracoccus denitrificans, which possesses a near-complete set of respiratory complexes analogous to mitochondria, serving as a key model for studying prokaryotic-mitochondrial . Further reductions to N₂O and N₂ involve additional reductases, generating proton motive comparable to aerobic . Photosynthetic prokaryotes, such as purple nonsulfur bacteria like Rhodobacter sphaeroides, incorporate reverse electron flow in their chromatophores—intracytoplasmic membrane vesicles housing the —to generate reducing equivalents under respiratory or semi-aerobic conditions. Driven by the proton motive force, electrons flow uphill from to NAD⁺ via (Complex I), supporting when light-driven cyclic is insufficient. This bidirectional capability allows seamless transitions between photosynthetic and respiratory modes, optimizing ATP production in fluctuating environments. Methanogenic represent an extreme adaptation, utilizing CO₂ as the terminal in , where hydrogenotrophic species like Methanosarcina acetivorans employ an involving methanophenazine to transfer electrons from H₂ oxidation to heterodisulfide reduction, coupled to proton or sodium translocation for ATP synthesis via . Cytochrome-containing methanogens enhance this with H⁺-translocating complexes, achieving energy yields akin to oxidative phosphorylation despite the , CO₂-reducing nature of the process. These diverse configurations underscore the prokaryotic ETC's role in global biogeochemical cycles.

Side Reactions and Inhibitors

Reactive Oxygen Species Production

During oxidative phosphorylation, (ROS) are generated as byproducts when electrons leak from the () and react with molecular oxygen, primarily forming anion (O₂⁻•). This leakage occurs at specific sites within the respiratory complexes, accounting for approximately 1-2% of electrons transferred through the chain under physiological conditions. The primary production sites are Complex I at the (FMN) site and Complex III at the Qo site, where semiquinone intermediates facilitate the one-electron reduction of O₂ to . Complex I produces ROS during both forward and reverse electron flow, while Complex III generation is prominent during forward transport from oxidation. Superoxide is the initial ROS formed, which undergoes dismutation—either spontaneously or enzymatically by (SOD)—to yield (H₂O₂). can further react with transition metals via the Fenton reaction to produce the highly reactive (•OH), exacerbating potential cellular damage. Mitochondrial (MnSOD) is the key enzyme catalyzing this dismutation in the matrix, converting O₂⁻• to H₂O₂ at bimolecular rate constants up to approximately 10^9 M⁻¹ s⁻¹. The structural features of Complexes I and III, including the FMN cofactor in Complex I and the Qo semiquinone pocket in Complex III, enable this electron diversion due to their proximity to ubiquinone binding sites. Several factors enhance ROS production during OXPHOS. Reverse electron transport (RET) at Complex I, driven by a highly reduced ubiquinone pool and high proton motive force, dramatically increases generation, often by 10-fold or more compared to forward flow. Elevated ΔpH across the further promotes RET-mediated ROS by influencing electron backflow from to NAD⁺. High (Δψ_m) also amplifies leakage, particularly under states of metabolic overload or substrate excess. These conditions highlight how OXPHOS efficiency inversely correlates with ROS output in dynamic cellular environments. Mitochondrial ROS exert dual roles in cellular physiology, causing oxidative damage while also serving as signaling molecules. Excess ROS oxidize proteins, lipids, and DNA, leading to mitochondrial dysfunction, mutagenesis, and apoptosis through thiol modifications and carbonyl formations. For instance, protein carbonylation disrupts ETC components, perpetuating a vicious cycle of ROS production. Conversely, controlled ROS levels stabilize hypoxia-inducible factor-1α (HIF-1α) by inhibiting prolyl hydroxylases, thereby activating adaptive gene expression for metabolism and angiogenesis. This signaling is evident in non-hypoxic contexts where mitochondrial ROS modulate redox-sensitive pathways. Cells mitigate ROS via mitochondrial antioxidants, primarily (GPx) and peroxiredoxins (Prx). , a , reduces H₂O₂ and lipid hydroperoxides using (GSH), preventing membrane peroxidation and maintaining thiol . Peroxiredoxins, such as Prx3 and Prx5, efficiently scavenge H₂O₂ with second-order rate constants around 10^7 M⁻¹ s⁻¹ via thioredoxin-dependent cycles, with Prx3 localized to the matrix and Prx5 to intermembrane spaces. These enzymes form a synergistic network with , ensuring rapid ROS detoxification and balance during OXPHOS.

Inhibitors and Uncouplers

Inhibitors of oxidative phosphorylation (OXPHOS) target specific components of the (ETC) or , blocking or proton translocation and thereby halting ATP production. These compounds are valuable tools for studying mitochondrial function and have applications in and . Uncouplers, in contrast, dissipate the proton motive force (PMF) across the without inhibiting , leading to generation instead of ATP . Complex I (NADH:ubiquinone ) inhibitors such as and piericidin A bind to the quinone-binding site (Q-site), preventing ubiquinone and from NADH. , derived from plant extracts, competitively inhibits the Q-site with high affinity, disrupting the proton-pumping activity of complex I. Piericidin A, a microbial antibiotic, similarly occupies the Q-site or an overlapping pocket, inhibiting proton translocation and serving as a key probe in mechanistic studies of complex I. These inhibitors have been instrumental in elucidating the and of complex I through cryo- analyses. For complex III (cytochrome bc1 complex), targets the Qi on the distal side of the ubiquinone reduction , blocking from to ubiquinone and halting the Q-cycle. Stigmatellin, a fungal , binds to the Qo on the proximal side, preventing oxidation and semiquinone formation, which inhibits proton translocation across the complex. Both compounds have been used in crystallographic studies to map pockets within the Rieske iron-sulfur protein and domains. Complex IV () is inhibited by and , which bind irreversibly to the binuclear (containing a3 and CuB), preventing oxygen reduction to and electron flow from . (CO) also targets this binuclear , acting as a competitive inhibitor with respect to oxygen and disrupting the enzyme's . These inhibitors have been employed in biochemical assays to quantify complex IV activity and investigate hypoxic signaling pathways. ATP synthase inhibitors like oligomycin and venturicidin act on the F0 subunit, blocking proton flow through the c-ring channel and preventing rotational catalysis for ATP synthesis. , a produced by , binds within the F0 domain to inhibit both ATP synthesis and hydrolysis in mitochondria and . Venturicidin, another Streptomyces-derived macrolide, similarly occludes the proton pathway in F0, showing potent activity against fungal and bacterial ATP synthases. Structural studies have revealed shared binding residues for these inhibitors, highlighting conserved mechanisms across species. Uncouplers such as carbonyl cyanide p-trifluoromethoxyphenyl (FCCP) and (DNP) function as protonophores, shuttling protons across the inner membrane to collapse the PMF and uncouple electron transport from ATP production, resulting in increased oxygen consumption and heat dissipation. FCCP, a synthetic , rapidly dissipates the proton gradient at low micromolar concentrations, making it a standard tool for assessing mitochondrial . DNP, historically used as a weight-loss agent, induces by promoting futile proton cycling, though its toxicity limits clinical use. Therapeutically, OXPHOS inhibitors find applications as antibiotics and pesticides due to their disruption of energy metabolism in pathogens and pests. and venturicidin exhibit antifungal and antibacterial properties by targeting microbial ATP synthases selectively. serves as an and , inhibiting complex I in and with minimal mammalian at low doses. These compounds underscore the potential of OXPHOS targeting in and agricultural strategies, informed by decades of biochemical research.

Physiological Adaptations

Responses to Hypoxia in Endotherms

Endotherms, particularly mammals, exhibit low tolerance to due to their high metabolic rates and dependence on continuous oxidative phosphorylation for ATP production. Severe leads to rapid cellular damage, primarily through calcium (Ca²⁺) overload in neurons and subsequent excitotoxic , which can cause irreversible within minutes. Upon reoxygenation following , a burst of (ROS) production exacerbates damage by triggering and mitochondrial dysfunction, further compromising integrity. Certain endotherms have evolved tolerance mechanisms to mitigate hypoxic stress, notably hibernating mammals like the 13-lined ground squirrel (Spermophilus tridecemlineatus). These animals upregulate defenses during to counteract potential ROS accumulation and protect against oxidative damage during from . Hypoxia-inducible factor-1 (HIF-1) plays a key role in downregulating the (ETC) activity, reducing oxygen consumption and preserving ATP levels during low-oxygen periods. At the molecular level, HIFs act as switches to adapt oxidative phosphorylation to by inhibiting the assembly of Complex IV () in the , thereby limiting electron flow and ROS generation under oxygen scarcity. This regulation enhances survival by shifting metabolism toward and conserving endogenous oxygen stores. Post-hypoxic reoxygenation triggers a ROS surge that can overwhelm cellular defenses, but activation of (AMPK) mitigates this by promoting and antioxidant responses, restoring function without excessive damage. Diving mammals, such as and whales, exemplify adaptive strategies through elevated concentrations in , which facilitates oxygen storage and diffusion to mitochondria during prolonged submergence. This enhances oxidative phosphorylation efficiency upon surfacing, allowing sustained aerobic metabolism despite intermittent , while also supporting higher mitochondrial densities for rapid ATP synthesis.

Responses to Hypoxia in Ectotherms

Ectotherms, or animals, exhibit diverse responses to that prioritize due to the oxygen-dependent nature of oxidative phosphorylation (OXPHOS), which becomes severely limited under low oxygen conditions. Unlike endotherms, which maintain high metabolic rates and face greater vulnerability to hypoxic stress, ectotherms often employ metabolic depression—a coordinated downregulation of ATP production and demand—to extend survival without oxygen. This strategy reduces reliance on OXPHOS by shifting to while suppressing ion transport and other energy-intensive processes, allowing species like amphibians and to endure prolonged . In less tolerant ectotherms, such as certain fish species, initial responses to include partial channel arrest, where voltage-gated channels in neuronal and muscle cells are downregulated to conserve ATP otherwise used for maintaining membrane potentials. However, this mechanism is insufficient for extended exposure, leading to imbalances, including disruptions in Na⁺ and K⁺ gradients, after several hours of , which can precipitate cellular damage and death. For instance, in exposed to , recovery phases show altered muscle and blood levels, highlighting the limits of this in intolerant species. Highly tolerant ectotherms, such as the wood frog Rana sylvatica, demonstrate profound metabolic depression during , reducing whole-body ATP demand by approximately 70% through widespread channel arrest and suppression of protein synthesis, cytoskeletal dynamics, and progression. This allows the frog to survive months of oxygen deprivation by matching limited glycolytic ATP supply to minimized demand, preserving endogenous fuels like for cytoprotection. Channel arrest specifically targets ion channels like Na⁺/K⁺-ATPase and Ca²⁺ pumps, preventing excitotoxic ion fluxes that would otherwise accelerate ATP depletion. Upon reoxygenation, anoxia-tolerant ectotherms like freshwater turtles (Trachemys scripta) exhibit minimal (ROS) production from mitochondria, avoiding oxidative damage that could impair OXPHOS recovery. This protection stems from upregulated (SOD) activity, which scavenges superoxide radicals generated at complex I, combined with enhanced purine nucleoside salvage pathways that recycle nucleotides to maintain ATP pools without excess purine catabolism. These adaptations ensure rapid restoration of OXPHOS efficiency post-anoxia. Key mechanisms underlying metabolic suppression in ectotherms include opioid signaling, where endogenous opioids like enkephalins bind receptors to inhibit release and downregulate energy expenditure in neural tissues during . In anoxia-tolerant species, this signaling pathway coordinates global metabolic arrest, complementing channel arrest to achieve hypometabolism. Additionally, controlled ROS levels serve as signals for revival, where low ROS during anoxia prevents damage, and a modest post-anoxic ROS burst may trigger responses and metabolic reactivation without overwhelming cellular defenses. A representative example is the (Carassius auratus), which tolerates chronic by suppressing metabolic rate up to 74% and favoring lipid oxidation over carbohydrates in mitochondria, thereby reducing OXPHOS flux while maintaining tissue function. Although glycolytic flux increases modestly, export to the environment helps prevent intracellular , allowing sustained survival without severe ion or pH disruptions.

Historical Development

Discovery of the Electron Transport Chain

In the early 1920s, Otto pioneered the study of by developing manometric techniques to quantify oxygen consumption in tissue slices, revealing that involves iron-containing pigments sensitive to inhibition. identified these pigments as key components of the respiratory process, particularly noting a CO-sensitive protein—later recognized as cytochrome a3 ( oxidase)—that directly interacts with oxygen to facilitate its reduction. His experiments demonstrated that inhibitors like CO block oxygen uptake at a terminal step, suggesting a sequential mechanism in aerobic . Building on Warburg's findings, David Keilin in 1925 employed low-temperature spectroscopy to observe distinct absorption bands in living cells, leading to the discovery of three (a, b, and c) as ubiquitous respiratory pigments in , , and . Using horse heart muscle and preparations, Keilin showed that these cytochromes exhibit characteristic spectra at temperatures, with displaying a prominent band at 550 nm when reduced. His work established cytochromes as electron carriers, oscillating between oxidized and reduced states during , and common across diverse organisms. Throughout , Keilin and Edward Hartree conducted key experiments on the succinoxidase system—responsible for succinate oxidation to fumarate—using inhibitors to delineate the transport sequence. By adding succinate to muscle preparations and applying or to block oxidase, they observed accumulation of reduced cytochromes upstream, confirming the pathway: a → oxygen. These spectroscopic studies with inhibitors like antimycin (affecting the b-c ) further revealed branch points and the linear flow of s from flavoproteins to proteins. In the 1940s and early 1950s, Britton Chance and G.R. Williams advanced analysis by inventing rapid dual-wavelength spectrophotometers to monitor cytochrome reduction states in real-time during steady-state . Their experiments on isolated mitochondria showed that addition (e.g., succinate) progressively reduces chain components in sequence, with oxidizing before c under aerobic conditions, supporting Keilin's proposed order. This quantitative approach quantified crossover points where oxidation rates balanced reduction, providing evidence for a branched yet ordered . A pivotal advance came in 1950 when E.C. Slater identified a soluble factor—now known as the component of Complex I—essential for from NADH to in the succinoxidase system. Through of beef heart extracts and assays measuring NADH oxidation rates, Slater demonstrated that this "Slater factor" was inhibited by fatty acids and distinct from , bridging the gap between NAD-linked substrates and the chain. His work clarified the for electrons from the tricarboxylic acid cycle into the transport sequence. By the 1960s, Youssef Hatefi achieved the isolation and purification of the four main respiratory complexes (I–IV) from bovine heart mitochondria, enabling reconstitution of the full electron transport chain. In 1962, Hatefi's group solubilized Complex I (NADH:coenzyme Q reductase) using detergents and purified it to homogeneity, confirming its role in transferring electrons from NADH to ubiquinone. According to the emerging chemiosmotic theory, it also pumps protons across the membrane. Subsequent purifications of Complexes II (succinate dehydrogenase), III (ubiquinol:cytochrome c reductase), and IV (cytochrome c oxidase) in the mid-1960s allowed demonstration of sequential electron flow upon reassembly, solidifying the modular architecture of the chain. These milestones shifted focus toward understanding energy coupling in the process.

Formulation of the Chemiosmotic Theory

In 1961, Peter Mitchell proposed the chemiosmotic hypothesis in a seminal paper, suggesting that the in the actively pumps protons into the , establishing a delocalized electrochemical proton gradient across the membrane rather than relying on high-energy chemical intermediates to couple oxidation to ATP synthesis. This delocalized proton motive force, comprising both a difference (ΔpH) and a (Δψ), was envisioned to drive ATP production via a reversible proton-translocating embedded in the same membrane. Mitchell's formulation unified the mechanisms of oxidative phosphorylation in mitochondria and photosynthetic phosphorylation in chloroplasts, challenging the prevailing views by emphasizing membrane topology and ion gradients over soluble chemical carriers. The chemiosmotic hypothesis encountered strong opposition from leading researchers in . Albert Lehninger, a prominent mitochondrial , favored a localized variant of , positing that energized protons remained confined to specific intramembrane domains or microenvironments near the respiratory complexes, rather than equilibrating in the bulk aqueous phases as Mitchell proposed. Similarly, Efraim Racker championed the chemical coupling hypothesis, which invoked transient high-energy phosphorylated intermediates (such as XP) formed during electron transport to directly transfer energy to ATP synthesis, drawing analogies to known substrate-level phosphorylations. These alternatives were supported by reconstitution experiments with isolated complexes, but they struggled to explain the obligatory role of intact membranes and the effects of ionophores. Key validations emerged in the 1960s, bolstering ideas. Experiments with inverted submitochondrial particles, prepared by of mitochondria, revealed ATP hydrolysis driving proton extrusion into the external medium, demonstrating the reversibility of the proposed proton-translocating and confirming vectorial proton movement coupled to . Concurrently, Jagendorf's landmark chloroplast studies showed that imposing an artificial pH gradient—by soaking thylakoids in an acidic medium followed by transfer to a basic containing ADP and Pi—induced ATP synthesis in the absence of , directly implicating a proton gradient as sufficient for energizing . By the 1970s, accumulating evidence from studies solidified the theory's acceptance; compounds like dissipated the proton gradient, uncoupling electron transport from ATP synthesis by stimulating while abolishing , precisely as predicted. Pioneers such as Fritz Lipmann, initially skeptical, acknowledged the paradigm shift through these and reconstitution experiments integrating proton pumps with . Mitchell received the 1978 Nobel Prize in Chemistry for the chemiosmotic theory, marking its transition from hypothesis to foundational principle and revolutionizing understanding of energy transduction in biology, supplanting earlier chemical intermediate models with a unified membrane-based .

References

  1. [1]
    Biochemistry, Oxidative Phosphorylation - StatPearls - NCBI Bookshelf
    Oxidative phosphorylation is a cellular process that harnesses the reduction of oxygen to generate high-energy phosphate bonds in the form of adenosine ...
  2. [2]
    The Mechanism of Oxidative Phosphorylation - The Cell - NCBI - NIH
    Most of the usable energy obtained from the breakdown of carbohydrates or fats is derived by oxidative phosphorylation, which takes place within mitochondria ...
  3. [3]
    Oxidative phosphorylation: regulation and role in cellular and tissue ...
    Dec 1, 2017 · Oxidative phosphorylation provides most of the ATP that higher animals and plants use to support life and is responsible for setting and maintaining metabolic ...
  4. [4]
    [PDF] Peter Mitchell - Nobel Lecture
    The fact that what began as the chemiosmotic hypothesis has now been acclaimed as the chemiosmotic theory - at the physiological level, even if not at the bio-.
  5. [5]
    Peter Mitchell – Facts - NobelPrize.org
    Peter Mitchell presented his theory in 1961. It states that the basis for the process is a flow of electrons and hydrogen ions through membranes in the ...Missing: hypothesis | Show results with:hypothesis
  6. [6]
    Inorganic Cation Transport and Energy Transduction in ...
    The relationship of these parameters is described by Δp = ΔΨ − ZΔpH, where ΔpH is the difference between the pH of the bulk medium and that of the cytosol and ...
  7. [7]
    Electron-Transport Chains and Their Proton Pumps - NCBI - NIH
    Some respiratory enzyme complexes pump one H per electron across the inner mitochondrial membrane, whereas others pump two. The detailed mechanism by which ...
  8. [8]
    The renaissance of mitochondrial pH | Journal of General Physiology
    May 28, 2012 · These measurements established that Δp ranges from 180 to 220 mV depending on the metabolic state of the mitochondria, with ΔΨm ranging from 150 ...Missing: typical | Show results with:typical<|control11|><|separator|>
  9. [9]
    Control Over the Contribution of the Mitochondrial Membrane ... - NIH
    The usual measured value of Δp is around 170–200 mV (2–9), although in older studies slightly higher values were encountered (10). This value may change ...
  10. [10]
    Mitochondrial membrane potential probes and the proton gradient
    This review will help illustrate both the strengths and potential pitfalls of common mitochondrial membrane potential dyes, and highlight best-usage approaches.
  11. [11]
    Guidelines on experimental methods to assess mitochondrial ...
    Dec 11, 2017 · The precise TMRM concentrations for quench or non-quench mode can be determined using FCCP—in quench mode, addition of ~10 μM FCCP (to collapse ...
  12. [12]
    Calibration and measurement of mitochondrial pH in intact adult rat ...
    May 13, 2021 · SNARF-1 AM is a pH-sensitive fluorophore for measuring mitochondrial pH; Mitochondrial pH measurements in live cells with confocal microscopy ...Materials And Equipment · Solution Preparation · Step-By-Step Method Details
  13. [13]
    The Regulation and Physiology of Mitochondrial Proton Leak
    Jun 1, 2011 · Protons are vectorially pumped against their electrochemical gradient into the mitochondrial intermembrane space, creating a protonmotive force ...
  14. [14]
    Mitochondrial Ion Channels - PMC - PubMed Central - NIH
    The energy dissipated by ion flux has a depolarizing influence on mitochondrial membrane potential (ΔΨm), which stimulates NADH oxidation, proton pumping, and ...
  15. [15]
    Transport Pathways—Proton Motive Force Interrelationship in ...
    In energized mitochondria, PIMAC functioning should be inhibited by ATP and high ΔΨ values; under these conditions ΔΨ allows for functioning of the AAC and ΔpH ...
  16. [16]
    Mitochondrial Uncoupling and Reactive Oxygen Species - PubMed
    Apr 4, 2018 · In a process called uncoupling, proton leak into the mitochondrial matrix independent of ATP production dissipates the pmf and energy is lost ...Missing: collapse | Show results with:collapse
  17. [17]
    Mitochondrial Uncoupling and Reactive Oxygen Species
    In a process called uncoupling, proton leak into the mitochondrial matrix independent of ATP production dissipates the pmf and energy is lost as heat.Missing: collapse | Show results with:collapse
  18. [18]
    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 ...
  19. [19]
    Regulation of Respiration and Apoptosis by Cytochrome c ... - Nature
    Nov 1, 2019 · Cytochrome c (Cytc) is a multifunctional protein, acting as an electron carrier in the electron transport chain (ETC), where it shuttles ...
  20. [20]
    Biochemistry of Mitochondrial Coenzyme Q Biosynthesis - PMC
    Coenzyme Q (CoQ, ubiquinone) is a redox active lipid produced across all domains of life that functions in electron transport and oxidative phosphorylation.
  21. [21]
    The Ubiquinone-Ubiquinol Redox Cycle and Its Clinical ...
    Jun 20, 2024 · The ubiquinone–ubiquinol redox cycle comprises two steps. In step 1, ubiquinone is reduced to ubiquinol in association with Complexes I and II of the ETC.
  22. [22]
    Cytochrome c phosphorylation: Control of mitochondrial electron ...
    Feb 2, 2020 · Cytc is an electron carrier in the mitochondrial electron transport chain (ETC) and thus central for aerobic energy production. Under conditions ...
  23. [23]
    Electron transport chains as a window into the earliest stages ... - NIH
    Aug 14, 2023 · The organic compound, ubiquinone, and the small protein, Cytochrome C, carry electrons between complexes. Ubiquinone itself contributes to ...
  24. [24]
    Review Mobility and function of Coenzyme Q (ubiquinone) in the ...
    The quinone behaves kinetically as a homogeneous pool freely diffusing in the lipid bilayer, thus setting the basis for the widely accepted random diffusion ...
  25. [25]
    EVALUATING CYTOCHROME C DIFFUSION IN THE ... - NIH
    The rate constant for cytochrome c permeation from the intracristal / intermembrane spaces to the external medium is a (min−1).
  26. [26]
    Electron Transport - an overview | ScienceDirect Topics
    In terms of quinone carriers, many bacteria use ubiquinone as used by eukaryotic mitochondria. Others employ menaquinone (vitamin K2). Some archaea, e.g. ...
  27. [27]
    Mitochondrial Respiratory Chain Supercomplexes: From Structure to ...
    Here, we review the structure, assembly, and functions of SCs, as well as the relationship between mitochondrial SCs and diseases.
  28. [28]
    High-resolution in situ structures of mammalian respiratory ... - Nature
    May 29, 2024 · We identify four main supercomplex organizations: I1III2IV1, I1III2IV2, I2III2IV2 and I2III4IV2, which potentially expand into higher-order ...
  29. [29]
    The functional significance of mitochondrial respiratory chain ...
    Oct 12, 2023 · In this review, we discuss the current knowledge on the functional significance of MRC supercomplexes, highlight experimental limitations, and suggest ...
  30. [30]
    Cardiolipin-Dependent Formation of Mitochondrial Respiratory ...
    Nov 9, 2013 · The mitochondrial signature phospholipid cardiolipin (CL) plays a central role in formation and stability of respiratory SCs from yeast to man.
  31. [31]
    Role of Cardiolipin in Mitochondrial Signaling Pathways - Frontiers
    CL deficiency causes an increase in mitochondrial ROS. CL is required for the structural assembly of the mitochondrial respiratory chain. CL deficiency causes a ...
  32. [32]
    Biochemistry, Electron Transport Chain - StatPearls - NCBI Bookshelf
    Sep 4, 2023 · ... intermembrane). ATP synthase, also called complex V, uses the ETC generated proton gradient across the inner mitochondrial membrane to form ATP.
  33. [33]
    Mitochondrial Respiratory Complex I: Structure, Function and ...
    Respiratory complex III is required to maintain complex I in mammalian mitochondria. Mol Cell. 2004;13:805–15. doi: 10.1016/s1097-2765(04)00124-8. [DOI] ...
  34. [34]
    Crystal structure of the entire respiratory complex I - PubMed Central
    It contains all the redox centres of the enzyme – non-covalently bound flavin mononucleotide (FMN) and nine iron-sulphur (Fe-S) clusters. NADH transfers two ...
  35. [35]
    Structure of respiratory complex I – An emerging blueprint for the ...
    Mar 19, 2022 · There are eight or nine FeS clusters, depending on the species, but only seven of them lie on the main pathway connecting the NADH and quinone.
  36. [36]
    Mössbauer Spectroscopy on Respiratory Complex I: The Iron–Sulfur ...
    Nov 28, 2011 · In mitochondria, complex I (NADH:quinone oxidoreductase) couples electron transfer to proton translocation across an energy-transducing ...
  37. [37]
    Architecture of complex I and its implications for electron transfer ...
    ... FMN and ubiquinone is formed by seven iron-sulfur clusters. Following N3 ... subunits and transmembrane segments form the proton pumping elements of complex I.
  38. [38]
    Electron tunneling in respiratory complex I - PMC - PubMed Central
    Distinct electron tunneling pathways between neighboring Fe/S clusters are identified; the pathways primarily consist of two cysteine ligands and one additional ...Fig. 1 · Results And Discussion · Electron Tunneling Pathways
  39. [39]
    Redox-Coupled Protonation of Respiratory Complex I - NIH
    FMN initially accepts two electrons from NADH as hydride; electrons are then transferred one at a time to a chain of seven iron-sulfur clusters spanning the ...
  40. [40]
    Mammalian Complex I Pumps 4 Protons per 2 Electrons at High and ...
    Previous studies have concluded that complex I pumps 4 protons per 2 electrons transferred (4H+/2e−) (7) but the observation that there are only 3 proton ...
  41. [41]
    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.
  42. [42]
    Functional Water Wires Catalyze Long-Range Proton Pumping in ...
    Dec 16, 2020 · Our findings suggest that the mammalian complex I pumps protons by tightly linked conformational and electrostatic coupling principles.<|control11|><|separator|>
  43. [43]
    Structure of mammalian respiratory complex I - PMC - PubMed Central
    Crystal structure of the entire respiratory complex I. Nature. 2013 ... bovine mitochondrial complex I. Biochim Biophys Acta. 2014;1837:929–939. doi ...
  44. [44]
    Leigh syndrome associated with mitochondrial complex I deficiency ...
    Mutations in the nuclear-encoded subunits of complex I of the mitochondrial respiratory chain are a recognized cause of Leigh syndrome (LS).
  45. [45]
    The p.M292T NDUFS2 mutation causes complex I-deficient Leigh ...
    Our results confirm that NDUFS2 is a mutational hotspot in Caucasian children with isolated complex I deficiency and recommend the routine diagnostic ...
  46. [46]
    Complex I deficiency and Leigh syndrome through the eyes of ... - NIH
    Mitochondrial complex I deficiency is associated with a wide range of clinical presentations, including Leigh syndrome. Its genetic causes are heterogeneous, ...
  47. [47]
    Complex II Biology in Aging, Health, and Disease - PubMed Central
    Jul 24, 2023 · Mitochondrial Complex II converts succinate to fumarate and plays an essential role in both the tricarboxylic acid (TCA) cycle and the electron transport chain ...
  48. [48]
    Structure of the human respiratory complex II - PNAS
    Apr 25, 2023 · Five prosthetic groups, FAD, [2Fe-2S], [4Fe-4S], [3Fe-4S], and a heme b are required for electron transfer flow from succinate to ubiquinone.
  49. [49]
    “Catalytic mechanisms of Complex II enzymes: A structural ...
    Electrons are transferred between these active sites via the three Fe:S clusters coordinated to the SdhB or FrdB subunit (Fig. 1). Fig. 3. The dicarboxylate- ...
  50. [50]
    Complex II subunit SDHD is critical for cell growth and metabolism ...
    Jun 12, 2021 · Complex II carries four protein subunits [2], all of which are encoded by nuclear genes. ... Due to the role of Complex II in the ETC and Krebs ...
  51. [51]
    Catalytic mechanisms of complex II enzymes: A structural perspective
    Complex II enzymes couple two distinct chemical reactions: the reversible oxidoreduction of succinate and fumarate, catalyzed in a soluble domain,
  52. [52]
    Crystal Structure of Mitochondrial Respiratory Membrane Protein ...
    The first crystal structure of porcine mitochondrial respiratory Complex II (succinate:ubiquinone oxidoreductase) has been determined at 2.4 Å resolution as ...The Hydrophobic Membrane... · Prosthetic Groups · Ubiquinone And Its Binding...<|separator|>
  53. [53]
    SDH-related Pheochromocytoma and paraganglioma - PMC - NIH
    In contrast, mutation in the catalytic subunit SDHA causes necrotizing encephalopathy, which is caused by other mutations in genes involved in energy metabolism ...
  54. [54]
    Crystal Structure of the Cytochrome bc1 Complex from Bovine Heart Mitochondria
    ### Summary of Complex III (Cytochrome bc1 Complex) Structure
  55. [55]
    The Q Cycle of Cytochrome bc Complexes: a Structure Perspective
    The mechanism then requires only one oxidation (turnover) of QpH2 to provide the single electron needed to form the quinol QnH2 and once primed, minimizes ...
  56. [56]
    Structure of the intact 14-subunit human cytochrome c oxidase
    Jul 20, 2018 · Here we obtained the 3.3 Å resolution structure of complex-IV ... heme molecules, 2 copper centers, and 1 magnesium ion (Fig. 1b) ...
  57. [57]
    Structures of Metal Sites of Oxidized Bovine Heart Cytochrome c ...
    Electron density distribution of the oxidized bovine cytochrome c oxidase at 2.8 Å resolution indicates a dinuclear copper center with an unexpected structure ...
  58. [58]
    Structural basis for functional properties of cytochrome c oxidase - NIH
    Mammalian CcO is a large integral membrane protein comprised of 13 subunits. It contains four redox active centers, CuA, heme a, and a heme a3/CuB binuclear ...
  59. [59]
    Oxygen Activation and Energy Conservation by Cytochrome c Oxidase
    Jan 19, 2018 · The key components of the BNC are two magnetically coupled redox-active metal centers, a high-spin heme (a3) and a copper ion (CuB). The high- ...
  60. [60]
    The electron distribution in the “activated” state of cytochrome c ...
    May 14, 2018 · The reaction is linked to translocation of four protons across the membrane for each O2 reduced to water. The free energy associated with ...
  61. [61]
    Cytochrome c oxidase: Intermediates of the catalytic cycle and their ...
    Mar 9, 2012 · A “catalytic cycle” of cytochrome oxidase involving complete reduction of the enzyme by 4 electrons followed by oxidation by O 2 is a chimera composed ...
  62. [62]
    Interconversions of P and F intermediates of cytochrome c oxidase ...
    Elucidation of the intermediate structures in the catalytic cycle is crucial for understanding both the mechanism of oxygen reduction and its coupling to proton ...Missing: APF | Show results with:APF
  63. [63]
    The Allosteric ATP-inhibition of Cytochrome C Oxidase Activity Is ...
    Jan 21, 2000 · It is suggested that after cAMP-dependent phosphorylation of cytochrome c oxidase mitochondrial respiration is controlled by the ATP/ADP-ratio ...
  64. [64]
    Alternative Oxidase: A Mitochondrial Respiratory Pathway to ...
    Alternative oxidase (AOX) is a non-energy conserving terminal oxidase in the plant mitochondrial electron transport chain.
  65. [65]
    The Active Site of the Cyanide-Resistant Oxidase From Plant ...
    Mar 27, 1995 · Using the known three-dimensional structures of other binuclear iron proteins, we have developed a structural model for the proposed catalytic ...
  66. [66]
    Alternative Oxidase: From Molecule and Function to Future Inhibitors
    Mar 4, 2024 · This enzyme has not been found only in Archaea, mammals, some yeasts and protists. ... We found that plants lacking AOX have increased concns.
  67. [67]
    Functional Coexpression of the Mitochondrial Alternative Oxidase ...
    Thermogenic tissues in the skunk cabbage spadices were identified using a high-resolution infrared thermal camera (Fig. 2). The thermogenic spadix is surrounded ...
  68. [68]
    Structure of electron transfer flavoprotein-ubiquinone ...
    ETF-QO is a 4Fe4S flavoprotein located in the inner mitochondrial membrane. It catalyzes ubiquinone (UQ) reduction by ETF, linking oxidation of fatty acids and ...Missing: review | Show results with:review
  69. [69]
    The electron transfer flavoprotein: Ubiquinone oxidoreductases
    ETF-QO is a component of the mitochondrial respiratory chain that together with electron transfer flavoprotein (ETF) forms a short pathway that transfers ...
  70. [70]
  71. [71]
    Localization and functional characterization of the alternative ...
    Mar 24, 2022 · Among eukaryotes, alternative oxidases have dispersed distribution and are found in plants, fungi, and protists, including Naegleria ssp.
  72. [72]
    Structure at 2.8 Â resolution of F1-ATPase from bovine ... - Nature
    Aug 25, 1994 · In the crystal structure of bovine mitochondrial F 1 -ATPase determined at 2.8 Å resolution, the three catalytic β-subunits differ in conformation and in the ...
  73. [73]
    The c-ring stoichiometry of ATP synthase is adapted to cell ...
    The c-ring stoichiometry determines the number of ions transferred during enzyme operation and has a direct impact on the ion-to-ATP ratio.
  74. [74]
    Unusual features of the c-ring of F1FO ATP synthases - Nature
    Dec 6, 2019 · We present the first high-resolution structure (2.3 Å) of the in meso crystallized c-ring of 14 subunits from spinach chloroplasts.
  75. [75]
    Arrangement of subunits in intact mammalian mitochondrial ATP ...
    Jul 2, 2012 · Here, we present the structure of intact bovine mitochondrial ATP synthase at ∼18 Å resolution by electron cryomicroscopy of single particles in ...Sign Up For Pnas Alerts · Results · Membrane-Bound F Region<|control11|><|separator|>
  76. [76]
    The structure of the membrane extrinsic region of bovine ATP synthase
    Dec 22, 2009 · The structure of the complex between bovine mitochondrial F 1 -ATPase and a stator subcomplex has been determined at a resolution of 3.2 Å.
  77. [77]
    The binding change mechanism for ATP synthase - PubMed - NIH
    The binding change mechanism for ATP synthase--some probabilities and possibilities. ... Author. P D Boyer. Affiliation. 1 Department of Chemistry and ...
  78. [78]
    Review The rotary binding change mechanism of ATP synthases
    The F0F1 ATP synthase functions as a rotary motor where subunit rotation driven by a current of protons flowing through F0 drives the binding changes in F1 ...
  79. [79]
    [PDF] Paul D. Boyer - Nobel Lecture
    These unusual features are energy-linked binding changes that include release of a tightly bound ATP, sequential conformational changes of three catalytic sites ...<|separator|>
  80. [80]
    ATP Synthesis and the Binding Change Mechanism
    Jun 9, 2006 · Thus according to Boyer's binding change mechanism for ATP synthesis, the three catalytic sites on the enzyme bind ADP and phosphate in ...
  81. [81]
    Oligomycin frames a common drug-binding site in the ATP synthase
    Aug 6, 2012 · We propose that oligomycin inhibits proton translocation in the F1Fo ATP synthase by locking the essential carboxyl in a semiclosed conformation ...Missing: F0 | Show results with:F0
  82. [82]
    The Overall Efficiency of Oxidative Phosphorylation – BIOC*2580
    This is ~33% of the theoretical maximum. However, under intracellular conditions, the free energy recovery is more than 60%!
  83. [83]
    Recent advances on the structure and function of NDH-1
    Here, we highlight progress in understanding the function of NDH-1 in the photosynthetic light reactions of both cyanobacteria and chloroplasts.
  84. [84]
  85. [85]
    Structural basis for energy transduction by respiratory alternative ...
    Apr 30, 2018 · Electron transfer in respiratory chains generates the electrochemical potential that serves as energy source for the cell. Prokaryotes can ...
  86. [86]
    The organisation of proton motive and non-proton motive redox ...
    The proton motive force (pmf) can be built up by different mechanisms like proton pumping, quinone/quinol cycling or by a redox loop.
  87. [87]
    Structure of Escherichia coli cytochrome bd-II type oxidase ... - Nature
    Nov 11, 2021 · The Escherichia coli aerobic respiratory chain contains three terminal oxidases, cytochrome bo3 oxidase, cytochrome bd-I oxidase (bd-I) and ...
  88. [88]
    Proton Pumping and Non-Pumping Terminal Respiratory Oxidases
    Oct 7, 2021 · In contrast to bd-type oxidases, HCOs generate the proton motive force not only by the transfer of electrons and protons to the catalytic ...
  89. [89]
    In Escherichia coli Ammonia Inhibits Cytochrome bo3 But Activates ...
    Dec 25, 2020 · Cytochrome bo3 and cytochrome bd-I are terminal oxidases in the aerobic respiratory chain of Escherichia coli [1]. Both enzymes catalyze the ...
  90. [90]
    Short-chain aurachin D derivatives are selective inhibitors of E. coli ...
    Dec 13, 2021 · At lower oxygen concentrations the efficiency of cytochrome bo3 decreases due to its lower oxygen affinity. To enable aerobic growth under ...
  91. [91]
    The sodium pumping NADH:quinone oxidoreductase (Na - PubMed
    Na(+)-NQR is a unique Na(+) pumping respiratory complex found only in prokaryotes, that plays a key role in the metabolism of marine and pathogenic bacteria.
  92. [92]
    p24183 · fdng_ecoli - UniProt
    Formate dehydrogenase allows E.coli to use formate as major electron donor during anaerobic respiration, when nitrate is used as electron acceptor.
  93. [93]
    Prokaryotic Nitrate Reduction: Molecular Properties and Functional ...
    Membrane-bound nitrate reductases are associated with denitrification and anaerobic nitrate respira- tion (Fig. 3). Although the most exhaustive biochemistry ...
  94. [94]
    Paracoccus denitrificans and the evolutionary origin of the ... - PubMed
    Paracoccus denitrificans and the evolutionary origin of the mitochondrion. ... Mitochondria / ultrastructure; Models, Biological*; Oxidative Phosphorylation ...
  95. [95]
    Modeling the electron transport chain of purple non-sulfur bacteria
    ... reverse electron flow under respiratory conditions. So far, the functional role of reverse electron flow has been mainly discussed for photosynthetic growth ...
  96. [96]
    Energy Conservation and Hydrogenase Function in Methanogenic ...
    Sep 18, 2019 · Cytochrome-containing methanogens are able to supplement the ion motive force generated by Mtr with an H+-translocating electron transport ...<|control11|><|separator|>
  97. [97]
    Production of Reactive Oxygen Species by Mitochondria - PubMed
    Sep 19, 2003 · In mitochondria, complex III is the principal site for ROS generation during the oxidation of complex I substrates, and rotenone protects by limiting electron ...
  98. [98]
    Mitochondrial Reactive Oxygen Species and Their Contribution in ...
    Although complex I and complex III are the primary production sites in mitochondria, complex II may produce ROS to a lesser extent. The FAD site of the complex ...
  99. [99]
    An Update on Mitochondrial Reactive Oxygen Species Production
    Complex I is denoted in purple since it uses both isopotential groups to form ROS and can produce ROS from both its flavin mononucleotide group and ubiquinone ...
  100. [100]
    Mitochondrial formation of reactive oxygen species - PMC - NIH
    Dismutation of O2−• (either spontaneously or through a reaction catalysed by superoxide dismutases) produces hydrogen peroxide (H2O2), which in turn may be ...
  101. [101]
    Mitochondrial Superoxide Dismutase - PubMed Central - NIH
    Critical Issues: The primary form of ROS produced by mitochondria is the superoxide radical anion. As a charged radical anion, superoxide is restricted in ...
  102. [102]
    Role of Mitochondrial Reverse Electron Transport in ROS Signaling
    Jun 26, 2017 · RET is produced when electrons from ubiquinol are transferred back to respiratory complex I, reducing NAD+ to NADH. This process generates a ...
  103. [103]
    Membrane potential and delta pH dependency of reverse electron ...
    Aug 17, 2018 · RET is dependent on mitochondrial membrane potential (Δψm) and transmembrane pH difference (ΔpH), components of the proton motive force (pmf); a ...
  104. [104]
    Physiologic Implications of Reactive Oxygen Species Production by ...
    Aug 6, 2019 · The mitochondrial protonmotive force (Δp) is a main driving force for RET. Therefore, decreasing the Δp removes a main driver of RET and ROS ...Missing: percentage | Show results with:percentage
  105. [105]
    Mitochondria in oxidative stress, inflammation and aging - Nature
    Jun 11, 2025 · It results in oxidative stress due to excessive reactive oxygen species (ROS) generation, which contributes to cell damage and death.
  106. [106]
    A targeted antioxidant reveals the importance of mitochondrial ...
    Reactive oxygen species (ROS) generation has been implicated in the stabilization of HIF-1α during this response, but this is still a matter of some debate. In ...
  107. [107]
    Redox signalling and mitochondrial stress responses
    Moderate increases in ROS can activate HIF-1α by oxidation of certain cysteine residues in HIF-1α regulatory proteins, whereas further oxidation of other ...
  108. [108]
    Mitochondrial Glutathione, a Key Survival Antioxidant
    Peroxiredoxin III, a mitochondrion-specific peroxidase, regulates apoptotic signaling by mitochondria. J Biol Chem 279: 41975–41984, 2004. Go to Citation.Mitochondrial Control Of... · Mgsh And Apoptosis · Mgsh In Pathologic Settings
  109. [109]
    Mitochondrial peroxiredoxin involvement in antioxidant defence and ...
    Prxs (peroxiredoxins) are a family of proteins that are extremely effective at scavenging peroxides. The Prxs exhibit a number of intriguing properties.
  110. [110]
    The Synergetic Coupling among the Cellular Antioxidants ... - Nature
    Sep 1, 2015 · In this paper we demonstrate that there is a synergetic coupling between GPxs, Prxs themselves and also with other antioxidants when the GPxs and Prxs are not ...
  111. [111]
    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.
  112. [112]
    Structure of inhibitor-bound mammalian complex I - Nature
    Oct 16, 2020 · Piericidin has been described to compete for the same or overlapping binding sites as the inhibitors rotenone and DQA, and to display partially- ...
  113. [113]
    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.
  114. [114]
    Inhibitor binding changes domain mobility in the iron–sulfur protein ...
    We cocrystallized the bovine bc1 complex with the Qi inhibitor antimycin A and with the Qo inhibitors myxothiazol, MOA-stilbene, UHDBT, and stigmatellin ...
  115. [115]
    The Q o site of the mitochondrial complex III is required for the ...
    Jun 11, 2007 · The Qo site of the mitochondrial complex III is required for the transduction of hypoxic signaling via reactive oxygen species production.Missing: semiquinone | Show results with:semiquinone
  116. [116]
    The effect of inhibitors on the oxygen kinetics of cytochrome c oxidase
    CO is strictly competitive, azide and formate are uncompetitive, and cyanide and sulfide are non-competitive inhibitors towards oxygen.
  117. [117]
    An overview of ATP synthase, inhibitors, and their toxicity - PMC
    Nov 20, 2023 · This review aims to provide a comprehensive overview of mitochondrial ATP synthase, its structural and functional features, and the most common inhibitors and ...
  118. [118]
    Oligomycin frames a common drug-binding site in the ATP synthase
    Aug 6, 2012 · Unlike oligomycin and ossamycin, venturicidin is a potent inhibitor of mitochondrial and bacterial ATP synthase (20–22). The residues identified ...
  119. [119]
    Venturicidin A, A Membrane-active Natural Product Inhibitor of ATP ...
    May 18, 2020 · On the other hand, VentA is specific to ATP synthase and it is expected to have non-covalent interactions like oligomycin and blocks the proton ...
  120. [120]
    Uncouplers of Mitochondrial Oxidative Phosphorylation Are Not ...
    Jan 1, 1998 · Uncouplers of mitochondrial oxidative phosphorylation, dinitrophenol (DNP) and carbonyl cyanide p-trifluoromethoxyphenylhydrazone (FCCP), were found to ...
  121. [121]
    Effects of mitochondrial uncouplers on intracellular calcium, pH and ...
    The inhibition of background K+ current may be related to the uncoupling of oxidative phosphorylation. Mitochondrial uncouplers have long been known to be ...
  122. [122]
    Uncouplers of Mitochondrial Oxidative Phosphorylation Are Not ...
    Uncouplers of mitochondrial oxidative phosphorylation, dinitrophenol (DNP) and carbonyl cyanideptrifluoromethoxyphenylhydrazone (FCCP), were found to stimulate ...
  123. [123]
    Inhibitors of ATP Synthase as New Antibacterial Candidates - PMC
    Mar 24, 2023 · Unfortunately, similarly to oligomycin A, venturicidine A inhibits both the mitochondrial and bacterial ATP synthase [70,71]. Although there are ...
  124. [124]
    Complex I inhibitors as insecticides and acaricides - ScienceDirect
    Rotenone and piericidin A were known for a long time as high-affinity inhibitors of proton-translocating NADH:Q oxidoreductase. Piericidin A was isolated from ...
  125. [125]
    Cytoskeletal Arrest: An Anoxia Tolerance Mechanism - PMC
    Aug 23, 2021 · Hypoxia in anoxia-intolerant mammals is characterized by hypoxia-inducible factor (HIF)-dependent perinuclear localization of mitochondria ...2. Cytoskeletal Dynamics... · 6. Cytosolic Calcium... · 8. Gasotransmitters Involved...
  126. [126]
    hibernation and death display different gene profiles - Hadj‐Moussa
    Feb 15, 2019 · Antioxidant defenses are also typically upregulated during torpor in ground squirrels in anticipation of large increases in reactive oxygen ...
  127. [127]
    Intracellular antioxidant enzymes are not globally upregulated ...
    Aug 6, 2025 · Intracellular antioxidant enzymes are not globally upregulated during hibernation in the major oxidative tissues of the 13-lined ground squirrel ...
  128. [128]
    HIF-1α regulation in mammalian hibernators: role of non ... - PubMed
    HIF-1α mRNA levels correlated with the protein increase in bat skeletal muscle and liver but not in squirrel skeletal muscle. Antisense HIF-1α transcripts were ...
  129. [129]
    Tribute to P. L. Lutz: putting life on `pause' – molecular regulation of ...
    May 15, 2007 · Cloning and expression of hypoxia-inducible factor 1α from the hibernating ground squirrel, Spermophilus tridecemlineatus. Biochim. Biophys ...
  130. [130]
    Environmental and behavioral regulation of HIF-mitochondria crosstalk
    Upon acute hypoxia exposure, HIFs play an important role in increasing oxidative phosphorylation efficiency by modulating complex IV activity. This effect is ...
  131. [131]
    Hypoxia Triggers AMPK Activation through Reactive Oxygen ... - NIH
    These findings reveal that hypoxia can trigger AMPK activation in the apparent absence of increased [AMP] through ROS-dependent CRAC channel activation.Missing: surge endotherms
  132. [132]
    Hypoxic activation of AMPK is dependent on mitochondrial ROS but ...
    Here we show that hypoxia activates AMPK through LKB1 without an increase in the AMP/ATP ratio. Hypoxia increased reactive oxygen species (ROS) levels.Missing: surge endotherms
  133. [133]
    Myoglobin and Mitochondria: A relationship bound by Oxygen and ...
    Myoglobin modulates mitochondrial function by regulating oxygen and nitric oxide levels, and they are intimately linked in their functional regulation.
  134. [134]
    Myoglobin oxygen affinity in aquatic and terrestrial birds and mammals
    Summary: Myoglobin oxygen affinity varies among terrestrial and aquatic birds and mammals, with long-duration diving species having greater myoglobin.
  135. [135]
    Oxygen conserving mitochondrial adaptations in the skeletal ...
    Sep 19, 2018 · The purpose of this study was to examine the capacity for oxidative metabolism in skeletal muscle of BHD compared to matched controls.
  136. [136]
    Hypoxia Tolerance in Reptiles, Amphibians, and Fishes
    As we outline next, ion channel arrest seems the most important first step in reducing ATP demand. Energy conservation: regulation of ionic conductance and ...Missing: imbalance | Show results with:imbalance
  137. [137]
    Anti-apoptotic response during anoxia and recovery in a freeze ...
    Mar 24, 2016 · The common wood frog, Rana sylvatica, utilizes freeze tolerance as a means of winter survival. Concealed beneath a layer of leaf litter and ...
  138. [138]
    Surviving anoxia: the maintenance of energy production and tissue ...
    Jul 10, 2020 · (1) In order to decrease ATP requirements during anoxia, animals can employ various strategies including reduced physical activity1, channel ...
  139. [139]
    Suppression of reactive oxygen species generation in heart ... - NIH
    Freshwater turtles (Trachemys scripta) are among the very few vertebrates capable of tolerating severe hypoxia and re-oxygenation without suffering from damage ...Missing: purine nucleoside salvage
  140. [140]
    Endogenous opioid system down-regulation during hibernation in ...
    Aug 14, 1989 · Collectively, these data suggest that endogenous opioid systems in these northern frogs are down-regulated during fall hibernation. Original ...
  141. [141]
    New insights into survival strategies to oxygen deprivation in anoxia ...
    In cold hypoxia, reduced red blood cell concentration of ATP plays a crucial role in increasing blood oxygen affinity and/or reducing oxygen unloading to ...
  142. [142]
    Goldfish Response to Chronic Hypoxia: Mitochondrial Respiration ...
    Mar 22, 2021 · The most striking physiological response to chronic hypoxia characterized here is a drastic downregulation of Na+/K+-ATPase in goldfish brain.
  143. [143]
    Warburg effect(s)—a biographical sketch of Otto Warburg and his ...
    Mar 8, 2016 · In this way, Warburg identified cytochrome a3 (cytochrome oxidase) as being the CO-sensitive respiratory enzyme, i.e., the one requiring oxygen.
  144. [144]
    Edward Charles Slater. 16 January 1917 — 26 March 2016 - Journals
    Mar 1, 2017 · Edward Charles Slater was a key bioenergetics player, discovering the Slater factor, building a successful lab, and shaping Dutch biochemistry.
  145. [145]
    Studies on the electron transfer system. XL. Preparation ... - PubMed
    1962 May:237:1676-80. Authors. Y HATEFI, A G HAAVIK, D E GRIFFITHS. PMID: 13905327. No abstract available. MeSH terms. Cell Respiration*; Coenzymes / metabolism ...Missing: isolation complexes
  146. [146]
    Chemiosmotic coupling in oxidative and photosynthetic ...
    Abstract. 50 years ago Peter Mitchell proposed the chemiosmotic hypothesis for which he was awarded the Nobel Prize for Chemistry in 1978. His comprehensive ...