Fact-checked by Grok 2 weeks ago

Chemical affinity

Chemical affinity, in the context of , is defined as the negative of the Gibbs energy with respect to the at constant and , providing a quantitative measure of the driving force for a ; it is positive for spontaneous processes and zero at . This concept, denoted as A = -\left( \frac{\partial G}{\partial \xi} \right)_{T,P}, where \xi is the , bridges classical with by indicating the direction and extent to which a will proceed under given conditions. The notion of chemical affinity originated in the 18th century as an empirical descriptor of the relative tendencies of substances to combine or displace one another in reactions, often compiled into affinity tables by chemists like Étienne-François Geoffroy to predict reaction outcomes. Over time, it evolved from these qualitative observations—rooted in alchemical ideas of attraction—into a rigorous thermodynamic quantity, particularly through the 19th-century work of , who integrated it with concepts to analyze . By the early , Théophile de Donder formalized its modern expression in , introducing the "degree of advancement" of a and linking to via inequalities involving uncompensated heat, laying the foundation for the Brussels school of irreversible processes. In contemporary usage, chemical affinity remains central to understanding reaction spontaneity, equilibrium constants, and the Gibbs free energy change (\Delta_r G = -A), where it effectively replaces or complements in discussions of reaction driving forces, especially in educational and applied contexts like and biochemical systems. Its evolution from a philosophical to a precise mathematical tool has profoundly influenced fields such as solution theory, , and , enabling predictions of reactivity without relying solely on empirical tables.

Historical Development

Early Theories and Alchemy

The concept of chemical affinity traces its philosophical roots to ancient Greek thought, particularly the pre-Socratic philosopher (c. 495–435 BCE), who proposed that all matter consists of four eternal "roots" or elements—earth, air, fire, and water—and that their combinations and separations are governed by two cosmic forces: (philia), which attracts and unites disparate elements into compounds, and Strife (neikos), which repels and divides them. This dualistic mechanism provided an early framework for understanding the attractions and repulsions underlying material transformations, influencing later interpretations of how substances interact and form new entities. Empedocles viewed these forces as cyclically dominant, driving the cosmos through phases of unity and dissolution, a notion that prefigured ideas of inherent tendencies in matter to combine or resist one another. In medieval Islamic , these elemental attractions evolved into more systematic notions of a "force" compelling substances to combine, transmute, or purify, as articulated by (c. 721–815 CE), known in Latin as . Jabir emphasized experimental observation to discern how metals and minerals interact, positing that all metals arise from the union of (conferring combustibility) and mercury (providing fluidity) in varying proportions, with an underlying affinity determining their stability and potential for transmutation into nobler forms like . His works, such as The Sum of Perfection, described chemical operations like and as means to manipulate these affinities, laying groundwork for viewing as driven by proportional balances rather than mere chance. This approach marked a shift toward empirical of attractive forces in , influencing European alchemists by attributing purposeful "pulls" to elemental constituents. The early 16th century saw further development in iatrochemistry through Theophrastus von Hohenheim, known as (1493–1541), who reframed affinities in a medical context by proposing the tria prima—sulfur (soul, combustible), mercury (spirit, volatile), and (body, fixed)—as the fundamental principles of all matter, with their interactions governed by natural correspondences and attractions akin to bodily humors. argued that diseases arise from imbalances in these principles, treatable by chemical remedies that exploit affinities between substances and the , such as using mercury compounds for their affinity to "draw out" impurities. , in particular, served as a "vinculum animae" or bonding agent with for both spiritual and corporeal realms, enabling alchemical preparations like the triumphal chariot of to harmonize disparate elements for therapeutic ends. His ideas bridged and , emphasizing selective attractions over universal elemental theories. By the , (1627–1691) critiqued and refined these notions in a more mechanistic vein, describing chemical "sympathies" and "antipathies" as observable tendencies for substances to unite or repel based on their corpuscular textures and motions, rather than qualities. In (1661), Boyle illustrated sympathies through examples like gold's ready colliquation with silver, , tin, lead, and , or quicksilver's with various metals, suggesting inherent compatibilities that form stable compounds without altering the particles' primitive natures. He contrasted this with antipathies, such as the violent ebullition and hissing when acid spirit of vitriol meets pot ashes or salt of , or the refusal of certain oils from the same source to mingle, attributing such behaviors to differing corpuscular arrangements that promote separation over union. Boyle's qualitative experiments, including distillations yielding selective salts from or sublimate, underscored these interactions as empirical phenomena, paving the way for 18th-century systematic affinity tables. Building on these ideas, (1642–1727) provided a physical interpretation in the 31st Query of his (1704), attributing chemical combinations to short-range attractive forces between the ultimate particles of matter, analogous to gravitational attraction but operating over very small distances. Newton suggested that these forces determine the affinities that cause particles to cohere into compounds, influencing the selective nature of chemical reactions and foreshadowing later mechanistic theories of chemistry.

18th- and 19th-Century Advances

In the early , Étienne-François Geoffroy introduced the first systematic approach to chemical affinity through his 1718 Table des différents rapports observés en chimie entre différentes substances, which ordered substances based on their relative tendencies to combine or displace one another in reactions. This table primarily focused on interactions between metals and acids, such as the displacement of one metal by another in acidic solutions, providing a hierarchical framework that allowed chemists to predict reaction outcomes empirically. Geoffroy's work marked a shift from qualitative descriptions to a semi-quantitative representation of affinities, influencing pharmaceutical preparations and experimental chemistry across Europe. Building on Geoffroy's foundation, Swedish chemist significantly expanded affinity tables in his 1775 Dissertation on Elective Attractions, creating comprehensive charts that encompassed nearly all known substances at the time. Bergman's tables, often presented as large fold-out diagrams with up to 59 columns and 50 rows, incorporated organic compounds alongside inorganics, establishing detailed hierarchies of reaction strengths based on observed displacements. These expansions enabled broader predictions of chemical behavior, including interactions involving alcohols and salts, and solidified affinity tables as standard tools in chemical education and practice until the mid-19th century. In the early 1800s, Humphry Davy's electrochemical experiments further illuminated the nature of affinity by demonstrating its connection to electricity. Using voltaic piles, Davy decomposed compounds like and in 1807, isolating elements such as and sodium, and proposed that chemical affinity arises from electrical forces between particles. His 1806 Bakerian Lecture detailed how electrical currents could overcome or reveal affinities, suggesting that the preferential reactivity of substances is governed by inherent electrical polarities. By the 1850s and 1860s, interpretations of affinity began incorporating thermal and kinetic dimensions. In the 1860s, Danish chemist Julius Thomsen and French chemist advanced the thermal theory of affinity, positing that the heat evolved during a serves as a quantitative measure of affinity strength. This view, developed amid mid-century advances in heat engines and , treated exothermic reactions as indicators of greater affinity, influencing thermochemical studies. Concurrently, Alexander Williamson advanced kinetic interpretations in the 1850s, proposing that drives reactions through continuous molecular collisions and exchanges, rather than static attractions. In his studies on etherification, Williamson described as a dynamic balance of and recombination, where molecules perpetually collide and reform, challenging earlier fixed-affinity models. This collision-based framework laid groundwork for understanding reaction rates in terms of molecular motion, bridging chemistry with nascent kinetic theories.

20th-Century Formalization

In the early , the concept of chemical affinity underwent a rigorous mathematical formalization within , shifting from its earlier qualitative and empirical interpretations as a metaphorical "force" to a precise that quantifies the driving force of chemical reactions. This transition was pioneered by Belgian physicist and chemist Théophile de Donder, who in 1922 introduced affinity as a measurable quantity linked to the progress of a reaction, building on the foundations of Gibbsian thermodynamics but extending it to irreversible processes. De Donder defined chemical affinity A for a reaction as A = -\sum \nu_i \mu_i, where \nu_i are the stoichiometric coefficients and \mu_i are the chemical potentials of the species involved, thereby establishing it as the negative of the Gibbs free energy change per unit extent of reaction at constant temperature and pressure. De Donder's formulation marked a pivotal departure from 19th-century views of affinity as an abstract attractive force, repositioning it as an intensive thermodynamic variable that vanishes at equilibrium and drives spontaneous change away from it. This work, detailed in his 1922 publications and later elaborated in his 1936 monograph Thermodynamic Theory of Affinity, integrated affinity into the variational principles of thermodynamics, influencing the Brussels school of nonequilibrium thermodynamics. By expressing affinity in terms of chemical potentials, de Donder provided a framework compatible with emerging quantum and statistical mechanics, though his primary focus remained on macroscopic descriptions. The formalization advanced significantly in the mid-20th century through the efforts of Ilya Prigogine, who extended de Donder's ideas to nonequilibrium systems during the 1940s and 1950s. In collaboration with Raymond Defay, Prigogine incorporated chemical affinity into the broader theory of irreversible processes, emphasizing its role in entropy production and the evolution of open systems far from equilibrium. Their 1954 treatise Chemical Thermodynamics formalized affinity as A = -(\partial G / \partial \xi)_{T,P}, where G is the Gibbs free energy and \xi is the reaction extent, and applied it to phenomena like dissipative structures, where affinity sustains steady states through continuous matter and energy exchange. This extension transformed affinity from an equilibrium-centric concept into a cornerstone of nonequilibrium thermodynamics, enabling analyses of complex chemical dynamics.

Conceptual Foundations

Qualitative Understanding

Chemical affinity, in qualitative terms, describes the thermodynamic driving force that determines the direction in which a proceeds toward under constant and . It quantifies the tendency of a to evolve spontaneously when there is an imbalance in the chemical potentials of reactants and products, with positive affinity favoring the forward reaction and zero affinity indicating . This intuitive concept arises from observations of reactions that release energy or increase disorder, leading to more stable states, as captured by the negative change in (ΔG = -A ξ, where ξ is the ). Unlike physical processes such as mixing, where no new form, chemical affinity pertains to transformations that alter molecular structures, driven by the system's quest for minimum . For instance, in the reaction of and oxygen to form (2H₂ + O₂ → 2H₂O), the strong driving force reflects the formation of stable O-H bonds, resulting in a large negative ΔG and spontaneous progression under standard conditions. This energy stabilization distinguishes affinity-driven reactions, providing a conceptual basis for predicting spontaneity without quantitative computation. Qualitatively, chemical serves as a tool for understanding feasibility, where a system's deviation from —measured by —guides the extent of . High indicates a strong propensity for the to advance, bridging empirical observations of reactivity with thermodynamic principles.

Relation to Reaction Driving Forces

Chemical represents the thermodynamic driving force that compels chemical systems to evolve toward by addressing imbalances in the chemical potentials of reactants and products. This concept, pioneered by Théophile de Donder in the early 1920s, conceptualizes as the impetus for spontaneous change, guiding along paths that minimize without altering the final state. In qualitative terms, it embodies the tendency for systems to progress in the direction that reduces energetic instability, particularly in processes where product formation lowers the overall Gibbs energy. In multi-component systems, chemical affinity functions as a directional force, determining the net flow of toward products when the affinity is positive, thereby dictating the reaction's progress in complex mixtures. This force ensures that reactions proceed to diminish the system's disequilibrium, with the magnitude of reflecting the degree of departure from equilibrium. For example, in acid-base neutralization, such as the reaction between and to form and salt, the driving force favors the formation of stable products over dissociated ions. Affinity also influences the selection of reaction pathways, particularly in catalyzed processes where the inherent driving force remains unchanged, but the catalyst lowers kinetic barriers to align with the system's thermodynamic favorability. In enzymatic catalysis, for instance, the substrate's interaction with the facilitates pathways that respect the overall , enhancing efficiency while preserving directional specificity. This interplay underscores 's role in initiating and guiding chemical transformations toward in diverse systems.

Thermodynamic Formulation

Equilibrium and Free Energy

In , the affinity of a reaches zero at , signifying that the system has minimized its under constant temperature and pressure conditions. This state implies that the change for the , ΔG, is also zero, as no net driving force exists to shift the in either direction. The chemical affinity A is formally defined as the negative of the G with respect to the ξ: A = -\left( \frac{\partial G}{\partial \xi} \right)_{T,P} This definition ensures that A > 0 for spontaneous forward reactions, promoting progress toward equilibrium until A = 0. To derive the expression for affinity, the Gibbs free energy change for a general reaction such as a\mathrm{A} + b\mathrm{B} \rightleftharpoons c\mathrm{C} + d\mathrm{D} is expressed as \Delta G = c\mu_\mathrm{C} + d\mu_\mathrm{D} - a\mu_\mathrm{A} - b\mu_\mathrm{B}, where \mu_i denotes the chemical potential of each species i. Consequently, the affinity is A = -\Delta G = -(c\mu_\mathrm{C} + d\mu_\mathrm{D} - a\mu_\mathrm{A} - b\mu_\mathrm{B}). The chemical potential for each species is given by \mu_i = \mu_i^\circ + RT \ln a_i, where \mu_i^\circ is the standard chemical potential, R is the gas constant (8.314 J mol⁻¹ K⁻¹), T is the absolute temperature, and a_i is the activity (approximating concentration or partial pressure for ideal systems). Substituting this into the expression for ΔG allows calculation of affinity based on the activities of reactants and products. At equilibrium, the activities satisfy \Delta G = 0, yielding A = 0 and the equilibrium constant K = ∏ (a_i)^{ν_i}, where ν_i are the stoichiometric coefficients (positive for products, negative for reactants). A negative value of ΔG corresponds to a positive affinity, indicating that the forward reaction is thermodynamically favored and proceeds spontaneously until equilibrium is reached. This relationship provides the quantitative criterion for spontaneity: reactions with A > 0 release , driving the system toward products, while A < 0 would favor the reverse direction. For instance, in the reaction \mathrm{H_2(g)} + \frac{1}{2}\mathrm{O_2(g)} \rightleftharpoons \mathrm{H_2O(l)} at standard conditions (298 K, 1 bar), the standard Gibbs free energy change is ΔG° = -237.1 kJ/mol, so the standard affinity A° = 237.1 kJ/mol (per mole of water formed). This large positive A° reflects the strong thermodynamic drive for water formation from its elements, with equilibrium lying far toward the product under standard conditions; actual equilibrium activities (e.g., very low hydrogen and oxygen partial pressures) would then balance to make ΔG = 0 and A = 0.

Non-Equilibrium Extensions

In non-equilibrium thermodynamics, chemical affinity extends the concept of driving force to irreversible processes and time-dependent systems, quantifying the departure from equilibrium in chemical reactions. Théophile de Donder formalized this extension in his 1922 work, defining affinity as a generalized thermodynamic potential that governs the direction and rate of reactions in systems not at rest. This framework builds on the equilibrium limit where affinity vanishes, but applies broadly to evolving systems where sustained fluxes occur. Unlike static equilibrium analyses, non-equilibrium affinity accounts for ongoing transformations, enabling the study of open systems far from balance. For systems near equilibrium, the affinity A is approximated as A = RT \ln \left( \frac{K}{Q} \right), where R is the gas constant, T the temperature, Q the reaction quotient reflecting instantaneous concentrations, and K the equilibrium constant. This expression captures how deviations in Q from K generate a driving force, with positive A indicating the potential for net forward progress when Q < K. De Donder's key relation links this affinity to the net reaction rate v = v_f - v_r, where v_f and v_r are the forward and reverse rates, respectively; near equilibrium, v \propto A, establishing a linear response between the thermodynamic force and the flux. This proportionality, derived from phenomenological coefficients, underscores affinity's role in predicting reaction velocities without detailed kinetic mechanisms. A central concept in these extensions is the dissipation function \sigma = \frac{A v}{T}, which measures the rate of entropy production due to the reaction. This bilinear form ensures \sigma \geq 0, aligning with the second law for irreversible processes, and quantifies the irreversible heat dissipation associated with non-zero affinity. In open systems, such as biological metabolism, affinity sustains continuous fluxes; for instance, in cellular pathways like , coupled reactions maintain non-zero affinities to drive against concentration gradients, preventing equilibrium and enabling life-sustaining cycles. These applications highlight how affinity integrates thermodynamics with dynamics in far-from-equilibrium environments, as seen in metabolic networks where multiple affinities couple to produce directed work.

Modern Applications and Interpretations

In Chemical Kinetics

In chemical kinetics, chemical affinity influences reaction rates primarily through its role in shaping the activation free energy barrier (ΔG‡) within (TST). TST posits that the rate constant for an elementary reaction is given by k = \frac{k_B T}{h} e^{-\Delta G^\ddagger / RT}, where the exponential term reflects the probability of reaching the transition state, and ΔG‡ represents the free energy difference between reactants and the activated complex. The overall chemical affinity, defined as A = -ΔG for the reaction, indirectly modulates ΔG‡ by determining the energetic landscape; for instance, according to the , in highly exergonic reactions (large positive A), the transition state structurally and energetically resembles the reactants more closely, often resulting in a lower ΔG‡ compared to endergonic counterparts. A classic example is combustion reactions, such as the oxidation of hydrocarbons, which exhibit high chemical affinity due to their strongly negative ΔG (approximately -800 kJ/mol for complete combustion of methane), driving the reaction strongly toward products. However, these reactions often display high activation energies (e.g., around 150-200 kJ/mol for initiation steps in hydrocarbon combustion), leading to slow rates at ambient conditions and necessitating an ignition source to overcome the barrier. This illustrates how affinity ensures thermodynamic favorability but does not guarantee rapid kinetics without sufficient energy input. In enzymatic catalysis, chemical affinity manifests in Michaelis-Menten kinetics, where the Michaelis constant (K_m) serves as an inverse measure of the enzyme's binding affinity for its substrate, reflecting the substrate concentration at half-maximal velocity (V_max). Specifically, K_m ≈ (k_{-1} + k_{cat}) / k_1, linking the equilibrium dissociation constant (K_d = k_{-1}/k_1, related to -ΔG_binding) to the catalytic turnover; enzymes with high substrate affinity (low K_m) achieve efficient rates at low concentrations by stabilizing the enzyme-substrate complex, thereby lowering the effective ΔG‡ for the transition state. For example, hexokinase has a low K_m (~0.1 mM) for glucose, indicating strong affinity that facilitates rapid phosphorylation in . Fundamentally, chemical predicts the direction and extent of a (spontaneous if A > 0), while , governed by ΔG‡, determines the timescale; a may have high affinity yet proceed slowly if the barrier is insurmountable under given conditions. In non-equilibrium systems, further acts as a direct driver of net rates, as per de Donder's relation, where near-equilibrium deviations from unity in the mass-action ratio amplify forward rates proportional to A/RT.

Quantum and Computational Perspectives

From a quantum mechanical viewpoint, chemical affinity arises from the interactions between electron densities and of reacting species, particularly through orbital overlaps that stabilize transition states and products. Kenichi Fukui's elucidates this by emphasizing that chemical reactivity—and hence affinity—is dominated by the highest occupied (HOMO) and lowest unoccupied molecular orbital (LUMO) of the reactants. For example, in electrophilic additions, the HOMO of the donates electrons to the LUMO of the , with the strength of this interaction quantified by the energy gap between these orbitals; smaller gaps enhance affinity and reaction propensity. This framework has been extended within density functional theory (DFT), where affinity is probed through electron density functionals that capture local reactivity. The Fukui function, f(\mathbf{r}) = \left( \frac{\partial \rho(\mathbf{r})}{\partial N} \right)_{v(\mathbf{r})}, defined as the derivative of electron density \rho(\mathbf{r}) with respect to the number of electrons N at fixed external potential v(\mathbf{r}), identifies regions of high reactivity: f^+(\mathbf{r}) for nucleophilic attack and f^-(\mathbf{r}) for electrophilic attack. In practice, these are approximated as the difference in electron densities between neutral and charged systems, providing a quantum basis for affinity by mapping electron transfer tendencies. Computational simulations further quantify affinity via DFT calculations of energies \Delta E, serving as proxies for the overall tendency, with thermodynamic affinity as its macroscopic counterpart. such as B3LYP are widely used for this purpose, offering balanced accuracy for \Delta E in molecular complexes; for instance, in protein-ligand , such calculations synthetic . Post-2000 advances have incorporated (ML) to predict affinities directly from quantum-derived features, accelerating by surpassing traditional DFT in speed and scalability. neural networks, trained on datasets like PDBbind (introduced 2004), model protein-ligand interactions via atomic graphs, achieving Pearson correlation coefficients r > 0.8 for free energies in ; notable examples include models like GraphDelta, which integrate 3D structural data for high-fidelity affinity forecasts without full quantum calculations. It is essential to distinguish this broader quantum interpretation of chemical affinity—which encompasses orbital and density-driven reaction tendencies—from , a specific property defined by the International Union of Pure and Applied Chemistry (IUPAC) as the energy change for the process \ce{X(g) + e^- -> X^-(g)}, typically the energy released upon attachment to a neutral gaseous atom.

References

  1. [1]
    affinity of reaction (A00178) - IUPAC Gold Book
    Negative partial derivative of Gibbs energy with respect to extent of reaction at constant pressure and temperature. It is positive for spontaneous reactions.<|control11|><|separator|>
  2. [2]
    Affinity, Expressed in Joules/Equivalent, Can Replace the Gibbs ...
    Mar 29, 2013 · The first measure would be to reintroduce “affinity”, defined by IUPAC (4) as A = −(dG/dξ)T,p, where ξ is the extent of reaction.
  3. [3]
  4. [4]
    Celebrating the birth of De Donder's chemical affinity (1922-2022)
    Jan 19, 2023 · The present article aims to reconstruct how De Donder elaborated his ideas and how he developed them by exploring his teaching activity.
  5. [5]
    Empedocles - Stanford Encyclopedia of Philosophy
    Sep 26, 2019 · Aristotle credits Empedocles with being the first to distinguish clearly these four elements (Metaphysics. 985a31–3). However, the fact that the ...Missing: chemical | Show results with:chemical
  6. [6]
    None
    ### Summary of Jabir ibn Hayyan, Paracelsus, and Robert Boyle on Chemical Combination, Forces, Sympathies, or Affinities in Alchemy
  7. [7]
  8. [8]
    [PDF] Ambix - Research
    Sulphur forms the intermediary joining Spirit to. Body. It is the vinculum animae that has affinity both to body and spirit- like the astral body of the ...
  9. [9]
    The Project Gutenberg eBook of The Sceptical Chymist, by Robert ...
    The Sceptical Chymist or Chymico-Physical Doubts & Paradoxes, Touching the Spagyrist's Principles Commonly call'd Hypostatical.
  10. [10]
    Etienne Francois Geoffroy's table of relations and the concept of affinity
    The paper describes Geoffroy's table of proportions. The context in which it was developed is discussed, and subsequent tables, particularly those of Bergman, ...Missing: sources | Show results with:sources
  11. [11]
    Torbern Bergman: A Man before His Time (Schufle, J. A.)
    ### Summary of Torbern Bergman's 1775 Affinity Tables
  12. [12]
    Humphry Davy | Science History Institute
    His electrochemical experiments led him to propose that the tendency of one substance to react preferentially with other substances—its “affinity”—is ...
  13. [13]
    [PDF] Deciphering the physical meaning of Gibbs's maximum work equation
    Feb 29, 2024 · A new concept of energy was needed. ΔHrxn and the thermal theory of affinity. In the mid-1800s, technical analyses of the steam engine led to ...
  14. [14]
    Associating Physics and Chemistry to Dissociate Molecules - jstor
    In the 1850s, Rudolf Clausius and Alexander William Williamson independently developed similar hypotheses at a time when physics and chemistry were beginning ...
  15. [15]
    Thermodynamic Theory of Affinity, Volume 1 - Google Books
    Thermodynamic Theory of Affinity, Volume 1. Front Cover. Théophile de Donder, Pierre Van Rysselberghe. Stanford University Press, 1936 - Chemical affinity - 142 ...<|control11|><|separator|>
  16. [16]
    Treatise on Thermodynamics: Chemical thermodynamics
    Authors, Ilya Prigogine, Raymond Defay ; Publisher, Longmans, Green, 1954 ; Original from, the University of Michigan ; Digitized, Nov 19, 2013.
  17. [17]
    [PDF] prigogine-lecture.pdf - Nobel Prize
    This is one of the main reasons why the applica- tions of thermodynamics were essentially limited to equilibrium. To extend thermodynamics to non-equilibrium ...Missing: 1940s 1950s
  18. [18]
    [PDF] PHYSICAL CHEMISTRY 1 (lecture notes) - Index of /
    Chemical affinity is the electronic property by which dissimilar chemical species are capable of forming chemical compounds. The following considerations ...
  19. [19]
    Why Does Sodium Form NaCl?
    Sodium forms NaCl due to the strong attraction between Na+ and Cl- ions, driven by the lattice energy of NaCl, not just chlorine's electron affinity.
  20. [20]
    [PDF] Studies concerning affinity
    The first theory about chemical affinity was advanced by the Swede Bergman in 1780, thus a t a time when the atomic theory was not yet developed. He assumes ...
  21. [21]
    Chemical Affinity and the Density of Energy Levels
    The concept of chemical affinity (A) was introduced into thermodynamics by Belgian physicist Théophile de Donder in 1922 to represent the driving force of a ...
  22. [22]
    [PDF] Celebrating the birth of De Donder's chemical affinity (1922-2022)
    De. Donder was not the first to apply methods of mathematical physics to thermodynamics and chemistry. He followed the work of Josiah Willard Gibbs and Pierre ...
  23. [23]
    REVIEW Thermodynamics and foundations of mass-action kinetics
    De Donder also supposed that the second law of thermodynamics could be ... Therefore, vector A is called the (chemical) affinity vector and Ap the (chemical) ...
  24. [24]
    The Precipitation Laws. | Chemical Reviews - ACS Publications
    ... chemical affinity of its formation reaction. Doklady Physical Chemistry 2013 ... Hydrochloric Acid Regeneration via Calcium Sulfate Crystallization for Non‐ ...
  25. [25]
    [PDF] Degree of rate control and De Donder relations - OSTI.GOV
    Two complementary formalisms for identifying rate- limiting transition states are broadly employed in chemical kinetics—De Donder relations based on measuring ...Missing: affinity | Show results with:affinity
  26. [26]
    Standard State and Enthalpy of Formation, Gibbs Free Energy of ...
    Definition and explanation of the terms standard state and standard enthalpy of formation, with listing of values for standard enthalpy and Gibbs free energy of ...
  27. [27]
    Nonequilibrium Thermodynamics of Chemical Reaction Networks
    Dec 22, 2016 · We build a rigorous nonequilibrium thermodynamic description for open chemical reaction networks of elementary reactions.
  28. [28]
    Nonequilibrium Thermodynamics in Biochemical Systems and Its ...
    In this review, we briefly introduce the history and current development of nonequilibrium thermodynamics, especially that in biochemical systems.
  29. [29]
    Role of Frontier Orbitals in Chemical Reactions - Science
    FUKUI, K, MOLECULAR ORBITAL THEORY OF ORIENTATION IN AROMATIC, HETEROAROMATIC, AND OTHER CONJUGATED MOLECULES, JOURNAL OF CHEMICAL PHYSICS 22: 1433 (1954).Missing: seminal | Show results with:seminal
  30. [30]
    Density functional theory, chemical reactivity, and the Fukui functions
    Jan 17, 2022 · We will also present a general introduction to the Fukui functions, and their relation with nucleophilicity and electrophilicity, with an ...Thomas--Fermi Approximation · Hardness And Softness · Fukui Functions
  31. [31]
    Pairwise Additivity and Three-Body Contributions for Density ...
    Feb 29, 2024 · In this work, we test the pairwise additivity approximation on the sulfotransferase-l-DOPA complex for 16 density functional theory (DFT) methods.
  32. [32]
    Modern machine‐learning for binding affinity estimation of protein ...
    Jun 11, 2024 · This review will highlight recent strides toward improved real world applicability of machine-learning based scoring, enabling a better understanding of the ...
  33. [33]
    electron affinity (E01977) - IUPAC Gold Book
    Energy required to detach an electron from the singly charged negative ion (energy for the process molecule).