Fact-checked by Grok 2 weeks ago

Standard molar entropy

Standard molar entropy, denoted as S_m^\circ, is the absolute of one of a substance in its at a specified , typically 298.15 , and a of 1 . For solids and liquids, the standard state is the pure substance at that temperature and pressure, while for gases, it is the hypothetical at the same conditions. The units are joules per per (J K^{-1} mol^{-1}). These values are determined experimentally based on the third law of thermodynamics, which posits that the of a perfect crystalline substance approaches zero as the temperature approaches (0 ). For solids and liquids, S_m^\circ is calculated by calorimetric measurements of (C_p) from near 0 to 298.15 , integrating \int_0^{298.15} (C_p / T) \, dT, and adding contributions from phase transitions such as and if applicable. For gases, spectroscopic methods can also provide values by considering molecular rotational, vibrational, and translational contributions. Standard molar entropies exhibit clear trends across s and molecular structures: they are lowest for solids, higher for liquids, and highest for gases due to increasing molecular mobility and disorder. Within the same phase, S_m^\circ generally increases with and complexity, as heavier or more intricate molecules possess more accessible microstates. For example, has a low S_m^\circ of 2.4 J K^{-1} mol^{-1} owing to its rigid structure, while gaseous iodine (I_2) has a much higher value of 260.7 J K^{-1} mol^{-1}. In , molar are essential for computing the change of a , \Delta S^\circ = \sum S_m^\circ (\text{products}) - \sum S_m^\circ (\text{reactants}), following the products-minus-reactants rule. This \Delta S^\circ contributes to the change, \Delta G^\circ = \Delta H^\circ - T \Delta S^\circ, enabling predictions of spontaneity under conditions. Tabulated values from sources like the NIST Chemistry WebBook facilitate these calculations across diverse applications in chemistry and .

Fundamentals

Definition and Units

Standard molar entropy, denoted as S_m^\circ, is the absolute entropy content of one mole of a pure substance in its standard state, defined at a temperature of 298.15 K and a pressure of 1 bar (100 kPa). Entropy itself is a thermodynamic state function that quantifies the degree of or in a at the molecular level. The notation S_m^\circ specifically indicates the standard molar value, distinguishing it from the total entropy S of an entire or the mass-specific entropy s (typically in J⋅kg⁻¹⋅K⁻¹). In the (SI), standard molar entropy is expressed in joules per per (J⋅mol⁻¹⋅K⁻¹). Prior to widespread adoption of SI units, values were commonly tabulated in calories per per (cal⋅mol⁻¹⋅K⁻¹), with the conversion factor 1 cal = 4.184 J.

Standard State Conditions

The standard molar entropy, denoted as S_m^\circ, is defined under specific conditions to ensure uniformity in thermodynamic data reporting. The reference temperature is fixed at 298.15 (25 °C), which serves as the conventional standard for compiling thermodynamic properties such as , reflecting conditions near ambient where many chemical processes occur. This choice facilitates direct comparisons of entropy values across diverse substances and experimental setups. The standard pressure is bar, equivalent to $10^5 , as recommended by the International Union of Pure and Applied Chemistry (IUPAC) in 1982. Prior to this recommendation, atm (101 325 ) was commonly used, but the shift to bar aligns with the (SI) and introduces only a minor difference of approximately 0.1% in pressure, resulting in negligible impacts on entropy values for most substances—particularly solids and liquids, where pressure effects are minimal. For gases, the standard state assumes ideal behavior at this pressure, treating the substance as a hypothetical . Pure liquids and solids are considered in their stable forms at bar and 298.15 K, without phase changes unless specified. For aqueous solutions and solutes, the standard state corresponds to a hypothetical dilute at a of 1 mol kg⁻¹ (unit ) and 1 bar pressure, extrapolated from properties at infinite dilution to account for non- behaviors at finite concentrations. These conditions collectively promote consistency and reproducibility in , enabling accurate predictions of reaction spontaneity and without ambiguity from varying environmental parameters.

Thermodynamic Foundations

Entropy as a State Function

is a fundamental thermodynamic property classified as a , meaning its value for a system depends solely on the initial and final states, independent of the or connecting them. This path independence arises because the differential form of entropy change, derived from the second law of thermodynamics, constitutes an , ensuring that the integral of dS over any closed cycle vanishes. For instance, in a reversible cyclic , the condition \oint \frac{dQ_\text{rev}}{T} = 0 confirms that entropy returns to its initial value, solidifying its status as a alongside , , and . This property allows entropy changes to be computed using any convenient reversible between states, even if the actual is irreversible. The foundational definition of entropy change stems from the work of , who in introduced the concept to quantify the unavailable in transformations. Clausius defined the entropy change \Delta S as \Delta S = \int \frac{dQ}{T}, where dQ is the infinitesimal absorbed by the system and T is the absolute temperature in ; for finite changes along a reversible path, this becomes \Delta S = \int \frac{dQ_\text{rev}}{T}. In reversible processes—idealized scenarios where the system and surroundings remain in quasi-static equilibrium throughout—the infinitesimal entropy change is precisely dS = \frac{dQ_\text{rev}}{T}, as no dissipative effects like or unrestrained occur to generate additional . This relation highlights 's role in measuring the directional tendency of flow from hot to cold bodies, with reversible processes serving as the benchmark for calculating \Delta S in practical applications, such as phase transitions or expansions of ideal gases. A statistical mechanical interpretation of entropy was later provided by in 1877, bridging macroscopic with microscopic behavior. expressed entropy as S = k \ln W, where k is Boltzmann's constant and W represents the number of microscopic configurations (microstates) consistent with a given macroscopic state. This formula posits entropy as a measure of disorder or multiplicity: systems naturally evolve toward states of higher W, increasing entropy, which aligns with the second law's prediction of spontaneous irreversibility. 's approach not only justified Clausius's thermodynamic entropy but also enabled the computation of absolute entropies from molecular statistics, reinforcing entropy's nature through probabilistic averaging. The establishment of an absolute entropy scale relies on the third law of thermodynamics, originally formulated as the by in 1906. This law states that the entropy of a perfect crystalline substance approaches zero as the approaches (T \to 0 ), with no residual disorder in the . Consequently, values can be determined absolutely by integrating dS = \frac{dQ_\text{rev}}{T} from 0 upward, providing a universal that positions standard molar entropy within a rigorous, path-independent framework for comparing thermodynamic across .

Relation to Absolute Entropy

The standard molar entropy, S^\circ, quantifies the absolute entropy content of one mole of a substance in its standard state at 298.15 K and 1 bar pressure, representing the total entropy increase from (0 K) to this temperature under reversible conditions. This value is derived from the third of thermodynamics, which establishes that the entropy of a perfect crystalline substance is zero at 0 K, providing an unequivocal zero point for measurements. In contrast to standard enthalpies of formation (\Delta H_f^\circ), which are relative quantities referenced to the elements in their standard states, S^\circ is inherently absolute, as the third eliminates the need for an arbitrary reference. The mathematical expression for the standard molar entropy at temperature T is: S^\circ(T) = \int_0^T \frac{C_p(T')}{T'} \, dT' + \sum \Delta S_{\text{phase changes}} + S_0 where C_p is the at constant , the integral accounts for the entropy contribution from heating, \Delta S_{\text{phase changes}} includes entropy increments from transitions (calculated as \Delta H / T at the transition ), and S_0 = 0 per the third . This formulation, rooted in the , ensures that S^\circ captures all configurational and thermal disorder from the upward. The absolute nature of S^\circ has profound implications for thermodynamic analysis, particularly in calculating the standard entropy change for a reaction, \Delta S^\circ_{\text{rxn}}, simply as the difference \sum S^\circ_{\text{products}} - \sum S^\circ_{\text{reactants}}, without introducing extraneous constants or reference adjustments. This direct subtractive approach stems from entropy's status as a state function and enables precise evaluations of spontaneity and equilibrium in chemical systems.

Determination Methods

Experimental Measurement

The primary experimental approach to determining the standard molar entropy S^\circ of a substance relies on to measure the at constant , C_{p,\text{m}}, over a range from near 0 to 298.15 . This employs a vacuum-jacketed , where the sample is thermally isolated to minimize exchange with the surroundings, allowing precise incremental heating and monitoring. Seminal designs for such instruments, capable of achieving accuracies better than 0.5% in measurements, were developed in the mid-20th century. The standard molar entropy is then computed on the third-law by integrating the data, incorporating corrections for any transitions encountered in the range: S^\circ(298.15\ \text{K}) = \int_0^{298.15} \frac{C_{p,\text{m}}}{T} \, dT + \sum_i \frac{\Delta_{\text{trans},i} H_m}{T_{\text{trans},i}} Here, the integral accounts for the vibrational and other contributions to , while the summation includes changes from transitions, such as or , calculated as the molar of transition \Delta_{\text{trans},i} H_m divided by the transition T_{\text{trans},i}; these enthalpies are obtained from separate calorimetric measurements of latent heats. This method yields absolute entropies with high reliability for pure crystalline solids, liquids, and gases in their standard states. Indirect experimental determination of S^\circ can also be achieved through equilibrium measurements, such as the temperature dependence of in electrochemical cells, which provides entropy changes for cell reactions via \Delta S = nF (\partial E / \partial T)_p, or from data analyzed using the Clapeyron equation to derive entropies of when combined with known reference values. These approaches are particularly useful for systems where direct is challenging, like aqueous ions or volatile compounds. Accuracy in calorimetric measurements of S^\circ is typically limited by factors such as sample impurities, which can introduce contributions, and deviations from ideal behavior at low temperatures, leading to uncertainties on the order of ±0.1 J⋅⁻¹⋅⁻¹ for well-characterized pure substances. Modern automated adiabatic calorimeters enhance precision by reducing manual errors and improving , often achieving overall uncertainties below ±0.5% for the integrated .

Computational Estimation

Computational estimation of standard molar entropy relies on theoretical frameworks from statistical mechanics and quantum chemistry, providing predictive tools when experimental data is lacking. In statistical mechanics, the standard molar entropy S^\circ for an ideal gas can be derived from the molecular partition function q, which encapsulates translational, rotational, vibrational, and electronic degrees of freedom. The expression is given by S^\circ = R \left[ \ln \left( \frac{q}{T} \right) + 1 + T \frac{\partial \ln q}{\partial T} \right], where R is the and the derivative is taken at constant volume; this formula arises from the relation between and the , with q computed as the product of contributions from each mode. Quantum chemistry methods enable ab initio calculation of the components by solving the for molecular wavefunctions, typically at the Hartree-Fock or post-Hartree-Fock levels. Software such as Gaussian performs these computations by optimizing , calculating vibrational frequencies via the , and evaluating rotational constants from the moments of inertia. The vibrational contribution dominates for polyatomic molecules and is obtained from the , while rotational and translational terms follow and particle-in-a-box models; electronic contributions are often negligible except for open-shell systems. These methods yield standard molar entropies accurate to within 1-5 J⋅mol⁻¹⋅K⁻¹ for small gas-phase molecules when using high-level basis sets like cc-pVTZ. For larger organic molecules, group additivity methods offer a semi-empirical alternative, decomposing the molecule into structural fragments and summing their pre-tabulated entropy contributions. Benson's group additivity scheme, developed in the 1960s, estimates S^\circ by assigning values to polyvalent atoms (e.g., C-(C)₂(H)₂ for a methylene group) and applying corrections for ring strain, stereochemistry, and gauche interactions. This approach is particularly effective for hydrocarbons and functionalized organics, achieving typical accuracies of 5-10 J⋅mol⁻¹⋅K⁻¹ for gas-phase standard molar entropies at 298 K. Despite these advances, computational estimation is most reliable for gases, where ideal behavior and separable degrees of freedom simplify partition function evaluation. For solids and liquids, additional corrections are necessary to account for lattice vibrations (phonons) using models like the Debye or Einstein approximation, intermolecular correlations, and phase-specific free volumes, as the independent-particle assumptions underlying molecular partition functions break down in condensed phases.

Applications and Values

Use in Reaction Thermodynamics

Standard molar entropy values are integral to evaluating the spontaneity of chemical reactions through the change, given by the equation \Delta G^\circ = \Delta H^\circ - T \Delta S^\circ, where \Delta S^\circ is calculated as the difference between the sum of the standard molar entropies of products and reactants, multiplied by their stoichiometric coefficients. At standard temperature (298 ), a negative \Delta G^\circ indicates a spontaneous reaction under standard conditions, with \Delta S^\circ often determining favorability when \Delta H^\circ is positive or small. The reaction entropy change \Delta S^\circ provides insight into disorder variations; for instance, in combustion reactions like the oxidation of (\ce{CH4(g) + 2O2(g) -> CO2(g) + 2H2O(g)}), \Delta S^\circ is slightly negative (approximately -5 J/mol·K) due to the comparable number of gas moles on both sides, reflecting minimal net change in translational freedom. In synthesis reactions involving gas-to-liquid or gas-to-solid transitions, such as the formation of , \Delta S^\circ is more markedly negative (around -198 J/mol·K), signifying a decrease in entropy as multiple gas molecules combine into fewer product molecules. Standard molar entropy data also informs the temperature dependence of equilibrium constants via the van't Hoff in its integrated form: \ln K = -\frac{\Delta H^\circ}{RT} + \frac{\Delta S^\circ}{R}, where \Delta S^\circ determines the entropy contribution to the equilibrium across temperatures, assuming \Delta H^\circ and \Delta S^\circ are approximately . This allows prediction of how reaction favorability shifts; for endothermic reactions with positive \Delta S^\circ, higher temperatures increase K, while the opposite holds for exothermic processes with negative \Delta S^\circ. A practical illustration is the Haber-Bosch ammonia synthesis (\ce{N2(g) + 3H2(g) -> 2NH3(g)}), where the negative \Delta S^\circ (from four gas moles to two) makes \Delta G^\circ less negative at higher temperatures, necessitating high pressures to favor the forward reaction by reducing the entropic penalty through . This entropy-driven strategy optimizes industrial yields despite the reaction's exothermicity.

Tabulated Standard Values

Standard molar entropy values, denoted as S^\circ, are compiled in comprehensive databases maintained by authoritative organizations. The National Institute of Standards and Technology (NIST) Chemistry WebBook provides extensively reviewed thermochemical data for thousands of substances, including absolute entropies derived from experimental and spectroscopic measurements. Similarly, the CRC Handbook of Chemistry and Physics offers tabulated values based on critically evaluated literature, while IUPAC recommendations ensure consistency across international standards. These compilations are periodically revised to incorporate new data and align with updated conventions. For instance, the 1982 IUPAC adoption of 1 as the (replacing 1 ) resulted in minor adjustments to gas-phase values, generally increasing them by ln(1 / 1 ) ≈ 0.11 J·⁻¹·⁻¹. The following table lists selected S^\circ values at 298.15 K and 1 for representative and compounds across , , and gas phases, illustrating the diversity in magnitudes from these sources.
SubstanceStateS^\circ (J·mol⁻¹·K⁻¹)Source
C (graphite)s5.74NIST Chemistry WebBook
NaCls72.1NIST Chemistry WebBook
H₂Ol69.9NIST Chemistry WebBook
H₂g130.7NIST Chemistry WebBook
CH₄g186.3NIST Chemistry WebBook
H₂Og188.8NIST Chemistry WebBook
O₂g205.1NIST Chemistry WebBook
CO₂g213.8NIST Chemistry WebBook
These examples span the common range of S^\circ values, typically 5–10 J·mol⁻¹·K⁻¹ for crystalline solids, around 70 J·mol⁻¹·K⁻¹ for liquids, and 130–220 J·mol⁻¹·K⁻¹ for gases, as documented in the referenced compilations.

Across the Periodic Table

Standard molar entropy values for elements exhibit clear trends across the periodic table, primarily influenced by and molecular properties at standard conditions (298 K, 1 ). Within groups, entropy generally increases from top to bottom due to larger radii and masses, which enable lower-frequency vibrational modes and greater positional disorder in the . This effect is evident in the alkali metals (), where the value rises progressively: at 29.12 J⋅mol⁻¹⋅K⁻¹, sodium at 51.21 J⋅mol⁻¹⋅K⁻¹, at 64.18 J⋅mol⁻¹⋅K⁻¹, at 69.45 J⋅mol⁻¹⋅K⁻¹, and cesium at 85.23 J⋅mol⁻¹⋅K⁻¹. Similar patterns occur in other groups, such as the alkaline earth metals, where beryllium's compact structure yields a low 9.50 J⋅mol⁻¹⋅K⁻¹ compared to barium's 62.8 J⋅mol⁻¹⋅K⁻¹. Across periods, trends for solid elements show that standard molar entropy generally increases with rising due to increasing mass that enhances vibrational contributions, but this is often tempered or reversed by strengthening and tighter atomic packing that restricts motion, particularly after the alkali metals. For instance, in 3 solids, sodium has 51.21 J⋅mol⁻¹⋅K⁻¹, decreasing to magnesium at 32.7 J⋅mol⁻¹⋅K⁻¹ and further to aluminum at 28.3 J⋅mol⁻¹⋅K⁻¹, reflecting a balance between mass gain and these structural factors. For gaseous elements, particularly diatomic molecules in later periods, entropy increases with molecular mass and , as seen in the : fluorine (F₂) at 202.79 J⋅mol⁻¹⋅K⁻¹ versus chlorine (Cl₂) at 223.08 J⋅mol⁻¹⋅K⁻¹, owing to more rotational and translational freedom in heavier molecules. However, these gaseous trends introduce complexity from additional absent in . Allotropes of the same can display markedly different standard molar entropies due to variations in and resulting disorder. Carbon provides a classic example: , with its rigid three-dimensional covalent network, has a low value of 2.38 J⋅mol⁻¹⋅K⁻¹, while graphite's layered structure permits greater interlayer vibrations and modes, yielding 5.74 J⋅mol⁻¹⋅K⁻¹. This difference arises from graphite's higher and softer vibrational spectrum compared to 's highly ordered, low-entropy . Several atomic and structural factors underpin these periodic variations. influences entropy through electronic contributions, which are periodic and more pronounced in transition metals due to degenerate orbitals, though typically small (~1-5 J⋅mol⁻¹⋅K⁻¹) relative to vibrational terms. Bonding type plays a key role: in elements like the alkali metals allows delocalized electrons and softer lattices, elevating entropy compared to covalent networks in carbon allotropes or semiconductors like (18.83 J⋅mol⁻¹⋅K⁻¹). The standard state further modulates values; for instance, bromine's liquid state (Br₂, 152.21 J⋅mol⁻¹⋅K⁻¹) confers higher entropy than iodine's solid (I₂, 116.14 J⋅mol⁻¹⋅K⁻¹), despite iodine's greater mass, because liquids exhibit more molecular freedom than solids.

Effects of Temperature and Pressure

The standard molar entropy S^\circ is defined at 298 and 1 , but for conditions at other s, the molar entropy S(T) can be calculated using the relation
S(T) = S^\circ + \int_{298}^{T} \frac{C_p}{T} \, dT,
where C_p is the at constant . This accounts for the increased as rises, allowing access to more vibrational, rotational, and translational microstates, which generally leads to higher values for T > 298 since C_p > 0. For many substances, C_p is approximately constant over moderate temperature ranges, simplifying the to \Delta S \approx C_p \ln(T / 298). Extensions of Kirchhoff's law principles apply this correction to reaction entropies or individual species, ensuring thermodynamic consistency across temperatures.
For pressure effects at constant temperature, the change in molar entropy for an ideal gas is given by
\Delta S = -R \ln \left( \frac{P}{P^\circ} \right),
where R is the gas constant and P^\circ = 1 bar, reflecting decreased entropy with compression due to reduced volume and fewer positional microstates. This effect is significant for gases but negligible for liquids and solids, where the partial derivative \left( \frac{\partial S}{\partial P} \right)_T = -\left( \frac{\partial V}{\partial T} \right)_P yields small values owing to low compressibility and thermal expansion.
As an example of temperature correction, the standard molar entropy of water vapor is S^\circ = 188.84 J mol^{-1} K^{-1} at 298 K. At 373 K, using an average C_p \approx 33.6 J mol^{-1} K^{-1}, the correction is approximately \Delta S \approx 33.6 \ln(373/298) \approx 7.6 J mol^{-1} K^{-1}, yielding S(373 \, \text{K}) \approx 196.4 J mol^{-1} K^{-1}. More precise values incorporate temperature-dependent C_p via Shomate equations from thermochemical databases. For real gases at high pressures, non-ideal behaviors require corrections beyond the formula, often using f as an effective that accounts for intermolecular forces and volume exclusions. The adjustment then involves \Delta S \approx -R \ln(f / P^\circ), where the fugacity coefficient \phi = f / P deviates from unity, typically computed from equations of state like Peng-Robinson. These corrections become essential above several bars, where effects can alter by several percent compared to ideal predictions.

References

  1. [1]
    13.6: The Third Law of Thermodynamics - Chemistry LibreTexts
    May 13, 2023 · ... standard molar entropy ( S o ), which is the entropy of 1 mol of a substance under standard pressure (1 bar). Often the standard molar entropy ...
  2. [2]
    16.7: Standard Molar Entropies - Chemistry LibreTexts
    Jul 19, 2023 · ... standard molar entropy, i.e., as the entropy of 1 mol of substance at the standard pressure of 1 atm (101.3 kPa) and given temperature.
  3. [3]
    6.2: Molar Entropies - Chemistry LibreTexts
    Apr 12, 2022 · The entropy at the temperature and pressure of interest, then, is the entropy change Δ ⁢ S = ∫ 0 T ′ d q / T of a reversible heating process at ...
  4. [4]
    16.9: Some Trends In Entropy Values - Chemistry LibreTexts
    Jul 20, 2022 · The most obvious feature of the table of molecular entropies is a general increase in the molar entropy as we move from solids to liquids to gases.
  5. [5]
    19.4: Entropy Changes in Chemical Reactions - Chemistry LibreTexts
    Jul 7, 2023 · Standard molar entropies are listed for a reference temperature (like 298 K) and 1 atm pressure (i.e. the entropy of a pure substance at 298 K ...
  6. [6]
    standard state (S05925) - IUPAC Gold Book
    Three standard states are recognized: For a gas phase it is the (hypothetical) state of the pure substance in the gaseous phase at the standard pressure \(p=p^ ...Missing: temperature | Show results with:temperature
  7. [7]
    IUPAC - standard pressure (S05921)
    - **Definition**: Standard pressure is a chosen value of pressure denoted by \( p^\circ \) or \( P^\circ \).
  8. [8]
  9. [9]
  10. [10]
    NIST Guide to the SI, Chapter 4: The Two Classes of SI Units and ...
    Jan 28, 2016 · Consider, for example, the quantity molar entropy: the unit J/ (mol · K) is obviously more easily understood than its SI base−unit equivalent, m ...
  11. [11]
    [PDF] Quantities, Units and Symbols in Physical Chemistry - IUPAC
    kelvin (symbol: K). The kelvin, unit of thermodynamic temperature, is the fraction 1/273.16 of the thermodynamic temperature of the triple point of water.
  12. [12]
  13. [13]
    18.4 Entropy Changes and the Third Law of Thermodynamics
    The third law of thermodynamics states that the entropy of any perfectly ordered, crystalline substance at absolute zero is zero. At temperatures greater than ...
  14. [14]
    [PDF] The NBS Tables of Chemical Thermodynamic Properties
    Values for chemical thermodynamic properties data are tabulated by convention as enthalpies of formation, Gibbs energies of formation, and entropies, all for ...
  15. [15]
    Calorimetric Determination of Heat Capacity, Entropy and Enthalpy ...
    The values of standard molar entropies at 298.15 K, S m(298.15), were derived from the low-temperature C p m data (LT fit) by numerical integration of the C p m ...
  16. [16]
    [PDF] Solution and Low-Temperature Heat Capacity Measurements
    Heat capacity data are used to calculate the entropy from the third law of thermo- dynamics. The heat capacity calorimeter is of the adiabatically shielded type ...
  17. [17]
  18. [18]
    None
    ### Summary of Entropy Calculation from Partition Functions for Gases
  19. [19]
    [PDF] Thermochemistry in Gaussian
    Jun 2, 2000 · Abstract. The purpose of this paper is to explain how various thermochemical values are computed in Gaussian. The paper documents what ...
  20. [20]
    Critical Evaluation of Thermochemical Properties of C1–C4 Species
    For these cases, the group additivity (GA) method developed by Benson6 can be used to estimate thermochemical data. Each “heavy” (nonhydrogen) atom in a ...
  21. [21]
    Group additivity values for entropy and heat capacities of C 2
    The group additivity (GA) method, developed by Benson and co-workers [1,2] is widely used to estimate thermochemical properties including heat of formation, ...
  22. [22]
    Philosophy of Statistical Mechanics
    Jan 10, 2023 · ... liquids and solids. But important applications of SM are to systems that are not dilute gases and so this is a significant limitation.<|control11|><|separator|>
  23. [23]
  24. [24]
    Gibbs free energy and spontaneity (article) - Khan Academy
    The second law of thermodynamics says that the entropy of the universe always increases for a spontaneous process: Δ S universe = Δ S system + Δ S surroundings ...
  25. [25]
    5.5 Entropy and the Second and Third Laws of Thermodynamics
    Use the data given to calculate the standard molar entropy change for the synthesis of ammonia from its elements: N2 (g) + 3 H2 (g) ⟶ 2 NH3 (g). S°298 for N ...
  26. [26]
    NIST Chemistry WebBook
    This site provides thermochemical, thermophysical, and ion energetics data compiled by NIST under the Standard Reference Data Program.A Guide to the NIST Chemistry... · NIST Organic... · Chemical Name Search · Author
  27. [27]
    Graphite
    ### Summary of Standard Molar Entropy S° for C Graphite at 298 K
  28. [28]
    sodium chloride
    - **Standard Molar Entropy (S°) for NaCl Solid at 298 K**:
  29. [29]
    Water
    - **Standard Molar Entropy (S°) for H2O Liquid at 298 K:**
  30. [30]
    Hydrogen
    - **Standard Molar Entropy (S°) for H₂ Gas at 298 K:**
  31. [31]
  32. [32]
    Water
    ### Summary of Standard Molar Entropy S° for H2O Gas at 298 K
  33. [33]
    Oxygen
    - **Standard Molar Entropy (S°) for O2 Gas at 298 K:**
  34. [34]
    Carbon dioxide
    ### Summary of Standard Molar Entropy (S°) for CO2 Gas at 298 K
  35. [35]
  36. [36]
  37. [37]
  38. [38]
  39. [39]
    Cesium
    ### Summary of Standard Molar Entropy for Solid Cs at 298 K
  40. [40]
  41. [41]
  42. [42]
  43. [43]
  44. [44]
  45. [45]
  46. [46]
    Graphite - the NIST WebBook
    Graphite. Formula: C; Molecular weight: 12.0107. IUPAC Standard InChI: InChI=1S/C Copy. InChI version 1.06. IUPAC Standard InChIKey: OKTJSMMVPCPJKN-UHFFFAOYSA-N
  47. [47]
  48. [48]
  49. [49]
    Bromine
    - **Standard Molar Entropy Value for Liquid Br₂ at 298 K:**
  50. [50]
    Iodine - the NIST WebBook
    Entropy of liquid at standard conditions (1 bar). S°solid, Entropy of solid at standard conditions. S°solid,1 bar, Entropy of solid at standard conditions (1 ...Condensed phase... · References
  51. [51]
    [PDF] 485 Chapter 13: Entropy and Applications
    in the data section at the back of this text. The temperature dependence of the entropy change for chemical reactions is given by Eq. 13.3.4 and 13.3.7 with ...
  52. [52]
    [PDF] J /mol K
    One mole of an ideal gas at 300.0 K is reversibly and isothermally compressed from a volume of 25.0 L to a volume of 10.0 L. Constant temperature is maintained.
  53. [53]
    [PDF] Physical chemistry I, 23421
    Jun 14, 2021 · To find entropy at a different temperature ∆S = R. T2. T1. (CP/T)dT. Reactions outside electrochemical cells are irreversible and so measuring ...<|control11|><|separator|>
  54. [54]
    [PDF] Thermodynamics equation
    ENTROPY CHANGE OF IDEAL GAS: CONSTANT SPECIFIC HEAT. S2 − S1 = cv ... S2 − S1 = S°2 − S°1 − R ln. P2. P1. (. kJ kg K. ) (Exact Analysis: for ...
  55. [55]
    [PDF] Chapter 16 Foundations of Thermodynamics Problems
    Dividing the Gibbs-Helmholtz equation by – R and specifying standard state pressure gives: ... change in entropy for an isothermal process in an ideal gas.
  56. [56]
    Water
    ### Summary of H2O Gas Entropy and Shomate Parameters
  57. [57]
    Water - the NIST WebBook
    Morita, Sato, et al., 1989, Uncertainty assigned by TRC = 0.25 bar; based on analysis of their obs. PVT and vapor pressure data some other data from literature; ...
  58. [58]
    [PDF] Non-Ideality Through Fugacity and Activity - University of Delaware
    We have defined the fugacity for pure gases and liquids, as well as for species ”i” in a gas and liquid mixture. • Fugacity is a direct measure of the chemical ...Missing: correction | Show results with:correction