Van der Waals equation

Last updated

The van der Waals equation, named for its originator, the Dutch physicist Johannes Diderik van der Waals, is an equation of state that extends the ideal gas law to include the non-zero size of gas molecules and the interactions between them (both of which depend on the specific substance). As a result the equation is able to model the phase change, liquid vapor. It also produces simple analytic expressions for the properties of real substances that shed light on their behavior. One way to write this equation is: [1] [2] [3]

Contents

where is pressure, is temperature, and is molar volume, with Avogadro's constant, the volume, and the number of molecules (the ratio is a physical quantity with base unit mole in the SI). In addition is the universal gas constant, is Boltzmann's constant, and and are experimentally determinable, substance-specific constants.

The force exerted by a molecule on another at a distance
r
{\displaystyle r}
is the negative of the slope of this curve at
r
{\displaystyle r}
. The force is repulsive, and large, for
r
<
s
{\displaystyle r<\sigma }
, and attractive when
r
>
r
0
>
s
{\displaystyle r>r_{0}>\sigma }
. Dimensionless Lenard Jones 6-12.png
The force exerted by a molecule on another at a distance is the negative of the slope of this curve at . The force is repulsive, and large, for , and attractive when .

The constant expresses the strength of the molecular interactions. It has dimension pressure times molar volume squared, [pv2] which is also molar energy times molar volume. The constant denotes an excluded molar volume due to a finite molecule size, since the centers of two molecules cannot be closer than their diameter, . A theoretical calculation of these constants for spherical molecules with an interparticle potential characterized by, , and a minimum energy, , as shown in the accompanying plot produces . Multiplying this by the number of moles, , gives the excluded volume as 4 times the volume of all the molecules. [4] This theory also produces where is a number that depends on the shape of the potential function, . [5]

In his book (see references [3] and [4]) Boltzmann wrote equations using (specific volume) in place of (molar volume) used here, Gibbs did as well, so do most engineers. Also the property, the reciprocal of number density, is used by physicists, but there is no essential difference between equations written with any of these properties. Equations of state written using molar volume contain , those using specific volume contain (the substance specific is the molar mass), and those written with number density contain .

Once and are experimentally determined for a given substance, the van der Waals equation can be used to predict the boiling point at any given pressure, the critical point (defined by pressure and temperature values, such that the substance cannot be liquefied either when no matter how low the temperature, or when no matter how high the pressure), and other attributes. These predictions are accurate for only a few substances. For most simple fluids they are only a valuable approximation. The equation also explains why superheated liquids can exist above their boiling point and subcooled vapors can exist below their condensation point.

Vdw isobars.png

The graph on the right is a plot of vs calculated from the equation at four constant pressure values. On the red isobar, , the slope is positive over the entire range, (although the plot only shows a finite quadrant). This describes a fluid as a gas for all , and is characteristic of all isobars The green isobar, , has a physically unreal negative slope, hence shown dotted gray, between its local minimum, , and local maximum, . This describes the fluid as two disconnected branches; a gas for , and a denser liquid for . [6]

The thermodynamic requirements of mechanical, thermal, and material equilibrium together with the equation specify two points on the curve, , and , shown as green circles that designate the coexisting boiling liquid and condensing gas respectively. Heating the fluid in this state increases the fraction of gas in the mixture; its , an average of and weighted by this fraction, increases while remains the same. This is shown as the dotted gray line, because it does not represent a solution of the equation; however, it does describe the observed behavior. The points above , superheated liquid, and those below it, subcooled vapor, are metastable; a sufficiently strong disturbance causes them to transform to the stable alternative. Consequently they are shown dashed. Finally the points in the region of negative slope are unstable. All this describes a fluid as a stable gas for , a stable liquid for , and a mixture of liquid and gas at , that also supports metastable states of subcooled gas and superheated liquid. It is characteristic of all isobars , where is a function of . [7] The orange isobar is the critical one on which the minimum and maximum are equal. The black isobar is the limit of positive pressures, although drawn solid none of its points represent stable solutions, they are either metastable (positive or zero slope) or unstable (negative slope. All this is a good explanation of the observed behavior of fluids.

Relationship to the ideal gas law

The ideal gas law follows from the van der Waals equation whenever is sufficiently large (or correspondingly whenever the molar density, , is sufficiently small), Specifically [8]

Putting these two approximations into the van der Waals equation when is large enough that both inequalities are satisfied reduces it to

which is the ideal gas law. [9] This is not surprising since the van der Waals equation was constructed from the ideal gas equation in order to obtain an equation valid beyond the limit of ideal gas behavior.

What is truly remarkable is the extent to which van der Waals succeeded. Indeed, Epstein in his classic thermodynamics textbook began his discussion of the van der Waals equation by writing: "In spite of its simplicity, it comprehends both the gaseous and the liquid state and brings out, in a most remarkable way, all the phenomena pertaining to the continuity of these two states". [10] Also in Volume 5 of his Lectures on Theoretical Physics Sommerfeld, in addition to noting that "Boltzmann [11] described van der Waals as the Newton of real gases", also wrote "It is very remarkable that the theory due to van der Waals is in a position to predict, at least qualitatively, the unstable [referring to superheated liquid, and subcooled vapor now called metastable] states" that are associated with the phase change process. [12]

Utility of the equation

The equation has been, and remains very useful because: [13]

The equation also plays an important role in the modern theory of phase transitions. [14] It depicts the liquid metals, Mercury and Cesium, quantitatively, and describes most real fluids qualitatively. [15] Consequently it can be regarded as one member of a family of equations of state, [16] that depend on a molecular parameter such as the critical compressibility factor, , or the Pitzer (acentric) factor, , where is a dimensionless saturation pressure, and log is the logarithm base 10. [17]

All this makes it a worthwhile pedagogical tool for physics, chemistry, and engineering lecturers, in addition to being a useful mathematical model which can aid student understanding.

History

In 1857 Rudolf Clasius published The Nature of the Motion which We Call Heat. In it he derived the relation for the pressure, , in a gas with particles per unit volume (number density), mass , and mean square speed . He then noted that using the classical laws of Boyle and Charles one could write with a constant of proportionality. Hence temperature was proportional to the average kinetic energy of the particles. [18] This article inspired further work based on the twin ideas that substances are composed of indivisible particles, and that heat is a consequence of the particle motion as governed by Newton's laws. The work, known as the kinetic theory of gases, was done principally by James Clerk Maxwell, and Ludwig Boltzmann. At about the same time J. Willard Gibbs also contributed, and advanced it by converting it into statistical mechanics. [19]

Van der Waals equation on a wall in Leiden VdWaalsLeiden2020.jpg
Van der Waals equation on a wall in Leiden

This environment influenced Johannes Diderik van der Waals. After initially pursuing a teaching credential, he was accepted for doctoral studies at the University of Leiden under Pieter Rijke. This led, in 1873, to a dissertation that provided a simple, particle based, equation that described the gas-liquid change of state, the origin of a critical temperature, and the concept of corresponding states. [20] [21] The equation is based on two premises, first that fluids are composed of particles with non-zero volumes, and second that at a large enough distance each particle exerts an attractive force on all other particles in its vicinity. These forces were called by Boltzmann van der Waals cohesive forces. [22]

In 1869 Irish professor of chemistry Thomas Andrews at Queen's University Belfast in a paper entitled On the Continuity of the Gaseous and Liquid States of Matter, [23] displayed an experimentally obtained set of isotherms of carbonic acid, HCO, that showed at low temperatures a jump in density at a certain pressure, while at higher temperatures there was no abrupt change; the figure can be seen here. Andrews called the isotherm at which the jump just disappeared the critical point. Given the similarity of the titles of this paper and van der Waals subsequent thesis one might think that van der Waals set out to develop a theoretical explanation of Andrews' experiments. However the opposite is true, van der Waals began work by trying to determine a mollecular attraction that appeared in Laplace's theory of capillarity, and only after establishing his equation tested it using Andrews results. [24] [25]

By 1877 sprays of both liquid oxygen and liquid nitrogen had been produced, and a new field of research, low temperature physics, had been opened. The van der Waals equation played a part in all this especially with respect to the liquefaction of hydrogen and helium which was finally achieved in 1908. [26] From measurements of and in two states with the same density, the van der Waals equation produces the values and . [27] Thus from two such measurements of pressure and temperature one could determine and , and from these values calculate the expected critical pressure, temperature, and molar volume. Goodstein summarized this contribution of the van der Waals equation as follows: [28]

All this labor required considerable faith in the belief that gas-liquid systems were all basically the same, even if no one had ever seen the liquid phase. This faith arose out of the repeated success of the van der Waals theory, which is essentially a universal equation of state, independent of the details of any particular substance once it has been properly scaled. ... As a result, not only was it possible to believe that hydrogen could be liquefied. but it was even possible to predict the necessary temperature and pressure.

Van der Waals was awarded the Nobel Prize in 1910, in recognition of the contribution of his formulation of this "equation of state for gases and liquids".

As noted previously, modern day studies of first order phase changes make use of the van der Waals equation together with the Gibbs criterion, equal chemical potential of each phase, as a model of the phenomenon. This model has an analytic coexistence (saturation) curve expressed parametrically, , that was first obtained by Plank, [29] known to Gibbs, and later derived in a beautifully simple and elegant manner by Lekner. [30] A summary of Lekner's solution is presented in a subsequent section, and a more complete discussion in the Maxwell Construction.

Critical point and corresponding states

Figure 1 shows four isotherms of the van der Waals equation (abbreviated as vdW) on a pressure, molar volume plane. The essential character of these curves is that:

Figure 1: Four isotherms of the van der Waals equation along with the black dash dot spinodal curve and the red dash dot coexistence (saturation) curve plotted using reduced (dimensionless) variables. VdW isotherms+2log.png
Figure 1: Four isotherms of the van der Waals equation along with the black dash dot spinodal curve and the red dash dot coexistence (saturation) curve plotted using reduced (dimensionless) variables.

1) at some critical temperature, the slope is negative, , everywhere except at a single point, the critical point, , where both the slope and curvature are zero, ;

2) at higher temperatures the slope of the isotherms is everywhere negative (values of for which the equation has 1 real root for );

3) at lower temperatures there are two points on each isotherm where the slope is zero (values of for which the equation has 3 real roots for )

Evaluating the two partial derivatives in 1) using the vdW equation and equating them to zero produces, , and using these in the equation gives . [31]

This calculation can be done algebraically by noting that the vdW equation can be written as a cubic in , which at the critical point is, . Moreover, at the critical point all three roots coalesce so it can also be written as

Then dividing the first by , and noting that these two cubic equations are the same when all their coefficients are equal gives three equations

whose solution produces the previous results for . [32] [33]

Using these critical values to define reduced properties renders the equation in the dimensionless form used to construct Fig. 1

This dimensionless form is a similarity relation; it indicates that all vdW fluids at the same will plot on the same curve. It expresses the law of corresponding states which Boltzmann described as follows: [34]

All the constants characterizing the gas have dropped out of this equation. If one bases measurements on the van der Waals units [Boltzmann's name for the reduced quantities here], then he obtains the same equation of state for all gases. ... Only the values of the critical volume, pressure, and temperature depend on the nature of the particular substance; the numbers that express the actual volume, pressure, and temperature as multiples of the critical values satisfy the same equation for all substances. In other words, the same equation relates the reduced volume, reduced pressure, and reduced temperature for all substances.

Obviously such a broad general relation is unlikely to be correct; nevertheless, the fact that one can obtain from it an essentially correct description of actual phenomena is very remarkable.

This "law" is just a special case of dimensional analysis in which an equation containing 6 dimensional quantities, , and 3 dimensions, [p], [v], [T], must be expressible in terms of 6-3=3 dimensionless groups. [35] Here is a characteristic molar volume, a characteristic pressure, and a characteristic temperature, and the 3 dimensionless groups are . The reduced properties defined previously are , , and . Recent research has suggested that there is a family of equations of state that depend on an additional dimensionless group, and this provides a more exact correlation of properties. Nevertheless, as Boltzmann observed, the van der Waals equation provides an essentially correct description.

The vdW equation produces , while for most real fluids . [36] Thus most real fluids do not satisfy this condition, and consequently their behavior is only described qualitatively by the vdW equation. However, the vdW equation of state is a member of a family of state equations based on the Pitzer (acentric) factor, , and the liquid metals, Mercury and Cesium, are well approximated by it. [15] [37]

Thermodynamic properties

The properties molar internal energy, , and entropy, , defined by the first and second laws of thermodynamics, hence all thermodynamic properties of a simple compressible substance, can be specified, up to a constant of integration, by two measurable functions, a mechanical equation of state, , and a constant volume specific heat, . [38] [39]

Internal energy and specific heat at constant volume

The internal energy is given by the energetic equation of state, [40] [41]

where is an arbitrary constant of integration.

Now in order for to be an exact differential, namely that be continuous with continuous partial derivatives, its second mixed partial derivatives must also be equal, . Then with this condition can be written simply as . Differentiating for the vdW equation gives , so . Consequently for a vdW fluid exactly as it is for an ideal gas. To keep things simple it is regarded as a constant here, with a number. Then both integrals can be easily evaluated and the result is

This is the energetic equation of state for a perfect vdW fluid. By making a dimensional analysis (what might be called extending the principle of corresponding states to other thermodynamic properties) it can be written simply in reduced form as

where and C is a dimensionless constant.

Enthalpy

The enthalpy is , and the product is just . Then is simply

This is the enthalpic equation of state for a perfect vdW fluid, or in reduced form

Entropy

The entropy is given by the entropic equation of state: [42] [43]

Using as before, and integrating the second term using we obtain simply

This is the entropic equation of state for a perfect vdW fluid, or in reduced form

Gibbs free energy

The Gibbs free energy is so combining the previous results gives

This is the Gibbs free energy for a perfect vdW fluid, or in reduced form

The thermodynamic derivatives: α, κT and cp

The two first partial derivatives of the vdW equation are

The first equation is , while the second is , where , the isothermal compressibility, is a measure of the relative increase of volume from an increase of pressure, at constant temperature, while , the coefficient of thermal expansion, is a measure of the relative increase of volume from an increase of temperature, at constant pressure. Therefore

In the limit while . Since the vdW equation in this limit becomes , finally . Both of these are the ideal gas values, which is consistent because, as noted earlier, the vdW fluid behaves like an ideal gas in this limit.

The specific heat at constant pressure, is defined as the partial derivative . However, it is not independent of , they are related by the Mayer equation, . [44] [45] Then the two partials of the vdW equation can be used to express as

Here in the limit , , which is also the ideal gas result as expected; however the limit gives the same result, which does not agree with experiments on liquids.

In this liquid limit we also find , namely that the vdW liquid is incompressible. Moreover, since , it is also mechanically incompressible, that is faster than .

Finally , and are all infinite on the curve . This curve, called the spinodal curve, is defined by , and is discussed at length in the next section.

Stability

According to the extremum principle of thermodynamics and , namely that at equilibrium the entropy is a maximum. This leads to a requirement that . [46] This mathematical criterion expresses a physical condition which Epstein [47] described as follows:

Figure 1 repeated VdW isotherms+2log.png
Figure 1 repeated

"It is obvious that this middle part, dotted in our curves [the place where the requirement is violated, dashed gray in Fig. 1 and repeated here], can have no physical reality. In fact, let us imagine the fluid in a state corresponding to this part of the curve contained in a heat conducting vertical cylinder whose top is formed by a piston. The piston can slide up and down in the cylinder, and we put on it a load exactly balancing the pressure of the gas. If we take a little weight off the piston, there will no longer be equilibrium and it will begin to move upward. However, as it moves the volume of the gas increases and with it its pressure. The resultant force on the piston gets larger, retaining its upward direction. The piston will, therefore, continue to move and the gas to expand until it reaches the state represented by the maximum of the isotherm. Vice versa, if we add ever so little to the load of the balanced piston, the gas will collapse to the state corresponding to the minimum of the isotherm"

While on an isotherm this requirement is satisfied everywhere so all states are gas, those states on an isotherm, which lie between the local minimum, , and local maximum, , for which (shown dashed gray in Fig. 1), are unstable and thus not observed. This is the genesis of the phase change; there is a range , for which no observable states exist. The states for are liquid and for are vapor; due to gravity the denser liquid lies below the vapor. The transition points, states with zero slope, are called spinodal points. Their locus is the spinodal curve that separates the regions of the plane for which liquid, vapor, and gas exist from a region where no observable states exist. This spinodal curve is obtained here from the vdW equation by differentiation (or equivalently from ) as

A projection of this space curve is plotted in Fig.1 as the black dash dot curve. It passes through the critical point which is also a spinodal point.

Saturation

Although the gap in delimited by the two spinodal points on an isotherm (e.g. shown in Fig. 1) is the origin of the phase change, the spinodal points do not represent its full extent, because both states, saturated liquid and saturated vapor coexist in equlilbrium; they both must have the same pressure as well as the same temperature. [48] Thus the phase change is characterized, at temperature , by a pressure that lies between that of the minimum and maximum spinodal points, and with molar volumes of liquid, and vapor .. Then from the vdW equation applied to these saturated liquid and vapor states

These two vdW equations contain 4 variables, , so another equation is required in order to specify the values of 3 of these variables uniquely in terms of a fourth. Such an equation is provided here by the equality of the Gibbs free energy in the saturated liquid and vapor states, . [49] This condition of material equilibrium can be obtained from a simple physical argument as follows: the energy required to vaporize a mole is from the second law at constant temperature , and from the first law at constant pressure . Equating these two, rearranging, and recalling that produces the result.

The Gibbs free energy is one of the 4 thermodynamic potentials whose partial derivatives produce all other thermodynamics state properties; [50] its differential is . Integrating this over an isotherm from to , noting that the pressure is the same at each endpoint, and setting the result to zero yields

Here because is a multivalued function, the integral must be divided into 3 parts corresponding to the 3 real roots of the vdW equation in the form, (this can be visualized most easily by imagining Fig. 1 rotated 90); the result is a special case of material equilibrium. [51] The last equality, which follows from integrating , is the Maxwell equal area rule which requires that the upper area between the vdW curve and the horizontal through be equal to the lower one. [52] This form means that the thermodynamic restriction that fixes is specified by the equation of state itself, . Using the equation for the Gibbs free energy obtained previously for the vdW equation applied to the saturated vapor state and subtracting the result applied to the saturated liquid state produces,

This is a third equation that along with the two vdW equations above can be solved numerically. This has been done given a value for either or , and tabular results presented; [53] [54] however, the equations also admit an analytic parametric solution obtained most simply and elegantly, by Lekner. [55] Details of this solution may be found in the Maxwell Construction; the results are

where

Figure 2: The dashed dot black curve is the stability limit (spinodal curve) and the dashed dot blue curve is the coexistence, or saturation curve, plotted in the
p
r
,
T
r
{\displaystyle p_{r},T_{r}}
plane. At every point in the region between the two curves there are two states, one stable and another metastable. The metastable states, superheated liquid, and subcooled vapor, are shown dotted in Fig. 1. Vdw stability-saturation.png
Figure 2: The dashed dot black curve is the stability limit (spinodal curve) and the dashed dot blue curve is the coexistence, or saturation curve, plotted in the plane. At every point in the region between the two curves there are two states, one stable and another metastable. The metastable states, superheated liquid, and subcooled vapor, are shown dotted in Fig. 1.

and the parameter is given physically by . The values of all other property discontinuities across the saturation curve also follow from this solution. [36] These functions define the coexistence curve which is the locus of the saturated liquid and saturated vapor states of the vdW fluid. The curve is plotted in Fig. 1 and Fig. 2, two projections of the state surface. These curves agree exactly with the numerical results referenced earlier.

Referring back to Fig. 1 the isotherms for are discontinuous. Considering as an example, it consists of the two separate green segments. The solid segment above the green circle on the left, and below the one on the right correspond to stable states, the dots represent the saturated liquid and vapor states that comprise the phase change, and the two green dotted segments below and above the dots are metastable states, superheated liquid and subcooled vapor, that are created in the process of phase transition, have a short lifetime, then devolve into their lower energy stable alternative.

In his treatise of 1898 in which he described the van der Waals equation in great detail Boltzmann discussed these states in a section titled "Undercooling, Delayed evaporation"; [56] they are now denoted subcooled vapor, and superheated liquid. Moreover, it has now become clear that these metastable states occur regularly in the phase transition process. In particular processes that involve very high heat fluxes create large numbers of these states, and transition to their stable alternative with a corresponding release of energy can be dangerous. Consequently there is a pressing need to study their thermal properties. [57]

In the same section Boltzmann also addressed and explained the negative pressures which some liquid metastable states exhibit (for example of Fig. 1). He concluded that such liquid states of tensile stresses were real, as did Tien and Lienhard many years later who wrote "The van der Waals equation predicts that at low temperatures liquids sustain enormous tension...In recent years measurements have been made that reveal this to be entirely correct." [58]

Even though the phase change produces a mathematical discontinuity in the homogeneous fluid properties, for example , there is no physical discontinuity. [51] As the liquid begins to vaporize the fluid becomes a heterogeneous mixture of liquid and vapor whose molar volume varies continuously from to according to the equation of state

where is the mole fraction of the vapor. This equation is called the lever rule and applies to other properties as well. [59] [51] The states it represents form a horizontal line connecting the same colored dots on an isotherm, but not shown in Fig. 1 as noted already since it is a distinct equation of state for the heterogeneous combination of liquid and vapor components.

Extended corresponding states behavior and the van der Waals Equation

Figure 3: The family of saturation curves showing the vdw curve as a member. The blue dots are calculated from Lekner's solution. Vapor pressure vs temperature.png
Figure 3: The family of saturation curves showing the vdw curve as a member. The blue dots are calculated from Lekner's solution.

The idea of corresponding states originated when van der Waals cast his equation in the dimensionless form, . However, as Boltzmann noted, such a simple representation could not correctly describe all substances. Indeed, the saturation analysis of this form produces , namely all substances have the same dimensionless coexistence curve. [60] In order to avoid this paradox an extended principle of corresponding states has been suggested in which where is a substance dependent dimensionless parameter related to the only physical feature associated with an individual substance, its critical point.

The first candidate for was the critical compressibility factor , but because that quantity is difficult to measure accurately, the acentric factor developed by Kenneth Pitzer, , is more useful. The saturation pressure in this situation is represented by a one parameter family of curves, . Several investigators have produced correlations of saturation data for a number of substances, the best is that of Dong and Lienhard, [61]

which has an rms error of over the range

Figure 4: A plot of the correlation including data from various substances. Tr vs Pitzer factor.jpg
Figure 4: A plot of the correlation including data from various substances.

Figure 3 is a plot of vs . for various values of as given by this equation. The ordinate is logarithmic in order to show the behavior at pressures far below the critical where differences among the various substances (indicated by varying values of ) are more pronounced.

Figure 4 is another plot of the same equation showing as a function of for various values of . It includes data from 51 substances including the vdW fluid over the range . This plot shows clearly that the vdW fluid () is a member of the class of real fluids; indeed it quantitatively describes the behavior of the liquid metals cesium () and mercury () whose values of are close to the vdW value. However, it describes the behavior of other fluids only qualitatively, because specific numerical values are modified by differing values of their Pitzer factor, .

Joule-Thomson coefficient and the inversion curve

The Joule-Thomson coefficient, , is of practical importance because the two end states of a throttling process () lie on a constant enthalpy curve. Although ideal gases, for which , do not change temperature in such a process, real gases do, and it is important in applications to know whether they heat up or cool down. [62]

This coefficient can be found in terms of the previously described derivatives as, [63]

so when is positive the gas temperature decreases when it passes through a throttle, and if it is negative the temperature increases. Therefore the condition defines a curve that separates the region of the plane where from the region where it is less than zero. This curve is called the inversion curve, and its equation is . Using the expression for derived previously for the van der Waals equation this is

Note that for there will be cooling for or in terms of the critical temperature . As Sommerfeld noted, "This is the case with air and with most other gases. Air can be cooled at will by repeated expansion and can finally be liquified." [64]

Figure 5: Curves of constant enthalpy in this plane have negative slope above this (green) inversion curve, positive slope below it and zero slope on it; they are S-shaped. A gas entering a throttle at a state corresponding to a point on this curve to the right of its maximum will cool if the final state is below the curve. The other (dashed purple) curve in the graph is the saturation curve. The graph on the right is the square (0,0),(1.1,1.1) of the left graph expanded to display the overlap between the inversion and saturation curves. Vdw inversion2.png
Figure 5: Curves of constant enthalpy in this plane have negative slope above this (green) inversion curve, positive slope below it and zero slope on it; they are S-shaped. A gas entering a throttle at a state corresponding to a point on this curve to the right of its maximum will cool if the final state is below the curve. The other (dashed purple) curve in the graph is the saturation curve. The graph on the right is the square (0,0),(1.1,1.1) of the left graph expanded to display the overlap between the inversion and saturation curves.

The inversion curve can be found by solving its equation for , and substituting into the vdW equation. This produces , where, for simplicity, have been replaced by .

The maximum of this, quadratic, curve occurs, with , for

which gives , or , and the corresponding . The zeros of the inversion curve , are, making use of the quadratic formula, , or and ( and ). In terms of the dimensionless variables, the zeros are at and , while the maximum is , and occurs at . Note from Fig. 5 that there is an overlap between the saturation curve and the inversion curve plotted there. This region is shown enlarged in the right hand graph of the figure. Therefore a van der Waals gas can be liquified by passing it through a throttle under the proper conditions; real gases are liquified in this way.

Compressibility factor

Figure 6: The isotherms, spinodal and coexistence curves here are the same as in Fig. 1. In addition the isotherm
T
r
=
27
/
8
{\displaystyle T_{r}=27/8}
, which has zero slope at the origin is plotted and the isotherm
T
r
=
[?]
{\displaystyle T_{r}=\infty }
. The abscissa here is
v
*
r
=
r
r
/
3
{\displaystyle v^{*}\rho =\rho _{r}/3}
which varies from 0 to 1. Vdw Z rho.png
Figure 6: The isotherms, spinodal and coexistence curves here are the same as in Fig. 1. In addition the isotherm , which has zero slope at the origin is plotted and the isotherm . The abscissa here is which varies from 0 to 1.
Figure 7: Generalized compressibility chart for a van der Waals gas. Vdw Z p r 1.png
Figure 7: Generalized compressibility chart for a van der Waals gas.

Real gases can be characterized by their difference from ideal. This is done by writing the mechanical equation of state in the form where , called the compressibility factor, is usually expressed either as a function of pressure and temperature, or density and temperature, and in each case in the limit or , , the ideal gas value. In the second case . [65] Thus for a van der Waals fluid the compressibility factor is , or in terms of reduced variables

where . At the critical point, , .

In the limit , (for finite ); the fluid behaves like an ideal gas, a point noted several times earlier. Note additionally that the derivative for , and when this is . The slope is positive or negative depending on whether is greater than or less than , and becomes infinitely large negative as approaches .

Figure 6 shows a plot of various isotherms of vs . Also shown are the spinodal and coexistence curves described previously. The subcritical isotherm consists of stable, metastable, and unstable segments, and are identified the same as they were in Fig. 1. Also included are the zero initial slope isotherm and the one corresponding to infinite temperature.

By plotting vs using as a parameter, one obtains the generalized compressibility chart for a vdW gas, which is shown in Fig. 7. Like all other vdW properties, this is not quantitatively correct for most gases but it has the correct qualitative features as can be seen by comparison with this figure which was produced from data using real gases. [66] [67] The two graphs are similar, including the caustic generated by the crossing isotherms; they are qualitatively very much alike.

Virial expansion

Statistical mechanics suggests that can be expressed by a power series called a virial expansion, [68]

The functions are the virial coefficients; the th term represents a particle interaction.

If we note in the expression for that for , the term can be expanded in an absolutely convergent series; this yields

This is the virial expansion for the van der Waals fluid. The first virial coefficient is the slope of at . Notice that it can be positive or negative depending on whether or not , which agrees with the result found previously by differentiation.

Gas mixtures

If a mixture of gases is being considered, and each gas has its own (attraction between molecules) and (volume occupied by molecules) values, then and for the mixture can be calculated as [69]

= total number of moles of gas present,
for each , = number of moles of gas present, and
[70] or [71]
[70] or [71]

and the rule of adding partial pressures becomes invalid if the numerical result of the equation is significantly different from the ideal gas equation .

Derivation

Textbooks in physical chemistry generally give two derivations of the title equation.[ who? ] One is the conventional derivation that goes back to Van der Waals, a mechanical equation of state that cannot be used to specify all thermodynamic functions; the other is a statistical mechanics derivation that makes explicit the intermolecular potential neglected in the first derivation.[ citation needed ] A particular advantage of the statistical mechanical derivation is that it yields the partition function for the system, and allows all thermodynamic functions to be specified (including the mechanical equation of state).[ citation needed ]

Conventional derivation

Consider one mole of gas composed of non-interacting point particles that satisfy the ideal gas law:(see any standard Physical Chemistry text, op. cit.)

Next, assume that all particles are hard spheres of the same finite radius r (the Van der Waals radius). The effect of the finite volume of the particles is to decrease the available void space in which the particles are free to move. We must replace V by V  b, where b is called the excluded volume (per mole) or "co-volume". The corrected equation becomes

The excluded volume is not just equal to the volume occupied by the solid, finite-sized particles, but actually four times the total molecular volume for one mole of a Van der waals' gas. To see this, we must realize that a particle is surrounded by a sphere of radius 2r (two times the original radius) that is forbidden for the centers of the other particles. If the distance between two particle centers were to be smaller than 2r, it would mean that the two particles penetrate each other, which, by definition, hard spheres are unable to do.

The excluded volume for the two particles (of average diameter d or radius r) is

,

which, divided by two (the number of colliding particles), gives the excluded volume per particle:

,

So b is four times the proper volume of the particle. It was a point of concern to Van der Waals that the factor four yields an upper bound; empirical values for b are usually lower. Of course, molecules are not infinitely hard, as Van der Waals thought, and are often fairly soft. To obtain the excluded volume per mole we just need to multiply by the number of molecules in a mole, i.e. by the avogadro number:

.

Next, we introduce a (not necessarily pairwise) attractive force between the particles. Van der Waals assumed that, notwithstanding the existence of this force, the density of the fluid is homogeneous; furthermore, he assumed that the range of the attractive force is so small that the great majority of the particles do not feel that the container is of finite size.[ citation needed ] Given the homogeneity of the fluid, the bulk of the particles do not experience a net force pulling them to the right or to the left. This is different for the particles in surface layers directly adjacent to the walls. They feel a net force from the bulk particles pulling them into the container, because this force is not compensated by particles on the side where the wall is (another assumption here is that there is no interaction between walls and particles, which is not true, as can be seen from the phenomenon of droplet formation; most types of liquid show adhesion). This net force decreases the force exerted onto the wall by the particles in the surface layer. The net force on a surface particle, pulling it into the container, is proportional to the number density. On considering one mole of gas, the number of particles will be NA

.

The number of particles in the surface layers is, again by assuming homogeneity, also proportional to the density. In total, the force on the walls is decreased by a factor proportional to the square of the density, and the pressure (force per unit surface) is decreased by:

,

so that:

Upon writing n for the number of moles and nVm = V, the equation obtains the second form given above:

It is of some historical interest to point out that Van der Waals, in his Nobel prize lecture, gave credit to Laplace for the argument that pressure is reduced proportional to the square of the density. [72]

Statistical thermodynamics derivation

The canonical partition function Z of an ideal gas consisting of N = nNA identical (non-interacting) particles, is: [73] [74]

where is the thermal de Broglie wavelength:

with the usual definitions: h is the Planck constant, m the mass of a particle, k the Boltzmann constant and T the absolute temperature. In an ideal gas z is the partition function of a single particle in a container of volume V. In order to derive the Van der Waals equation we assume now that each particle moves independently in an average potential field offered by the other particles. The averaging over the particles is easy because we will assume that the particle density of the Van der Waals fluid is homogeneous. The interaction between a pair of particles, which are hard spheres, is taken to be:

r is the distance between the centers of the spheres and d is the distance where the hard spheres touch each other (twice the Van der Waals radius). The depth of the Van der Waals well is .

Because the particles are not coupled under the mean field Hamiltonian, the mean field approximation of the total partition function still factorizes:

,

but the intermolecular potential necessitates two modifications to z. First, because of the finite size of the particles, not all of V is available, but only V − Nb', where (just as in the conventional derivation above):

.

Second, we insert a Boltzmann factor exp[ - Φ/2kT] to take care of the average intermolecular potential. We divide here the potential by two because this interaction energy is shared between two particles. Thus:

The total attraction felt by a single particle is:

where we assumed that in a shell of thickness dr there are N/V 4πr2dr particles. This is a mean field approximation; the position of the particles is averaged. In reality the density close to the particle is different than far away as can be described by a pair correlation function. Furthermore, it is neglected that the fluid is enclosed between walls. Performing the integral we get:

Hence, we obtain:

From statistical thermodynamics we know that:

,

so that we only have to differentiate the terms containing . We get:

Application to compressible fluids

The equation is also usable as a PVT equation for compressible fluids (e.g. polymers). In this case specific volume changes are small and it can be written in a simplified form:

where p is the pressure, V is specific volume, T is the temperature and A, B, C are parameters.

See also

Related Research Articles

<span class="mw-page-title-main">Brownian motion</span> Random motion of particles suspended in a fluid

Brownian motion is the random motion of particles suspended in a medium.

<span class="mw-page-title-main">Equation of state</span> An equation describing the state of matter under a given set of physical conditions

In physics and chemistry, an equation of state is a thermodynamic equation relating state variables, which describe the state of matter under a given set of physical conditions, such as pressure, volume, temperature, or internal energy. Most modern equations of state are formulated in the Helmholtz free energy. Equations of state are useful in describing the properties of pure substances and mixtures in liquids, gases, and solid states as well as the state of matter in the interior of stars.

<span class="mw-page-title-main">Fick's laws of diffusion</span> Mathematical descriptions of molecular diffusion

Fick's laws of diffusion describe diffusion and were first posited by Adolf Fick in 1855 on the basis of largely experimental results. They can be used to solve for the diffusion coefficient, D. Fick's first law can be used to derive his second law which in turn is identical to the diffusion equation.

<span class="mw-page-title-main">Ideal gas law</span> Equation of the state of a hypothetical ideal gas

The ideal gas law, also called the general gas equation, is the equation of state of a hypothetical ideal gas. It is a good approximation of the behavior of many gases under many conditions, although it has several limitations. It was first stated by Benoît Paul Émile Clapeyron in 1834 as a combination of the empirical Boyle's law, Charles's law, Avogadro's law, and Gay-Lussac's law. The ideal gas law is often written in an empirical form:

<span class="mw-page-title-main">Ideal gas</span> Mathematical model which approximates the behavior of real gases

An ideal gas is a theoretical gas composed of many randomly moving point particles that are not subject to interparticle interactions. The ideal gas concept is useful because it obeys the ideal gas law, a simplified equation of state, and is amenable to analysis under statistical mechanics. The requirement of zero interaction can often be relaxed if, for example, the interaction is perfectly elastic or regarded as point-like collisions.

<span class="mw-page-title-main">Compressibility</span> Measure of the relative volume change of a fluid or solid as a response to a pressure change

In thermodynamics and fluid mechanics, the compressibility is a measure of the instantaneous relative volume change of a fluid or solid as a response to a pressure change. In its simple form, the compressibility may be expressed as

<span class="mw-page-title-main">Boltzmann equation</span> Equation of statistical mechanics

The Boltzmann equation or Boltzmann transport equation (BTE) describes the statistical behaviour of a thermodynamic system not in a state of equilibrium; it was devised by Ludwig Boltzmann in 1872. The classic example of such a system is a fluid with temperature gradients in space causing heat to flow from hotter regions to colder ones, by the random but biased transport of the particles making up that fluid. In the modern literature the term Boltzmann equation is often used in a more general sense, referring to any kinetic equation that describes the change of a macroscopic quantity in a thermodynamic system, such as energy, charge or particle number.

The virial expansion is a model of thermodynamic equations of state. This model expresses the pressure P of a gas in local equilibrium as a power series of the density. This equation may be represented in terms of the compressibility factor, , as

In chemical thermodynamics, the fugacity of a real gas is an effective partial pressure which replaces the mechanical partial pressure in an accurate computation of chemical equilibrium. It is equal to the pressure of an ideal gas which has the same temperature and molar Gibbs free energy as the real gas.

In atmospheric dynamics, oceanography, asteroseismology and geophysics, the Brunt–Väisälä frequency, or buoyancy frequency, is a measure of the stability of a fluid to vertical displacements such as those caused by convection. More precisely it is the frequency at which a vertically displaced parcel will oscillate within a statically stable environment. It is named after David Brunt and Vilho Väisälä. It can be used as a measure of atmospheric stratification.

In physics, the Einstein relation is a previously unexpected connection revealed independently by William Sutherland in 1904, Albert Einstein in 1905, and by Marian Smoluchowski in 1906 in their works on Brownian motion. The more general form of the equation in the classical case is

<span class="mw-page-title-main">Lattice Boltzmann methods</span> Class of computational fluid dynamics methods

The lattice Boltzmann methods (LBM), originated from the lattice gas automata (LGA) method (Hardy-Pomeau-Pazzis and Frisch-Hasslacher-Pomeau models), is a class of computational fluid dynamics (CFD) methods for fluid simulation. Instead of solving the Navier–Stokes equations directly, a fluid density on a lattice is simulated with streaming and collision (relaxation) processes. The method is versatile as the model fluid can straightforwardly be made to mimic common fluid behaviour like vapour/liquid coexistence, and so fluid systems such as liquid droplets can be simulated. Also, fluids in complex environments such as porous media can be straightforwardly simulated, whereas with complex boundaries other CFD methods can be hard to work with.

In nonideal fluid dynamics, the Hagen–Poiseuille equation, also known as the Hagen–Poiseuille law, Poiseuille law or Poiseuille equation, is a physical law that gives the pressure drop in an incompressible and Newtonian fluid in laminar flow flowing through a long cylindrical pipe of constant cross section. It can be successfully applied to air flow in lung alveoli, or the flow through a drinking straw or through a hypodermic needle. It was experimentally derived independently by Jean Léonard Marie Poiseuille in 1838 and Gotthilf Heinrich Ludwig Hagen, and published by Hagen in 1839 and then by Poiseuille in 1840–41 and 1846. The theoretical justification of the Poiseuille law was given by George Stokes in 1845.

<span class="mw-page-title-main">Maxwell construction</span>

In thermodynamic equilibrium, a necessary condition for stability is that pressure, , does not increase with molar volume, ; this is expressed mathematically as , where is the temperature.

In thermodynamics a residual property is defined as the difference between a real fluid property and an ideal gas property, both considered at the same density, temperature, and composition, typically expressed as

In physics and thermodynamics, the Redlich–Kwong equation of state is an empirical, algebraic equation that relates temperature, pressure, and volume of gases. It is generally more accurate than the van der Waals equation and the ideal gas equation at temperatures above the critical temperature. It was formulated by Otto Redlich and Joseph Neng Shun Kwong in 1949. It showed that a two-parameter, cubic equation of state could well reflect reality in many situations, standing alongside the much more complicated Beattie–Bridgeman model and Benedict–Webb–Rubin equation that were used at the time. The Redlich–Kwong equation has undergone many revisions and modifications, in order to improve its accuracy in terms of predicting gas-phase properties of more compounds, as well as in better simulating conditions at lower temperatures, including vapor–liquid equilibria.

<span class="mw-page-title-main">Diffusion</span> Transport of dissolved species from the highest to the lowest concentration region

Diffusion is the net movement of anything generally from a region of higher concentration to a region of lower concentration. Diffusion is driven by a gradient in Gibbs free energy or chemical potential. It is possible to diffuse "uphill" from a region of lower concentration to a region of higher concentration, as in spinodal decomposition. Diffusion is a stochastic process due to the inherent randomness of the diffusing entity and can be used to model many real-life stochastic scenarios. Therefore, diffusion and the corresponding mathematical models are used in several fields beyond physics, such as statistics, probability theory, information theory, neural networks, finance, and marketing.

<span class="mw-page-title-main">Rayleigh–Plesset equation</span> Ordinary differential equation

In fluid mechanics, the Rayleigh–Plesset equation or Besant–Rayleigh–Plesset equation is a nonlinear ordinary differential equation which governs the dynamics of a spherical bubble in an infinite body of incompressible fluid. Its general form is usually written as

The shear viscosity of a fluid is a material property that describes the friction between internal neighboring fluid surfaces flowing with different fluid velocities. This friction is the effect of (linear) momentum exchange caused by molecules with sufficient energy to move between these fluid sheets due to fluctuations in their motion. The viscosity is not a material constant, but a material property that depends on temperature, pressure, fluid mixture composition, local velocity variations. This functional relationship is described by a mathematical viscosity model called a constitutive equation which is usually far more complex than the defining equation of shear viscosity. One such complicating feature is the relation between the viscosity model for a pure fluid and the model for a fluid mixture which is called mixing rules. When scientists and engineers use new arguments or theories to develop a new viscosity model, instead of improving the reigning model, it may lead to the first model in a new class of models. This article will display one or two representative models for different classes of viscosity models, and these classes are:

Cubic equations of state are a specific class of thermodynamic models for modeling the pressure of a gas as a function of temperature and density and which can be rewritten as a cubic function of the molar volume.

References

  1. van der Waals, J.D., Edited and with an Introduction by J.S. Rowlinson, On the Continuity of the Gaseous and Liquid States, Dover, N.Y. (1988) p. 174
  2. Epstein, P.S., Textbook of Thermodynamics, John Wiley & Sons, NY, p 9, (1937)
  3. Boltzmann, L., Translated by S.G. Brush, Lectures on Gas Theory, Dover, N.Y. (1995) p 231
  4. Boltzmann, p 221-224
  5. Tien, C.L., and Lienhard, J.H., Statistical Thermodynamics Revised Printing, Hemisphere Publishing, NY, p.250, (1979)
  6. Truesdale, C. and Bharatha, S., Classical Thermodynamics as a Theory of Heat Engines, Springer-Verlag, NY, pp 13-15, (1977)
  7. Epstein, p. 11
  8. Epstein, p.10
  9. Epstein, p.10
  10. Epstein, P.S., p. 10
  11. Boltzmann, L. Enzykl. der Mathem. Wiss., Vol. V. 1, p. 550
  12. Sommerfeld, A., Thermodynamics and Statistical Mechanics Lectures on Theoretical Physics Volume V, Edited by Bopp, F. and Meixner, J., Translated by Kestin, J., Academic Press, N.Y., pp 55, 66, (1956)
  13. Sommerfeld, pp. 55-68
  14. Goodstein, D.L., States of Matter, Dover, N.Y., pp 443-452, (1975)
  15. 1 2 Lienhard, J.H., The Properties and Behavior of Superheated Liquids, Lat. Am. J. Heat and Mass Transfer, 10, 169-187, (1986)
  16. Peck, R.E., The Assimilation of van der Waals Equation in the Corresponding States Family, Can. J. Chem. Eng., 60, 446-449 (1982)
  17. Pitzer, K.S., Lippman D.Z., Curl R.F., Huggins C.M. and Peterson D.E., The Volumetric and Thermodynamic Properties of Fluids. II. Compressibility Factor, Vapor Pressure and Entropy of Vaporization, J. Am. Chem. Soc., 77, 3433-3440, (1955)
  18. Weinberg, S., Foundations of Modern Physics, Cambridge University Press, Cambridge, U.K., pp 4-5 (2021)
  19. Gibbs, J.W. The Collected Works of J. Willard Gibbs Volume II Part One Elementary Principles in Statistical Mechanics, Yale University Press, New Haven, pp vii-xii, (1901 reprinted in 1948)
  20. van der Waals, J.D., Over de Continuïteit van den Gas en Vloeistoftoestand, Ph.D. Thesis, Leiden Univ. (1873)
  21. van der Waals, J.D., Rowlinson, J.S. Ed On the Continuity of the Gaseous and Liquid States, Dover, N.Y. (1988) pp.121-282
  22. Boltzmann, p 218
  23. Andrews, Thomas, The Bakerian Lecture: On the Continuity of the Gaseous and Liquid States of Matter, Philosophical Transactions of the Royal Society of London, {\bf 159}, pp. 575-590, (1869)
  24. Klein, M. J., The Historical Origins of the Van der Waals Equation, Physica, {\bf 73}, p. 31, (1974)
  25. van der Waals, J.D., pp.191-194
  26. Goodstein, D.L., p 450-451
  27. Boltzmann, L., p. 232-233
  28. Goodstein, D.L., p. 452
  29. Rowlinson Ed., p. 19
  30. Lekner, J. Parametric solution of the van der Waals liquid-vapor coexistence curve, Am. J. Phys., 50, 161-163, (1982)
  31. Sommerfeld, pp. 56-57
  32. Goodstein, p 449
  33. Boltzmann, p 237-238
  34. Boltzmann, L., pp 239-240
  35. Barenblatt, G.I., Scaling, Cambridge University Press, pp 22-26, (2003)
  36. 1 2 Johnston, D.C., Thermodynamic Properties of the van der Waals Fluid, arXiv:1402.1205v1, (2014)
  37. Dong, W.G., and Lienhard, J.H., Corresponding States Correlation of Saturated and Metastable Properties, Canad J Chem Eng, 64, 159-161, (1986)
  38. Whitman, A.M., Thermodynamics: Basic Principles and Engineering Applications, Springer, N.Y., pp 155, 203, 204 (2023)
  39. Moran, M.J., and Shapiro, H.N., Fundamentals of Engineering Thermodynamics 4th Edition, pp. 574, 580 (2000)
  40. Whitman, p 155
  41. Moran and Shapiro, p 574
  42. Whitman, p. 203
  43. Moran and Shapiro, p. 574
  44. Whitman, p. 204
  45. Moran and Shapiro, p. 580
  46. Callen, H.B., Thermodynamics, John Wiley \& Sons, N.Y., pp 131-135, (1960)
  47. Epstein, p 10
  48. Callen, pp. 37-44
  49. Callen, p. 153
  50. Callen, pp. 85-101
  51. 1 2 3 Callen, pp 146-156
  52. Maxwell, J.C. On the Dynamical Evidence of the Molecular Constitution of Bodies, Nature, 11 (279), 357-359, (1875)
  53. Shamsundar, N. and Lienhard, J.H., Saturation and Metastable Properties of the van der Waals Fluid, Canad J Chem Eng, 61, 876-880, (1983)
  54. Barrufet, M.A., Eubank, P.T., Generalized Saturation Properties of Pure Fluids Via Cubic Equations of State, ASEE Chemical Engineering Education, 168-175, (1989)
  55. Lekner, J.
  56. Boltzmann, pp. 248-250
  57. Lienhard. J.H., Shamsundar, N., and Biney, P.O., Spinodal Lines and Equations of State: A Review, Nuclear Engineering and Design, 95, 297-314, (1986)
  58. Tien, C.L., and Lienhard, J.H., p.254
  59. Sommerfeld, p 66
  60. Rowlinson, p. 22
  61. Dong, W.G., and Lienhard, J.H., Corresponding States Correlation of Saturated and Metastable Properties, Canad J Chem Eng, 64, 159-161, (1986)
  62. Sommerfeld, pp 61-63
  63. Sommerfeld, pp 60-62
  64. Sommerfeld, p 61
  65. Van Wylen, G.J., and Sonntag, R.E., Fundamentals of Classical Thermodynamics Second Edition, Wiley, N.Y., p. 49, (1973)
  66. Su, G.J., Modified Law of Corresponding States for Real Gases, Ind. Eng. Chem., 38, 803-806, (1946)
  67. Moran, M.J., and Shapiro, H.N., p. 113
  68. Tien and Lienhard, p. 247-248
  69. Redlich, Otto; Kwong, J. N. S. (1 February 1949). "On the Thermodynamics of Solutions. V. An Equation of State. Fugacities of Gaseous Solutions" (PDF). Chemical Reviews. 44 (1): 233–244. doi:10.1021/cr60137a013 . Retrieved 2 April 2024.
  70. 1 2 Niemeyer, Kyle. "Mixture properties". Computational Thermodynamics. Archived from the original on 2 April 2024. Retrieved 2 April 2024.
  71. 1 2 Hewitt, Nigel. "Who was Van der Waals anyway and what has he to do with my Nitrox fill?". Maths for Divers. Archived from the original on 11 March 2020. Retrieved 1 February 2019.
  72. van der Waals, Johannes Diderik. "J. D. van der Waals - Nobel lecture". The Nobel Prize. Nobel Prize Outreach. Archived from the original on 15 November 2023. Retrieved 5 December 2023.
  73. Hill, Terrell L. (1960). An Introduction to Statistical Thermodynamics. Courier Corporation. p. 77.
  74. Denker, John (2014). "Chapter 26.9, Derivation: Particle in a Box". Modern Thermodynamics. CreateSpace Independent Publishing Platform. ISBN   978-1502530356. Archived from the original on 18 August 2016. Retrieved 23 August 2016.

Further reading