The Q cycle is a key biochemical mechanism in the cytochrome bc₁ complex of the mitochondrial and bacterial respiratory electron transport chain, as well as in the analogous cytochrome b₆f complex of the photosynthetic electron transport chain in chloroplasts and cyanobacteria, where it couples the oxidation of ubiquinol (QH₂) or plastoquinol to the reduction of cytochrome c (or plastocyanin) while translocating protons across the inner mitochondrial membrane or thylakoid membrane to establish a proton motive force for ATP synthesis.[1][2] This process enhances energy conservation by effectively doubling the proton-to-electron ratio compared to a linear transfer, achieving a net translocation of 2 H⁺ per electron transferred.[3]Proposed by Peter Mitchell in 1975 as part of the chemiosmotic hypothesis, the Q cycle resolves discrepancies in early observations of oxidant-induced reduction of cytochrome b and provides a unified explanation for proton pumping in diverse energy-transducing systems.[4][1] The mechanism operates through two quinone-binding sites: the Qo site (also called QP or quinol oxidation site) on the positive (p) side of the membrane and the Qi site (or QN, quinone reduction site) on the negative (n) side.[2][3] At the Qo site, QH₂ undergoes bifurcation upon oxidation by the Rieske iron-sulfur protein ([2Fe-2S] cluster), releasing two protons to the p-side; one electron travels via the high-potential chain through the Rieske protein to cytochrome c₁ (or f) and then to cytochrome c, while the second electron moves across the membrane via the low-potential chain of hemes b_L and b_H to the Qi site.[1][2]The full cycle requires two turnovers of QH₂ oxidation to reduce one ubiquinone (Q) to QH₂ at the Qi site, where the semiquinone intermediate is stabilized transiently before the second electron and two protons (from the n-side) complete the reduction.[3][1] This results in a net reaction of QH₂ + 2 cyt c (oxidized) + 2 H⁺ (n-side) → Q + 2 cyt c (reduced) + 4 H⁺ (p-side), with the cytochromebc₁ complex functioning as a homodimer but operating via independent monomeric units for catalysis.[2] Structural studies, including X-ray crystallography of the bovine and yeastbc₁ complexes, have revealed the dimeric architecture with inter-monomer cavities facilitating quinol/quinone exchange and precise heme orientations (e.g., edge-to-edge distances of ~7 Å between b_L and b_H), confirming the bifurcation pathway and inhibitor binding sites like stigmatellin at Qo.[1][3]In physiological contexts, the Q cycle's efficiency peaks near ambient temperatures (~298 K), optimizing thermodynamic yield under varying environmental conditions, and it minimizes reactive oxygen species production by modulating the Rieske protein's "spring-loaded" movement. Disruptions, such as those caused by mutations or inhibitors (e.g., antimycin at Qi), impair ATP production and are linked to pathologies like mitochondrial diseases.[2] Overall, the Q cycle exemplifies evolutionary conservation across domains of life, underpinning bioenergetic processes in both aerobic respiration and oxygenic photosynthesis.[1]
Overview and Context
Definition and Purpose
The Q cycle is a cyclic electron transfer mechanism operating within the cytochrome bc₁ complex, where ubiquinol (QH₂) is oxidized and ubiquinone (Q) is reduced, resulting in the translocation of four protons across the membrane for every two electrons transferred to cytochrome c. This process bifurcates the two electrons from each QH₂ molecule oxidized at the Qo site: one electron follows the high-potential chain through the Rieske iron-sulfur protein and cytochrome c₁ to reduce cytochrome c, while the other traverses the low-potential chain via cytochromes b_L and b_H to reduce a Q molecule at the Qi site, forming a semiquinone intermediate that completes the cycle upon a second turnover.[1] The net outcome of the full cycle is the oxidation of one QH₂ to Q, the reduction of two cytochrome c molecules, the uptake of two protons from the matrix side, and the release of four protons into the intermembrane space, thereby establishing a proton gradient.The primary purpose of the Q cycle is to amplify the proton motive force (PMF) generated during respiration and photosynthesis, exceeding what would be achieved by a simple linear redox reaction between QH₂ and cytochrome c.[1] By coupling electron bifurcation to proton translocation, it enhances the efficiency of energy conservation, directly contributing to ATP synthesis through oxidative phosphorylation in mitochondria or photophosphorylation in chloroplasts. This mechanism ensures that the energy from quinol oxidation is maximally harnessed to drive the PMF, which powers ATP synthase, while also minimizing wasteful side reactions like superoxide production under physiological conditions.[5]The concept was coined by Peter Mitchell in 1975 as part of his chemiosmotic theory, proposed to reconcile experimental observations of proton pumping stoichiometry in respiratory chains that could not be explained by direct scalar proton release alone.[4] Mitchell's formulation addressed the need for a cyclic pathway to account for the observed 2 H⁺/e⁻ translocation ratio in the cytochrome bc₁ complex, building on earlier discoveries like oxidant-induced reduction of cytochrome b.[4] Subsequent refinements, informed by kinetic and structural studies, solidified the Q cycle as a cornerstone of bioenergetics, with the modern "modified Q cycle" widely accepted based on evidence from bacterial and mitochondrial systems.[5]
Biological Location
The Q cycle operates primarily within the cytochrome bc₁ complex, known as complex III of the electron transport chain, which is embedded in the inner mitochondrial membrane of eukaryotic cells. This localization positions the complex to facilitate the transfer of electrons from ubiquinol to cytochrome c while contributing to the proton gradient essential for ATP synthesis. The inner mitochondrial membrane provides a lipid bilayer environment that supports the dimeric structure of the bc₁ complex and enables the spatial separation of quinone binding sites necessary for the cycle's bifurcated electron pathway.[6][7][8]In prokaryotes, the Q cycle occurs in analogous cytochrome bc₁ complexes located in the plasma membrane of respiring bacteria, such as Paracoccus denitrificans, where it functions in the aerobic respiratory chain. Photosynthetic prokaryotes and eukaryotes, including cyanobacteria and plants, feature a related Q cycle mechanism within the cytochrome b₆f complex situated in the thylakoidmembranes of chloroplasts or bacterial chromatophores, supporting cyclic electronflow around photosystem I to generate proton motive force without net oxygen evolution. These membrane localizations reflect the conservation of the Q cycle's core architecture across diverse cellular compartments optimized for energy transduction.[7][8]The Q cycle is widely distributed among aerobic respiring organisms and certain photosynthetic lineages that possess bc₁ or b₆f complexes, spanning bacteria, fungi, animals, and plants, but it is absent in strict anaerobes lacking these complexes, such as many fermentative bacteria. This distribution underscores its role in oxygen-dependent respiration and photophosphorylation, where it enhances efficiency by doubling proton translocation per electron pair. In plants, alternative oxidases in the mitochondrial electron transport chain can bypass the Q cycle to mitigate reactive oxygen species accumulation under low-oxygen conditions, such as in flooded roots, by directly oxidizing excess ubiquinol and reducing electron pressure on the bc₁ complex.[9]
Molecular Components
Cytochrome bc1 Complex Structure
The cytochrome bc1 complex is a dimeric integral membrane protein embedded in the inner mitochondrial membrane, with each monomer consisting of 11 distinct subunits and a molecular weight of approximately 240 kDa, yielding a total dimer mass of about 480 kDa.[10] The core functional subunits include cytochrome b (a polytopic membrane protein with eight transmembrane helices), the extrinsic Rieske iron-sulfur protein (with a [2Fe-2S] cluster), and the peripheral cytochrome c1 (containing a c-type heme), which together form the catalytic core responsible for redox reactions.[11] The remaining eight accessory subunits—such as core proteins 1 and 2, and smaller polypeptides like Qcr6 to Qcr10 in yeast homologs—contribute to structural stability, dimerization, and regulatory interactions without direct involvement in catalysis.[10]Key structural domains of the complex include the Qo site, located on the intermembrane space (positive) side near the Rieske protein and cytochrome c1, where ubiquinol oxidation occurs, and the Qi site, situated on the matrix (negative) side proximal to cytochrome b, facilitating ubiquinone reduction.[11] The bifurcation point, formed by the transmembrane helices of cytochrome b (specifically helices A to H), serves as the structural junction where electrons from ubiquinol bifurcate toward either the Rieske [2Fe-2S] cluster or the low-potential heme bL. These sites are lined by conserved residues from cytochrome b and coordinated by lipids and water molecules, creating hydrophobic pockets for quinone binding.[11]The first high-resolution crystal structure of the bovine mitochondrial bc1 complex was determined by Xia et al. in 1997 at 2.9 Å resolution, unveiling the overall architecture and the "clamp-like" conformational mobility of the Rieske protein head domain, which swings between positions near the Qo site and cytochrome c1. Subsequent structures, including the complete 11-subunit model by Zhang et al. in 1998, refined the subunit arrangement and transmembrane topology.[10] Further advancements, such as the 1.9 Å structure of the yeast cytochrome bc₁ complex by Solmaz and Hunte in 2008, detailed the ubiquinone binding pockets at both Qo and Qi sites, highlighting interactions with inhibitors and substrates that inform site specificity.[12]The dimeric nature of the bc1 complex is crucial for Q cycle operation, as the extensive interface between monomers—spanning over 100 residues and involving accessory subunits—allows conformational changes at the Qo site of one monomer to propagate signals influencing quinone reduction at the Qi site of the adjacent monomer, ensuring efficient and regulated catalysis.[11] This inter-monomeric communication, evident in crystal structures, prevents uncoupled electron flow and stabilizes the reactive semiquinone intermediates.
Key Redox Molecules
Ubiquinol (QH₂), the reduced form of ubiquinone, functions as a lipid-soluble two-electron carrier embedded in the inner mitochondrial membrane, where it participates in electron transfer during the Q cycle. Upon oxidation at the Qₒ site, it releases two protons into the intermembrane space, contributing to the proton gradient essential for ATP synthesis. In mammals, ubiquinol features a chemical formula of C₅₉H₉₂O₄, characterized by a redox-active benzoquinone ring attached to a variable isoprenoid tail, typically comprising 10 units for enhanced membrane solubility.[13][14]Ubiquinone (Q), the oxidized counterpart to ubiquinol, serves as an electron acceptor at the Qᵢ site during the Q cycle, where it is reduced by taking up two protons from the mitochondrial matrix. A critical intermediate in this process is the semiquinone anion (Q•⁻), a stabilized radicalspecies formed at the Qᵢ site that facilitates sequential one-electron transfers. The interconversion between ubiquinone and ubiquinol enables the cyclic movement of electrons and protons across the membrane, with the quinone pool maintaining a dynamic redox equilibrium.[14][15]Cytochrome b, a core subunit of the bc₁ complex, harbors two b-type hemes that enable transmembrane electron transfer in the low-potential branch of the Q cycle. The proximal heme b_L possesses a low midpoint redox potential of approximately -60 mV, while the distal heme b_H has a higher potential of about +50 mV; these distinct potentials drive sequential electron flow from the Qₒ site toward the Qᵢ site, preventing backflow and supporting efficient charge separation across the membrane.[14][15]The Rieske iron-sulfur cluster, a [2Fe-2S] center within the mobile extrinsic domain of the Rieske protein, exhibits a high midpoint redox potential of roughly +280 mV, positioning it as an effective one-electron acceptor from ubiquinol at the Qₒ site. This cluster's histidine coordination and pH-dependent protonation enhance its role in bifurcated electron transfer, directing one electron to the high-potential chain while coupling to proton release.[16]Cytochrome c₁ contains a covalently bound c-type heme with a midpoint redox potential of approximately +230 mV, enabling it to receive electrons from the reduced Rieske cluster and subsequently donate them to soluble cytochrome c in the intermembrane space. This heme's axial ligation by histidine and methionine residues optimizes its redox tuning for rapid turnover in the Q cycle's high-potential pathway.[17][14]
Mechanism
Two-Site Hypothesis
The two-site hypothesis for the Q cycle was originally proposed by Peter Mitchell in 1975 to explain observed non-integer ratios of protons translocated per electron (H⁺/e⁻) in submitochondrial particles, which deviated from expectations of simple linear electron transfer in the cytochrome bc₁ complex. This model posited distinct binding sites for ubiquinol (QH₂) oxidation and ubiquinone (Q) reduction within the complex, enabling bifurcated electron flow and enhanced proton motive force generation beyond a 1:1 H⁺/e⁻ stoichiometry. In 1984, Peter R. Rich further refined the hypothesis by analyzing physicochemical constraints on quinoneredox potentials and site-specific interactions, emphasizing how differential semiquinone stabilities at each site drive the cyclic mechanism.[18]At the core of the two-site hypothesis, QH₂ oxidation occurs at the Qo site (positive or outer side of the membrane), where electron bifurcation directs one electron to the high-potential chain—via the Rieske iron-sulfur protein ([2Fe-2S]), cytochrome c₁, and ultimately cytochrome c—while the second electron travels through the low-potential chain involving cytochromes b_L and b_H to reduce Q at the Qi site (negative or inner side). This asymmetry ensures that the semiquinone intermediate (Q•⁻) is stabilized at the Qi site for subsequent two-electron reduction to QH₂ (with proton uptake from the matrix), but is highly unstable and short-lived at the Qo site, preventing reactive oxygen species formation and propelling the cycle. The distinct Qo and Qi sites, identified through structural studies of the cytochrome bc₁ complex, facilitate vectorial proton translocation across the membrane.Supporting evidence for the two-site model derives from inhibitor binding studies, where stigmatellin selectively blocks the Qo site by mimicking quinol and preventing Rieske protein displacement, thus inhibiting initial QH₂ oxidation without affecting Qi-mediated reduction. Conversely, antimycin A binds near the Qi site, blocking electron transfer from b_H to Q and accumulating reduced cytochromesb, which confirms site-specific bifurcation. Electron paramagnetic resonance (EPR) spectroscopy has directly detected the stable Q•⁻ radical at the Qi site under conditions favoring its accumulation, such as partial reduction with antimycin, validating semiquinone involvement in the cycle's second arm.Early debates in the 1970s and 1980s contrasted linear (non-cyclic) electron transfer models, which predicted only 1 H⁺/e⁻, against cyclic schemes like the Q cycle requiring 2 H⁺/e⁻ for bc₁ activity. These alternatives were largely dismissed following proton flux measurements in isolated complexes and chromatophores, which demonstrated the predicted 2:1 stoichiometry consistent with the two-site bifurcation.
Step-by-Step Electron and Proton Transfer
The [Q cycle](/page/Q cycle) in the cytochrome bc₁ complex proceeds through two sequential half-cycles, each initiating with the oxidation of ubiquinol (QH₂) at the Qo site on the intermembrane space (p-side) of the inner mitochondrial membrane, resulting in bifurcated electron transfer and proton translocation.[14]In the first half-cycle, QH₂ binds to the Qo site and undergoes oxidation, releasing two protons into the intermembrane space. One electron is transferred to the oxidized Rieske iron-sulfur protein ([2Fe-2S] cluster), which undergoes a rapid conformational change—displacing its extrinsic domain by approximately 30 Å to dock with cytochrome c₁—enabling subsequent electron delivery to cytochrome c via the high-potential chain. The second electron follows the low-potential chain, reducing heme b_L and then heme b_H, before crossing the dimer interface to the Qi site on the adjacent monomer, where it reduces bound ubiquinone (Q) to a semiquinone radical anion (Q•⁻); this reduction is coupled to the uptake of one proton from the matrix. The Rieske conformational change occurs on a timescale of 10–100 μs, ensuring efficient bifurcation without significant back-reaction. The semiquinone at the Qi site exhibits a lifetime of approximately 1 ms under physiological conditions.[19][14]The second half-cycle mirrors the first, with another QH₂ molecule binding to the Qo site and undergoing bifurcated oxidation, again releasing two protons to the intermembrane space. One electron travels the high-potential chain to reduce a second cytochrome c molecule, while the other electron traverses the low-potential chain (b_L to b_H) to the Qi site semiquinone (Q•⁻), providing the second electron for its reduction to ubiquinol (QH₂) and coupling to uptake of one proton from the matrix. This completes the quinone reduction at the Qi site (n-side, matrix-facing).[14]The full Q cycle requires functional cooperation between the two monomers of the dimeric bc₁ complex, as the low-potential electron transfer from b_H in one monomer supplies the Qi site of the other, enabling the double turnover necessary for net ubiquinol oxidation.[19] The overall net reaction is: QH₂ + 2 cytochrome c (oxidized) + 2 H⁺ (matrix) → Q + 2 cytochrome c (reduced) + 4 H⁺ (intermembrane space).[14]Inhibitors disrupt specific steps: myxothiazol binds near the Qo site, blocking QH₂ oxidation and preventing electron bifurcation by stabilizing the Rieske protein in a non-docked position; antimycin A binds at the Qi site, inhibiting Q reduction and trapping the semiquinone intermediate by blocking electron arrival from heme b_H.[14]
Energetics and Function
Proton Translocation Efficiency
The Q cycle mechanism in the cytochrome bc₁ complex achieves a proton translocation stoichiometry of 4 H⁺ per 2 electrons transferred from ubiquinol (QH₂) to two molecules of cytochrome c. This includes two scalar protons released upon QH₂ deprotonation at the Qo site on the positive side of the membrane and two vectorial protons taken up during QH₂ formation at the Qi site on the negative side, contrasting with non-cyclic models that translocate only 2 H⁺ per 2 e⁻ through scalar release alone.[20][21]The redox energetics driving this process span a total potential difference (ΔE) of approximately 190 mV between the ubiquinol/ubiquinone couple (E_m ≈ +60 mV) and cytochrome c (E_m ≈ +250 mV), providing the free energy for electron transfer. The Q cycle amplifies the generation of the proton motive force (PMF) by bifurcating electrons at the Qo site, which enables translocation of 4 H⁺ using the same overall redox span as a linear mechanism—effectively doubling the H⁺/e⁻ ratio and enhancing energy conservation. The ΔpH component of the PMF typically contributes 30–60 mV (corresponding to ΔpH ≈ 0.5–1 unit) in mitochondria. This improved efficiency arises from the bifurcation of electrons at the Qo site, enabling the low-potential b hemes to recycle one electron back to the Qi site for quinol reduction, thereby amplifying proton displacement without requiring extra redox span.[20][22]The net free energy change (ΔG) for the Q cycle reaction, accounting for the redox driving force and the opposing ΔpH, can be expressed as:\Delta G = -n F \Delta E + 2.3 R T \Delta \mathrm{pH}where n = 2 (electrons transferred), F is the Faraday constant, R is the gas constant, and T is temperature in Kelvin; this formulation highlights the balance between redox energy input and the chemical proton gradient back-pressure for the scalar protons. Overall coupling efficiency to PMF is approximately 78%, reflecting the fraction of redox energy conserved as trans-membrane proton gradient rather than dissipated as heat.[23][20]Experimental validation of this 4 H⁺/2 e⁻ stoichiometry comes from reconstitution assays of purified bc₁ complex into liposomes, where steady-state proton extrusion coupled to electron transfer from decylubiquinol to cytochrome c yielded ratios approaching 4 H⁺ per 2 e⁻ under controlled conditions, confirming the Q cycle's role in vectorial proton movement.[20]In the analogous cytochrome b₆f complex of photosynthesis, the energetics are similar, with plastoquinol/plastoquinone (E_m ≈ +100 mV) to plastocyanin (E_m ≈ +360 mV) providing a ΔE of ~260 mV, supporting 4 H⁺/2 e⁻ translocation adapted for the linear electron flow between photosystems, though with adjustments for the thylakoid ΔpH often exceeding 2 units under illumination.[1]
Integration with Electron Transport Chain
The Q cycle operates within complex III (cytochrome bc₁ complex) of the mitochondrial electron transport chain (ETC), serving as a critical link between the oxidation of ubiquinol (QH₂) and the reduction of cytochrome c. Upstream, QH₂ is generated by the transfer of electrons from reducing equivalents such as NADH via complex I (NADH:ubiquinone oxidoreductase) or from succinate via complex II (succinate:ubiquinone oxidoreductase), populating the ubiquinone pool in the inner mitochondrial membrane.[17][24] This positions the Q cycle to receive electrons from both NADH-linked substrates (entering at complex I) and FADH₂-linked substrates (entering at complex II), enabling flexible routing of electrons into the chain.[17]Downstream, the Q cycle facilitates the transfer of electrons to cytochrome c, which then delivers them to complex IV (cytochrome c oxidase) for the ultimate reduction of molecular oxygen to water.[17][24] In the context of full ETC coupling from NADH oxidation, the Q cycle at complex III contributes 4 protons translocated per 2 electrons (out of a total of 10 protons across the chain: 4 from complex I, 4 from complex III, and 2 from complex IV), thereby enhancing the protonmotive force that drives ATP synthesis.[24] In certain organisms, such as plants, alternative pathways like the alternative oxidase can bypass complexes III and IV, directly oxidizing QH₂ to reduce O₂ without proton translocation at these sites, which lowers overall respiratory efficiency but helps manage excess reducing power.[25]The integration of the Q cycle is further regulated through the formation of respiratory supercomplexes, which assemble complexes I, III, and IV into higher-order structures that promote efficient substrate channeling between the ubiquinone pool and cytochrome c, thereby minimizing electron leakage and reducing reactive oxygen species (ROS) production.[26] This organization optimizes electron flow and protects against oxidative stress, underscoring the Q cycle's role in maintaining balanced ETC function.[26]
Physiological and Evolutionary Aspects
Role in Cellular Respiration
The Q cycle, integral to the cytochrome bc1 complex (complex III) in the mitochondrial electron transport chain, contributes to aerobic cellular respiration by enhancing proton translocation across the inner mitochondrial membrane, thereby bolstering the proton motive force (PMF) that powers ATP synthesis. During the cycle, oxidation of ubiquinol at the Qo site and reduction at the Qi site result in the net translocation of four protons to the intermembrane space per two electrons transferred to cytochrome c, amplifying the electrochemical gradient essential for oxidative phosphorylation. This PMF drives ATP synthase (complex V) to convert ADP and inorganic phosphate into ATP, with the Q cycle contributing the translocation of 4 protons per two electrons transferred, which supports the overall efficiency of approximately 2.5 ATP per NADH oxidized in eukaryotic mitochondria (based on ~4 H⁺ per ATP).A notable side effect of the Q cycle is its potential to generate reactive oxygen species (ROS), primarily through the unstable semiquinone intermediate formed at the Qo site during ubiquinol bifurcation, which can reduce molecular oxygen to superoxide, with leakage rates typically below 1% of electron flux under physiological conditions but up to several percent under stress. Superoxide production is largely confined to the matrix or intermembrane space, where it is rapidly neutralized by superoxide dismutase (SOD) enzymes, converting it to hydrogen peroxide and oxygen to safeguard cellular components from oxidative stress. This controlled ROS leakage may also serve signaling roles in respiration but becomes detrimental when dysregulated.[27]Pathologically, disruptions to the Q cycle via mutations in the Rieske iron-sulfur protein or cytochrome b subunits impair electron flow and proton pumping, manifesting as exercise intolerance due to diminished ATP supply and heightened ROS output in skeletal muscle mitochondria. In ischemic tissues, Q cycle inhibition during oxygen deprivation exacerbates reperfusion injury by promoting semiquinone accumulation and superoxide bursts upon reoxygenation, contributing to myocardial or neuronal damage.[28][29][30]Advancements in the 2020s, including cryo-EM structures of the cytochrome bc1 complex resolved at near-atomic resolution in 2022-2023, have illuminated dynamic gating at the Qo site that stabilizes semiquinone under low-oxygen (hypoxic) conditions, thereby minimizing ROS generation and preserving respiratory efficiency during metabolic stress.[31]
Variations Across Organisms
In bacterial systems, the cytochrome bc1 complex typically consists of three core subunits—cytochrome b, cytochrome c1, and the Rieske iron-sulfur protein—representing a simpler architecture compared to eukaryotic homologs, yet it operates the canonical Q cycle mechanism for proton translocation during respiration or photosynthesis.[11] For instance, in the purple phototrophic bacterium Rhodobacter sphaeroides, the bc1 complex facilitates cyclic electron flow around the photosynthetic reaction center, oxidizing ubiquinol at the Qo site and reducing cytochrome c2, thereby generating a proton motive force essential for ATP synthesis under anaerobic photosynthetic conditions.[31] This adaptation underscores the complex's role in linking photosynthetic and respiratory electron transport in facultative anaerobes.The chloroplast cytochrome b6f complex exhibits a mechanistically analogous Q cycle but substitutes plastocyanin (or cytochrome c6 in some cyanobacteria) for cytochrome c as the electron acceptor at the high-potential chain, enabling electron transfer from photosystem II to photosystem I.[32] In linear electron flow, the b6f complex can sometimes bypass the full Q cycle, operating in a scalar mode that translocates only 2 H⁺ per 2 electrons rather than the full 4 H⁺, resulting in reduced proton pumping efficiency compared to the obligatory Q cycle in bc1; this flexibility supports balanced NADPH/ATP production under varying light conditions.[33] Structural features, such as a shorter n-side proton uptake pathway facilitated by heme cn near the Qi site, further optimize proton transfer in the thylakoid membrane despite these variations.[34]The Q cycle mechanism traces its evolutionary origins to a bacterial ancestor around 2.7–3 billion years ago, with the emergence of oxygenic photosynthesis in ancient cyanobacteria, and later diversification of respiratory chains in proteobacteria around the Great Oxidation Event (~2.4 billion years ago).[35] It is absent in archaea, where any instances arise from lateral genetransfer from bacterial donors rather than vertical inheritance.[36] However, the mechanism persists in certain anaerobicbacteria through variants employing menaquinone instead of ubiquinone as the quinone substrate, allowing Q cycle operation in low-oxygen environments such as those of purple nonsulfur bacteria during fermentative or sulfur-based metabolism.[37]Recent structural studies on cyanobacterial cytochrome b6f complexes, including high-resolution cryo-EM analyses from 2023, reveal modifications at the Qi site that enhance environmental adaptability, such as improved stability under fluctuating redox conditions akin to those in sulfidic habitats where cyanobacteria perform anoxygenic photosynthesis.[32] These adaptations, including altered quinone binding affinities, contribute to sulfide tolerance by minimizing inhibitory interactions at the reduction site during transitions between oxic and anoxic states.[38]