Fact-checked by Grok 2 weeks ago

Spontaneous process

A spontaneous process is a thermodynamic in which a physical or occurs naturally under given conditions without the need for continuous external input, driven by an inherent tendency to increase the total of the . This concept is central to the second law of thermodynamics, which states that in an , spontaneous processes proceed in the direction that leads to an increase in disorder or (ΔS > 0). Common examples include the expansion of a gas into a , flowing from a hot object to a colder one, and the melting of ice at temperatures above 0°C. In , the spontaneity of a process at constant and pressure is quantitatively predicted using (G), defined as G = H - TS, where H is , T is , and S is . A process is spontaneous if the change in (ΔG) is negative (ΔG < 0), as this indicates a decrease in the system's free energy and favors the forward direction; at equilibrium, ΔG = 0, and if ΔG > 0, the process is nonspontaneous and requires external work. The relationship is expressed by the equation ΔG = ΔH - TΔS, where both enthalpic (ΔH) and entropic (TΔS) contributions determine the outcome—endothermic processes can still be spontaneous if the increase is sufficiently large. Importantly, thermodynamic spontaneity describes the feasibility of a but does not imply its rate or speed, which is governed by ; for instance, the conversion of to is spontaneous yet extremely slow due to high . This distinction underscores that while spontaneous processes align with the universe's tendency toward greater disorder, practical observations may require catalysts or specific conditions to occur at observable timescales.

Introduction

Definition

A spontaneous process in refers to a physical or chemical change that occurs without the need for continuous external input and proceeds in the direction that results in an increase in the total of the , as dictated by law of . This means that for such a to be feasible, the sum of the change of the system and its surroundings must be positive (ΔS_ > 0). The concept emphasizes the natural tendency of systems to evolve toward states of greater disorder or probability without ongoing intervention. Unlike reversible processes, which are idealized and occur through a series of states requiring only driving forces to maintain balance, spontaneous processes are inherently irreversible and driven by finite imbalances that propel the system away from . In practice, spontaneous processes cannot be reversed without external work, distinguishing them from the theoretical reversibility that assumes no dissipative losses like or across finite gradients. Understanding spontaneous processes requires familiarity with thermodynamic systems—isolated (no exchange of or ), closed ( exchange only), or open (exchange of both)—and state functions such as , , and , which are path-independent properties used to assess process feasibility. The term emerged in the context of 19th-century , with foundational formalization by through his introduction of and the second law in the 1850s–1860s, and further refined by J. Willard Gibbs in the 1870s via criteria involving for constant-temperature and pressure conditions.

Historical Development

The concept of spontaneous processes in thermodynamics traces its roots to 18th-century investigations into and . , a Scottish chemist, advanced early understandings through his discoveries of specific and in the 1760s, observing that substances absorb without temperature change during phase transitions, which laid groundwork for quantifying flow in natural processes. Concurrently, and developed the in the 1780s, positing as an indestructible fluid (calorique) that flows from hotter to cooler bodies, explaining affinity-driven reactions like as spontaneous due to this directional transfer. These ideas framed spontaneous events as natural tendencies without external work, influencing later thermodynamic frameworks. A pivotal shift occurred with Sadi Carnot's 1824 publication Réflexions sur la puissance motrice du feu, which introduced the —a reversible ideal operating between two temperatures—distinguishing engineered efficiency from irreversible, spontaneous heat flows in nature, such as or expansion without work. This work marked the term "spontaneous" acquiring a precise thermodynamic , emphasizing processes that proceed unidirectionally without sustained input. Building on Carnot, formalized the second law in 1850 through Über die bewegende Kraft der Wärme, asserting that heat cannot spontaneously flow from cold to hot bodies, and later introduced in 1865 as a measure of energy unavailability, quantifying the irreversibility of spontaneous changes. In the 1870s, extended these principles in his On the Equilibrium of Heterogeneous Substances (1876–1878), defining (now ) as the maximum reversible work available at constant temperature and pressure, providing a criterion for predicting spontaneity in chemical systems. Simultaneously, incorporated the concept into via his H-theorem (1872) and formula S = k \ln W (1877), interpreting spontaneous processes as probable increases in microscopic disorder among molecular configurations. refined this synthesis in the early 20th century, particularly through his 1900 quantum hypothesis for , which reconciled classical with atomic discreteness and clarified 's role in irreversible processes.

Thermodynamic Principles

Second Law of Thermodynamics

The second law of thermodynamics provides the foundational principle for understanding spontaneous processes, stating that in any spontaneous process, the entropy of the universe increases, with ΔS_universe > 0, while equality holds for reversible processes. This law establishes the directionality of natural processes, ensuring that systems evolve toward states of greater disorder without external intervention. The key equation encapsulating this is ΔS_universe = ΔS_system + ΔS_surroundings > 0 for spontaneous processes, where the total entropy change across the system and its surroundings determines spontaneity. Equivalent formulations of the second law highlight its implications for and work. The Kelvin-Planck statement, articulated in 1851, asserts the impossibility of a device operating in a that receives from a single reservoir and produces an equivalent amount of work, thereby prohibiting machines of the second kind. The Clausius statement, from 1854, declares that cannot spontaneously flow from a colder body to a hotter one without some other change occurring in the universe. These principles imply a universal for thermodynamic processes, as the second law forbids the complete reversal of spontaneous events and underscores the irreversibility inherent in nature, such as the dissipation of useful into unusable forms. The law was originally formulated by in 1854 and William Thomson () in 1851, building on earlier work by Sadi Carnot to reconcile observations of engines with emerging principles.

Entropy Concept

Entropy, denoted as S, is a thermodynamic state function that quantifies the degree of energy dispersal or microscopic disorder within a system, representing the portion of the system's internal energy that is unavailable for useful work. This concept, introduced by Rudolf Clausius in 1865, arises from the observation that certain processes lead to an irreversible degradation of energy quality, even as total energy is conserved. At the microscopic level, provided a statistical interpretation in 1877, linking to the number of possible microstates W corresponding to a macroscopic state: S = k \ln W where k is Boltzmann's constant ($1.380649 \times 10^{-23} J/K). This illustrates how measures the multiplicity of ways can be distributed among particles, with higher W indicating greater and thus higher . Macroscopically, entropy changes are described differentially. For a reversible , the infinitesimal change is dS = \frac{\delta Q_\text{rev}}{T}, where \delta Q_\text{rev} is the reversible and T is the absolute in . In irreversible processes, the entropy change of the system exceeds the heat transfer divided by temperature: \Delta S > \int \frac{\delta Q}{T}. This inequality reflects the generation of entropy due to irreversibilities, such as friction or mixing. Entropy is measured in joules per kelvin (J/K) in the (SI). Absolute entropy values for substances can be determined using the third law of thermodynamics, formulated by around 1906, which states that the entropy of a perfect crystalline substance approaches zero as temperature approaches (0 K). This provides a reference point for calculating standard entropies at other temperatures via integration of heat capacities. In isolated systems, entropy reaches its maximum value at , driving spontaneous processes toward states of greater disorder.

Criteria for Spontaneity

Entropy-Based Criterion

The entropy-based criterion for spontaneity, derived from the second law of , states that a process is spontaneous if the total entropy change of the is positive (ΔS_universe > 0), non-spontaneous if negative (ΔS_universe < 0), and at equilibrium if zero (ΔS_universe = 0). To apply this criterion, the total entropy change is calculated as ΔS_universe = ΔS_system + ΔS_surroundings. The system's entropy change (ΔS_system) is determined experimentally through methods like calorimetry or from standard thermodynamic tables. For processes at constant temperature, the surroundings' entropy change is approximated as ΔS_surroundings = -ΔH_system / T, where ΔH_system is the enthalpy change of the system and T is the absolute temperature in kelvin, assuming the heat transfer to the surroundings is reversible. Consider the example of ice at its standard melting point of 0°C (273 ). Here, ΔS_system is positive at approximately 22.0 J/· due to the phase transition from solid to liquid, while ΔS_surroundings is negative because heat is absorbed from the surroundings (ΔH_fusion ≈ 6.02 kJ/). Above 0°C, such as at 10°C (283 ), ΔS_surroundings ≈ -21.3 J/·, yielding ΔS_universe ≈ +0.7 J/· > 0, so is spontaneous; below 0°C, such as at -10°C (263 ), ΔS_universe ≈ -0.9 J/· < 0, so it is non-spontaneous. This criterion applies universally to all processes since the universe can be treated as an isolated system, but it is often cumbersome for non-isolated systems where precise measurement of surroundings' entropy requires comprehensive data on heat exchanges and temperature variations, making it impractical for complex real-world scenarios. As an alternative for constant-temperature and pressure conditions, the change provides a more convenient system-focused assessment of spontaneity.

Free Energy-Based Criteria

Free energies provide a practical means to assess the spontaneity of processes under controlled conditions by incorporating both enthalpic and entropic contributions into a single thermodynamic potential, thereby simplifying the general requirement that the total entropy change of the universe must be positive (\Delta S_{\text{universe}} > 0). These potentials are particularly useful in laboratory settings where temperature is held constant, allowing chemists and physicists to predict whether a or change will proceed without calculating the changes of the surroundings explicitly. The , denoted G, is the appropriate criterion for processes at constant (T) and (P), defined as G = [H](/page/Enthalpy) - TS, where H is the , T is the absolute , and S is the . The change in Gibbs free energy for a process is thus \Delta G = \Delta H - T \Delta S (at constant T). This function was introduced by in his seminal 1876 paper "On the Equilibrium of Heterogeneous Substances." For systems at constant temperature and volume (V), the , denoted A, serves as the spontaneity indicator, defined as A = U - TS, where U is the . The corresponding change is \Delta A = \Delta U - T \Delta S (at constant T). developed this concept in his 1882 lecture "On the Thermodynamics of Chemical Processes." A negative change in these free energies signals spontaneity: \Delta G < 0 at constant T and P, or \Delta A < 0 at constant T and V, while \Delta G = 0 or \Delta A = 0 indicates equilibrium. For the Gibbs free energy, this criterion derives from the second law: at constant T and P, \Delta S_{\text{universe}} = \Delta S_{\text{sys}} + \Delta S_{\text{surr}} > 0, where \Delta S_{\text{surr}} = -\Delta H / T (assuming reversible heat transfer to the surroundings). Substituting yields \Delta S_{\text{universe}} = \Delta S - \Delta H / T > 0; multiplying through by −T (and reversing the inequality) gives \Delta H - T \Delta S < 0, or \Delta G < 0. Analogously, for the Helmholtz free energy at constant T and V, \Delta S_{\text{surr}} = -\Delta U / T, leading to \Delta S_{\text{universe}} = \Delta S - \Delta U / T > 0, and thus \Delta A < 0 upon multiplication by −T.

Applications and Examples

Physical Processes

Spontaneous physical processes encompass a range of changes that occur without external intervention, driven by thermodynamic principles such as increases in . One classic example is the free expansion of an ideal gas into a vacuum, where the gas rushes to fill an evacuated space, resulting in no heat exchange or work done by the system. For an ideal gas, this isothermal process features zero change in internal energy (ΔU = 0) and enthalpy (ΔH = 0), yet it is irreversible and spontaneous because the system's increases according to ΔS = nR \ln(V_f / V_i) > 0, leading to a net increase in the of the . Although the process does not occur at constant , the of the system decreases (ΔG < 0), consistent with spontaneity, as the final state has lower chemical potential due to the reduced . Diffusion represents another fundamental spontaneous physical process, where particles—such as gases or solutes—spread from regions of higher concentration to lower concentration, eliminating gradients and maximizing disorder. This mixing is driven by an entropy gradient, as the dispersal of molecules increases the configurational entropy of the system without a corresponding energy cost for ideal cases. For instance, when two distinct ideal gases are initially separated in a container and then allowed to mix freely, the process proceeds spontaneously, with the entropy change given by ΔS_mix = -nR (x_1 \ln x_1 + x_2 \ln x_2) > 0, where x_i are fractions, reflecting the increased number of microstates. No external work or heat is required, underscoring that the driving force is purely entropic, aligning with the second law of . Phase changes, such as and , provide clear examples of spontaneous physical processes under appropriate conditions, often evaluated using criteria at constant temperature and pressure. Consider melting at 25°C: this (ΔH > 0) is spontaneous because the increase upon transitioning to liquid water dominates, yielding ΔG = ΔH - TΔS < 0, where the TΔS term outweighs ΔH at temperatures above the melting point. Similarly, evaporation of a liquid into its vapor phase can be spontaneous if the partial pressure of the vapor is below the equilibrium vapor pressure. The Gibbs free energy change for this process is described by the equation \Delta G = \Delta G^\circ + RT \ln Q where \Delta G^\circ is the standard free energy change, R is the gas constant, T is temperature, and Q is the activity quotient (often the ratio of the vapor's partial pressure to the standard pressure for pure substances). If Q < 1 such that ΔG < 0, evaporation proceeds spontaneously until equilibrium is reached. This application of the Gibbs free energy criterion highlights how phase transitions balance enthalpic and entropic contributions to determine spontaneity.

Chemical Reactions

In chemical reactions, spontaneity is determined by the Gibbs free energy change (), where a negative indicates a spontaneous process under constant temperature and pressure. Exothermic reactions, characterized by a large negative enthalpy change (), often exhibit negative values, driving the reaction forward. For instance, the combustion of methane (CH₄ + 2O₂ → CO₂ + 2H₂O) is highly exothermic with ΔH° ≈ -890 kJ/mol, resulting in ΔG° ≈ -818 kJ/mol at 298 K, making it spontaneous due to the dominant enthalpic contribution despite a modest entropy change.) However, spontaneity can also occur in endothermic reactions if the entropy increase (ΔS) is sufficiently large to yield a negative ΔG, as per ΔG = ΔH - TΔS. The dissolution of ammonium nitrate (NH₄NO₃(s) → NH₄⁺(aq) + NO₃⁻(aq)) is endothermic (ΔH > 0, approximately +25 kJ/mol), yet spontaneous at because the positive ΔS from hydration and increased overcomes the enthalpic barrier, leading to ΔG < 0. This process is commonly used in cold packs, where the cooling effect highlights the endothermic nature while confirming overall spontaneity. The extent of spontaneity in chemical reactions is quantitatively linked to the equilibrium constant (K) through the relation K = e^{-ΔG°/RT}, where R is the gas constant and T is temperature in Kelvin; a large K (>1) corresponds to a highly negative ΔG°, favoring product formation./7:_Equilibrium_and_Thermodynamics/7.11:_Gibbs_Free_Energy_and_Equilibrium) In electrochemical reactions, such as those in processes, spontaneity ties directly to cell potential via the Nernst equation's foundation: ΔG = -nFE, where n is the number of moles of electrons transferred, F is Faraday's constant, and E is the cell potential; a positive E yields negative ΔG, indicating spontaneity./Electrochemistry/Nernst_Equation) A classic redox example is the rusting of iron (4Fe + 3O₂ + 6H₂O → 4Fe(OH)₃), which is spontaneous under moist air conditions at ambient temperatures, with ΔG° < 0 driven by the exothermic oxidation despite kinetic barriers that slow the process. This reaction exemplifies how environmental factors like moisture enable spontaneity in corrosion processes central to materials science.

Biological Systems

In biological systems, spontaneous processes are essential for maintaining life, often occurring through coupled reactions in open systems that exchange energy and matter with their surroundings. A key mechanism involves the coupling of exergonic reactions, such as the hydrolysis of adenosine triphosphate (ATP) to adenosine diphosphate (ADP) and inorganic phosphate, which releases energy with a standard free energy change of ΔG° = -30.5 kJ/mol under physiological conditions. This energy drives endergonic processes like protein synthesis, where the formation of peptide bonds requires approximately +20 to +40 kJ/mol; the overall coupled reaction becomes spontaneous as the negative ΔG from ATP hydrolysis outweighs the positive ΔG of the biosynthetic step. Such coupling ensures that non-spontaneous anabolic reactions proceed efficiently in the cellular environment. Metabolic pathways exemplify how spontaneity is achieved across multiple steps, even when individual reactions vary in their free energy changes. In glycolysis, the conversion of glucose to pyruvate consists of ten enzyme-catalyzed steps, some of which are endergonic under standard conditions but are pulled forward by subsequent exergonic steps, resulting in an overall negative free energy change of approximately -96 kJ/mol in human erythrocytes under steady-state conditions. This net exergonic nature makes the pathway spontaneous, producing ATP and enabling energy extraction from glucose while maintaining flux through the sequence. The pathway's design highlights how biological systems optimize spontaneity by integrating reactions into cascades where local equilibria are displaced by product removal or coupling. Spontaneous processes also underpin homeostasis in biological systems, particularly through the maintenance of ion gradients across membranes. Active transport, such as the , creates non-spontaneous electrochemical gradients (endergonic, requiring ATP), but these are sustained spontaneously in the broader context of open systems where the overall process dissipates heat, increasing total entropy. This heat dissipation, arising from metabolic inefficiencies, ensures compliance with the while allowing ordered structures like membrane potentials to persist. 's theory of dissipative structures, recognized with the 1977 Nobel Prize in Chemistry, provides the framework for understanding this far-from-equilibrium spontaneity in living organisms, where continuous influx of energy (e.g., from nutrients) and efflux of entropy (as heat and waste) enable self-organization and stability. A specific illustration of spontaneity in biology is DNA replication, which is thermodynamically favorable under intracellular conditions despite requiring enzymatic catalysis. The polymerization step involves phosphodiester bond formation between deoxyribonucleoside triphosphates (dNTPs), with a standard ΔG° of about +25 kJ/mol (endergonic due to the release of pyrophosphate). However, in the cellular milieu—with high dNTP concentrations (millimolar range) and low pyrophosphate levels maintained by pyrophosphatases—the actual ΔG becomes negative (approximately -20 to -30 kJ/mol per nucleotide), rendering the process spontaneous and driving faithful genome duplication. DNA polymerase enzymes lower the activation energy without altering the overall thermodynamics, ensuring rapid and accurate replication.

Equilibrium and Limitations

Approach to Equilibrium

In spontaneous processes, the thermodynamic driving force dictates the direction toward equilibrium, while the rate of approach is governed by kinetic factors. The change in Gibbs free energy, ΔG, serves as the indicator of this direction: for a process at constant temperature and pressure, ΔG is negative when the forward reaction is spontaneous, zero at equilibrium, and positive for the reverse direction. As the system evolves, ΔG progressively decreases until it reaches zero, signifying that the system has attained the state of minimum Gibbs free energy, where no further net change occurs. Le Chatelier's principle elucidates how systems at or near equilibrium respond to perturbations, ensuring the maintenance of spontaneity in the direction that restores balance. If a stress such as a change in concentration, pressure, or temperature is applied, the equilibrium shifts spontaneously to counteract the disturbance—for instance, an increase in reactant concentration prompts a shift toward products to reduce the excess. This principle underscores that such adjustments are inherently spontaneous, driven by the system's tendency to minimize deviations from the equilibrium state. The precise direction of spontaneity in chemical reactions is quantified by comparing the reaction quotient Q, which reflects current concentrations or activities, to the equilibrium constant K, which defines the equilibrium composition. \text{If } Q < K, \text{ the forward reaction is spontaneous; if } Q > K, \text{ the reverse reaction is spontaneous.} At equilibrium, Q = K, and ΔG = 0, halting net progress. This relationship holds for reversible processes, guiding the system inexorably toward the point. For multi-component systems, the Gibbs phase rule provides insight into the constraints at , specifying the F available to vary intensive variables like , , and without disrupting phase coexistence: F = C - P + 2 Here, C is the number of independent components, and P is the number of phases. This rule indicates that at equilibrium, the system's spontaneity is confined within these degrees of freedom, ensuring stable phase relations in complex mixtures such as alloys or solutions. In closed systems, where no matter or energy exchanges occur with the surroundings, spontaneous processes inevitably cease upon reaching the minimum , as any further change would require an increase in , violating of . This equilibrium state represents the most probable configuration, with maximized under the given constraints.

Non-Equilibrium Considerations

In open systems, spontaneous processes can sustain steady states far from global by maintaining continuous fluxes of and , resulting in dissipative structures that exhibit spatiotemporal order. These structures emerge through local increases in , which drive the system toward configurations that maximize dissipation while adhering to the second law of thermodynamics. A classic example is Rayleigh-Bénard , where a horizontal layer heated from below spontaneously forms ordered hexagonal convection cells once the surpasses a critical , enhancing efficiency. Irreversibility in such non-equilibrium dynamics underpins the thermodynamic , as the cumulative in open systems enforces a preferred direction for evolution, distinguishing past from future even in near-equilibrium regimes. This temporal asymmetry arises from the statistical improbability of reversing microscopic motions to undo macroscopic changes, linking local spontaneity to global irreversibility. However, classical thermodynamic criteria for spontaneity, such as minimization, are inherently limited to quasi-static processes near and break down in rapid or strongly nonlinear far-from-equilibrium scenarios, where gradients and fluxes become non-local and interdependent. To address these shortcomings, extended incorporates higher moments of the and relaxation times, providing a more robust framework for describing in such regimes. A foundational element of near-equilibrium analysis within this domain is Onsager's reciprocal relations, established in 1931, which assert the symmetry of phenomenological coefficients relating thermodynamic forces to conjugate fluxes, ensuring consistency with . Far-from-equilibrium chemical systems further illustrate these principles through spontaneous oscillatory behavior, as seen in the Belousov-Zhabotinsky reaction, an open-system process involving the oxidation of by in the presence of a metal catalyst, which generates self-sustained cycles of color changes and propagating waves without external forcing. These oscillations highlight how local spontaneity can perpetuate dynamic steady states, far exceeding the predictions of equilibrium thermodynamics.

References

  1. [1]
    Spontaneity – Chemistry - JMU Libraries Pressbooks
    A spontaneous process is one that occurs naturally under certain conditions. A nonspontaneous process, on the other hand, will not take place unless it is “ ...
  2. [2]
    Entropy and the 2nd & 3rd Laws of Thermodynamics
    Chemical Thermodynamics ; Second Law: In an isolated system, natural processes are spontaneous when they lead to an increase in disorder, or entropy.
  3. [3]
    Gibb's Free Energy - Chemistry 301
    If this standard free energy change is negative at a given temperature then we would call it "spontaneous." If it is positive, then it is not spontaneous (the ...
  4. [4]
    Spontaneity & Entropy
    A spontaneous process is one that, once started, continues on its own without input of energy. A non-spontaneous process needs a continual input of energy.
  5. [5]
    Entropy - CHEM 101
    The observed direction of a process is called the spontaneous direction and constitutes what we call a spontaneous process, and its unobserved opposite is the ...
  6. [6]
    [PDF] Spontaneous Processes - Columbia University
    So melting or vaporization increases entropy; freezing or condensation decrease entropy. Likewise, expansion of a gas increases entropy, compression of a gas ...<|separator|>
  7. [7]
    CHEM 245 - The Second Law of Thermodynamics
    The equality only holds for a special theoretical case - a reversible process. All other processes are irreversible which is the same as a spontaneous process.<|control11|><|separator|>
  8. [8]
    [PDF] Deciphering the physical meaning of Gibbs's maximum work equation
    Feb 29, 2024 · Spontaneous Process can take place only in such a direction as to cause dim- inution of free energy. It was Helmholtz who gave us the ...
  9. [9]
    Joseph Black and Latent Heat - American Physical Society
    The latent heat that Black discovered greatly slows the melting of snow and ice. He gave the first account of this work on April 23, 1762 at the University of ...
  10. [10]
    A History of Thermodynamics: The Missing Manual - PMC
    We present a history of thermodynamics. Part 1 discusses definitions, a pre-history of heat and temperature, and steam engine efficiency, which motivated ...
  11. [11]
    The Historical Development of Thermodynamics - ResearchGate
    Dec 16, 2015 · Thermodynamics as a wide branch of physics had a long historical development from the ancient times to the 20th century.
  12. [12]
    Boltzmann's Work in Statistical Physics
    Nov 17, 2004 · Particularly famous is his statistical explanation of the second law of thermodynamics. The celebrated formula S = k logW, expressing a relation ...
  13. [13]
    Max Planck and the birth of the quantum hypothesis - AIP Publishing
    Sep 1, 2016 · One of the most interesting episodes in the history of science was Max Planck's introduction of the quantum hypothesis, at the beginning of the 20th century.Missing: refinements | Show results with:refinements
  14. [14]
    [PDF] LECTURE NOTES ON THERMODYNAMICS
    May 17, 2025 · These are lecture notes for AME 20231, Thermodynamics, a sophomore-level undergraduate course taught in the Department of Aerospace and ...
  15. [15]
    Lord Kelvin | On the Dynamical Theory of Heat
    Lord Kelvin's paper extends the dynamical theory of heat, showing modifications to Carnot's work, and points out the significance of steam observations.
  16. [16]
    [PDF] Rudolf Clausius, “Concerning Several Conveniently ... - Le Moyne
    For each body there are found two quantities, the transformation value of its heat content and its disgregration, the sum of which is its entropy. This ...<|separator|>
  17. [17]
    Translation of Ludwig Boltzmann's Paper “On the Relationship ...
    Boltzmann shows that the statistical mechanical quantity he denotes by Ω (multiplied by 2/3) is equal to the thermodynamic quantity entropy (S) as defined by ...Missing: sources | Show results with:sources
  18. [18]
    6.5 Irreversibility, Entropy Changes, and ``Lost Work'' - MIT
    The entropy of a system can be altered in two ways: (i) through heat exchange and (ii) through irreversibilities.Missing: δQ/ | Show results with:δQ/
  19. [19]
    Albert Einstein and Walther Nernst's Heat Theorem, 1911–1916
    This paper discusses the early history of Walther Nernst's Heat Theorem and the first stages of its development into the Third Law of Thermodynamics.
  20. [20]
    19.4: Criteria for Spontaneous Change: The Second Law of ...
    Jul 12, 2023 · Key Concepts and Summary. The second law of thermodynamics states that a spontaneous process increases the entropy of the universe, Suniv > 0.Definition: The Second Law of... · Gibbs Energy and Changes of...
  21. [21]
    The Second and Third Laws of Thermodynamics - Lumen Learning
    As ΔSuniv < 0 at each of these temperatures, melting is not spontaneous at either of them. The given values for entropy and enthalpy are for NaCl at 298 K. It ...
  22. [22]
    Gibbs Free Energy
    Driving Forces and Gibbs Free Energy. Some reactions are spontaneous because they give off energy in the form of heat ( delta H < 0). Others are spontaneous ...Driving Forces & Gibbs Free... · The Effect of Temperature on...
  23. [23]
    4. Free Energy and Equilibrium
    Here we will introduce a new measure of thermodynamic energy that already factors in the entropy changes of the surroundings—the Gibbs free energy G, which is a ...
  24. [24]
    [PDF] On the equilibrium of heterogeneous substances : first [-second] part
    Equilibrium of Heterogeneous Substances. Em. 8. developable surface formed on the sheets (A) and (B), the surface ...
  25. [25]
    [PDF] Physical Memoirs Selected and Translated from Foreign Sources
    von HELMHOLTZ. ... On the assumption of the unrestricted applica bility of Clausius's law, it would therefore be the value of the free energy, not that of the ...
  26. [26]
    [PDF] Entropy and 2nd Law of Thermodynamics Reading Spontaneity
    ■ Spontaneous: “Occurring without. outside intervention.” ■ A reaction or change of state is said to. be spontaneous if it is thermodynamically allowed. ■ For ...
  27. [27]
  28. [28]
    Example: Free energy of an expanding gas - Nexus Wiki - ComPADRE
    Apr 22, 2019 · To get this, we will need to figure out what the entropy change is for an ideal gas whose temperature remains constant but whose volume changes.
  29. [29]
    [PDF] Chapter 14 (and 15.4): Entropy and Free Energy
    Second Law of Thermodynamics: the entropy of the universe increases in a spontaneous process and remains unchanged in an equilibrium process. Spontaneous: ...Missing: ideal | Show results with:ideal
  30. [30]
    [PDF] 10. Diffusion
    Aug 4, 2018 · Diffusion refers to the phenomenon by which concentration and temperature gradients spontaneously disappear with time, and the properties of the ...Missing: gases | Show results with:gases
  31. [31]
    Entropy, Enthalpy and Free Energy. What is a Spontaneous Process?
    A spontaneous process increases the entropy of the universe, and is defined by ΔG < 0, where ΔG is the Gibbs Free Energy.Missing: calculation | Show results with:calculation<|control11|><|separator|>
  32. [32]
    Free Energy – Chemistry - UH Pressbooks
    Key Equations. ΔG = ΔH − TΔS; ΔG = ΔG° + RT ln Q; ΔG° = −RT ln K. Chemistry ... The evaporation of one mole of water at 298 K has a standard free energy ...
  33. [33]
    Standard enthalpy of formation, Gibbs energy of formation, entropy ...
    The table includes standard enthalpy of formation (ΔH°f), Gibbs free energy of formation (ΔG°f), entropy (S°), and molar heat capacity (C_p) for organic ...
  34. [34]
    Taking entropy changes further - Chemguide
    The entropy change to the surroundings will be negative because of the cooling caused by the ammonium nitrate dissolving, but this is more than made up for by ...
  35. [35]
    16.1 Spontaneity | General College Chemistry II - Lumen Learning
    Iron exposed to the earth's atmosphere will corrode, but rust is not converted to iron without intentional chemical treatment. A spontaneous process is one that ...
  36. [36]
    ATP: Adenosine Triphosphate - OpenEd CUNY
    The calculated ∆G for the hydrolysis of one ATP mole into ADP and Pi is −7.3 kcal/mole (−30.5 kJ/mol). Since this calculation is true under standard conditions, ...
  37. [37]
  38. [38]
    Glycolysis - PMC - NIH
    The Gibbs free energy, ΔG, has to be either negative or close to 0 if glycolysis is to proceed in the forward direction. The ΔG°′ and ΔG of all 10 reactions are ...
  39. [39]
    Improved bounds on entropy production in living systems - PNAS
    Apr 27, 2021 · To perform essential functions, from sensing to replication and locomotion, organisms consume energy and dissipate heat. The rate at which they ...
  40. [40]
    Press release: The 1977 Nobel Prize in Chemistry - NobelPrize.org
    Prigogine has demonstrated that a new form of ordered structures can exist under such conditions, and he has given them the name ”dissipative structures” to ...
  41. [41]
    Dissipative structures in biological systems: bistability, oscillations ...
    From the very beginning of his scientific work, Ilya Prigogine devoted his attention to non-equilibrium self-organization in chemical, physical and biological ...<|control11|><|separator|>
  42. [42]
    18.6: Spontaneity and Equilibrium - Chemistry LibreTexts
    Jul 4, 2022 · Conversely, if Q > K, then the reaction proceeds spontaneously to the left as written, resulting in the net conversion of products to reactants.
  43. [43]
    19.6: Gibbs Energy Change and Equilibrium - Chemistry LibreTexts
    Jul 12, 2023 · In a spontaneous change, Gibbs energy always decreases and never increases. An important consequence of the one-way downward path of the free ...
  44. [44]
    Le Chatelier's Principle Fundamentals - Chemistry LibreTexts
    Jan 29, 2023 · Hence, Le Châtelier's principle states that any change to a system at equilibrium will adjust to compensate for that change. In 1884 the French ...
  45. [45]
    13.1: The Gibbs Phase Rule for Multicomponent Systems
    Apr 12, 2022 · The number of degrees of freedom is the maximum number of intensive properties of the equilibrium system we may independently vary, or fix at ...
  46. [46]
    [PDF] 1 General Gibbs Minimization as an Approach to Equilibrium Most of ...
    Gibbs minimization finds equilibrium by minimizing Gibbs free energy, which is a function of temperature, pressure, and composition, until a minimum is reached.
  47. [47]
    Benard Cells - an overview | ScienceDirect Topics
    Unlike a heat engine, the Bénard convection cells emerge spontaneously as the system responds to the throughput of energy. The configuration of the cells is ...
  48. [48]
    Time, Irreversibility and Entropy Production in Nonequilibrium Systems
    The aim of this review is to shed light on time and irreversibility, in order to link macroscopic to microscopic approaches to these complicated problems.
  49. [49]
    Extended Thermodynamics - an overview | ScienceDirect Topics
    One obvious limitation of Thermodynamics is ... Classical Thermodynamics or Equilibrium Thermodynamics has been shown according to different formulations.
  50. [50]
    In command of non-equilibrium - RSC Publishing
    May 5, 2016 · A well-known example is the Belousov–Zhabotinsky reaction, fuelled by the oxidation of malonic acid in the presence of bromate which in the ...