Fact-checked by Grok 2 weeks ago

Exergonic reaction

An is a in which the change in (ΔG) is negative (ΔG < 0), meaning the free energy of the products is lower than that of the reactants, resulting in a net release of free energy to the surroundings. These reactions are spontaneous, occurring without the need for continuous external energy input, though they may still face an activation energy barrier that affects their rate. The change is calculated using the equation ΔG = ΔH - TΔS, where ΔH is the enthalpy change, T is the absolute temperature, and ΔS is the entropy change; a negative ΔG arises from favorable enthalpy (exothermic) or entropy (increased disorder) contributions, or both. In biological systems, exergonic reactions play a crucial role in metabolism, powering endergonic processes through coupling, such as the hydrolysis of adenosine triphosphate (ATP) where ΔG ≈ -7.3 kcal/mol under standard conditions. Catabolic pathways, like the breakdown of glucose in cellular respiration, exemplify exergonic reactions that release energy to drive anabolic synthesis. Unlike endergonic reactions (ΔG > 0), which require energy input and are non-spontaneous, exergonic processes contribute to the overall directionality of metabolic networks, ensuring efficient energy flow in living organisms. While spontaneity indicates thermodynamic favorability, the actual speed of an exergonic reaction depends on kinetic factors, such as enzyme catalysis in biology, and not all such reactions occur rapidly without facilitation.

Fundamentals

Definition

An exergonic reaction is a chemical process or reaction that releases to the surroundings, rendering it energetically favorable. This release occurs as the system transitions to a lower state, with the excess becoming available for other processes. The defining thermodynamic criterion for an exergonic reaction is a negative change in (ΔG < 0), evaluated under standard conditions of constant temperature and pressure. Gibbs free energy encapsulates the balance between enthalpy and entropy, determining the reaction's directionality without implying speed or mechanism. The term "exergonic" originates from the Greek prefix "ex-" meaning "out" and "ergon" meaning "work," reflecting the outward release of energy capable of performing work; it was first proposed in 1940 by chemist in the context of thermodynamic processes. Although analogous to that liberate heat, exergonic processes specifically emphasize the liberation of free energy rather than thermal energy alone, distinguishing them from mere heat-releasing events.

Thermodynamic Basis

The thermodynamic basis of exergonic reactions is rooted in the concept of Gibbs free energy, a state function that quantifies the maximum reversible work available from a system at constant temperature and pressure. The change in Gibbs free energy, denoted as \Delta G, for a process is given by the equation \Delta G = \Delta H - T \Delta S, where \Delta H is the change in enthalpy, T is the absolute temperature in Kelvin, and \Delta S is the change in entropy. For an exergonic reaction, \Delta G < 0, indicating that the process releases free energy and is thermodynamically favorable. This negative \Delta G arises when the enthalpy decrease outweighs the entropy term or when a positive \Delta S contributes sufficiently at higher temperatures. The full expression for the free energy change in a reaction is \Delta G = G_{\text{products}} - G_{\text{reactants}}, where a value less than zero signifies that the products possess lower free energy than the reactants, driving the reaction toward completion under the specified conditions. This criterion applies specifically to systems maintained at constant temperature and pressure, which are common in chemical and biological contexts. For systems at constant temperature and volume, the analogous quantity is the , defined as A = U - TS, where U is the internal energy; spontaneity is indicated by \Delta A < 0. Exergonic processes align with the second law of thermodynamics, which states that spontaneous changes increase the total entropy of the universe. At constant temperature and pressure, the Gibbs free energy change relates directly to this through the derivation \Delta G = -T \Delta S_{\text{universe}}, where \Delta S_{\text{universe}} = \Delta S_{\text{system}} + \Delta S_{\text{surroundings}} and \Delta S_{\text{surroundings}} = -\Delta H / T; thus, \Delta G < 0 implies \Delta S_{\text{universe}} > 0. In standard thermodynamic notation, the Gibbs free energy change \Delta G^\circ is evaluated under standard conditions of 1 and 298 K (25°C), providing a reference for comparing reaction favorability across different systems.

Properties

Spontaneity

Exergonic reactions are characterized by a negative change in (\Delta G < 0), which indicates that they are thermodynamically spontaneous in the forward direction under constant temperature and pressure conditions. This spontaneity arises because the free energy of the products is lower than that of the reactants, allowing the reaction to proceed without external energy input once initiated. However, this thermodynamic favorability does not guarantee that the reaction will occur rapidly in practice. A key distinction exists between the thermodynamics and kinetics of exergonic reactions: while thermodynamics determines the direction and feasibility based on \Delta G, kinetics governs the rate at which the reaction proceeds. An exergonic reaction may be thermodynamically favorable yet kinetically hindered, meaning it could remain slow or stalled without an appropriate catalyst to facilitate the process. Catalysts accelerate such reactions by providing an alternative pathway that lowers the required energy input, but they do not alter the overall \Delta G value. The primary kinetic barrier in exergonic reactions is the activation energy (E_a), defined as the energy difference between the reactants and the high-energy transition state. Even though the reaction is spontaneous overall, molecules must acquire sufficient energy to reach this transition state, often through collisions or thermal agitation. Catalysts reduce E_a by stabilizing the transition state, thereby increasing the reaction rate without affecting the thermodynamic spontaneity. Under standard conditions, exergonic reactions drive toward equilibrium, but those with a large negative \Delta G (strong exergonicity) are effectively irreversible, as the reverse reaction becomes negligible due to the energetic favorability of the forward path. The magnitude of \Delta G thus determines the extent to which the reaction proceeds to completion before equilibrium is approached. Temperature influences the spontaneity of exergonic reactions through its effect on the term in the Gibbs free energy equation, \Delta G = \Delta H - T\Delta S, where an increase in temperature amplifies the T\Delta S contribution. For reactions where increases (\Delta S > 0), higher temperatures can enhance spontaneity by making \Delta G more negative, potentially shifting borderline cases toward exergonic behavior. Conversely, if decreases, elevated temperatures might reduce spontaneity, though most exergonic processes remain favorable across physiological ranges.

Equilibrium Implications

Exergonic reactions, characterized by a negative standard change (ΔG° < 0), exhibit an equilibrium constant K greater than 1, indicating that the position of strongly favors the formation of products over reactants. This relationship is quantitatively described by the equation ΔG° = -RT ln K, where R is the gas constant (8.314 J/mol·K) and T is the absolute temperature in Kelvin; the negative ΔG° thus corresponds to a positive ln K, confirming K > 1 and a thermodynamic drive toward product accumulation. In strongly exergonic reactions, where ΔG° is significantly negative (e.g., large magnitudes leading to K ≫ 1), the position shifts far toward the products, resulting in very low concentrations of reactants at . This product dominance arises because the minimum occurs at a state enriched in products, minimizing the system's overall Gibbs energy. The equilibria of exergonic reactions are subject to , which predicts that perturbations such as changes in concentration, , or will elicit responses that counteract the disturbance and restore . For instance, increasing reactant concentrations shifts the equilibrium toward products to consume the excess, while changes affect the position based on the reaction's : exothermic exergonic reactions shift toward reactants upon heating to absorb the added energy. Although exergonic reactions proceed spontaneously in the forward direction, they remain reversible at the molecular level, with both forward and reverse microscopic rates existing; however, the net flux is overwhelmingly forward due to the energetic favorability, ensuring minimal back-conversion under standard conditions. Under non-standard conditions, the direction of spontaneity is determined by comparing the reaction quotient (the ratio of product to reactant activities raised to stoichiometric powers) to : if < , the forward reaction is spontaneous, driving the system toward equilibrium by consuming excess reactants or producing more products as needed.

Examples

Chemical Reactions

Exergonic reactions are prevalent in inorganic and organic chemistry, where they drive processes ranging from energy release in combustion to spontaneous redox transformations. These reactions exhibit a negative change in (ΔG < 0), making them thermodynamically favorable under standard conditions. Representative examples illustrate how bond formation, electron transfer, and structural rearrangements contribute to exergonicity in abiotic systems. One classic example is the combustion of hydrocarbons, such as methane, which releases substantial energy and is central to energy production. The reaction is: \ce{CH4(g) + 2O2(g) -> CO2(g) + 2H2O(l)} The standard change for this process is ΔG° ≈ -818 kJ/mol, indicating a highly exergonic reaction primarily driven by the large negative change (ΔH° ≈ -890 kJ/mol), which outweighs the unfavorable decrease (ΔS° ≈ -243 J/mol·K) associated with the net reduction in gas moles from three to one. Acid-base neutralizations between strong acids and bases also exemplify exergonic processes through the formation of stable ionic bonds and . For instance, the reaction between and proceeds as: \ce{HCl(aq) + NaOH(aq) -> NaCl(aq) + H2O(l)} This yields ΔG° ≈ -80 kJ/mol, negative due to the favorable of hydrating the ions and forming the strong O-H bonds in , with the process being spontaneous and exothermic (ΔH° ≈ -57 kJ/mol). Redox reactions often proceed exergonically via spontaneous , as seen in the displacement between and copper(II) ions: \ce{Zn(s) + Cu^{2+}(aq) -> Zn^{2+}(aq) + Cu(s)} Here, ΔG° ≈ -213 kJ/mol, reflecting the higher of Cu²⁺/Cu (E° = +0.34 V) compared to Zn²⁺/Zn (E° = -0.76 V), resulting in a positive potential (E°_cell = 1.10 V) that drives the forward without external input. Disproportionation reactions, where a single species is simultaneously oxidized and reduced, can also be exergonic, though often kinetically hindered. The decomposition of hydrogen peroxide serves as an example: \ce{2H2O2(l) -> 2H2O(l) + O2(g)} This reaction has ΔG° ≈ -233 kJ/mol, making it thermodynamically favorable due to the stability of the products, but it proceeds slowly without a catalyst owing to a high activation energy barrier. In industrial applications, exergonic reactions power energy technologies like fuel cells, where the spontaneity (ΔG < 0) of oxidation-reduction processes converts chemical energy directly into electrical energy. For example, in hydrogen-oxygen fuel cells, the overall reaction 2H₂(g) + O₂(g) → 2H₂O(l) has ΔG° ≈ -474 kJ/mol, enabling efficient electricity generation with minimal waste heat.

Biological Processes

Exergonic reactions are fundamental to biological energy metabolism, particularly in catabolic processes that release free energy to sustain cellular functions. Cellular respiration exemplifies this, where the complete oxidation of glucose follows the overall reaction C_6H_{12}O_6 + 6O_2 \rightarrow 6CO_2 + 6H_2O, with a standard free energy change \Delta G^{\circ\prime} \approx -2870 kJ/mol, making it highly spontaneous. This process is divided into stages, including glycolysis, which partially oxidizes glucose to pyruvate with a \Delta G^{\circ\prime} of approximately -85 kJ/mol, and the Krebs cycle (also known as the citric acid cycle), where acetyl-CoA is further oxidized, contributing the majority of the energy release through exergonic decarboxylation and dehydrogenation steps. These stages collectively harness the large negative \Delta G to generate reducing equivalents like NADH and FADH_2, which fuel subsequent energy capture. A key exergonic reaction in biology is the hydrolysis of (ATP), represented as ATP + H_2O \rightarrow ADP + P_i, with a standard free energy change \Delta G^{\circ\prime} \approx -30.5 kJ/mol; under typical cellular conditions, this value is more negative, around -57 kJ/mol, due to non-standard concentrations of reactants and products. This reaction provides immediate energy for diverse processes, such as active transport, muscle contraction, and biosynthesis, by releasing phosphate under physiological pH and ion concentrations. In oxidative phosphorylation, exergonic redox reactions in the (ETC) drive the process: electrons from and are transferred through protein complexes, creating a proton gradient across the inner mitochondrial membrane with stepwise \Delta G releases, such as the reduction of by NADH (potential difference of approximately +360 mV). These exergonic steps indirectly power via , though the gradient itself is an intermediate energy form. In anaerobic conditions, fermentation pathways serve as alternative exergonic routes for energy extraction. For instance, lactic acid fermentation converts glucose to two molecules of lactate via pyruvate, with an overall \Delta G^{\circ\prime} \approx -198 kJ/mol, yielding far less energy than aerobic respiration but allowing rapid ATP production without oxygen. This process regenerates NAD^+ essential for glycolysis continuation, as seen in muscle cells during intense exercise. Exergonic catabolism, including these pathways, holds evolutionary significance as the foundation for heterotrophic life, enabling organisms to capture and utilize energy from organic food sources through oxidation, a trait conserved across bacteria, archaea, and eukaryotes since early Earth. This catabolic efficiency supported the diversification of life by providing a reliable energy source independent of abiotic synthesis.

Relations to Other Concepts

Comparison with Endergonic Reactions

Endergonic reactions are the thermodynamic opposites of exergonic reactions, characterized by a positive change in Gibbs (ΔG > 0), which indicates that they require an input of to proceed and are non-spontaneous in the forward direction under standard conditions. In contrast, exergonic reactions have ΔG < 0, making them spontaneous and energy-releasing. This opposition extends to equilibrium behavior: exergonic reactions favor product formation with an equilibrium constant > 1, as derived from the relationship ΔG° = -T ln , where is the and T is temperature in ; conversely, endergonic reactions have < 1, favoring reactants, such that the reverse reaction is exergonic. Thus, the reverse of an endergonic process is exergonic and drives the system toward equilibrium, favoring the reactants in the forward direction. The interplay of (ΔH) and (ΔS) further distinguishes these reactions, governed by the equation ΔG = ΔH - TΔS. Exergonic reactions are often driven by exothermic enthalpy changes (ΔH < 0) and/or positive changes (ΔS > 0), releasing and increasing , while endergonic reactions typically involve endothermic enthalpy (ΔH > 0) or negative changes (ΔS < 0), absorbing or decreasing . modulates this balance, as the TΔS term amplifies entropy's influence at higher temperatures. In practical terms, exergonic reactions underpin catabolic processes that release for cellular work, whereas endergonic reactions support anabolic processes that build complex molecules from simpler ones, necessitating to occur. Common misconceptions include assuming all exergonic reactions proceed rapidly, overlooking that determines feasibility while governs rate via barriers; similarly, endergonic reactions are not inherently impossible but require external input to shift the net ΔG negative.

Coupling in Metabolic Pathways

In metabolic pathways, exergonic reactions are frequently coupled to endergonic reactions to enable the progression of thermodynamically unfavorable processes, where the negative change in (ΔG) from the exergonic step compensates for the positive ΔG of the endergonic step, resulting in an overall negative ΔG for the coupled system. This coupling often occurs through shared chemical intermediates or enzymes that facilitate the transfer of without dissipation into uncoupled forms. For instance, the of a high- links the two reactions, ensuring that the released drives the or assembly required for cellular functions. A prime example of this mechanism is the role of (ATP) as the universal energy currency in cells, where its —an exergonic reaction with ΔG ≈ -30.5 kJ/mol under standard conditions—provides the energy to power endergonic processes such as , across membranes, and of macromolecules. In , for example, the sodium-potassium pump uses the energy from to move ions against their concentration gradients, a process that would otherwise be endergonic. This coupling is mediated by enzymes that form phosphorylated intermediates, transiently linking the phosphate group transfer from ATP to the target reaction. Specific metabolic pathways illustrate this integration vividly. In , the reversal of —an endergonic pathway overall—is driven by coupling to exergonic oxidative phosphorylation steps that generate ATP and NADH, bypassing irreversible glycolytic reactions through alternative enzymes like and . Similarly, in , the endergonic fixation of CO₂ in the is powered by ATP and NADPH produced in the , where exergonic electron flow from water to NADP⁺ creates a proton gradient for ATP synthesis. These examples highlight how coupling ensures directional flux in anabolic pathways. The efficiency of such coupling is not absolute, typically ranging from 40% to 60% in cellular systems due to inevitable heat loss and , though mitochondrial electron transport achieves up to 80-90% thermodynamic efficiency through tight coupling of reactions to proton translocation. Thermodynamic limits, governed by the second law, prevent 100% yield, as some must be dissipated to maintain irreversibility and directionality. In pathway integration, exergonic steps often serve as control points for feedback regulation; for example, the exergonic reaction in is allosterically inhibited by downstream ATP, modulating flux to prevent overproduction and align with cellular demands. This regulatory role ensures metabolic across interconnected pathways.

References

  1. [1]
    6.1: Endergonic and Exergonic Reactions - Biology LibreTexts
    Aug 23, 2021 · Exergonic reactions have lower Gibbs energy products and are spontaneous. Endergonic reactions have higher Gibbs energy products and are not ...<|control11|><|separator|>
  2. [2]
    Free energy | Endergonic vs exergonic reactions (article)
    Reactions that have a negative ∆G release free energy and are called exergonic reactions. (Handy mnemonic: EXergonic means energy is EXiting the system.) A ...
  3. [3]
    Gibbs Free Energy
    Reactions can also be classified as exergonic ( delta G < 0) or endergonic ( delta G > 0) on the basis of whether the free energy of the system decreases or ...
  4. [4]
    Energy and enzymes | Biological Principles
    If ΔG < 0, a chemical reaction is exergonic, releases free energy, and can progress spontaneously, with no input of additional energy. This negative ΔG does not ...
  5. [5]
    Basics of Metabolic Reactions - PMC - NIH
    Exergonic reactions have a negative change in the Gibbs free energy and, thus, release energy. In contrast, endergonic reactions have a positive change in the ...
  6. [6]
    Gibbs (Free) Energy - Chemistry LibreTexts
    Jan 29, 2023 · Gibbs free energy, denoted G , combines enthalpy and entropy into a single value. The change in free energy, ΔG , is equal to the sum of the ...
  7. [7]
    Gibbs Free Energy - an overview | ScienceDirect Topics
    When Δ G is negative, the reaction is exergonic, and releases energy. This includes exothermic reactions in which the entropy increases, or exothermic reactions ...
  8. [8]
  9. [9]
    18.5: Second Law and Gibbs Free Energy - Chemistry LibreTexts
    Jan 19, 2025 · The Second Law of Thermodynamics states that a spontaneous process occurs when there is an increase in the entropy of the universe.Introduction · Derivation of Gibbs Free Energy · Driving Chemical Reactions
  10. [10]
    Chemical Reactivity - MSU chemistry
    Although spontaneous transformations all have negative ΔGº s, not all exergonic processes are spontaneous, due to activation energy barriers to reaction.
  11. [11]
    7.6: Energetics and Kinetics - Chemistry LibreTexts
    Aug 6, 2019 · If examining Gibbs free energy, the reaction would be described instead as “exergonic” and would spontaneously proceed to the right (favors ...
  12. [12]
    [PDF] Klein 6 thermo-kinetics
    When we refer to the kinetics of a reaction, we refer to the difference in energy between the highest transition state and starting materials. If a reaction can ...
  13. [13]
    [PDF] Notes LECTURE 5 Biochemistry 9.17.04
    Sep 17, 2004 · The activation energy, Ea, is the energy difference between the reactants and the transition state. Thus, for the reaction to proceed, energy is ...<|separator|>
  14. [14]
    Potential, Kinetic, Free, and Activation Energy – Biology
    Exergonic and endergonic reactions result in changes in Gibbs free energy. Exergonic reactions release energy; endergonic reactions require energy to proceed.<|control11|><|separator|>
  15. [15]
    Regulation of glycolysis - CHEM 245
    Feb 23, 2019 · The magnitude of the ΔG for these steps makes them essentially irreversible. The most important point of control is at the reaction catalyzed ...Missing: delta | Show results with:delta
  16. [16]
    [PDF] CHAPTER 6 AN INTRODUCTION TO METABOLISM
    An exergonic reaction proceeds with a net release of free energy; ∆G is negative. ○ The magnitude of ∆G for an exergonic reaction is the maximum amount of work ...Missing: delta | Show results with:delta<|separator|>
  17. [17]
    Gibbs Free Energy
    For a spontaneous process at constant temperature and pressure, DG must be negative. In many cases, we can predict the sign of from the signs of DH and DS.Missing: exergonic | Show results with:exergonic
  18. [18]
    [PDF] Chemical equilibrium
    At constant temperature and pressure: If ΑrG < 0, the forward reaction is spontaneous – exergonic (work-producing). If ΑrG > 0, the reverse reaction is ...
  19. [19]
    [PDF] Foundations of Biochemistry
    Recall that if delta. G is < 0, the reaction is spontaneous (or Exergonic). If delta G is 0, the reaction is at equilibrium, and if delta G is > 0, the reaction ...
  20. [20]
    Chapter 6: Enzyme Principles and Biotechnological Applications
    If they did, the reaction would essentially stop (i.e. the forward and reverse reactions would balance each other), and there would be no net flux through the ...
  21. [21]
    [PDF] Thermodynamics, Chemical Equilibrium, and Gibbs Free Energy
    Jan 21, 2014 · If we add more reactant or more product, the reaction will proceed spontaneously (without external help) as long as the value for Gr decreases.Missing: large delta
  22. [22]
    Methane - the NIST WebBook
    Methane (CH4) has a molecular weight of 16.0425 and is also known as marsh gas, methyl hydride, and biogas.Missing: gibbs free
  23. [23]
  24. [24]
  25. [25]
  26. [26]
  27. [27]
  28. [28]
  29. [29]
  30. [30]
  31. [31]
    [PDF] Fuel Cell Handbook (Seventh Edition)
    ... Reactions in Fuel Cells ... exothermic overall reaction so that the stack cooling requirements are reduced. Further system simplification would occur ...
  32. [32]
    Gibbs free energy for glucose oxidation with O2
    The free energy for glucose oxidation with O2 (ΔG'0= -2870 kJ/mol glucose) exceeds that for glucose fermentation (ΔG'0= -218 kJ/mol glucose)."
  33. [33]
    Metabolic Energy - The Cell - NCBI Bookshelf - NIH
    The reaction yields a large amount of free energy: ΔG°′= -686 kcal/mol. To harness this free energy in usable form, glucose is oxidized within cells in a ...
  34. [34]
    How Cells Obtain Energy from Food - NCBI - NIH
    In total, the complete oxidation of a molecule of glucose to H2O and CO2 is used by the cell to produce about 30 molecules of ATP. In contrast, only 2 ...
  35. [35]
    Biology 2e, The Cell, Metabolism, ATP: Adenosine Triphosphate
    In fact, the ∆G for one ATP mole's hydrolysis in a living cell is almost double the value at standard conditions: –14 kcal/mol (−57 kJ/mol).
  36. [36]
    [PDF] Scientific Background - FSU Biology
    CO2 is driven by the free-energy loss associated with the oxidation of NADPH. ... associated with the complete combustion of a mole of glucose is 673 kcal mol. - ...
  37. [37]
    Biochemistry, Anaerobic Glycolysis - StatPearls - NCBI Bookshelf - NIH
    Jul 31, 2023 · Anaerobic glycolysis serves as a means of energy production in cells that cannot produce adequate energy through oxidative phosphorylation. In ...
  38. [38]
    Early bioenergetic evolution - PMC - PubMed Central
    Life is a net exergonic chemical reaction, it releases energy to go forward. Many settings have been proposed as the site for the chemical synthesis for life's ...
  39. [39]
    Metabolism - PMC - PubMed Central
    The importance and the evolutionary advantage of these pathways can be seen as many remain unchanged by animals, plants, fungi, and bacteria.
  40. [40]
    4.1 Energy and Metabolism – Concepts of Zoology – Hawaiʻi Edition
    Figure 4.3 Catabolic pathways are those that generate energy by breaking down larger molecules. Anabolic pathways are those that require energy to synthesize ...Metabolic Pathways · Thermodynamics · Enzymes
  41. [41]
    [PDF] Exergonic-Vs-Endergonic-Reactions.pdf
    One hallmark of exergonic reactions is a negative change in Gibbs free energy (ΔG < 0). Gibbs free energy is a thermodynamic quantity that combines enthalpy ...
  42. [42]
    6.9: ATP - Adenosine Triphosphate - Biology LibreTexts
    Nov 22, 2024 · Cells couple the exergonic reaction of ATP hydrolysis with endergonic reactions to harness the energy within the bonds of ATP.
  43. [43]
    ATP cycle and reaction coupling | Energy (article) - Khan Academy
    This strategy is used in many metabolic pathways in the cell, providing a way for the energy released by converting ATP to ADP to drive other reactions forward.
  44. [44]
    13.3: Gluconeogenesis - Biology LibreTexts
    Aug 23, 2025 · Explain the principle of reaction coupling and how unfavorable reactions are driven forward by coupling with strongly exergonic steps.
  45. [45]
    15.2: Photosynthesis - Chemistry LibreTexts
    Jan 20, 2025 · During noncyclic photophosphorylation, the generation of ATP is coupled to a one-way flow of electrons from H 2 ⁡ O to N ⁢ A ⁢ D ⁢ P + . We will ...
  46. [46]
    Thermodynamic efficiency, reversibility, and degree of coupling in ...
    Aug 18, 2020 · We find that the thermodynamic efficiency is about 80–90% and that the degree of coupling between the redox and proton translocation reactions is very high ...
  47. [47]
    14.1: Regulation of Metabolic Pathways - Biology LibreTexts
    Aug 23, 2025 · Exquisite mechanisms have evolved that control the flux of metabolites through metabolic pathways to insure that the output of the pathways ...Regulation of single enzymes... · Which Enzymes to Regulate...