Fact-checked by Grok 2 weeks ago

Van der Waals equation

The Van der Waals equation is a fundamental thermodynamic that modifies the to describe the behavior of real gases and liquids by incorporating the finite volume of molecules and the attractive forces between them. Developed by in his 1873 doctoral dissertation, it takes the form \left(P + \frac{an^2}{V^2}\right)(V - nb) = nRT, where P is the pressure, V is the volume, n is the number of moles, T is the absolute temperature, R is the universal gas constant, a represents the strength of intermolecular attractions, and b accounts for the due to molecular size. This equation provides a unified framework for understanding the continuity between gaseous and liquid phases, addressing key deviations from ideal behavior observed at high pressures and low temperatures. Van der Waals derived the equation by building on Rudolf Clausius's , recognizing that the assumptions of point-like molecules with no interactions fail for real substances. He introduced the pressure correction term \frac{an^2}{V^2} to compensate for the reduced exerted by molecules on walls due to mutual , and the volume correction nb to reflect the actually unavailable for molecular motion. These adjustments were motivated by experimental observations, such as the non-linearity in PV versus P plots for compressed gases, and aimed to explain processes without invoking a sharp distinction between gas and liquid states. Van der Waals's work, initially met with skepticism, gained prominence through experimental validations by researchers like and earned him the in 1910 for advancing the understanding of equations of state for gases and liquids. One of the equation's most notable features is its prediction of a critical point, beyond which the liquid-gas phase distinction disappears, marking the end of the two-phase coexistence region. For a given gas, the critical parameters are T_c = \frac{8a}{27Rb}, P_c = \frac{a}{27b^2}, and V_c = 3nb, derived directly from the equation's cubic form in volume. This leads to the law of corresponding states, which posits that all gases exhibit similar behavior when expressed in reduced variables (P_r = P/P_c, V_r = V/V_c, T_r = T/T_c), yielding a universal equation \left(P_r + \frac{3}{V_r^2}\right)(3V_r - 1) = 8T_r. The equation also models phase transitions via the Maxwell construction, where the area under the van der Waals isotherm loop equals zero to enforce thermodynamic stability during liquid-vapor equilibrium. While approximate and empirical in nature—requiring gas-specific a and b values determined experimentally—it laid the groundwork for more advanced equations of state and remains widely used in for predicting real-gas properties.

Introduction and Formulation

Equation Statement

The van der Waals equation provides a more accurate description of real gas behavior compared to the ideal gas law by incorporating corrections for intermolecular attractions and the finite volume of gas molecules./16%3A_The_Properties_of_Gases/16.02%3A_van_der_Waals_and_Redlich-Kwong_Equations_of_State) For one mole of gas, the standard form of the equation is \left( P + \frac{a}{V_m^2} \right) (V_m - b) = RT, where P is the pressure, V_m is the molar volume, T is the temperature, R is the universal gas constant, a accounts for attractive forces between molecules, and b accounts for the repulsive volume occupied by the molecules themselves./02%3A_Gas_Laws/2.12%3A_Van_der_Waals%27_Equation) An equivalent form expresses the equation in terms of reduced variables, which normalize the thermodynamic properties relative to their critical values and reveal universal behavior across different gases: P_r = \frac{8 T_r}{3 V_r - 1} - \frac{3}{V_r^2}, where the reduced pressure is P_r = P / P_c, the reduced temperature is T_r = T / T_c, and the reduced volume is V_r = V_m / V_c, with P_c, T_c, and V_c denoting the critical , , and , respectively./16%3A_The_Properties_of_Gases/16.02%3A_van_der_Waals_and_Redlich-Kwong_Equations_of_State) The constants a and b are specific to each gas and are typically reported in units of \mathrm{L^2 \cdot [bar](/page/Bar) \cdot mol^{-2}} for a and \mathrm{L \cdot [mol](/page/Mol)^{-1}} for b; for , representative values are a = 3.59 \, \mathrm{L^2 \cdot [bar](/page/Bar) \cdot [mol](/page/Mol)^{-2}} and b = 0.043 \, \mathrm{L \cdot [mol](/page/Mol)^{-1}}.

Physical Parameters and Interpretation

The parameter a in the Van der Waals equation accounts for the attractive intermolecular forces between gas molecules, which reduce the effective exerted on the walls of the . These forces arise from long-range attractions that pull molecules back toward the interior of the gas, effectively diminishing the transferred to the surface during collisions. In his original formulation, van der Waals approximated these attractions using a mean-field approach, treating the gas as having a uniform average and reducing the distributed attractive forces to an equivalent acting inward on the molecules near the wall. This correction conceptually modifies the law's term by adding a positive contribution proportional to the square of the molecular , reflecting how attractions counteract the outward push in denser gases. The parameter b, on the other hand, represents the finite volume occupied by the gas themselves, which excludes a portion of the total container volume from being available for molecular motion. Unlike the assumption of point particles, real have a non-zero size, leading to an effective free volume of V - nb for n of gas. Van der Waals derived b as approximately four times the actual volume of the per , based on a simple geometric model where each excludes a spherical region around itself to avoid overlap with neighbors. This correction addresses the repulsive interactions at close range due to molecular hardness, becoming particularly significant at high densities where molecular crowding limits the gas's . Qualitatively, these parameters provide a physical bridge from the to behavior: a enhances the relative to the measured external to account for attractive "pulling back," while b diminishes the effective volume to incorporate the "occupied space" of molecules, as introduced in the equation's formulation earlier. This dual correction captures the molecular origins of deviations from ideality without relying on detailed quantum or , relying instead on empirical constants tailored to specific gases.

Historical Context

Discovery by van der Waals

In 1873, submitted his doctoral dissertation at , marking a pivotal moment in the study of gases and liquids. Titled Over de Continuïteit van den Gas- en Vloeistoftoestand (On the Continuity of the Gaseous and Liquid States), the thesis addressed fundamental discrepancies in the behavior of real gases under conditions approaching . was particularly motivated by experimental observations of deviations from the , such as those reported by in 1869, which demonstrated the existence of critical temperatures and pressures where gases transition continuously to liquids without abrupt changes. These experiments highlighted how the model failed to account for the and behavior of substances like at higher densities. The core insight of van der Waals' work stemmed from his recognition that real gases exhibit non-ideal behavior primarily due to two factors: the finite occupied by molecules themselves, which reduces the effective space available for gas motion, and the attractive forces between molecules, which diminish the exerted on container walls. Drawing inspiration from ' 1857 treatise on the kinetic theory of heat, van der Waals sought to extend molecular concepts to denser states, proposing that gases and liquids represent a continuum rather than distinct phases. This perspective led him to formulate an that incorporated corrections for these molecular effects, providing the first theoretical framework to unify gaseous and liquid properties under varying conditions of , , and . Van der Waals' 1873 dissertation not only resolved key inconsistencies in existing gas laws but also laid the groundwork for understanding intermolecular interactions. For his groundbreaking contributions to the equation of state for gases and liquids, he was awarded the in .

Subsequent Developments

In the years following the introduction of the equation in 1873, himself contributed to its broader thermodynamic implications through lectures that emphasized the concept of corresponding states, positing that all fluids exhibit similar behavior when their pressure, volume, and temperature are scaled by their critical values. This idea, rooted in the equation's structure accounting for intermolecular forces and finite molecular size, provided a framework for comparing diverse substances and was later formalized in subsequent theoretical developments, influencing the understanding of phase equilibria across gases and liquids. A significant extension came in 1875 when James Clerk Maxwell applied the Van der Waals equation to , addressing the unphysical oscillations in subcritical isotherms by introducing the equal-area rule—also termed the common tangent construction—which ensures thermodynamic consistency by equating the areas above and below the coexistence pressure in the pressure-volume diagram. This construction, proposed in Maxwell's analysis of behavior, resolved the equation's prediction of metastable states and laid groundwork for modeling liquid-vapor transitions, with further refinements by contemporaries like J. Willard Gibbs in the late 1870s. In 1881, initiated experimental verifications of the Van der Waals equation, motivated by its predictions for low-temperature gas properties, through precise measurements of isotherms for various gases that tested deviations from ideal behavior and the law of corresponding states. Establishing a cryogenic laboratory at , Onnes' early work focused on monatomic and diatomic gases, providing empirical data that confirmed the equation's qualitative accuracy for real gases under compressed conditions, though quantitative discrepancies highlighted needs for refinements like virial expansions. By the early 1900s, his ongoing experiments extended these validations, notably through his production of starting in 1906, enabling the first liquefaction of in 1908, and subsequent experiments at low temperatures.

Relation to Ideal Gas Law

Modifications for Molecular Interactions

The Van der Waals equation modifies the to incorporate the effects of intermolecular attractions and repulsions in real gases, providing a more accurate description under conditions where molecular interactions are significant. These adjustments address two primary deviations from ideal behavior: the finite size of gas molecules, which reduces the effective available for , and the attractive forces between molecules, which diminish the exerted on the walls. The correction accounts for intermolecular attractions that pull molecules toward each other, reducing the force of collisions with the walls compared to an . To compensate, the observed P is augmented by a term \frac{a}{V_m^2}, where V_m is the and a is a positive constant reflecting the strength of these attractions; higher a values indicate stronger forces, as seen in gases like CO₂ with a = 3.59 L² atm/mol². This effective P + \frac{a}{V_m^2} is then used in the equation, ensuring the model captures how attractions lower the measured , particularly at higher densities. The volume correction addresses the repulsive interactions arising from the physical size of molecules, which occupy space and exclude a portion of the total volume from being available for free movement. The effective volume is thus V_m - b, where b is an empirical constant representing the per due to molecular co-volume; for example, CO₂ has b = 0.0427 L/mol. This subtraction prevents overestimation of the available space, becoming more critical at high pressures where molecules are closely packed. Combining these corrections, the Van der Waals equation for one of gas is given by: \left(P + \frac{a}{V_m^2}\right)(V_m - b) = RT At low densities (high V_m or low pressure), both correction terms become negligible, and the equation approaches the PV_m = RT, as \frac{a}{V_m^2} \approx 0 and b \ll V_m. However, deviations grow near conditions, where attractions dominate and cause the gas to be less compressible than predicted by the ideal law, while the volume correction further highlights the impact of repulsions in dense states.

Derivation from Kinetic Theory

The derivation of the van der Waals equation begins with the , PV = nRT, which arises from kinetic theory assuming point-like molecules with no intermolecular forces and collisions contributing to via to the container walls. To account for the finite size of molecules, which violates the point-particle assumption, the effective free volume available for molecular motion is reduced by the occupied by the molecules themselves; this leads to the correction P(V - nb) = nRT, where b represents the excluded volume per mole due to hard-sphere repulsions. Intermolecular attractions further modify the term in kinetic theory. Molecules approaching the container wall experience a backward pull from attractions to neighboring molecules, reducing the rate of wall collisions and thus the observed ; in the mean-field , this reduction is \delta P \approx -a n^2 / V^2, where a quantifies the average pairwise attraction strength proportional to the square of molecular n/V. The kinetic is therefore the measured plus this correction, yielding the full van der Waals equation: \left( P + \frac{a n^2}{V^2} \right) (V - n b) = n R T This form assumes isotropic attractions acting equally in all directions, hard-sphere-like repulsions for volume exclusion, and a mean-field treatment that averages interactions while neglecting density fluctuations and correlations. The parameters a and b thus interpret the strength of attractions and the scale of repulsive exclusions, respectively.

Critical Phenomena

Critical Point Definition

The critical point of a substance, as described by the Van der Waals equation of , represents the specific , , and volume conditions where the and vapor phases become indistinguishable, terminating the line of vapor- equilibrium in the . This point signifies the boundary beyond which increasing cannot induce , resulting in a with no . Mathematically, the critical point occurs at the of the - isotherm, where the first and second partial derivatives of with respect to at constant vanish: \left( \frac{\partial P}{\partial V} \right)_T = 0, \quad \left( \frac{\partial^2 P}{\partial V^2} \right)_T = 0. These conditions are applied to the Van der Waals equation \left( P + \frac{a}{V^2} \right) (V - b) = RT, yielding the critical parameters in terms of the equation's constants a (related to intermolecular attractions), b (related to molecular ), and the R: the critical molar V_c = 3b, the critical P_c = \frac{a}{27b^2}, and the critical T_c = \frac{8a}{27Rb}. For (CO₂), experimental measurements align closely with these predictions when a and b are fitted accordingly, giving a critical T_c \approx 304 and critical pressure P_c \approx 73.8 bar.

Principle of Corresponding States

The principle of corresponding states, first articulated by in his 1873 doctoral thesis, asserts that the thermodynamic properties of all fluids are universal when expressed using reduced variables scaled by their respective critical parameters. This means that substances at the same reduced T_r = T / T_c, reduced pressure P_r = P / P_c, and reduced V_r = V / V_c exhibit identical behavior, regardless of their molecular differences, provided they follow a similar . The critical T_c, pressure P_c, and volume V_c represent the conditions at the fluid's critical point, where the distinction between liquid and gas phases vanishes. This universality emerges from the structural homogeneity of the van der Waals equation of state, \left( P + \frac{a}{V_m^2} \right) (V_m - b) = R T, where a and b are substance-specific constants accounting for intermolecular attractions and repulsions. Substituting the reduced variables and expressing a and b in terms of the critical parameters—specifically, V_c = 3b, T_c = \frac{8a}{27 R b}, and P_c = \frac{a}{27 b^2}—eliminates the dependence on a and b. The resulting reduced equation is \left( P_r + \frac{3}{V_r^2} \right) (3 V_r - 1) = 8 T_r, which is identical for all fluids and reveals that pressure-volume isotherms in reduced coordinates coincide across substances. This derivation underscores how the van der Waals model captures a fundamental similarity in fluid behavior near critical conditions. A direct consequence of this principle is its application to the critical compressibility factor, Z_c = P_c V_c / (R T_c), which quantifies the deviation from ideal gas behavior at the critical point. Inserting the van der Waals critical relations into the definition yields Z_c = 3/8 = 0.375, implying a universal value for all gases modeled by the equation. These critical parameters are obtained from the inflection point of the critical isotherm, where the first and second derivatives of pressure with respect to volume vanish.

Isothermal Behavior

Isotherm Shapes and Van der Waals Loops

The pressure-volume (P-V) isotherms of the reveal non-ideal gas behavior through their dependence on temperature relative to the critical temperature T_c = \frac{8a}{27Rb}, where a and b are the equation's empirical constants accounting for intermolecular attractions and repulsions, respectively. For temperatures above T_c, the isotherms are smooth and monotonically decreasing in as decreases, with a single real positive root for the at any given . These curves gradually approach the hyperbolic form of the at high temperatures or large , where molecular interactions become negligible compared to . At the critical temperature T_c, the isotherm exhibits an at the critical V_c = 3b, where the curve has a horizontal , signifying the boundary beyond which distinct and gas phases cannot coexist. This point corresponds to the critical pressure P_c = \frac{a}{27b^2}, with the cubic nature of the equation yielding a triple root at V_c. Below T_c, the isotherms take on an S-shaped form, characterized by a Van der Waals loop that includes a region of negative slope \left( \frac{dP}{dV} \right)_T < 0. This unphysical feature, where pressure would increase with increasing volume, indicates mechanical instability in the intermediate volume range, typically between approximately $2b and (3 + \sqrt{5})b. The loop emerges from the interplay of attractive forces (parameter a), which lower the pressure by pulling molecules together in the mid-volume regime, and repulsive exclusions (parameter b), which cause a sharp pressure rise at low volumes due to molecular crowding.

Maxwell Construction for Phase Transitions

The Maxwell construction provides a method to resolve the unphysical oscillations, or Van der Waals loops, that appear in the isotherms of the below the critical temperature, replacing them with a flat line segment representing the two-phase coexistence region. This construction, originally proposed by in 1875, involves identifying the saturation pressure P_\text{sat} for a given temperature such that a horizontal line at this pressure divides the area enclosed by the loop into two equal parts: one above the line (where the Van der Waals pressure exceeds P_\text{sat}) and one below (where it is less). Mathematically, this equal-area rule is expressed as the integral condition \int_{V_l}^{V_g} (P_\text{vdW}(V) - P_\text{sat}) \, dV = 0, where V_l and V_g are the specific volumes of the liquid and gas phases at coexistence, ensuring the net area vanishes over the loop. Physically, this procedure enforces thermodynamic equilibrium in the two-phase region by balancing mechanical equilibrium, where the pressure is uniform across both liquid and vapor phases (P_\text{liquid} = P_\text{vapor} = P_\text{sat}), and chemical equilibrium, where the chemical potentials of the two phases are equal (\mu_\text{liquid} = \mu_\text{vapor}). The equal-area condition arises from the requirement that the per mole, g = \mu, must be the same for both phases at coexistence, which can be derived from the thermodynamic relation dg = v \, dp (at constant temperature), leading to equal areas under the pressure-volume curve when integrated appropriately. This interpretation connects the graphical rule directly to the stability criteria for phase transitions in real fluids. To implement the Maxwell construction mathematically, one solves simultaneously for P_\text{sat}, V_l, and V_g such that the Van der Waals pressure equation P_\text{vdW}(V_l) = P_\text{sat} = P_\text{vdW}(V_g) holds at the coexistence volumes, while also satisfying the equal-area integral \int_{V_l}^{V_g} (P_\text{vdW}(V) - P_\text{sat}) \, dV = 0. For the reduced Van der Waals equation \tilde{p}(\tilde{v}) = \frac{8 \tilde{T}}{3 \tilde{v} - 1} - \frac{3}{\tilde{v}^2}, this typically requires numerical solution of a transcendental equation, often by iterating on trial values of \tilde{p}_\text{eq} until the areas balance, yielding the coexistence curve in the phase diagram. This approach accurately predicts the vapor-liquid equilibrium pressures for temperatures below the critical point.

Thermodynamic Derivatives

Internal Energy and Entropy Expressions

The internal energy U of a van der Waals gas can be derived from the fundamental thermodynamic relation \left( \frac{\partial U}{\partial V} \right)_T = T \left( \frac{\partial P}{\partial T} \right)_V - P, applied to the equation of state P = \frac{n R T}{V - n b} - \frac{a n^2}{V^2}, where a and b are the van der Waals constants accounting for intermolecular attractions and excluded volume, respectively, n is the number of moles, R is the gas constant, V is the volume, and T is the temperature. Computing the partial derivative yields \left( \frac{\partial P}{\partial T} \right)_V = \frac{n R}{V - n b}, so \left( \frac{\partial U}{\partial V} \right)_T = T \cdot \frac{n R}{V - n b} - \left( \frac{n R T}{V - n b} - \frac{a n^2}{V^2} \right) = \frac{a n^2}{V^2}. This Maxwell relation arises directly from the attractive term in the equation of state. Integrating with respect to V at constant T, and noting that U approaches the ideal gas value U^\text{ideal} as V \to \infty, gives the departure from ideal behavior as -\frac{a n^2}{V}. Thus, the total internal energy is U = U^\text{ideal}(T) - \frac{a n^2}{V}, where U^\text{ideal}(T) depends only on temperature, reflecting the mean-field treatment of attractions that contributes a volume-dependent but temperature-independent potential energy shift. The entropy S follows from the relation \left( \frac{\partial S}{\partial V} \right)_T = \left( \frac{\partial P}{\partial T} \right)_V = \frac{n R}{V - n b}, which captures the reduced configurational freedom due to the finite molecular size parameterized by b. Integrating at constant T and comparing to the ideal gas entropy S^\text{ideal} = n C_V^\text{ideal} \ln T + n R \ln V + S_0, where C_V^\text{ideal} is the ideal molar heat capacity at constant volume and S_0 is a reference constant, yields the correction S = S^\text{ideal} + n R \ln \left( 1 - \frac{n b}{V} \right). This logarithmic term accounts for the effective free volume V - n b available to the molecules, with no additional contribution from the a parameter since the attractions do not affect the entropy in this model. The temperature independence of the attractive correction in U implies that the heat capacity at constant volume is unaffected by intermolecular forces or excluded volume effects, so C_V = \left( \frac{\partial U}{\partial T} \right)_V = C_V^\text{ideal}(T), mirroring the ideal gas behavior despite the modifications to the equation of state.

Enthalpy, Helmholtz, and Gibbs Free Energies

The enthalpy H, Helmholtz free energy A, and Gibbs free energy G for a van der Waals gas are derived via Legendre transforms from the internal energy U and entropy S, incorporating corrections for molecular attractions and excluded volume beyond ideal gas behavior. These potentials provide key insights into thermodynamic stability and phase equilibria, with the van der Waals parameters a (related to intermolecular attractions) and b (related to molecular volume) modifying the ideal gas forms. The enthalpy is given by the Legendre transform H = U + PV. For the van der Waals gas, substituting the expressions for U and the pressure from the equation of state yields H = H_{\mathrm{ideal}} - 2 a n^2 / V, where the term -2 a n^2 / V arises from the attractive interactions contributing to both the internal energy and the PV product, while the excluded volume effect from b is neglected in this leading-order correction. The Helmholtz free energy is obtained as A = U - TS. Its van der Waals form is A = A_{\mathrm{ideal}} - a n^2 / V - n R T \ln\left(1 - b n / V\right), where the logarithmic term accounts for the reduced effective volume due to b, and the -a n^2 / V term reflects the mean-field attraction energy. This expression, derived from statistical mechanics considerations of the partition function, highlights how attractions lower the free energy relative to the ideal case. The Gibbs free energy follows from the transform G = A + PV, or equivalently G = H - TS. For the van der Waals gas, G = A + PV, and since the system is a single component, the chemical potential is \mu = G / n, which plays a central role in determining phase equilibrium conditions via equality of \mu and pressure across phases. These free energies are crucial for assessing thermodynamic stability. The Helmholtz free energy A must be convex in volume for mechanical stability, but below the critical temperature, its non-convexity in the van der Waals model signals instability, resolved by the Maxwell construction that enforces a common tangent. Similarly, second derivatives of G with respect to pressure or composition ensure chemical and phase stability, with the concave nature of G after the transform eliminating metastable regions.

Transport and Response Properties

Compressibility Factor and Virial Coefficients

The compressibility factor Z, defined as Z = \frac{P V_m}{R T} where V_m is the molar volume, P is pressure, R is the gas constant, and T is temperature, quantifies deviations from ideal gas behavior in the Van der Waals equation. Substituting the Van der Waals equation of state (P + \frac{a}{V_m^2})(V_m - b) = R T yields the explicit form Z = \frac{V_m}{V_m - b} - \frac{a}{R T V_m}, where a accounts for attractive intermolecular forces and b for the finite volume of molecules. This expression highlights how Z > 1 at high densities due to repulsive effects (via the b term) and Z < 1 at low densities due to attractions (via the a term). To connect the Van der Waals equation with the behavior of low-density gases, it is expanded in the virial form Z = 1 + \frac{B}{V_m} + \frac{C}{V_m^2} + \cdots, where B, C, etc., are the second, third, and higher virial coefficients that depend on temperature. For the Van der Waals model, the second virial coefficient is B = b - \frac{a}{R T}, reflecting the competition between excluded volume (+b) and attractive interactions (-\frac{a}{R T}). The third virial coefficient simplifies to C = b^2, arising solely from the repulsive volume correction in this approximation. Higher-order coefficients beyond C are not captured in the basic Van der Waals framework, limiting its accuracy at very high densities. Physically, the second virial coefficient B encapsulates pairwise molecular interactions: positive B indicates net repulsion (dominating at high T), while negative B signals net attraction (at low T). This aligns with statistical mechanical interpretations where B derives from the two-body potential. The Van der Waals virial expansion matches experimental compressibility factors well for moderate densities, where pairwise effects dominate, but deviates at higher densities due to neglected many-body interactions./16%3A_The_Properties_of_Gases/16.05%3A_The_Second_Virial_Coefficient)

Joule-Thomson Coefficient

The Joule-Thomson coefficient, \mu_{JT}, quantifies the temperature change of a gas during an isenthalpic process, such as throttling, and is defined as \mu_{JT} = \left( \frac{\partial T}{\partial P} \right)_H. This coefficient arises in applications like gas liquefaction and refrigeration, where real gas behavior leads to either cooling or heating depending on conditions. Thermodynamically, it can be expressed using the relation \mu_{JT} = \frac{1}{C_p} \left[ T \left( \frac{\partial V}{\partial T} \right)_P - V \right], where C_p is the heat capacity at constant pressure; for an ideal gas, this yields \mu_{JT} = 0, but for real gases, deviations occur due to intermolecular attractions and repulsions. For a van der Waals gas, substituting the equation of state into the thermodynamic expression yields \mu_{JT} = \frac{2 a (1 - b \rho)^2 / (R T) - b}{C_p}, where \rho = 1/V_m is the molar density, V_m is the molar volume, a and b are the van der Waals constants, and R is the gas constant. This form captures the competition between attractive forces (proportional to a) that promote cooling and repulsive effects (proportional to b) that promote heating; the coefficient is positive when the attractive term dominates, leading to a temperature decrease as pressure drops in throttling. The derivation stems from the enthalpy expression for the van der Waals gas, which includes a term -2a/V_m reflecting intermolecular attractions. The inversion curve represents the locus of states where \mu_{JT} = 0, marking the boundary between cooling (\mu_{JT} > 0) and heating (\mu_{JT} < 0) regions in the P-T plane. Setting the van der Waals expression for \mu_{JT} to zero gives the condition $2 a (1 - b \rho)^2 / (R T) = b, which defines this curve; it starts at the origin and reaches a maximum temperature before closing. For the van der Waals model, the maximum inversion temperature is approximately T_{inv} \approx \frac{27}{4} T_c, where T_c = \frac{8 a}{27 R b} is the . This approximation aligns with the low-pressure limit of the inversion curve and explains why most gases exhibit cooling during throttling at ambient conditions, as room temperature lies below their inversion temperature for typical van der Waals parameters.

Applications to Mixtures

Mixing Rules for Parameters

To extend the van der Waals equation to multicomponent mixtures, mixing rules are employed to determine effective parameters a_\text{mix} and b_\text{mix} from the pure-component values a_i and b_i, based on the mole fractions x_i of each component i. These rules originate from the foundational work on fluid mixtures and are designed to account for intermolecular interactions in non-ideal systems. The excluded volume parameter b follows a linear mixing rule, given by b_\text{mix} = \sum_i x_i b_i, which assumes additive contributions from the individual molecular sizes without cross-term corrections. This simple averaging reflects the hard-sphere-like repulsion in the van der Waals model and is widely adopted for its consistency with statistical mechanics for the repulsive part. For the attractive parameter a, a quadratic mixing rule is used to capture pairwise interactions: a_\text{mix} = \sum_i \sum_j x_i x_j \sqrt{a_i a_j}. The cross terms employ the geometric mean \sqrt{a_i a_j}, known as the , which approximates the interaction energy between unlike molecules based on London dispersion forces; this is combined with the Lorentz rule for sizes in broader contexts but simplifies here due to the linear b treatment. The full expression thus incorporates all binary interactions, enabling the prediction of deviations from ideal mixing when pure-component parameters differ significantly. In the mixture equation of state, the molar volume is defined as V_m = V / n_\text{total}, where V is the total volume and n_\text{total} is the total number of moles, allowing the single-component form to be applied directly with the mixed parameters. Differences in a_i and b_i across components introduce non-ideality, such as excess volumes or enthalpies of mixing, without requiring additional adjustable terms in the basic model.

Phase Equilibrium in Binary Mixtures

Phase equilibrium in binary mixtures described by the Van der Waals equation requires the equality of temperature T, pressure P, and chemical potentials \mu_i for each component i between the coexisting vapor (V) and liquid (L) phases, i.e., \mu_i^V = \mu_i^L for i = 1, 2. This condition ensures thermodynamic consistency at the interface, where the chemical potential of each species is derived from the partial molar Gibbs free energy. For Van der Waals mixtures, the parameters a and b are typically obtained using mixing rules, such as the quadratic form for a_m = \sum_i \sum_j x_i x_j a_{ij} and linear for b_m = \sum_i x_i b_i, which incorporate binary interaction parameters to account for unlike-pair attractions. To compute vapor-liquid equilibria (VLE), the chemical potentials are calculated from the equation of state by integrating the Helmholtz free energy or directly from the Gibbs free energy expression for the mixture. A practical graphical method involves plotting the molar Gibbs free energy G_m(T, P, x) versus mole fraction x for a fixed T and P, where the coexisting compositions are found by the common tangent construction: the line tangent to both the liquid-like and vapor-like branches of the G_m(x) curve determines the tie line connecting the equilibrium compositions x^L and x^V. This construction minimizes the total free energy, with the slope of the tangent equal to the common chemical potential difference \Delta \mu = \mu_1 - \mu_2. Numerical solutions often employ iterative methods to solve the equality conditions simultaneously with the equation of state. The Van der Waals model can predict s in binary mixtures when the interaction parameters lead to significant deviations from Raoult's law, such as positive deviations (minimum-boiling ) if the unlike-pair attraction a_{12} is less than \sqrt{a_{11} a_{22}}, causing the vapor and liquid compositions to coincide at a specific x. For example, in mixtures like ethanol-water analogs modeled with Van der Waals parameters, negative deviations arise when a_{12} > \sqrt{a_{11} a_{22}}, resulting in a maximum-boiling where the shows a closed loop in P-x space. These predictions highlight the model's utility in capturing non-ideal behavior, though accuracy depends on the choice of interaction parameters fitted to experimental data.

Limitations and Extensions

Validity Across Gas Conditions

The van der Waals equation provides reasonable qualitative predictions for the phase behavior of fluids near the critical point, capturing the existence of a critical isotherm and the onset of , though quantitative accuracy is limited. For instance, the equation yields a critical Z_c = 0.375, compared to experimental values around 0.27–0.28 for many real gases, resulting in an overestimation by approximately 35–40%. This discrepancy arises from the simplistic modeling of intermolecular forces, yet the equation remains useful for illustrating without requiring advanced computational methods. At low densities, corresponding to high temperatures and low pressures, the van der Waals equation aligns well with experimental data by reproducing the second virial coefficient of the , effectively accounting for pairwise attractions while higher-order terms become negligible. This makes it suitable for dilute gas conditions where deviations from ideality are small. However, as density increases toward moderate levels, the equation offers only qualitative insights into isotherms, overestimating pressures due to its assumption of a constant parameter b. For high densities, such as in the liquid phase, it significantly underpredicts densities by 10–30% for many substances, failing to capture the nuanced effects of molecular packing and short-range repulsions. The equation performs poorly for quantum gases like at low temperatures, where quantum mechanical effects such as and exchange interactions dominate, rendering the classical treatment inadequate without additional corrections. In supercritical regimes, above the critical temperature, the van der Waals model provides approximate descriptions of properties for some fluids like CO₂ but generally fails to predict and thermodynamic responses accurately without parameter adjustments, as it overlooks enhanced fluctuations and non-classical scaling near the critical region.

Comparisons with Other Equations of State

The Van der Waals equation represents a significant departure from the , PV = nRT, by incorporating two key corrections for real gas behavior: the finite volume occupied by molecules, accounted for by the parameter b, and the reduction in effective pressure due to intermolecular attractions, captured by the parameter a. These adjustments address assumptions in the that treat molecules as point particles with no interactions, leading to notable deviations at high pressures—where molecular volume becomes significant—and low temperatures, where attractions dominate and cause gases to liquefy more readily. For instance, for oxygen at 298 K and moderate pressures, the overestimates pressure by less than 2%, but the Van der Waals corrections provide a closer match to experimental data. Compared to the virial equation of state, which expands pressure deviations from ideality as an infinite series in density (P/kT = \rho + B_2(T)\rho^2 + B_3(T)\rho^3 + \cdots) and excels for low-density gases, the Van der Waals equation offers a compact cubic form that approximates only the second virial coefficient while ignoring higher-order terms. This makes it suitable for qualitative insights into gas-liquid transitions but less precise for dense or moderately dense systems, where the full virial expansion, informed by experimental coefficients, captures more nuanced interactions. The Van der Waals model derives its second virial coefficient as B_2(T) = b - a/RT, linking it directly to the virial framework at low densities. Modern , such as the Redlich-Kwong and Peng-Robinson models, build directly on the Van der Waals framework by retaining its repulsive term (RT/(V_m - b)) while refining the attractive term for greater accuracy. The Redlich-Kwong equation introduces temperature dependence in the attractive parameter (a/\sqrt{T}), yielding markedly better predictions for factors and behavior above the critical temperature and at high pressures compared to Van der Waals, which often overestimates these properties. The Peng-Robinson equation further enhances estimates and vapor-liquid equilibria for hydrocarbons by adjusting the attractive term's functional form, achieving critical factors closer to experimental values (0.307 versus Van der Waals' 0.375). However, both successors share limitations with Van der Waals in handling associating fluids like or alcohols, where advanced models such as statistical associating fluid theory (SAFT) or perturbed-chain SAFT (PC-SAFT) are preferred for their explicit treatment of hydrogen bonding. Historically, the Van der Waals equation, introduced in , established the for semi-empirical equations of through its elegant balance of —using just two substance-specific parameters—and its pioneering qualitative prediction of gas-liquid continuity, influencing all subsequent developments in thermodynamic modeling for fluids.

References

  1. [1]
    [PDF] Johannes D. van der Waals - Nobel Lecture
    Now that I am privileged to appear before this distinguished gathering to speak of my theoretical studies on the nature of gases and liquids, I must.Missing: Diderik | Show results with:Diderik
  2. [2]
    Johannes van der Waals
    Van der Waals proposed that we correct for the fact that the volume of a real gas is too large at high pressures by subtracting a term from the volume of the ...
  3. [3]
    None
    ### Summary of Van der Waals Equation Significance
  4. [4]
    [PDF] Van der Waals equation, Maxwell construction, and Legendre ...
    The van der Waals equation is the best known example of an equation of state that exhibits a first-order phase transition with a critical end point.
  5. [5]
    Non-ideal gas - Van der Waal's Equation and Constants
    If a gas behaves ideally, both a and b are zero, and van der Waals equations approaches the ideal gas law P V = n R T. The constant a provides a correction for ...
  6. [6]
    Johannes Diderik van der Waals – Biographical - NobelPrize.org
    The latter, who in 1913 received the Nobel Prize for his low-temperature studies and his production of liquid helium, wrote in 1910 “that Van der Waals' ...
  7. [7]
    The Nobel Prize in Physics 1910 - NobelPrize.org
    The Nobel Prize in Physics 1910 was awarded to Johannes Diderik van der Waals for his work on the equation of state for gases and liquids.
  8. [8]
    [PDF] Physics 1910 JOHANNES DIDERIK VAN DER WAALS
    I intend to discuss in sequence: (1) the broad outlines of my equation of state and how I arrived at it;. (2) what my attitude was and still is to that equation ...
  9. [9]
    How van der Waals first linked liquids and gases | Opinion
    Jun 6, 2023 · Van der Waals presented the first explanation for (as his title put it) this 'continuity of the gaseous and liquid states'. Formation of the ...Missing: original | Show results with:original
  10. [10]
    Heike Kamerlingh Onnes – Biographical - NobelPrize.org
    In particular he had in mind the establishment of a cryogenic laboratory which would enable him to verify Van der Waals' law of corresponding states over a ...Missing: equation | Show results with:equation
  11. [11]
    [PDF] Heike Kamerlingh Onnes - Nobel Lecture
    The study of the equation of state related to the monumental theories of. Van der Waals has always been a main item of research in the Leyden labor- atory. In ...
  12. [12]
    9.6 Non-Ideal Gas Behavior - Chemistry 2e | OpenStax
    Feb 14, 2019 · The van der Waals equation improves upon the ideal gas law by adding two terms: one to account for the volume of the gas molecules and another ...
  13. [13]
    [PDF] History of the Kinetic Theory of Gases* by Stephen G. Brush** Table ...
    mathematical research instrument by the Scottish physicist James Clerk Maxwell (1831-1879). ... Van der Waals did not specify a particular form for this force law ...
  14. [14]
    [PDF] V.C The Second Virial Coefficient & van der Waals Equation - MIT
    Remarks and observations: (1) The tail of the van der Waals attractive potential (∝ r−6) extends to very long sep- arations.<|control11|><|separator|>
  15. [15]
    Derivation of van der Waals Equation of State
    (See Exercise 2.) is four times the volume of a hard-sphere molecule. , appearing in this equation of state, to the inter-molecular potential.
  16. [16]
  17. [17]
    Carbon dioxide - the NIST WebBook
    Critical pressure. Ptriple, Triple point pressure. Tc, Critical temperature. Ttriple, Triple point temperature. Vc, Critical volume. ΔsubH, Enthalpy of ...Phase change data · References
  18. [18]
    Principle of Corresponding States (PCS) | PNG 520
    The principle of Corresponding States (PCS) was stated by van der Waals and reads: “Substances behave alike at the same reduced states. Substances at same ...
  19. [19]
  20. [20]
    [PDF] Statistical Mechanics I - MIT OpenCourseWare
    Tr = T/Tc, a reduced version of the van der Waals equation is obtained as. 8. Tr. 3. Pr = 3 vr − 1/3. − vr. 2 . (V.79). We have thus constructed a universal ...<|control11|><|separator|>
  21. [21]
    [PDF] Mathematical Analysis of the van der Waals Equation
    Abstract. The parametric cubic van der Waals polynomial pV 3 − (RT + bp)V 2 + aV − ab is analysed mathematically and some new generic features ...Missing: primary source<|control11|><|separator|>
  22. [22]
    Phase Transformations in Van der Waals Fluid
    Although the van der Waals equation of state associates some pressures with more than one molar volume, the thermodynamically stable state is that with the ...Missing: explanation | Show results with:explanation
  23. [23]
    [PDF] The gas laws pV = nRT
    The van der Waals loops occur when both terms in the equation have similar magnitudes. The first term arises from the kinetic energy of the molecules and ...<|separator|>
  24. [24]
    [PDF] Is the Maxwell construction correct in predicting the van der Waals ...
    Apr 19, 2016 · For the latter, the van der. Waals equation of state played a special role, as it lead to not only deviations from ideal gas behavior, but also.
  25. [25]
    [PDF] Thermodynamics of Real Gases
    • Note that the second term in the internal energy for a Van der Waals gas depends on V and not on T. It is therefore not correct to say that the internal ...
  26. [26]
    None
    **Summary of Internal Energy Departure Function Derivation for Van der Waals Equation:**
  27. [27]
    [PDF] Entropy and Partial Differential Equations - UC Berkeley Mathematics
    Van der Waals fluid. A van der Waals fluid is a simple fluid with the equation of state. P = RT. V − b. − a. V 2. (V >b, N = 1). (10) for constants a, b > 0 ...
  28. [28]
    [PDF] Thermodynamic Properties of the van der Waals Fluid
    All thermodynamic properties are formulated in terms of reduced parameters that result in many laws of corresponding states, which by definition are the same ...
  29. [29]
    [PDF] LECTURE NOTES ON INTERMEDIATE THERMODYNAMICS
    Mar 28, 2025 · ... Helmholtz free energy, (4.97) τ1,2 = y − ψ1x1 − ψ2x2 = g = g(P, T) = u + Pv − T s, Gibbs free energy. (4.98). It has already been shown ...<|control11|><|separator|>
  30. [30]
    [PDF] 1 Chemistry 312 Spring 2019 Equation Guide Constants h = 6.626 x ...
    (van der Waals). Connections with B. Z = PV. RT. Z = 1 +B'P + C'P2 + ... (virial). B = B'RT. Z =1+. B. V. +. C. V 2 + ... (virial). B = b – a/RT. (with van der ...Missing: third | Show results with:third
  31. [31]
    [PDF] Chemistry 431 Problem Set 10 Fall 2023
    where a and b are constants. 6. Expand the van der Waals equation of state ... You may terminate the series after the third virial coefficient. 1. Page ...
  32. [32]
    Joule-Thomson effect - Physical Chemistry Lab Course
    ... Van der Waals equation in equation (25) yields the Joule-Thomson coefficient. μ = 2 a R T − b C m , p \mu = \frac{\frac{2a}{RT}-b}{C_{\text{m},p}} μ=Cm,p​RT2a​− ...
  33. [33]
    [PDF] Joule-thomson inversion curves and related coefficients for several ...
    The inversion curve and the isothermal Joule-Thomson coefficient are shown to comply with the principle of corresponding states. The isenthalpic Joule- Thomson ...
  34. [34]
    Taking Another Look at the van der Waals Equation of State–Almost ...
    Aug 2, 2019 · The van der Waals (vdW) equation of state has long fascinated researchers and engineers, largely because of its simplicity and engineering flexibility.Excess Properties from the van... · Mixing and Combining Rules... · ReferencesMissing: primary | Show results with:primary
  35. [35]
    [PDF] Testing the Lorentz- Berthelot Mixing Rules with Ab Initio ... - DTIC
    The Lorentz-Berthelot mixing rules are extensively used for estimating intermolecular potential parameters between pairs of non-identical molecules (ij) ...
  36. [36]
    [PDF] The van der Waals mixture construct of a p-x two-component gas ...
    Feb 24, 2022 · Abstract. The van der Waals model, when applied to a mixture, is used to construct tie-line points on a liquid *) vapor two phase p − x diagram.
  37. [37]
    [PDF] On the calculation of critical liquid-vapor lines of binary mixtures
    The critical lines of binary mixtures of conformal fluids are calculated on the basis of the one-fluid Van der Waals theory.
  38. [38]
    Critical lines and phase equilibria in binary van der Waals mixtures
    This paper presents theoretical studies of phase equilibria in binary mixtures obeying the van der Waals equation, especially liquid-liquid equilibria that can ...
  39. [39]
    New algorithm for calculation of azeotropes from equations of state
    Azeotropic loci of several mixtures that are well-described by a simple generalized van der Waals equation of state have been calculated successfully with the ...
  40. [40]
    Azeotrope-Breaking Potential of Binary Mixtures in Phase Equilibria ...
    The diagrams are depicted in the space of the equation of state parameters, e.g., van der Waals a and b parameters. The specific points and lines of global ...
  41. [41]
    Mechanical Instabilities and the Mathematical Behavior of van der ...
    We explore the mathematical behavior of van der Waals gases at temperatures where classical descriptions are inadequate due to emerging quantum effects.<|control11|><|separator|>
  42. [42]
    1.5: Ideal vs. real gases and the van der Waals equation
    Jun 5, 2024 · The uncorrected ideal gas computation has less than a 2% error when compared to this "corrected" van der Waals calculation. But it does give ...
  43. [43]
    16.2: van der Waals and Redlich-Kwong Equations of State
    Mar 8, 2025 · The van der Waals Equation of State is an equation relating the density of gases and liquids to the pressure, volume, and temperature conditions ...