Fact-checked by Grok 2 weeks ago

Bioenergetics

Bioenergetics is the branch of biochemistry and that examines the flow and transformation of within , encompassing the mechanisms by which organisms capture, convert, and utilize to sustain life processes such as , growth, and reproduction. This field integrates principles from and to explain how from external sources—like in or chemical bonds in food—is harnessed to drive cellular activities. Central to bioenergetics are the , which govern transformations in biological contexts. states that is conserved, meaning it can neither be created nor destroyed but only converted from one form to another, such as to during . The second law introduces the concept of , asserting that transfers increase disorder in the and that usable decreases over time, resulting in inefficiencies like loss in metabolic reactions. These principles ensure that biological systems, while open and exchanging with their environment, operate with a net increase in , yet maintain through coupled reactions that link energy-releasing (exergonic) processes to energy-requiring (endergonic) ones. Adenosine triphosphate (ATP) serves as the primary energy currency in cells, facilitating energy transfer through its hydrolysis, which releases free energy under standard biological conditions (ΔG°′ ≈ -30.5 kJ/mol). ATP is synthesized mainly via cellular respiration in mitochondria or photosynthesis in chloroplasts, where electron transport chains create proton gradients that power ATP synthase. Key metabolic pathways, including glycolysis, the citric acid cycle, and oxidative phosphorylation, exemplify catabolic processes that break down nutrients to generate ATP, while anabolic pathways like protein synthesis consume it. Electron carriers such as nicotinamide adenine dinucleotide (NAD⁺/NADH) and flavin adenine dinucleotide (FAD/FADH₂) play crucial roles in shuttling high-energy electrons, linking catabolism to ATP production and enabling efficient energy coupling. Bioenergetics extends beyond basic to influence broader physiological and pathological phenomena. In multicellular organisms, it underpins processes like uptake, transport across membranes, and , all of which rely on gradients. Disruptions in bioenergetic pathways, such as mitochondrial dysfunction, are implicated in aging and diseases including neurodegeneration, highlighting the field's relevance to and therapeutic development. Overall, bioenergetics reveals how life's complexity arises from precise , from prokaryotic cells to human physiology.

Fundamentals

Definition and Scope

Bioenergetics is the branch of biochemistry that examines the transformation, flow, and utilization of energy within living organisms, particularly how cells acquire, store, and release energy to sustain biological processes. This field focuses on the mechanisms by which energy from external sources, such as or nutrients, is converted into usable forms to drive cellular activities. The scope of bioenergetics spans multiple scales, from molecular events like the of (ATP), the central currency of cells, to organismal phenomena such as metabolic rates that determine overall budgets. It encompasses both catabolic pathways, which release through the breakdown of complex molecules, and anabolic pathways, which require input to synthesize macromolecules essential for growth and repair. These processes ensure the maintenance of life by balancing demands across cellular, , and whole-organism levels. Bioenergetics originated in the early through investigations into energetics by Otto Meyerhof and , who linked chemical reactions to mechanical work, and studies on by Arthur Harden and William Young, revealing the role of in energy transfer. The field advanced significantly with Hans Krebs's elucidation of the in 1937, providing a key framework for energy-yielding metabolism, and Peter Mitchell's chemiosmotic theory in 1961, which explained energy coupling in membranes and earned him the 1978 . As an interdisciplinary domain, bioenergetics integrates principles from , biochemistry, and to model dynamics in biological systems. Its applications extend to , where disruptions in bioenergetic pathways contribute to metabolic disorders like mitochondrial diseases, and to , informing models of flow through ecosystems that underpin trophic interactions and .

Thermodynamic Principles

Bioenergetics is fundamentally governed by the laws of thermodynamics, which dictate the conservation and transformation of energy in living systems. The first law of thermodynamics, also known as the law of conservation of energy, states that the total energy of an isolated system remains constant; energy can neither be created nor destroyed, but only converted from one form to another. In biological contexts, this principle applies to cells and organisms, where chemical energy from nutrients is transformed into mechanical work, heat, or stored forms like ATP, without net loss or gain in the overall energy balance. For instance, during metabolism, the energy input from food equals the sum of outputs in work, heat, and waste products. The second law of thermodynamics introduces the concept of , stating that in any , the total of an and its surroundings increases, leading to greater over time. Living systems, however, appear to defy this trend by maintaining highly ordered structures, such as organized proteins and membranes, which represent low- states. This is possible because biological systems are not isolated; they continuously import low- energy (e.g., from or ) and export high- waste (e.g., and CO₂), thereby increasing the of the surroundings while locally decreasing their own . This dissipative process ensures compliance with the second law on a universal scale. A key thermodynamic quantity in bioenergetics is the Gibbs free energy change (ΔG), which predicts the spontaneity of reactions under constant temperature and pressure. Defined by the equation \Delta G = \Delta H - T \Delta S where ΔH is the enthalpy change, T is the absolute temperature, and ΔS is the entropy change, ΔG determines if a process is exergonic (spontaneous, ΔG < 0) or endergonic (non-spontaneous, ΔG > 0). The standard free energy change (ΔG°) refers to conditions of 1 atm pressure, 25°C, and 1 M concentrations for solutes, providing a benchmark for reaction favorability. In biology, negative ΔG values drive essential processes like glycolysis, while positive values require coupling to exergonic reactions for feasibility. The relationship between ΔG° and the (K_eq) is given by \Delta G^\circ = -RT \ln K_{eq} where R is the (8.314 J/mol·K) and T is in . This equation links to , as a large positive K_eq (favoring products) corresponds to a negative ΔG°, indicating a highly favorable . In biological systems, this helps predict the and extent of reactions, such as enzyme-catalyzed conversions, where deviations from conditions adjust the actual ΔG via the . Biological systems operate as open systems, exchanging both matter and with their environment, which allows them to sustain non-equilibrium steady states far from . Unlike closed systems that inevitably approach equilibrium and maximum , open systems like cells can self-organize through dissipative structures, where energy throughput maintains order and enables functions like growth and repair. This framework, developed in , explains how living organisms achieve without violating the second law. In biological reactions, (ΔH) reflects heat changes associated with bond breaking and formation, often exothermic (ΔH < 0) in energy-releasing steps like oxidation. Entropy (ΔS) quantifies changes in molecular disorder, such as increased freedom in unfolded proteins or decreased order in assemblies like lipid bilayers. For example, protein folding typically involves a negative ΔS due to reduced conformational flexibility, balanced by favorable ΔH from hydrophobic interactions, resulting in an overall negative ΔG that stabilizes the native state. These contributions highlight how bioenergetic processes balance energetic and probabilistic factors to drive cellular functions.

Reaction Types

Exergonic and Endergonic Processes

In bioenergetics, reactions are classified as exergonic or endergonic based on the change in Gibbs free energy (ΔG), which determines spontaneity under cellular conditions. Exergonic reactions occur spontaneously and release free energy when ΔG is negative (ΔG < 0), providing usable energy for cellular work. A representative example is the oxidation of during catabolism, which liberates energy that can be harnessed for . Endergonic reactions, in contrast, are non-spontaneous and require an input of , as indicated by a positive ΔG (ΔG > 0). These processes cannot proceed alone in cells but must be coupled to s to drive them forward. For instance, the of into proteins is endergonic, necessitating energy from other metabolic pathways to form bonds. A prototypical in is the of (ATP) to (ADP) and inorganic phosphate (Pi), with a standard change (ΔG°') of approximately -30.5 kJ/mol under biochemical conditions (pH 7, 25°C, 1 mM Mg²⁺). This reaction powers numerous endergonic processes through energy coupling, making ATP the primary currency in cells. Many bioenergetic reactions are reversible, and their directionality can shift based on reactant and product concentrations via the mass action ratio. The actual free energy change is given by the equation: \Delta G = \Delta G^{\circ\prime} + RT \ln Q where \Delta G^{\circ\prime} is the standard free energy change, R is the gas constant, T is the absolute temperature, and Q is the reaction quotient (the ratio of product to reactant concentrations). In cellular environments, non-standard concentrations often make ΔG more negative than ΔG°', favoring exergonic directions. Biologically, exergonic catabolic pathways, such as breakdown, generate that fuels endergonic anabolic processes, like , ensuring a net negative ΔG for overall . This coupling maintains cellular by balancing production and consumption, supporting growth, repair, and response to environmental changes.

Redox Reactions in Biology

Redox reactions, fundamental to bioenergetics, involve the transfer of electrons between molecules, where oxidation is defined as the loss of electrons and as the gain of electrons. These processes always occur simultaneously in paired s: the oxidation releases electrons from a donor, while the accepts those electrons at an acceptor. To balance a complete redox equation, the number of electrons lost in the oxidation half must equal those gained in the half, often requiring multiplication of half-reactions by appropriate coefficients to conserve charge and mass. In biological systems, such reactions underpin capture and transfer, enabling the oxidation of nutrients to release electrons for subsequent use. The driving force of a redox reaction is quantified by the standard reduction potential (E°'), which measures a half-reaction's tendency to gain electrons under standard conditions ( 7, 25°C, 1 M concentrations for solutes). Positive E°' values indicate a strong tendency for , while negative values favor oxidation; the overall potential (ΔE°') is the difference between the reduction potential of the acceptor and the donor, with spontaneous reactions yielding positive ΔE°'. Under non-standard biological conditions, the adjusts the potential:
E = E^{\circ\prime} - \frac{RT}{nF} \ln Q
where R is the , T is temperature in , n is the number of electrons transferred, F is Faraday's constant, and Q is the (ratio of reduced to oxidized species concentrations). This equation predicts how cellular concentrations influence reaction direction and energy yield.
Key biological electron carriers facilitate electron shuttling in these reactions, often as coenzymes or proteins with embedded centers. (NAD⁺/NADH) operates at E°' ≈ -0.32 V, accepting a hydride ion (H⁻, equivalent to two electrons and a proton) in catabolic dehydrogenations. (FAD/FADH₂) has E°' ≈ -0.22 V (varying with protein binding), also handling two-electron transfers but suited for higher-potential substrates. , iron-heme proteins, exhibit positive potentials (e.g., at +0.25 V), enabling stepwise electron relay toward oxygen. These carriers link low-potential donors like NADH to high-potential acceptors, harnessing the energy difference. The table below summarizes select potentials:
Half-ReactionE°' (V)
NAD⁺ + H⁺ + 2e⁻ → NADH-0.320
FAD + 2H⁺ + 2e⁻ → FADH₂-0.219
Cytochrome c (Fe³⁺) + e⁻ → Fe²⁺+0.254
In energy transduction, redox couples from catabolic pathways—such as NADH generated in glycolysis and the tricarboxylic acid cycle—provide high-energy electrons that flow to acceptors like oxygen (E°' +0.82 V), releasing free energy to drive ATP synthesis. This electron flow creates a potential gradient exploited in membrane-bound systems, converting catabolic redox energy into usable chemical bonds without direct coupling to non-redox steps. For instance, dehydrogenases like alcohol dehydrogenase oxidize substrates while reducing NAD⁺ to NADH, fueling downstream metabolism. However, incomplete electron transfer can produce reactive byproducts, such as superoxide (O₂⁻•), formed when O₂ prematurely accepts a single electron during aerobic respiration, contributing to oxidative stress if not scavenged.

Energy Coupling

Reaction Coupling

In bioenergetics, reaction coupling refers to the linkage of an (with positive change, ΔG > 0) to an (ΔG < 0) such that the overall process yields a negative ΔG, making it thermodynamically favorable. This principle relies on shared chemical intermediates or carriers that transfer between the reactions, preventing the dissipation of energy as heat and ensuring efficient cellular function. Coupling occurs in two primary types: direct and indirect. Direct coupling involves the immediate transfer of a chemical group, such as in where a from an is directly transferred to a in an endergonic step, as seen in certain glycolytic intermediates. Indirect coupling, by contrast, utilizes high-energy intermediates like ATP to mediate energy transfer across separate reactions, allowing spatial and temporal separation while maintaining overall feasibility. A representative example is , the synthesis of glucose from non-carbohydrate precursors like pyruvate, which is inherently endergonic but rendered viable through to . The conversion of two pyruvate molecules to glucose requires the hydrolysis of six ATP equivalents (four ATP and two GTP), providing the necessary energy input; the overall ΔG under physiological conditions for the pathway is approximately -16 kJ/mol, negative due to this , enabling net glucose production under physiological conditions. ATP plays a central role in such s by serving as a universal energy currency, though its specific mechanisms are detailed elsewhere. Evolutionarily, reaction confers a significant advantage by enabling the of complex macromolecules from simpler precursors, fostering metabolic versatility and the emergence of sophisticated cellular networks essential for . This mechanism minimizes energy waste, allowing organisms to allocate resources efficiently for growth and adaptation in diverse environments. However, uncoupled reactions can occur, leading to pitfalls such as energy dissipation as heat rather than productive work. In , for instance, uncoupling protein 1 () intentionally decouples electron transport from ATP synthesis, channeling proton motive force into for non-shivering heat , which is adaptive in environments but represents a controlled loss of coupling .

Phosphorylation and Group Transfer

Phosphorylation serves as a central mechanism in bioenergetics for transferring energy through the addition of phosphate groups to molecules, enabling the storage and release of energy in cellular processes. (ATP), the universal energy carrier, consists of an base linked to a sugar and a chain of three phosphate groups, where the bonds between the β and γ phosphates, as well as α and β phosphates, are high-energy phosphoanhydride linkages that store potential energy due to electrostatic repulsion and stabilization upon . The of ATP to () and inorganic phosphate (Pi) is exergonic, with a standard free energy change (ΔG°') of approximately -30.5 kJ/mol at pH 7 and 25°C, reflecting the release of energy that drives endergonic reactions in the cell. This reaction, ATP + H₂O → + Pᵢ, is highly favorable because the products are more stable than the reactant, primarily due to the relief of charge repulsion in the triphosphate chain and increased in and Pi. Two primary types of phosphorylation facilitate ATP production and energy transfer: and . In , a group is directly transferred from a high-energy to via an , without involving electron transport; a key example occurs in , where catalyzes the transfer from 1,3-bisphosphoglycerate to , forming ATP. , in contrast, couples transfer to with the oxidation of electron donors in the , generating ATP through a proton gradient across membranes, though the direct transfer mechanism is mediated by . These processes highlight phosphorylation's role in converting from diverse sources into the ATP pool. Beyond ATP, other phosphorylated compounds exhibit higher group transfer potentials, making them useful for specific energy storage and transfer roles. Phosphoenolpyruvate (PEP), an intermediate in , has a particularly high free energy of (ΔG°' ≈ -61.9 kJ/mol), allowing it to phosphorylate even under conditions where ATP synthesis might otherwise be unfavorable. phosphate, prevalent in muscle and tissues, serves as a rapid ATP reservoir with a ΔG°' of around -43.1 kJ/mol; it donates its phosphate to via during high-energy demand, such as , to quickly replenish ATP without relying on slower metabolic pathways. The effectiveness of phosphorylation in bioenergetics stems from the group transfer potential of phosphate, which quantifies the tendency of a phosphorylated compound to donate its phosphate to an acceptor based on the free energy of hydrolysis—the more negative the ΔG°', the higher the potential. This ordered transfer, often enzyme-catalyzed, exploits differences in these potentials to drive reactions forward; for instance, the high potential of PEP or ATP enables direct phosphorylation of substrates with lower potentials, like glucose in hexokinase-catalyzed reactions. In enzymes such as ATPases and kinases, the release or transfer of phosphate induces conformational changes, transitioning the protein between active and inactive states to perform mechanical work or catalyze subsequent steps, as seen in the power stroke of myosin during muscle contraction. Regulation of phosphorylation states is crucial for maintaining cellular and integrating bioenergetic processes with signaling. Protein kinases catalyze the addition of groups from ATP to target proteins, often activating or inhibiting enzymatic activity, while protein phosphatases reverse this by hydrolyzing the ester bonds, ensuring dynamic control. This kinase-phosphatase balance is pivotal in pathways, where cascades amplify extracellular signals to modulate metabolic fluxes, such as in the insulin-mediated activation of glycogen synthase kinase-3.
CompoundReactionΔG°' (kJ/mol)
ATP (to ADP + Pᵢ)Hydrolysis of γ-phosphate-30.5
PhosphoenolpyruvatePEP → Pyruvate + Pᵢ-61.9
Creatine phosphateCreatine phosphate → Creatine + Pᵢ-43.1

Key Bioenergetic Processes

Cellular Respiration

Cellular respiration is the primary catabolic process in aerobic organisms, involving the complete oxidation of organic molecules, such as glucose, to produce adenosine triphosphate (ATP) as the main energy currency. This multistage pathway occurs primarily in eukaryotic cells, with glycolysis in the cytosol and subsequent stages in the mitochondria, ultimately yielding up to 32 ATP molecules per glucose molecule under typical conditions. The overall reaction is represented by the equation: \text{C}_6\text{H}_{12}\text{O}_6 + 6\text{O}_2 \rightarrow 6\text{CO}_2 + 6\text{H}_2\text{O} + \text{energy (ATP + heat)} This process releases approximately 686 kcal/mol of free energy from glucose, with about 40% captured in ATP and the remainder dissipated as heat. The first stage, glycolysis, takes place in the cytosol and converts one glucose molecule into two pyruvate molecules, generating a net yield of 2 ATP through substrate-level phosphorylation and 2 NADH molecules as electron carriers. The key reaction is: \text{Glucose} + 2\text{NAD}^+ + 2\text{ADP} + 2\text{P}_i \rightarrow 2\text{pyruvate} + 2\text{NADH} + 2\text{H}^+ + 2\text{ATP} + 2\text{H}_2\text{O} Pyruvate oxidation follows in the mitochondrial matrix, where each pyruvate is decarboxylated to form acetyl-CoA, producing 1 NADH per pyruvate (2 NADH total per glucose) and releasing CO₂. No ATP is directly produced here, but the NADH contributes to later ATP synthesis. The tricarboxylic acid (TCA) cycle, also known as the Krebs or citric acid cycle, oxidizes the two molecules in the , yielding 2 ATP (via GTP equivalents) through , 6 NADH, and 2 FADH₂ per glucose. The overall TCA cycle reaction per is: \text{Acetyl-CoA} + 3\text{NAD}^+ + \text{FAD} + \text{GDP} + \text{P}_i + 2\text{H}_2\text{O} \rightarrow 2\text{CO}_2 + 3\text{NADH} + 3\text{H}^+ + \text{FADH}_2 + \text{GTP} + \text{CoA} These reduced carriers (totaling 10 NADH and 2 FADH₂ from all stages) donate electrons to the in the during , the final stage that accounts for the majority of ATP production. The consists of four protein complexes (I–IV), mobile carriers ubiquinone (coenzyme Q) and , and uses O₂ as the terminal , reducing it to . Electrons from NADH enter at Complex I (), pumping protons across the membrane, while those from FADH₂ enter at Complex II (), bypassing Complex I. Ubiquinone shuttles electrons from Complexes I and II to Complex III (cytochrome bc₁), and transfers them to Complex IV (), where O₂ is reduced. The key NADH oxidation in the can be summarized as: \text{NADH} + \frac{1}{2}\text{O}_2 + \text{H}^+ \rightarrow \text{NAD}^+ + \text{H}_2\text{O} Proton pumping creates an electrochemical gradient that drives ATP synthesis, with a P/O ratio (ATP per oxygen atom reduced) of approximately 2.5 for NADH and 1.5 for FADH₂, leading to a theoretical maximum of 36–38 ATP per glucose but an actual yield of 30–32 ATP due to factors like proton leaks and shuttle inefficiencies. In anaerobic conditions, cells rely on to regenerate NAD⁺ for continuation, as the cannot function without O₂. in muscle cells converts pyruvate to lactate, while alcoholic fermentation in produces and CO₂, both yielding only the net 2 ATP from with no additional . These pathways maintain minimal ATP production but allow survival in oxygen-limited environments.

Photosynthesis

Photosynthesis is the process by which autotrophic organisms, primarily , algae, and cyanobacteria, convert solar energy into chemical energy stored in carbohydrates. This bioenergetic pathway occurs in chloroplasts and consists of two main stages: the , which capture light energy to generate ATP and NADPH, and the light-independent reactions (Calvin-Benson cycle), which use these products to fix atmospheric CO₂ into organic molecules. The overall reaction is 6CO₂ + 6H₂O + light energy → C₆H₁₂O₆ + 6O₂, representing a reversal of and serving as the primary energy input for Earth's . In the , occurring in the membranes, I (PSI) and II (PSII) absorb photons to initiate electron transport. PSII, with its reaction center P680, absorbs light at approximately 680 nm and oxidizes through the , splitting it via the reaction 2H₂O → 4H⁺ + 4e⁻ + O₂, releasing oxygen as a byproduct and providing electrons for the chain. These electrons pass through , the cytochrome b₆f complex, and to PSI, which absorbs light at 700 nm and boosts the electrons to reduce NADP⁺ to NADPH. This non-cyclic electron flow, also known as the Z-scheme, generates a proton gradient across the membrane for ATP synthesis via . Cyclic electron flow around PSI, involving and the cytochrome b₆f complex, produces additional ATP without NADPH or O₂ evolution, helping balance the ATP/NADPH ratio needed for carbon fixation. The Z-scheme illustrates the changes, starting from the high potential of (+0.82 V) at PSII, dropping through carriers, and boosted twice by absorption to reach the low potential of NADP⁺/NADPH (-0.32 V) at . This pathway requires input from two photons per , overcoming the endergonic nature of NADPH formation; the standard free energy change (ΔG°') for H₂O + NADP⁺ → ½O₂ + NADPH + H⁺ is approximately +220 kJ/mol, supplied by the absorbed . The ensures efficient charge separation and minimizes loss, with traversing a potential difference of about 1.14 V overall. The light-independent reactions, or Calvin-Benson cycle, take place in the stroma and convert CO₂ into (G3P), a precursor to glucose. The cycle begins with carbon fixation, where ribulose-1,5-bisphosphate carboxylase/oxygenase (), the most abundant enzyme on Earth, catalyzes the addition of CO₂ to ribulose 1,5-bisphosphate (RuBP), forming two molecules of 3-phosphoglycerate (3-PGA); this reaction has ΔG°' = -51.9 kJ/mol and is the primary CO₂-fixing step in plants. In the reduction phase, ATP phosphorylates 3-PGA to 1,3-bisphosphoglycerate, and NADPH reduces it to G3P, yielding one net G3P per three CO₂ fixed after consuming 9 ATP and 6 NADPH. The regeneration phase uses the remaining G3P and additional ATP to reform RuBP through a series of rearrangements involving enzymes like phosphoribulokinase. For every three turns of the cycle, one G3P exits for carbohydrate synthesis. C3 photosynthesis, dominant in most , directly fixes CO₂ via in mesophyll cells but suffers from when O₂ competes with CO₂, leading to the oxygenation of RuBP and a wasteful release of CO₂, reducing net fixation by 25-30% under high light and temperature. C4 photosynthesis, evolved in like and , enhances efficiency in hot, dry environments by spatially separating initial CO₂ fixation (using in mesophyll cells to form a C4 ) from the (in bundle sheath cells), concentrating CO₂ around to minimize , though at the cost of extra ATP. This adaptation allows C4 to achieve higher productivity in arid conditions. Overall, photosynthesis converts only about 1-2% of incident into in typical field conditions, limited by factors such as light absorption spectra, electron transport losses, and . Theoretical maximum efficiencies reach 4-6% under optimal CO₂ and temperature, but real-world losses, including excess heat dissipation and incomplete light utilization, constrain practical yields. Enhancing this efficiency remains a key target for improving global .

Membrane Mechanisms

Cotransport Systems

Cotransport systems, also known as secondary active transporters, facilitate the movement of ions and molecules across biological membranes by coupling the transport of one species to the downhill movement of another, harnessing pre-existing electrochemical gradients to drive otherwise unfavorable processes. These systems are classified into symporters, which transport two species in the same direction, and antiporters (or exchangers), which move them in opposite directions. A prominent example of a symporter is the sodium-glucose linked transporter (SGLT), particularly SGLT1 and SGLT2, which co-transport Na⁺ and glucose into cells, enabling glucose uptake against its concentration gradient in the intestinal epithelium and renal proximal tubule. The energy for cotransport derives from ion electrochemical gradients, typically established by primary active transporters such as the Na⁺/K⁺-ATPase, which hydrolyzes ATP to pump Na⁺ out of the cell, creating a low intracellular Na⁺ concentration and a negative that favors Na⁺ influx. This secondary active transport mechanism allows symporters and antiporters to power the uphill transport of substrates like nutrients or without direct ATP usage. In the , H⁺/amino acid symporters, such as those from the proton-coupled oligopeptide transporter (POT) family like PEPT1, facilitate the absorption of and peptides by coupling their influx to the proton gradient across the apical membrane of enterocytes. For export functions, ATP-binding cassette (ABC) transporters, while primarily ATP-driven, can integrate with membrane gradients in certain contexts to expel toxins and metabolites, contributing to cellular . The of cotransport is governed by the change for , where the overall ΔG must be negative for the coupled process to proceed spontaneously. For an or neutral solute, the is given by: \Delta G_{\text{transport}} = RT \ln \left( \frac{C_{\text{out}}}{C_{\text{in}}} \right) + zF\Delta\psi Here, R is the , T is , C_{\text{out}} and C_{\text{in}} are extracellular and intracellular concentrations, z is the charge, F is the , and \Delta\psi is the ; coupling to a driving with a sufficiently negative ΔG enables favorable uphill of the . Physiologically, cotransport systems are essential for nutrient uptake, such as glucose and in the intestine and , and for via transporters like the Na⁺/K⁺/2Cl⁻ (NKCC). Defects in these systems underlie diseases; for instance, results from mutations in the SLC3A1 or SLC7A9 genes encoding the rBAT/b⁰,⁺AT heterodimeric transporter, impairing renal of cystine and dibasic , leading to cystine stone formation. These mechanisms often utilize gradients, including proton gradients generated by chemiosmotic processes, to sustain cellular demands.

Chemiosmotic Theory

The chemiosmotic theory was proposed by Peter Mitchell in 1961 as an alternative to the prevailing chemical hypothesis, which posited the existence of high-energy chemical intermediates to couple electron transport to ATP phosphorylation in oxidative and photosynthetic processes. Mitchell's hypothesis suggested that the energy from reactions is instead stored as an electrochemical proton gradient across a coupling membrane, directly driving ATP synthesis without soluble intermediates. This revolutionary idea faced significant initial resistance but gained acceptance through experimental validation, earning Mitchell the in 1978. At its core, the theory describes how proton translocation establishes a proton motive force (PMF), the primary energy for ATP , quantified as \Delta p = \Delta \psi - \frac{2.3RT}{F} \Delta \mathrm{pH}, where \Delta \psi is the electrical , \Delta \mathrm{pH} is the transmembrane pH , R is the , T is the absolute temperature, and F is the . This PMF arises from the vectorial pumping of protons across the during transport in respiratory chains or photosynthetic flow. The components involve redox-driven proton extrusion from the matrix or , creating a gradient that ATP synthase harnesses by permitting controlled proton influx to catalyze ATP formation from ADP and inorganic . Key evidence supporting the theory includes the effects of uncouplers like (DNP), which increase membrane permeability to protons, thereby dissipating the PMF and stimulating electron transport while abolishing ATP synthesis—demonstrating that the gradient, not direct chemical , is essential.90579-7) Experiments by Mitchell and colleagues further confirmed stoichiometric proton translocation linked to , with ratios matching observed efficiencies. The chemiosmotic framework applies universally to energy-transducing membranes in mitochondria, chloroplasts, and prokaryotes, where proton gradients power ATP synthesis across diverse organisms. In certain systems, such as bacterial plasma membranes, the process operates in reverse, with driving proton extrusion to generate PMF for secondary or motility.00228-3)

Binding Change Mechanism

The binding change mechanism, proposed by Paul D. Boyer, describes how converts the proton motive force into through rotational catalysis and conformational changes in its catalytic sites. This model, for which Boyer shared the 1997 , posits that ATP synthesis occurs not through direct energy input to form the ATP bond but via sequential alterations in nucleotide binding affinity at three catalytic sites on the . ATP synthase consists of two main domains: the membrane-embedded F0 portion, which includes a rotating c-ring of 8–15 c-subunits depending on the organism, and the peripheral F1 portion, a soluble hexameric complex of three α-subunits and three β-subunits that house the catalytic sites, connected by a central γ-subunit rotor and a peripheral . In the mechanism, proton translocation through the F0 domain drives of the c-ring and attached γ-shaft, which mechanically interacts with the F1 domain to induce conformational changes among the three β-subunits' catalytic sites: open (O), loose (L), and tight (T). As the γ-subunit rotates in 120° steps powered by the proton motive force, it sequentially alters the conformations, causing and inorganic (Pi) to bind loosely at one site, tighten to form ATP without net energy input for bond formation, and open to release the product at high affinity. A full 360° of the , typically driven by approximately 10 protons in bacterial systems (yielding about 3 ATP molecules), completes one across all three sites, with the energy from proton flow primarily used to alter binding affinities rather than directly synthesizing ATP. This rotary process ensures efficient energy coupling, with elastic deformations in the -stator linkage transmitting while buffering stress.55940-1/fulltext) The of this involve stepwise s observable at the single-molecule level, where each 120° turn correlates with ATP or , and the overall rate is modulated by proton flux and substrate availability. inhibits the process by binding to the F0 c-ring, blocking proton conduction and halting , which prevents ATP in mitochondria and . Variations in ATP synthase structure and function occur across organisms; for instance, eukaryotic mitochondrial enzymes feature an 8-c-subunit ring requiring about 2.7 protons per ATP, while bacterial versions often have 10–12 c-subunits, adjusting the H+/ATP stoichiometry. The binding change mechanism shares evolutionary and structural with the bacterial flagellar motor, where similar rotary elements driven by fluxes enable generation for , highlighting a conserved of -powered in bioenergetics.

System Integration

Energy Balance

In bioenergetic pathways, energy balance refers to the stoichiometric accounting of energy inputs, such as from oxidation, against outputs like ATP production, dissipation, and biosynthetic demands. The complete oxidation of one molecule of glucose in eukaryotic cells theoretically yields 30 to 32 ATP molecules, accounting for the proton motive force across the mitochondrial membrane and inefficiencies in shuttle systems for cytoplasmic NADH. Earlier estimates suggested up to 38 ATP, but modern assessments adjust downward due to partial uncoupling and alternative uses of the proton gradient. These yields represent the net gain after substrate-level phosphorylations in and the , with the majority derived from . Significant portions of the chemical energy from glucose oxidation are lost as heat or through membrane leaks, reducing overall efficiency. In cellular respiration, approximately 60% of the free energy is dissipated as heat, primarily during electron transport and proton translocation, while the remainder is captured in ATP. Proton leaks across the further diminish ATP synthesis by allowing protons to re-enter without driving , contributing to basal metabolic heat production. These losses maintain mitochondrial integrity and cellular but impose thermodynamic constraints on . Bomb calorimetry provides a for total content by substrates in excess oxygen to measure heat release, yielding about 686 kcal/mol for glucose, which represents the gross change. In contrast, physiological efficiency in biological systems is far lower, capturing only around 40% of this in usable forms like ATP, with the rest lost to and non-productive processes. This discrepancy highlights how bioenergetic pathways prioritize controlled, stepwise energy release over maximal , enabling coupling to anabolic reactions. Balancing catabolic energy generation with anabolic demands is critical, particularly in growing cells where a substantial fraction of ATP is diverted to . In rapidly dividing bacterial cells, protein synthesis alone can consume up to 50% of total , with additional allocation to and formation pushing biosynthetic demands higher, often exceeding 75% of catabolic output in nutrient-rich conditions. This allocation shifts dynamically; in steady-state non-growing cells, more supports , whereas phases favor production, optimizing resource use under varying environmental pressures. Isotope tracing techniques enable precise measurement of energy es in bioenergetic systems by tracking labeled atoms through metabolic pathways. The use of ¹³C-labeled glucose or allows quantification of carbon into the and via , revealing steady-state rates of ATP production and substrate utilization. Similarly, ¹⁸O tracing, often via ¹⁸O₂ exposure, monitors oxygen incorporation into metabolic water and ATP phosphates, providing insights into efficiency and proton handling without isotopic dilution from environmental sources. Steady-state models integrate these data to compute balances, distinguishing active energy pathways from leaks or alternative routes. Disruptions to energy balance, such as in mitochondrial diseases, severely impair ATP yield by compromising respiratory chain complexes. Mutations in genes encoding electron transport components, like those in complex I or , can reduce efficiency by 50% or more, leading to systemic energy deficits and reliance on . exacerbates this by limiting oxygen availability, which diminishes ATP production through inhibited electron transport and increased lactate fermentation, often halving cellular energy output within minutes. These conditions underscore the fragility of bioenergetic , where even partial failures cascade into metabolic imbalances.02131-2/fulltext)

Efficiency and Regulation

Bioenergetic systems achieve high efficiency through precise capture of available free energy, quantified thermodynamically as the ratio of captured Gibbs free energy (\Delta G_{\text{captured}}) to the total available (\Delta G_{\text{available}}), often reaching 60-70% in mitochondrial oxidative phosphorylation under optimal conditions. This efficiency is reflected in the respiratory control ratio (RCR), defined as the ratio of ADP-stimulated (state 3) to resting (state 4) oxygen consumption rates in isolated mitochondria, with values exceeding 6 indicating healthy, tightly coupled function. Deviations below this threshold signal inefficiencies, such as proton leaks that reduce overall energy balance. Regulation of bioenergetics occurs via allosteric mechanisms, where adenine nucleotides modulate key enzymes; for instance, allosterically activates , enhancing electron transport and ATP synthesis in response to energy demand. Hormonal signals further integrate systemic control, as insulin promotes in muscle and adipose tissues by translocating transporters to the plasma membrane, thereby fueling glycolytic and oxidative pathways. Cells adapt bioenergetic fluxes to environmental stresses through transcription factors like hypoxia-inducible factor (HIF-1α), which under low oxygen conditions upregulates glycolytic enzymes such as and , shifting metabolism from to to sustain ATP production. In contrast, uncoupling protein 1 (UCP1) in dissipates the proton gradient across the , diverting energy from ATP synthesis to heat generation for non-shivering , particularly during cold exposure. Emerging research highlights quantum coherence in photosynthetic light-harvesting complexes, where wavelike delocalization enables near-unity in energy transfer, as demonstrated in studies of bacterial reaction centers persisting into the 2020s. approaches enhance bioenergetic yields by engineering microbial pathways, such as optimizing enzymes in to boost photosynthetic CO₂ fixation rates by up to 25%. As of 2025, synthetic pathways have been engineered to function alongside the native in plants, potentially boosting carbon fixation further. Despite these optimizations, bioenergetic systems face limitations from (ROS), which arise primarily from leaks and induce , damaging proteins, lipids, and DNA to impair efficiency. Evolutionarily, trade-offs prioritize rapid energy flux over maximal efficiency, as seen in fast-twitch muscle fibers where high glycolytic speeds support burst activity at the cost of lower thermodynamic yields compared to oxidative fibers.

References

  1. [1]
    Bioenergetics and Life's Origins - PMC - PubMed Central
    Bioenergetics is central to our understanding of living systems, yet has attracted relatively little attention in origins of life research.
  2. [2]
    [PDF] Chapter 6 Topic: Bioenergetics Main concepts
    Main concepts: •Energy is defined as the capacity of a system to do work or the capacity to cause change, such as synthesizing molecules or moving objects.
  3. [3]
    [PDF] Bioenergetics
    The laws of bioenergetics can enable you to understand why these energy transfers occur. Muscle contraction: chemical energy to mechanical energy. Vitamin D ...
  4. [4]
    Bioenergetics and Metabolism: A Bench to Bedside Perspective - PMC
    Molecular biology is influenced by bioenergetics and cell energy metabolism. Cells obviously depend on energy and systems have developed to produce energy, ...
  5. [5]
    Bioenergetics - an overview | ScienceDirect Topics
    Bioenergetics is defined as the study of energy transduction in living organisms, encompassing the capture of light energy and its storage as chemical bond ...
  6. [6]
    Bioenergetics: The transformation of free energy in living systems
    Jul 16, 2015 · Bioenergetics: The transformation of free energy in living systems ... Your browser can't play this video. ... Visit us (http://www.khanacademy.org/science/ ...
  7. [7]
    ATP cycle and reaction coupling | Energy (article) - Khan Academy
    Adenosine triphosphate, or ATP, is a small, relatively simple molecule. It can be thought of as the main energy currency of cells, much as money is the main ...Missing: central | Show results with:central
  8. [8]
  9. [9]
    The dawn of muscle energetics: contraction before and after ...
    Dec 1, 2023 · Muscle contraction was thought to be fueled by a chemical reaction. The dawn of muscle energetics began in the early twentieth century when Otto ...
  10. [10]
    Hallmarks of a new era in mitochondrial biochemistry - PMC
    1937—Hans Krebs publishes a paper on the Krebs cycle (Krebs and Johnson 1937) ... Peter Mitchell finally solved the mystery in 1961 with his radical new ...
  11. [11]
    Peter Mitchell – Biographical - NobelPrize.org
    Sir Hans Krebs Lecture and Medal of the Federation of European Biochemical Societies, 1978. Honorary Degree of Doctor of Science, University of Chicago, 1978 ...
  12. [12]
    Chapter 11: Bioenergetics: The Role of ATP - AccessPharmacy
    BIOMEDICAL IMPORTANCE. ++. Bioenergetics, or biochemical thermodynamics, is the study of the energy changes accompanying biochemical reactions. Biologic systems ...Missing: definition | Show results with:definition
  13. [13]
    Bioenergetic medicine - PMC - PubMed Central
    Bioenergetic medicine involves manipulating a bioenergetic pathway to increase or decrease its associated flux. This can affect cell energy stores as well as ...Missing: scope interdisciplinary applications
  14. [14]
    [PDF] Thermodynamics and Biological Systems - Yaşar Demirel
    This chapter explores the involvement of thermodynamics in living systems. Thermodynamics is an exact science of energy analysis with its first and second laws.<|control11|><|separator|>
  15. [15]
    [PDF] Biological Thermodynamics - Casegroup
    The free energy must be negative for a reaction to be spontaneous! Page 17. The Gibbs free energy (ΔG). Biological Thermodynamics.
  16. [16]
    Energy and enzymes | Biological Principles
    Gibbs Free Energy. Gibbs free energy is a measure of the amount of work that is potentially obtainable. Instead of absolute quantities, what is usually ...
  17. [17]
    [PDF] Chem 352 - Lecture 3 The Energetics of Life
    The First Law of Thermodynamics states that the total energy of an isolated system is a constant. • Since the universe is an isolated system, this means that.
  18. [18]
  19. [19]
    The Theory of Open Systems in Physics and Biology - jstor
    the characteristic state of the living organism is that of an open system. A system is closed if no material enters or leaves it; it is open if.
  20. [20]
    CHEM 245 - Bioenergetics:
    The free energy change for biochemical standard state - symbolized as ΔG°′ - depends on a standard state, or a set of standard conditions.<|control11|><|separator|>
  21. [21]
    [PDF] Biochemistry II (BI/CH 422 & BI/CH 622) Chemical Reactivity
    Energetics of Some Chemical Reactions:​​ Isomerization reactions have smaller free-energy changes. Complete oxidation of reduced compounds (redox reactions) is ...
  22. [22]
    Bioenergetics of early life: Coupling of reaction networks and ... - NIH
    Jul 21, 2022 · ... endergonic reactions and pair them with exergonic processes. Enzymes facilitate this process by keeping all substrates and energy‐rich ...
  23. [23]
    [PDF] lehninger-ch13_small.pdf
    FIGURE 13-1 Chemical basis for the large free-energy change asso- ciated with ATP hydrolysis. ... The standard free energy of hydrolysis of ATP is. -30.5 kJ/mol.
  24. [24]
    [PDF] Cellular Respiration & Metabolism Metabolism Coupled Reactions
    Exergonic reaction: reaction releases energy • Endergonic reaction: reaction requires energy • Coupled bioenergetic reactions: the energy released by the ...
  25. [25]
    4.1 Energy and Metabolism – Concepts of Zoology – Hawaiʻi Edition
    Reactions that have a negative change in free energy and consequently release free energy are called exergonic reactions. Think: exergonic means energy is ...
  26. [26]
    [PDF] Redox Chemistry Handout | MIT Department of Biology
    Redox reactions involve the transfer of electrons (usually abbreviated e−) from one molecule to the other. Reduction is when a molecule gains electrons.
  27. [27]
    Nernst Equation - EdTech Books
    ... equation relating free energy change to cell potential yields the Nernst equation: − n F E cell = − n F E cell ° + R T ln Q − n F E cell = − n F E cell ° + R T ...
  28. [28]
    ATP and Oxidative Phosphorylation Reactions - csbsju
    Apr 15, 2016 · This chapter will discuss the properties that make ATP so useful biologically, and how exergonic biological oxidation reactions are coupled to the synthesis of ...
  29. [29]
  30. [30]
    ROS Function in Redox Signaling and Oxidative Stress - PMC
    Reactive oxygen species (ROS) are byproducts of aerobic metabolism. ROS include the superoxide anion (O2−), hydrogen peroxide (H2O2), and hydroxyl ...
  31. [31]
    Catalysis and the Use of Energy by Cells - NCBI - NIH
    How reaction coupling is used to drive energetically unfavorable reactions. ... A mechanical model illustrating the principle of coupled chemical reactions.
  32. [32]
    The Essence of ATP Coupling - PMC - NIH
    The traditional explanation of ATP coupling is based on the raising of the equilibrium constants of the biochemical reactions.
  33. [33]
    Uncoupling Protein 1 of Brown Adipocytes, the Only Uncoupler - NIH
    The recent renewal of the interest in human brown adipose tissue makes UCP1 as a potential target for strategies of treatment of metabolic disorders. Keywords: ...
  34. [34]
    How Cells Obtain Energy from Food - NCBI - NIH
    Figure 2-75 compares the high-energy phosphoanhydride bonds in ATP with other phosphate bonds, several of which are generated during glycolysis.
  35. [35]
    [PDF] Phosphate Transfer Potentials ATP can make various types of ...
    +30.5 kJ/mol. ======================================= PEP + ADP → Pyruvate + ... Large free energy change typical of kinase due to the phosphoryl transfer ...
  36. [36]
    Respiration, chemiosmosis and oxidative phosphorylation
    Cells can make ATP in either of two ways: either by substrate-level phosphorylation of ADP or by oxidative phosphorylation of ADP.
  37. [37]
    [PDF] Note that ∆G0 for hydrolysis of ATP is given in the footnote of Table ...
    The standard Gibbs free energy of hydrolysis of phosphocreatine at 25°C and pH 7 is -43.1 kJ mol-1 : phosphocreatine + H2O → creatine + phosphate ... ΔG ...
  38. [38]
    A prebiotic basis for ATP as the universal energy currency - PMC
    Oct 4, 2022 · There is nothing particularly special about the “high-energy” phosphoanhydride bonds in ATP. Rather, its ability to drive phosphorylation or ...
  39. [39]
    How Enzymes Handle the Energy Derived from the Cleavage of ...
    These included phosphocreatine, phosphoenolpyruvate, and the γ- and β-phosphate phosphoanhydride bonds of ATP.
  40. [40]
    Lecture 19 - houghton biology site
    C6H12O6 + 6 O2 ---> 6 CO2 + 6 H2O + energy (heat and light). The same equation applies for the biological, metabolic use of glucose. This process, however, has ...
  41. [41]
    [PDF] Cellular Respiration: Harvesting Chemical Energy
    Oxidative phosphorylation produces almost 90% of the ATP generated by respiration. Some ATP is also formed directly during glycolysis and the citric acid cycle ...
  42. [42]
    Glycolysis
    The overall reaction for glycolysis is: glucose (6C) + 2 NAD+ 2 ADP +2 inorganic phosphates (Pi) yields 2 pyruvate (3C) + 2 NADH + 2 H+ + 2 net ATP. 5.
  43. [43]
    Energy-Releasing Pathways - Concepts of Biology
    Aug 20, 2004 · Pyruvate Oxidation. Pyruvate from glycolysis is shuttled into the mitochondrion. As it enters, both of the pyruvate are oxidized (another redox ...Missing: stages | Show results with:stages<|control11|><|separator|>
  44. [44]
    The Citric Acid (Krebs) Cycle
    The overall reaction for the citric acid cycle is: 2 acetyl groups + 6 NAD+ + 2 FAD + 2 ADP + 2 Pi yields 4 CO2 + 6 NADH + 6 H+ + 2 FADH2 + 2 ATP. 5. The citric ...Missing: FADH2 | Show results with:FADH2
  45. [45]
    CHEM 440 - Electron transport chain
    Dec 23, 2016 · Two mobile electron carriers shuttle electrons between these complexes, coenzyme Q (ubiquinone) and cytochrome c. Three of the four complexes ...
  46. [46]
    Theoretical ATP Yield for Aerobic Respiration
    The theoretical maximum yield of ATP generated per glucose is 36 to 38, depending on how the 2 NADH generated in the cytoplasm during glycolysis enter the ...Missing: actual | Show results with:actual
  47. [47]
    [PDF] A New Approach to Calculating Theoretical P/O Ratios - EliScholar
    Nov 15, 2006 · Schoolworth, 1979), and the overall yield would be 31 ATP per glucose. While more attention to detail is granted than most would give (i.e. ...Missing: actual | Show results with:actual
  48. [48]
    Fermentation, mitochondria, and regulation - Biological Principles
    These reactions produce ethanol in yeast and lactic acid in mammalian cells, specifically in muscle cells under oxygen deficit and in most tumor cells (see ...
  49. [49]
    Photosynthesis - PMC - PubMed Central - NIH
    Oct 26, 2016 · The light reactions occur in the chloroplast thylakoid membrane and involve the splitting of water into oxygen, protons and electrons. The ...Missing: seminal | Show results with:seminal
  50. [50]
    [PDF] Evolution of the Z-scheme of photosynthesis: a perspective
    3 A 1964–1965 scheme of two light reactions (I and II) and two photosystems (PSI and PSII), which included basic reactions of Calvin-Benson cycle, and of ...Missing: seminal | Show results with:seminal
  51. [51]
    Light induced oxidative water splitting in photosynthesis: Energetics ...
    This mini-review briefly summarizes our state of knowledge on energetics, kinetics and mechanism of oxidative water splitting.Missing: paper | Show results with:paper<|control11|><|separator|>
  52. [52]
    [PDF] Melvin Calvin - Nobel Lecture
    Calvin and A. A. Benson, The Path of Carbon in Photosynthesis, Science, 107 ... Calvin, The Photosynthetic Cycle: CO, dependent Tran- sients, J. Am. Chem ...
  53. [53]
    The Path from C3 to C4 Photosynthesis - PMC - NIH
    The C4 photosynthetic carbon cycle is an elaborated addition to the C3 photosynthetic pathway. It evolved as an adaptation to high light intensities, high ...
  54. [54]
    improving photosynthetic efficiency for sustainable crop production
    This review explores the evidence to date that manipulation of the Calvin–Benson cycle, photorespiration, and electron transport can lead to improvement in ...
  55. [55]
    [PDF] Solar energy conversion efficiencies in photosynthesis - eScholarship
    Jul 1, 2009 · solar energy conversion efficiency. Thus, the best-case solar-to- biomass energy conversion efficiency was estimated to be 8–10%. (Fig. 3 ...
  56. [56]
    What is the maximum efficiency with which photosynthesis ... - PubMed
    The maximum conversion efficiency of solar energy to biomass is 4.6% for C3 photosynthesis at 30 degrees C and today's 380 ppm atmospheric [CO2], but 6% for C4 ...
  57. [57]
    General principles of secondary active transporter function - PMC
    Secondary active transporters couple the spontaneous influx of a “driving” ion such as Na + or H + to the flux of the substrate.
  58. [58]
    Uniporters, Symporters and Antiporters
    Nov 1, 1994 · Simple symport systems are often just referred to as cotransporters, whereas simple antiport systems are often referred to as exchangers.Abstract · Acknowledgements · Welcoming Emily Baird And...
  59. [59]
    Na+, K+-ATPase: Ubiquitous Multifunctional Transmembrane ...
    The transport of 3 Na+ for 2 K+ across the membrane, through the means of the sodium pump, maintains transmembrane gradients for the ions and produces a ...
  60. [60]
    H+-coupled nutrient, micronutrient and drug transporters in the ... - NIH
    These transporters are responsible for the initial stage in absorption of a remarkable variety of essential and non-essential nutrients and micronutrients.
  61. [61]
    ABC transporters: The power to change - PMC - PubMed Central
    A given ABC transporter may function as either an importer or an exporter, moving molecules in or out of cells, respectively, but no example is known of an ABC ...
  62. [62]
    The structural basis of secondary active transport mechanisms
    In a symport mechanism, the transport cycle is defined by the inclusion of binding and release steps of the co-substrate, which would be similar to those of the ...4. Coupling In Secondary... · 4.2. In Antiport, Substrate... · 4.3. Antiport Of Charged...
  63. [63]
    Na+‐K+‐2Cl− Cotransporter (NKCC) Physiological Function in ...
    Apr 1, 2018 · NKCC is therefore viewed as a secondary active transport mechanism. Due to the strict coupling of substrates for this type of carrier, the ...
  64. [64]
    Cystinuria - StatPearls - NCBI Bookshelf - NIH
    Cystine stones are due to an inherited defect in the transport of the amino acid cystine, leading to excessive excretion in the kidney, causing cystinuria.
  65. [65]
    [PDF] Peter Mitchell - Nobel Lecture
    My original proposal (Mitchell, 1961b, c) for the protonmotive. ATPase, reproduced in Fig. 10, corresponds to the group-translocation system of Fig. 9 B ...
  66. [66]
    [PDF] Paul D. Boyer - Nobel Lecture
    During net ATP synthesis the three catalytic sites on the enzyme, acting in sequence, first bind ADP and phosphate, then undergo a conformational change so as ...Missing: original | Show results with:original
  67. [67]
    Subunit composition of ATP synthase | MRC Mitochondrial Biology ...
    The F1 catalytic domain of the mitochondrial enzyme is a globular assembly of five different proteins, α, β, γ, δ and ε with the stoichiometry 3:3:1:1:1 [1].
  68. [68]
    Structure of a bacterial ATP synthase - eLife
    Feb 6, 2019 · Bacterial ATP synthase has three αβ pairs in the F1 region and 10 c-subunits in the F0 region, with a rotor subcomplex and a peripheral stalk.<|control11|><|separator|>
  69. [69]
    The Rotary Mechanism of the ATP Synthase - PMC - PubMed Central
    This review addresses the role of rotation in catalysis of ATP synthesis/hydrolysis and the transport of protons or sodium.Fig. 1 · Rotation In The Catalytic... · Partial Reaction And...
  70. [70]
    The six steps of the complete F1-ATPase rotary catalytic cycle - Nature
    Aug 3, 2021 · F1Fo ATP synthase interchanges phosphate transfer energy and proton motive force via a rotary catalysis mechanism.
  71. [71]
    ATP synthase: From sequence to ring size to the P/O ratio - PNAS
    Sep 21, 2010 · The P/O ratio is the number of ATP molecules synthesized per pair of electrons, fundamental for understanding ATP yield from cell fuels.<|control11|><|separator|>
  72. [72]
    Oligomycin frames a common drug-binding site in the ATP synthase
    Aug 6, 2012 · Oligomycin has been recognized as a potent inhibitor of the mitochondrial ATP synthase since 1958 when it was reported by Henry Lardy et al. (1) ...
  73. [73]
    Number of protons that are required by the F - BioNumbers
    The bioenergetic cost of the enzyme making an ATP is 2.7 protons in vertebrates and probably in invertebrates also.
  74. [74]
    ATP synthase: Evolution, energetics, and membrane interactions
    Sep 23, 2020 · In this review, we present an overview of several structural and functional features of the F-type ATPases that vary across taxa and are purported to be ...
  75. [75]
    Flagellar export apparatus and ATP synthetase: Homology ...
    May 16, 2021 · We report evidence further supporting homology between proteins in the F1FO-ATP synthetase and the bacterial flagellar motor (BFM).
  76. [76]
    Oxidative phosphorylation | Biology (article) - Khan Academy
    However, most current sources estimate that the maximum ATP yield for a molecule of glucose is around 30-32 ATP 2 , 3 , 4 ‍ . This range is lower than previous ...
  77. [77]
    ATP synthesis and storage - PMC - PubMed Central - NIH
    The oxidation process results in free energy production ... In the first process, when glucose is converted into pyruvate, the amount of ATP produced is low.
  78. [78]
    [PDF] Glucose and Fat Oxidation: Bomb Calorimeter Be Damned
    In terms of anaerobic and aerobic energy loss, glycolytic ATP production is where respiration and bomb calorimetry (combustion) part ways. In the 1970s ...<|separator|>
  79. [79]
    Oxidative phosphorylation: regulation and role in cellular and tissue ...
    This is a high, about 80%, overall coupling efficiency. The about 20% loss is as heat released when oxygen is reduced to the bound peroxide intermediate of ...
  80. [80]
    CALCULATION OF THE ENERGY CONTENT OF FOODS
    The total combustible energy content (or theoretical maximum energy content) of a food can be measured using bomb calorimetry. Not all combustible energy is ...
  81. [81]
    [PDF] CONCEPTS ON EFFICIENCY IN BIOLOGICAL CALORIMETRY AND ...
    On the one hand, the enthalpy efficiency is relevant in the context of biological calorimetry in relation to uncoupling and the integration of aerobic and.
  82. [82]
    Energy allocation theory for bacterial growth control in and out of ...
    In the case of bacteria, the rate of energy used for growth is proportional to the growth rate of the cell. The hypothesis for maximizing the growth power is ...
  83. [83]
    Evaluation of 13C isotopic tracers for metabolic flux analysis in ...
    We computationally evaluated specifically labeled 13 C glucose and glutamine tracers for their ability to precisely and accurately estimate fluxes in central ...
  84. [84]
    Isotopic labeling of metabolic water with 18 O2 - PubMed
    Mar 30, 2023 · As oxygen gas is reduced during respiration, H2 18 O is produced. The rate of H2 18 O production and the synthesis of biomolecules that ...
  85. [85]
    A guide to 13 C metabolic flux analysis for the cancer biologist - Nature
    Apr 16, 2018 · The isotopic tracing strategies and 13C-MFA methods reviewed here present powerful tools for elucidating metabolic flux rewiring in cancer cells ...
  86. [86]
    Measurement of ATP production in mitochondrial disorders - PubMed
    Mitochondrial disease leads to impaired respiratory chain function and reduced ATP production. The aim of this study was to compare disturbances in ...
  87. [87]
    Hypoxia. 2. Hypoxia regulates cellular metabolism - PubMed Central
    Hypoxia diminishes ATP production in part by lowering the activity of the electron transport chain through activation of the transcription factor hypoxia- ...
  88. [88]
    Assessing mitochondrial dysfunction in cells - PMC - PubMed Central
    High RCR indicates good function, and low RCR usually indicates dysfunction. The utility of RCR rests on its complexity: virtually any change in oxidative ...
  89. [89]
    Bioenergetic profiles and respiratory control in mitochondrial ...
    Aug 1, 2025 · Bioenergetic profiling unravels the complexities of mitochondrial respiratory control on the basis of a consistent comparative database. 1 ...
  90. [90]
    Protein Phosphorylation and Prevention of Cytochrome Oxidase ...
    Jun 8, 2011 · We show that mitochondrial energy metabolism regulation by phosphorylation of COXIV-1 is coupled with prevention of COX allosteric inhibition by ATP.
  91. [91]
    Insulin regulation of gluconeogenesis - PMC - PubMed Central - NIH
    Here, we review some of the molecular mechanisms through which insulin modulates hepatic gluconeogenesis, thus controlling glucose production by the liver.Missing: bioenergetics | Show results with:bioenergetics
  92. [92]
    Regulation of glycolysis by the hypoxia-inducible factor (HIF)
    Under hypoxia, cells shift to increased glycolysis, regulated by HIF-1α, which induces glycolytic enzymes. This is key for maintaining ATP levels.
  93. [93]
    Beneficial and Detrimental Effects of Reactive Oxygen Species on ...
    Decreased resistance to oxidative stress may be an indication of elevated baseline levels of ROS; however, increased levels of ROS can also cause upregulation ...
  94. [94]
    Power and Efficiency in Living Systems - MDPI
    A power/efficiency approach promises to provide much insight into the tempo, functioning, and evolution of living systems, small and large.