You are on page 1of 45

EQUATIONS OF STATES IN THERMODYNAMICS

EQUATION OF STATE
Matter is a substance which has weight, occupies space and has definite shape or takes the shape of the container holding it. It can be broadly classified into 4 types, namely,

1. Solid: The particles (ions, atoms or molecules) are packed closely together. The forces between particles are strong enough so that the particles cannot move freely but can only vibrate. As a result, a solid has a stable, definite shape, and a definite volume. Solids can only change their shape by force, as when broken or cut. 2. Liquid: A liquid is a nearly incompressible fluid that conforms to the shape of its container but retains a (nearly) constant volume independent of pressure. The volume is definite if the temperature and pressure are constant. When a solid is heated above its melting point, it becomes liquid, given that the pressure is higher than the triple point of the substance. Intermolecular (or interatomic or interionic) forces are still important, but the molecules have enough energy to move relative to each other and the structure is mobile. This means that the shape of a liquid is not definite but is determined by its container. The volume is usually greater than that of the corresponding solid, the most well known exception being water, H2O. The highest temperature at which a given liquid can exist is its critical temperature. 3. Gas: A gas is a compressible fluid. Not only will a gas conform to the shape of its container but it will also expand to fill the container. In a gas, the molecules have enough kinetic energy so that the effect of intermolecular forces is small (or zero for an ideal gas) is a compressible fluid. Not only will a gas conform to the shape of its container but it will also expand to fill the container. In a gas, the molecules have enough kinetic energy so that the effect of intermolecular forces is small (or zero for an ideal gas), and the typical distance between neighbouring molecules is much greater than the molecular size. A gas has no definite shape or volume, but occupies the entire container in which it is confined. A liquid may be converted to a gas by heating at constant pressure to the boiling point, or else by reducing the pressure at constant temperature. At temperatures below its critical temperature, a gas is also called a vapour, and can be liquefied by compression alone without cooling. A vapour can exist in equilibrium with a liquid (or solid), in which case the gas pressure equals the vapour pressure of the liquid (or solid).

DOS IN MATERIALS SCIENCE, M.Tech

EQUATIONS OF STATES IN THERMODYNAMICS

A supercritical fluid (SCF) is a gas whose temperature and pressure are above the critical temperature and critical pressure respectively. In this state, the distinction between liquid and gas disappears. A supercritical fluid has the physical properties of a gas, but its high density confers solvent properties in some cases, which leads to useful applications. For example, supercritical carbon dioxide is used to extract caffeine in the manufacture of decaffeinated coffee. 4. Plasma: Like a gas, plasma does not have definite shape or volume. Unlike gases, plasmas are electrically conductive, produce magnetic fields and electric currents, and respond strongly to electromagnetic forces. Positively charged nuclei swim in a "sea" of freely-moving disassociated electrons, similar to the way such charges exist in conductive metal. In fact it is this electron "sea" that allows matter in the plasma state to conduct electricity. The plasma state is often misunderstood, but it is actually quite common on Earth, and the majority of people observe it on a regular basis without even realizing it. Lightning, electric sparks, fluorescent lights, neon lights, plasma televisions, and the Sun are all examples of illuminated matter in the plasma state. A gas is usually converted to plasma in one of two ways, either from a huge voltage difference between two points, or by exposing it to extremely high temperatures. When a gas is heated to extremely high temperatures, such as during nuclear fusion, electrons begin to leave the atoms, resulting in the presence of free electrons. At very high temperatures, such as those present in stars, it is assumed that essentially all electrons are "free," and that very high-energy plasma is essentially bare nuclei swimming in a sea of electrons. In this chapter we consider the State of Matter to be Gaseous. And hence well be completely dealing with the Gas Laws, Gas Equations, Degree of Freedom, Energy Distribution, Paths, Critical Constants and Co-efficients. Derivation of Gas Equation: The state of an amount of gas is determined by its pressure, volume, and temperature. The modern form of the equation relates these simply in two main forms. The temperature used in the equation of state is an absolute temperature: in the SI system of units, kelvin; in the Imperial system, degrees Rankine.

DOS IN MATERIALS SCIENCE, M.Tech

EQUATIONS OF STATES IN THERMODYNAMICS

Common form The most frequently introduced form is

where P is the pressure of the gas, V is the volume of the gas, n is the amount of substance of gas (also known as number of moles), T is the temperature of the gas and R is the ideal, or universal, gas constant, equal to the product of Boltzmann's constant and Avogadro's constant. In SI units, P is measured in pascals, V is measured in cubic metres, n is measured in moles, and T in kelvin (273.15 Kelvin = 0.01 degrees Celsius). R has the value 8.314 JK1mol1 or 0.08206 Latmmol1K1 if using pressure in standard atmospheres (atm) instead of pascals, and volume in litres instead of cubic metres. Molar form How much gas is present could be specified by giving the mass instead of the chemical amount of gas. Therefore, an alternative form of the ideal gas law may be useful. The chemical amount (n) (in moles) is equal to the mass (m) (in grams) divided by the molar mass (M) (in grams per mole):

By replacing n with m / M, and subsequently introducing density = m/V, we get:

Defining the specific gas constant Rspecific as the ratio R/M,

This form of the ideal gas law is very useful because it links pressure, density, and temperature in a unique foLrmula independent of the quantity of the considered gas. Alternatively, the law may be written in terms of the specific volume v, the reciprocal of density, as

DOS IN MATERIALS SCIENCE, M.Tech

EQUATIONS OF STATES IN THERMODYNAMICS

It is common, especially in engineering applications, to represent the ideal gas constant by the symbol R. In such cases, the universal gas constant is usually given a different symbol such as R to distinguish it. In any case, the context and/or units of the gas constant should make it clear as to whether the universal or specific gas constant is being referred to. Statistical mechanics In statistical mechanics the following molecular equation is derived from first principles:

where P is the absolute pressure of the gas measured in Pascals; V is the volume (in this equation the volume is expressed in meters cubed, as Pascal times cubic meter equals one Joule); N is the number of particles in the gas; k is the Boltzmann constant relating temperature and energy; and T is the absolute temperature. The actual number of molecules contrasts to the other formulation, which uses n, the number of moles. This relation implies that Nk = nR, and the consistency of this result with experiment is a good check on the principles of statistical mechanics. From this we can notice that for an average particle mass of times the atomic mass constant mu (i.e., the mass is u)

and since = m/V, we find that the ideal gas law can be rewritten as:

In SI units, P is measured in pascals; V in cubic metres; N is a dimensionless number; and T in kelvin. k has the value 1.381023 JK1 in SI units.

Gas laws: The early gas laws were developed at the end of the 18th century, when scientists began to realize that relationships between the pressure, volume and temperature of a sample of gas
DOS IN MATERIALS SCIENCE, M.Tech 4

EQUATIONS OF STATES IN THERMODYNAMICS

could be obtained which would hold for all gases. Gases behave in a similar way over a wide variety of conditions because to a good approximation they all have molecules which are widely spaced, and nowadays the equation of state for an ideal gas is derived from kinetic theory. The earlier gas laws are now considered as special cases of the ideal gas equation, with one or more of the variables held constant. Boyle's law: Boyle's law shows that, at constant temperature, the product of an ideal gas's pressure and volume is always constant. It was published in 1662. It can be determined experimentally using a pressure gauge and a variable volume container. It can also be found through the use of logic; if a container, with a fixed number of molecules inside, is reduced in volume, more molecules will hit the sides of the container per unit time, causing a greater pressure. As a mathematical equation, Boyle's law is:

where P is the pressure (Pa), V the volume (m3) of an gas, and k1 (measured in joules) is the constant from this equationit is not the same as the constants from the other equations below.

This is known as Boyle's law which states: the volume of a given mass of gas is inversely proportional to its pressure, if the temperature remains constant. Mathematically this is:

where k is a constant (i.e. NOT Boltzmann's constant) Charles's law Charles's Law, or the law of volumes, was found in 1787 by Jacques Charles. It says that, for an ideal gas at constant pressure, the volume is directly proportional to its temperature.

DOS IN MATERIALS SCIENCE, M.Tech

EQUATIONS OF STATES IN THERMODYNAMICS

Gay-Lussac's law Gay-Lussac's law, or the pressure law, was found by Joseph Louis Gay-Lussac in 1809. It states that the pressure exerted on a container's sides by an ideal gas is proportional to its temperature.

Avogadro's law Avogadro's law states that the volume occupied by an ideal gas is proportional to the number of moles (or molecules) present in the container. This gives rise to the molar volume of a gas, which at STP is 22.4 dm3 (or litres). The relation is given by

where n is equal to the number of moles of gas (the number of molecules divided by Avogadro's Number).

Combined and ideal gas laws The combined gas law or general gas equation is formed by the combination of the three laws, and shows the relationship between the pressure, volume, and temperature for a fixed mass of gas:

With the addition of Avogadro's law, the combined gas law develops into the ideal gas law:

where P is pressure V is volume n is the number of moles


DOS IN MATERIALS SCIENCE, M.Tech 6

EQUATIONS OF STATES IN THERMODYNAMICS

R is the universal gas constant T is temperature (K) where the constant, now named R, is the gas constant with a value of .08206 (atmL)/(molK). An equivalent formulation of this law is:

where P is the absolute pressure V is the volume N is the number of gas molecules k is the Boltzmann constant (1.3811023 JK1 in SI units) T is the temperature (K)

These equations are exact only for an ideal gas, which neglects various intermolecular effects (see real gas). However, the ideal gas law is a good approximation for most gases under moderate pressure and temperature. This law has the following important consequences:

1. If temperature and pressure are kept constant, then the volume of the gas is directly proportional to the number of molecules of gas. 2. If the temperature and volume remain constant, then the pressure of the gas changes is directly proportional to the number of molecules of gas present. 3. If the number of gas molecules and the temperature remain constant, then the pressure is inversely proportional to the volume. 4. If the temperature changes and the number of gas molecules are kept constant, then either pressure or volume (or both) will change in direct proportion to the temperature.

DOS IN MATERIALS SCIENCE, M.Tech

EQUATIONS OF STATES IN THERMODYNAMICS

Avagadro's Hypothesis Avogadro's law (sometimes referred to as Avogadro's hypothesis or Avogadro's principle) is a gas law which states that, under the same condition of temperature and pressure, equal volumes of all gases contain the same number of molecules. The law is named after Amedeo Avogadro who, in 1811,[1] hypothesized that two given samples of an ideal gas, of the same volume and at the same temperature and pressure, contain the same number of molecules. Thus, the number of molecules or atoms in a specific volume of ideal gas is independent of their size or the molar mass of the gas. As an example, equal volumes of molecular hydrogen and nitrogen contain the same number of molecules when they are at the same temperature and pressure, and observe ideal gas behavior. In practice, real gases show small deviations from the ideal behavior and the law holds only approximately, but is still a useful approximation for scientists. Mathematical definition Avogadro's law is stated mathematically as:

Where: V is the volume of the gas(es). n is the amount of substance of the gas. k is a proportionality constant. The most significant consequence of Avogadro's law is that the ideal gas constant has the same value for all gases. This means that:

Where: p is the pressure of the gas in the cell T is the temperature in kelvin of the gas
DOS IN MATERIALS SCIENCE, M.Tech 8

EQUATIONS OF STATES IN THERMODYNAMICS

Ideal gas law A common rearrangement of this equation is by letting R be the proportionality constant, and rearranging as follows:

This equation is known as the ideal gas law. Molar volume Taking STP to be 101.325 kPa and 273.15 K, we can find the volume of one mole of a gas:

For 100.000 kPa and 273.15 K, the molar volume of an ideal gas is 22.414 dm3mol-1. Graham's Law of Diffusion Graham's law, known as Graham's law of effusion, was formulated by Scottish physical chemist Thomas Graham in 1846. Graham found experimentally that the rate of effusion of a gas is inversely proportional to the square root of the mass of its particles. This formula can be written as:

where: Rate1 is the rate of effusion of the first gas (volume or number of moles per unit time). Rate2 is the rate of effusion for the second gas. M1 is the molar mass of gas 1 M2 is the molar mass of gas 2.

DOS IN MATERIALS SCIENCE, M.Tech

EQUATIONS OF STATES IN THERMODYNAMICS

Graham's law states that the rate of effusion of a gas is inversely proportional to the square root of its molecular weight. Thus, if the molecular weight of one gas is four times that of another, it would diffuse through a porous plug or escape through a small pinhole in a vessel at half the rate of the other. A complete theoretical explanation of Graham's law was provided years later by the kinetic theory of gases. Graham's law provides a basis for separating isotopes by diffusion a method that came to play a crucial role in the development of the atomic bomb. Graham's law is most accurate for molecular effusion which involves the movement of one gas at a time through a hole. It is only approximate for diffusion of one gas in another or in air, as these processes involve the movement of more than one gas. History Graham's research on the diffusion of gases was triggered by his reading about the observation of German chemist Johann Dbereiner that hydrogen gas diffused out of a small crack in a glass bottle faster than the surrounding air diffused in to replace it. Graham measured the rate of diffusion of gases through plaster plugs, through very fine tubes, and through small orifices. In this way he slowed down the process so that it could be studied quantitatively. He first stated the law as we know it today in 1831. Graham went on to study the diffusion of substances in solution and in the process made the discovery that some apparent solutions actually are suspensions of particles too large to pass through a parchment filter. He termed these materials colloids, a term that has come to denote an important class of finely divided materials. At the time Graham did his work the concept of molecular weight was being established, in large part through measurements of gases. Italian physicist Amadeo Avogadro had suggested in 1811 that equal volumes of different gases contain equal numbers of molecules. Thus, the relative molecular weights of two gases are equal to the ratio of weights of equal volumes of the gases. Avogadro's insight together with other studies of gas behaviour provided a basis for later theoretical work by Scottish physicist James Clerk Maxwell to explain the properties of gases as collections of small particles moving through largely empty space. Perhaps the greatest success of the kinetic theory of gases, as it came to be called, was the discovery that for gases, the temperature as measured on the Kelvin (absolute) temperature scale is directly proportional to the average kinetic energy of the gas molecules. The kinetic energy of any object is equal to one-half its mass times the square of its velocity. Thus, to
DOS IN MATERIALS SCIENCE, M.Tech 10

EQUATIONS OF STATES IN THERMODYNAMICS

have equal kinetic energies, the velocities of two different molecules would have to be in inverse proportion to the square roots of their masses. The rate of effusion is determined by the number of molecules entering an aperture per unit time, and hence by the average molecular velocity. Graham's law for diffusion could thus be understood as a consequence of the molecular kinetic energies being equal at the same temperature. Example Let gas 1 be H2 and gas 2 be O2.

Therefore, hydrogen molecules effuse four times faster than those of oxygen. Graham's Law can also be used to find the approximate molecular weight of a gas if one gas is a known species, and if there is a specific ratio between the rates of two gases (such as in the previous example). The equation can be solved for either one of the molecular weights provided the subscripts are consistent.

Graham's law was the basis for separating 235U from 238U found in natural uraninite (uranium ore) during the Manhattan project to build the first atomic bomb. The United States government built a gaseous diffusion plant at the then phenomenal cost of $100 million in Clinton, Tennessee. In this plant, uranium from uranium ore was first converted to uranium hexafluoride and then forced repeatedly to diffuse through porous barriers, each time becoming a little more enriched in the slightly lighter 235U isotope. Degree of Freedom: A degree of freedom is an independent physical parameter, often called a dimension, in the formal description of the state of a physical system. The set of all dimensions of a system is known as a phase space.

DOS IN MATERIALS SCIENCE, M.Tech

11

EQUATIONS OF STATES IN THERMODYNAMICS

Definition In physics, a degree of freedom of a system is a formal description of a parameter that contributes to the state of a physical system. It can also be defined as the minimum number of coordinates required to specify the position of a particle or system of particles. In mechanics, a point particle can move independently in the three directions of space. Thus, the momentum of a particle consists of three components, each called a degree of freedom. A system of N independent particles, therefore, has the total of 3N degrees of freedom. Similarly in statistical mechanics, a degree of freedom is a single scalar number describing the microstate of a system. The specification of all microstates of a system is a point in the system's phase space. A degree of freedom may be any useful property that is not dependent on other variables. For example, in the 3D ideal chain model, two angles are necessary to describe each monomer's orientation. Degrees of freedom of gas molecules

Different ways of visualizing the 3 degrees of freedom of a dumbbell-shaped diatomic molecule. (CM: center of mass of the system, T: translational motion, R: rotational motion, V: vibrational motion.) In three-dimensional space, three degrees of freedom are associated with the movement of a mechanical particle. A diatomic gas molecule thus has 6 degrees of freedom. This set may be

DOS IN MATERIALS SCIENCE, M.Tech

12

EQUATIONS OF STATES IN THERMODYNAMICS

decomposed in terms of translations, rotations, and vibrations of the molecule. The center of mass motion of the entire molecule accounts for 3 degrees of freedom. In addition, the molecule has two rotational degrees of motion and one vibrational mode. The rotations occur around the two axes perpendicular to the line between the two atoms. The rotation around the atom-atom bond is not counted. This yields, for a diatomic molecule, a decomposition of: 3N = 6 = 3+2+1. For a general (non-linear) molecule with N > 2 atoms, all 3 rotational degrees of freedom are considered, resulting in the decomposition: 3N = 3 + 3 + (3N - 6) which means that an N-atom molecule has 3N - 6 vibrational degrees of freedom for N > 2. As defined above one can also count degrees of freedom using the minimum number of coordinates required to specify a position. This is done as follows: 1. For a single particle we need 2 coordinates in a 2-D plane to specify its position and 3 coordinates in 3-D plane. Thus its degree of freedom in a 3-D plane is 3. 2. For a body consisting of 2 particles (ex. a diatomic molecule) in a 3-D plane with constant distance between them (let's say d) we can show (below) its degree of freedom to be 5. Let's say one particle in this body has coordinates (x1,y1,z1) and the other has x-coordinate(x2) and y-coordinate(y2). Application of the formula for distance between two coordinates (

) results in one equation with one unknown, in which we can solve for z2. (Note:Here any one of x1, x2, y1, y2, z1, or z2 can be unknown.) Contrary to the classical equipartition theorem, at room temperature, the vibrational motion of molecules typically makes negligible contributions to the heat capacity. This is because these degrees of freedom are frozen because the spacing between the energy eigenvalues exceeds the energy corresponding to ambient temperatures (kT). In the following table such degrees of freedom are disregarded because of their low effect on total energy. However, at very high temperatures they cannot be neglected.

DOS IN MATERIALS SCIENCE, M.Tech

13

EQUATIONS OF STATES IN THERMODYNAMICS

Monatomic Linear molecules Non-Linear molecules Position (x, y and z) 3 Rotation (x, y and z) 0 Vibration Total 0 3 3 2 3N - 5 3N 3 3 3N - 6 3N

Independent degrees of freedom The set of degrees of freedom of a system is independent if the energy

associated with the set can be written in the following form:

where

is a function of the sole variable and

. is the associated energy:

example: if

are two degrees of freedom, and

If If

, then the two degrees of freedom are independent. , then the two degrees of freedom are not and is a coupling term,

independent. The term involving the product of

that describes an interaction between the two degrees of freedom. At thermodynamic equilibrium, are all statistically independent of each other. is distributed according to the

For i from 1 to N, the value of the ith degree of freedom

Boltzmann distribution. Its probability density function is the following:

, In this section, and throughout the article the brackets enclose.


DOS IN MATERIALS SCIENCE, M.Tech 14

denote the mean of the quantity they

EQUATIONS OF STATES IN THERMODYNAMICS

The internal energy of the system is the sum of the average energies associated to each of the degrees of freedom:

Demonstrations A system exchanges energy in the form of heat with its surroundings and the number of particles in the system remains fixed. This corresponds to studying the system in the canonical ensemble. Note that in statistical mechanics, a result that is demonstrated for a system in a particular ensemble remains true for this system at the thermodynamic limit in any ensemble. In the canonical ensemble, at thermodynamic equilibrium, the state of the system is distributed among all micro-states according to the Boltzmann distribution. If the system's temperature and is

is Boltzmann's constant, then the probability density

function associated to each micro-state is the following:

, The denominator in the above expression plays an important role.[1] This expression immediately breaks down into a product of terms depending of a single degree of freedom:

The existence of such a breakdown of the multidimensional probability density function into a product of functions of one variable is enough by itself to demonstrate that statistically independent from each other. Since each function is normalized, it follows immediately that , for i from 1 to N. is the probability density are

function of the degree of freedom

Finally, the internal energy of the system is its mean energy. The energy of a degree of freedom is a function of the sole variable . Since are independent from

each other, the energies each other.

are also statistically independent from

DOS IN MATERIALS SCIENCE, M.Tech

15

EQUATIONS OF STATES IN THERMODYNAMICS

The total internal energy of the system can thus be written as:

Quadratic degrees of freedom A degree of freedom can be written as , where is a linear combination of other quadratic degrees of freedom. and are two degrees of freedom, and is the associated energy: is quadratic if the energy terms associated to this degree of freedom

example: if

If independent and non-quadratic.

, then the two degrees of freedom are not

If non-quadratic.

, then the two degrees of freedom are independent and

If independent but are quadratic.

, then the two degrees of freedom are not

If quadratic.

, then the two degrees of freedom are independent and

For example, in Newtonian mechanics, the dynamics of a system of quadratic degrees of freedom are controlled by a set of homogeneous linear differential equations with constant coefficients. Quadratic and independent degree of freedom are quadratic and independent degrees of freedom if the energy associated to a microstate of the system they represent can be written as:

DOS IN MATERIALS SCIENCE, M.Tech

16

EQUATIONS OF STATES IN THERMODYNAMICS

Equipartition theorem In the classical limit of statistical mechanics, at thermodynamic equilibrium, the internal energy of a system of N quadratic and independent degrees of freedom is:

Here, the mean energy associated with a degree of freedom is:

Since the degrees of freedom are independent, the internal energy of the system is equal to the sum of the mean energy associated with each degree of freedom, which demonstrates the result. In classical statistical mechanics, the equipartition theorem is a general formula that relates the temperature of a system with its average energies. The equipartition theorem is also known as the law of equipartition, equipartition of energy, or simply equipartition. The original idea of equipartition was that, in thermal equilibrium, energy is shared equally among all of its various forms; for example, the average kinetic energy per degree of freedom in the translational motion of a molecule should equal that of its rotational motions. The equipartition theorem makes quantitative predictions. Like the virial theorem, it gives the total average kinetic and potential energies for a system at a given temperature, from which the system's heat capacity can be computed. However, equipartition also gives the average values of individual components of the energy, such as the kinetic energy of a particular particle or the potential energy of a single spring. For example, it predicts that every atom in a monoatomic ideal gas has an average kinetic energy of (3/2)kBT in thermal equilibrium, where kB is the Boltzmann constant and T is the (thermodynamic) temperature. More generally, it can be applied to any classical system in thermal equilibrium, no matter how complicated.

DOS IN MATERIALS SCIENCE, M.Tech

17

EQUATIONS OF STATES IN THERMODYNAMICS

The equipartition theorem can be used to derive the ideal gas law, and the DulongPetit law for the specific heat capacities of solids. It can also be used to predict the properties of stars, even white dwarfs and neutron stars, since it holds even when relativistic effects are considered. Although the equipartition theorem makes very accurate predictions in certain conditions, it becomes inaccurate when quantum effects are significant, such as at low temperatures. When the thermal energy kBT is smaller than the quantum energy spacing in a particular degree of freedom, the average energy and heat capacity of this degree of freedom are less than the values predicted by equipartition. Such a degree of freedom is said to be "frozen out" when the thermal energy is much smaller than this spacing. For example, the heat capacity of a solid decreases at low temperatures as various types of motion become frozen out, rather than remaining constant as predicted by equipartition. Such decreases in heat capacity were among the first signs to physicists of the 19th century that classical physics was incorrect and that a new, more subtle, scientific model was required. Along with other evidence, equipartition's failure to model black-body radiationalso known as the ultraviolet catastropheled Max Planck to suggest that energy in the oscillators in an object, which emit light, were quantized, a revolutionary hypothesis that spurred the development of quantum mechanics and quantum field theory. Basic concept and simple examples

Figure 2. Probability density functions of the molecular speed for four noble gases at a temperature of 298.15 K (25 C). The four gases are helium (4He), neon (20Ne), argon (40Ar)

DOS IN MATERIALS SCIENCE, M.Tech

18

EQUATIONS OF STATES IN THERMODYNAMICS

and xenon (132Xe); the superscripts indicate their mass numbers. These probability density functions have dimensions of probability times inverse speed; since probability is dimensionless, they can be expressed in units of seconds per meter. The name "equipartition" means "equal division," as derived from the Latin equi from the antecedent, quus ("equal or even"), and partition from the antecedent, partitionem ("division, portion"). The original concept of equipartition was that the total kinetic energy of a system is shared equally among all of its independent parts, on the average, once the system has reached thermal equilibrium. Equipartition also makes quantitative predictions for these energies. For example, it predicts that every atom of a noble gas, in thermal equilibrium at temperature T, has an average translational kinetic energy of (3/2)kBT, where kB is the Boltzmann constant. As a consequence, since kinetic energy is equal to 1/2(mass)(velocity)2, the heavier atoms of xenon have a lower average speed than do the lighter atoms of helium at the same temperature. Figure 2 shows the MaxwellBoltzmann distribution for the speeds of the atoms in four noble gases. In this example, the key point is that the kinetic energy is quadratic in the velocity. The equipartition theorem shows that in thermal equilibrium, any degree of freedom (such as a component of the position or velocity of a particle) which appears only quadratically in the energy has an average energy of 12kBT and therefore contributes 12kB to the system's heat capacity. This has many applications. Translational energy and ideal gases The (Newtonian) kinetic energy of a particle of mass m, velocity v is given by

where vx, vy and vz are the Cartesian components of the velocity v. Here, H is short for Hamiltonian, and used henceforth as a symbol for energy because the Hamiltonian formalism plays a central role in the most general form of the equipartition theorem. Since the kinetic energy is quadratic in the components of the velocity, by equipartition these three components each contribute 12kBT to the average kinetic energy in thermal equilibrium. Thus the average kinetic energy of the particle is (3/2)kBT, as in the example of noble gases above.

DOS IN MATERIALS SCIENCE, M.Tech

19

EQUATIONS OF STATES IN THERMODYNAMICS

More generally, in an ideal gas, the total energy consists purely of (translational) kinetic energy: by assumption, the particles have no internal degrees of freedom and move independently of one another. Equipartition therefore predicts that the average total energy of an ideal gas of N particles is (3/2) NkBT. It follows that the heat capacity of the gas is (3/2) NkB and hence, in particular, the heat capacity of a mole of such gas particles is (3/2)NAkB = (3/2)R, where NA is the Avogadro constant and R is the gas constant. Since R 2 cal/(molK), equipartition predicts that the molar heat capacity of an ideal gas is roughly 3 cal/(molK). This prediction is confirmed by experiment. The mean kinetic energy also allows the root mean square speed vrms of the gas particles to be calculated:

where M = NAm is the mass of a mole of gas particles. This result is useful for many applications such as Graham's law of effusion, which provides a method for enriching uranium.[4] Rotational energy and molecular tumbling in solution A similar example is provided by a rotating molecule with principal moments of inertia I1, I2 and I3. The rotational energy of such a molecule is given by

where 1, 2, and 3 are the principal components of the angular velocity. By exactly the same reasoning as in the translational case, equipartition implies that in thermal equilibrium the average rotational energy of each particle is (3/2)kBT. Similarly, the equipartition theorem allows the average (more precisely, the root mean square) angular speed of the molecules to be calculated. The tumbling of rigid moleculesthat is, the random rotations of molecules in solution plays a key role in the relaxations observed by nuclear magnetic resonance, particularly protein NMR and residual dipolar couplings.[6] Rotational diffusion can also be observed by other biophysical probes such as fluorescence anisotropy, flow birefringence and dielectric spectroscopy.
DOS IN MATERIALS SCIENCE, M.Tech 20

EQUATIONS OF STATES IN THERMODYNAMICS

Potential energy and harmonic oscillators Equipartition applies to potential energies as well as kinetic energies: important examples include harmonic oscillators such as a spring, which has a quadratic potential energy

where the constant a describes the stiffness of the spring and q is the deviation from equilibrium. If such a one dimensional system has mass m, then its kinetic energy Hkin is

where v and p = mv denote the velocity and momentum of the oscillator. Combining these terms yields the total energy[8]

Equipartition therefore implies that in thermal equilibrium, the oscillator has average energy

where the angular brackets

denote the average of the enclosed quantity,[9]

This result is valid for any type of harmonic oscillator, such as a pendulum, a vibrating molecule or a passive electronic oscillator. Systems of such oscillators arise in many situations; by equipartition, each such oscillator receives an average total energy kBT and hence contributes kB to the system's heat capacity. This can be used to derive the formula for JohnsonNyquist noise[10] and the DulongPetit law of solid heat capacities. The latter application was particularly significant in the history of equipartition.

DOS IN MATERIALS SCIENCE, M.Tech

21

EQUATIONS OF STATES IN THERMODYNAMICS

Figure 3. Atoms in a crystal can vibrate about their equilibrium positions in the lattice. Such vibrations account largely for the heat capacity of crystalline dielectrics; with metals, electrons also contribute to the heat capacity. Specific heat capacity of solids An important application of the equipartition theorem is to the specific heat capacity of a crystalline solid. Each atom in such a solid can oscillate in three independent directions, so the solid can be viewed as a system of 3N independent simple harmonic oscillators, where N denotes the number of atoms in the lattice. Since each harmonic oscillator has average energy kBT, the average total energy of the solid is 3NkBT, and its heat capacity is 3NkB. By taking N to be the Avogadro constant NA, and using the relation R = NAkB between the gas constant R and the Boltzmann constant kB, this provides an explanation for the DulongPetit law of specific heat capacities of solids, which stated that the specific heat capacity (per unit mass) of a solid element is inversely proportional to its atomic weight. A modern version is that the molar heat capacity of a solid is 3R 6 cal/(molK). However, this law is inaccurate at lower temperatures, due to quantum effects; it is also inconsistent with the experimentally derived third law of thermodynamics, according to which the molar heat capacity of any substance must go to zero as the temperature goes to absolute zero.[10] A more accurate theory, incorporating quantum effects, was developed by Albert Einstein (1907) and Peter Debye (1911).[11] Many other physical systems can be modeled as sets of coupled oscillators. The motions of such oscillators can be decomposed into normal modes, like the vibrational modes of a piano string or the resonances of an organ pipe. On the other hand, equipartition often breaks down for such systems, because there is no exchange of energy between the normal modes. In an extreme situation, the modes are independent and so their energies are independently conserved. This shows that some sort of mixing of energies, formally called ergodicity, is important for the law of equipartition to hold. Sedimentation of particles Potential energies are not always quadratic in the position. However, the equipartition theorem also shows that if a degree of freedom x contributes only a multiple of xs (for a fixed real number s) to the energy, then in thermal equilibrium the average energy of that part is kBT/s.
DOS IN MATERIALS SCIENCE, M.Tech 22

EQUATIONS OF STATES IN THERMODYNAMICS

There is a simple application of this extension to the sedimentation of particles under gravity.[12] For example, the haze sometimes seen in beer can be caused by clumps of proteins that scatter light.[13] Over time, these clumps settle downwards under the influence of gravity, causing more haze near the bottom of a bottle than near its top. However, in a process working in the opposite direction, the particles also diffuse back up towards the top of the bottle. Once equilibrium has been reached, the equipartition theorem may be used to determine the average position of a particular clump of buoyant mass mb. For an infinitely tall bottle of beer, the gravitational potential energy is given by

where z is the height of the protein clump in the bottle and g is the acceleration due to gravity. Since s = 1, the average potential energy of a protein clump equals kBT. Hence, a protein clump with a buoyant mass of 10 MDa (roughly the size of a virus) would produce a haze with an average height of about 2 cm at equilibrium. The process of such sedimentation to equilibrium is described by the MasonWeaver equation. History For an approximate conversion to the corresponding SI unit of J/(molK), such values should be multiplied by 4.2 J/cal. The equipartition of kinetic energy was proposed initially in 1843, and more correctly in 1845, by John James Waterston. In 1859, James Clerk Maxwell argued that the kinetic heat energy of a gas is equally divided between linear and rotational energy. In 1876, Ludwig Boltzmann expanded on this principle by showing that the average energy was divided equally among all the independent components of motion in a system. Boltzmann applied the equipartition theorem to provide a theoretical explanation of the DulongPetit law for the specific heat capacities of solids.

DOS IN MATERIALS SCIENCE, M.Tech

23

EQUATIONS OF STATES IN THERMODYNAMICS

Figure 4. Idealized plot of the molar specific heat of a diatomic gas against temperature. It agrees with the value (7/2)R predicted by equipartition at high temperatures (where R is the gas constant), but decreases to (5/2)R and then (3/2)R at lower temperatures, as the vibrational and rotational modes of motion are "frozen out". The failure of the equipartition theorem led to a paradox that was only resolved by quantum mechanics. For most molecules, the transitional temperature Trot is much less than room temperature, whereas Tvib can be ten times larger or more. A typical example is carbon monoxide, CO, for which Trot 2.8 K and Tvib 3103 K. For molecules with very large or weakly bound atoms, Tvib can be close to room temperature (about 300 K); for example, Tvib 308 K for iodine gas, I2. The history of the equipartition theorem is intertwined with that of specific heat capacity, both of which were studied in the 19th century. In 1819, the French physicists Pierre Louis Dulong and Alexis Thrse Petit discovered that the specific heat capacities of solid elements at room temperature were inversely proportional to the atomic weight of the element. Their law was used for many years as a technique for measuring atomic weights. However, subsequent studies by James Dewar and Heinrich Friedrich Weber showed that this Dulong Petit law holds only at high temperatures; at lower temperatures, or for exceptionally hard solids such as diamond, the specific heat capacity was lower. Experimental observations of the specific heat capacities of gases also raised concerns about the validity of the equipartition theorem. The theorem predicts that the molar heat capacity of simple monoatomic gases should be roughly 3 cal/(molK), whereas that of diatomic gases should be roughly 7 cal/(molK). Experiments confirmed the former prediction, but found that molar heat capacities of diatomic gases were typically about 5 cal/(molK), and fell to about 3 cal/(molK) at very low temperatures. Maxwell noted in 1875 that the disagreement between experiment and the equipartition theorem was much worse than even these numbers suggest; since atoms have internal parts, heat energy should go into the motion of these internal parts, making the predicted specific heats of monoatomic and diatomic gases much higher than 3 cal/(molK) and 7 cal/(molK), respectively. A third discrepancy concerned the specific heat of metals. According to the classical Drude model, metallic electrons act as a nearly ideal gas, and so they should contribute (3/2) NekB to the heat capacity by the equipartition theorem, where Ne is the number of electrons. Experimentally, however, electrons contribute little to the heat capacity: the molar heat capacities of many conductors and insulators are nearly the same.

DOS IN MATERIALS SCIENCE, M.Tech

24

EQUATIONS OF STATES IN THERMODYNAMICS

Several explanations of equipartition's failure to account for molar heat capacities were proposed. Boltzmann defended the derivation of his equipartition theorem as correct, but suggested that gases might not be in thermal equilibrium because of their interactions with the ether. Lord Kelvin suggested that the derivation of the equipartition theorem must be incorrect, since it disagreed with experiment, but was unable to show how. In 1900 Lord Rayleigh instead put forward a more radical view that the equipartition theorem and the experimental assumption of thermal equilibrium were both correct; to reconcile them, he noted the need for a new principle that would provide an "escape from the destructive simplicity" of the equipartition theorem. Albert Einstein provided that escape, by showing in 1906 that these anomalies in the specific heat were due to quantum effects, specifically the quantization of energy in the elastic modes of the solid. Einstein used the failure of equipartition to argue for the need of a new quantum theory of matter. Nernst's 1910 measurements of specific heats at low temperatures supported Einstein's theory, and led to the widespread acceptance of quantum theory among physicists. General formulation of the equipartition theorem The most general form of the equipartition theorem states that under suitable assumptions (discussed below), for a physical system with Hamiltonian energy function H and degrees of freedom xn, the following equipartition formula holds in thermal equilibrium for all indices m and n:

Here mn is the Kronecker delta, which is equal to one if m = n and is zero otherwise. The averaging brackets is assumed to be an ensemble average over phase space or, under an

assumption of ergodicity, a time average of a single system. The general equipartition theorem holds in both the microcanonical ensemble,[9] when the total energy of the system is constant, and also in the canonical ensemble, when the system is coupled to a heat bath with which it can exchange energy. Derivations of the general formula are given later in the article.

DOS IN MATERIALS SCIENCE, M.Tech

25

EQUATIONS OF STATES IN THERMODYNAMICS

The general formula is equivalent to the following two:

1.

2. If a degree of freedom xn appears only as a quadratic term anxn2 in the Hamiltonian H, then the first of these formulae implies that

which is twice the contribution that this degree of freedom makes to the average energy

Thus the equipartition theorem for systems with quadratic energies follows easily from the general formula. A similar argument, with 2 replaced by s, applies to energies of the form anxns. The degrees of freedom xn are coordinates on the phase space of the system and are therefore commonly subdivided into generalized position coordinates qk and generalized momentum coordinates pk, where pk is the conjugate momentum to qk. In this situation, formula 1 means that for all k,

Using the equations of Hamiltonian mechanics, these formulae may also be written

Similarly, one can show using formula 2 that

and

DOS IN MATERIALS SCIENCE, M.Tech

26

EQUATIONS OF STATES IN THERMODYNAMICS

Relation to the virial theorem The general equipartition theorem is an extension of the virial theorem (proposed in 1870), which states that

where t denotes time. Two key differences are that the virial theorem relates summed rather than individual averages to each other, and it does not connect them to the temperature T. Another difference is that traditional derivations of the virial theorem use averages over time, whereas those of the equipartition theorem use averages over phase space. Mean free path In physics, the mean free path is the average distance travelled by a moving particle (such as an atom, a molecule, a photon) between successive impacts (collisions) which modify its direction or energy or other particle properties. Derivation

Figure: Slab of target Imagine a beam of particles being shot through a target, and consider an infinitesimally thin slab of the target (Figure). The atoms (or particles) that might stop a beam particle are shown in red. The magnitude of the mean free path depends on the characteristics of the system the particle is in:

DOS IN MATERIALS SCIENCE, M.Tech

27

EQUATIONS OF STATES IN THERMODYNAMICS

Where is the mean free path, n is the number of target particles per unit volume, and the effective cross sectional area for collision. The area of the slab is and its volume is

is

. The typical number of stopping atoms in . The probability that a beam

the slab is the concentration n times the volume, i.e.,

particle will be stopped in that slab is the net area of the stopping atoms divided by the total area of the slab.

where

is the area (or, more formally, the "scattering cross-section") of one atom.

The drop in beam intensity equals the incoming beam intensity multiplied by the probability of the particle being stopped within the slab

This is an ordinary differential equation

whose solution is known as Beer-Lambert law and has the form

, where x is

the distance travelled by the beam through the target and I0 is the beam intensity before it entered the target; is called the mean free path because it equals the mean distance traveled by a beam particle before being stopped. To see this, note that the probability that a particle is absorbed between x and x + dx is given by

Thus the expectation value (or average, or simply mean) of x is

DOS IN MATERIALS SCIENCE, M.Tech

28

EQUATIONS OF STATES IN THERMODYNAMICS

The fraction of particles that are not stopped (attenuated) by the slab is called transmission

where x is equal to the thickness of the slab x = dx. Mean free path in kinetic theory In kinetic theory the mean free path of a particle, such as a molecule, is the average distance the particle travels between collisions with other moving particles. The formula still holds for a particle with a high velocity relative to the velocities of an ensemble of identical particles with random locations. If, on the other hand, the velocities of the identical particles have a Maxwell distribution, the following relationship applies:

and it may be shown that the mean free path, in meters, is:

where kB is the Boltzmann constant in J/K, T is the temperature in K, p is pressure in Pascals, and d is the diameter of the gas particles in meters. Following table lists some typical values for air at different pressures and at room temperature. Pressure in hPa (mbar) 1013 300 1 1 103 103 107 107 1012 <1012

Vacuum range
Ambient pressure

Molecules / cm3 2.7 1019 1019 1016 1016 1013 1013 109 109 104 <104

Molecules / m3 Mean free path 2.7 1025 1025 1022 1022 1019 1019 1015 1015 1010 <1010 68 nm[4] 0.1 100 m 0.1 100 mm 10 cm 1 km 1 km 105 km >105 km

Low vacuum Medium vacuum High vacuum


Ultra high vacuum

Extremely high vacuum

DOS IN MATERIALS SCIENCE, M.Tech

29

EQUATIONS OF STATES IN THERMODYNAMICS

Mean free path in radiography

Mean free path for photons in energy range from 1 keV to 20 MeV for Elements Z = 1 to 100. Based on data from. The discontinuities are due to low density of gas elements. Six bands correspond to neighborhoods of six noble gases. Also shown are locations of absorption edges In gamma-ray radiography the mean free path of a pencil beam of mono-energetic photons is the average distance a photon travels between collisions with atoms of the target material. It depends on the material and the energy of the photons:

where is the linear attenuation coefficient, / is the mass attenuation coefficient and is the density of the material. The Mass attenuation coefficient can be looked up or calculated for any material and energy combination using the NIST databases In X-ray radiography the calculation of the mean free path is more complicated, because photons are not mono-energetic, but have some distribution of energies called spectrum. As photons move through the target material they are attenuated with probabilities depending on their energy, as a result their distribution changes in process called Spectrum Hardening. Because of Spectrum Hardening the mean free path of the X-ray spectrum changes with distance.

DOS IN MATERIALS SCIENCE, M.Tech

30

EQUATIONS OF STATES IN THERMODYNAMICS

Sometimes one measures the thickness of a material in the number of mean free paths. Material with the thickness of one mean free path will attenuate 37% (1/e) of photons. This concept is closely related to Half-Value Layer (HVL); a material with a thickness of one HVL will attenuate 50% of photons. A standard x-ray image is a transmission image, a minus log of it is sometimes referred as number of mean free paths image. Mean free path in particle physics In particle physics the concept of the mean free path is not commonly used, being replaced by the similar concept of attenuation length. In particular, for high-energy photons, which mostly interact by electron-positron pair production, the radiation length is used much like the mean free path in radiography. Mean free path in nuclear physics Independent particle models in nuclear physics require the undisturbed orbiting of nucleons within the nucleus before they interact with other nucleons. Blatt and Weisskopf, in their 1952 textbook "Theoretical Nuclear Physics" (p. 778) wrote "The effective mean free path of a nucleon in nuclear matter must be somewhat larger than the nuclear dimensions in order to allow the use of the independent particle model. This requirement seems to be in contradiction to the assumptions made in the theory. We are facing here one of the fundamental problems of nuclear structure physics which has yet to be solved." (quoted by Norman D. Cook in "Models of the Atomic Nucleus" Ed.2 (2010) Springer, in Chapter 5 "The Mean Free Path of Nucleons in Nuclei"). Mean free path in optics If one takes a suspension of non light absorbing particles of diameter d with a volume fraction , the mean free path [9] of the photons is:

where Qs is the scattering efficiency factor. Qs can be evaluated numerically for spherical particles thanks to the Mie theory calculation

DOS IN MATERIALS SCIENCE, M.Tech

31

EQUATIONS OF STATES IN THERMODYNAMICS

Mean free path in acoustics In an otherwise empty cavity, the mean free path of a single particle bouncing off the walls is:

where V is volume of the cavity and S is total inside surface area of cavity. This relation is used in the derivation of the Sabine equation in acoustics, using a geometrical approximation of sound propagation.[10] Examples A classic application of the mean free path is to estimate the size of atoms or molecules. Another important application is in estimating the resistivity of a material from the mean free path of its electrons. For example, for sound waves in an enclosure, the mean free path is the average distance the wave travels between reflections off the enclosure's walls. In aerodynamics, the mean free path is in the same order of magnitude as the shockwave thickness at mach numbers greater than one.

Van der Waals equation The van der Waals equation is an equation of state for a fluid composed of particles that have a non-zero volume and a pairwise attractive inter-particle force (such as the van der Waals force). It was derived in 1873 by Johannes Diderik van der Waals, who received the Nobel prize in 1910 for "his work on the equation of state for gases and liquids". The equation is based on a modification of the ideal gas law and approximates the behavior of real fluids, taking into account the nonzero size of molecules and the attraction between them. Equation

DOS IN MATERIALS SCIENCE, M.Tech

32

EQUATIONS OF STATES IN THERMODYNAMICS

The van der Waals isotherms: the model correctly predicts a mostly incompressible liquid phase, but the oscillations in the phase transition zone do not fit experimental data. The equation uses the following state variables: the pressure of the fluid p, total volume of the container containing the fluid V, number of moles n, and absolute temperature of the system T. One form of the equation is

where

is the volume of the container shared between each particle (not the velocity of a particle),

is the total number of particles, and

is Boltzmann's constant, given by the universal gas constant R and Avogadro's constant NA. Extra parameters are introduced: a is a measure for the attraction between the particles, and b is the average volume excluded from v by a particle. The equation can be cast into the better known form

where

DOS IN MATERIALS SCIENCE, M.Tech

33

EQUATIONS OF STATES IN THERMODYNAMICS

is a measure of the attraction between the particles,

is the volume excluded by a mole of particles. A careful distinction must be drawn between the volume available to a particle and the volume of a particle. In particular, in the first equation v refers to the empty space available per particle. That is, v is the volume V of the container divided by the total number nNA of particles. The parameter b, on the other hand, is proportional to the proper volume of a single particle - the volume bounded by the atomic radius. This is the volume to be subtracted from v because of the space taken up by one particle. In Van der Waals' original derivation, given below, b is four times the proper volume of the particle. Observe further that the pressure p goes to infinity when the container is completely filled with particles so that there is no void space left for the particles to move. This occurs when V = nb. Validity Above the critical temperature the van der Waals equation is an improvement of the ideal gas law, and for lower temperatures the equation is also qualitatively reasonable for the liquid state and the low-pressure gaseous state. However, the Van der Waals model is not appropriate for rigorous quantitative calculations, remaining useful only for teaching and qualitative purposes.[1] In the first-order phase transition the range of (P, V, T), where the liquid phase and the gas phase are in equilibrium, it does not exhibit the empirical fact that p is constant as a function of V for a given temperature: although this behavior can be easily inserted into the van der Waals model (see Maxwell's correction below), the result is no longer a simple analytical model, and others (such as those based on the principle of corresponding states) achieve a better fit with roughly the same work. Derivation Most textbooks give two different derivations. One is the conventional derivation that goes back to van der Waals and the other is a statistical mechanics derivation. The latter has the major advantage that it makes explicit the intermolecular potential, which is neglected in the first derivation. The conventional Van der Waals equation is a mechanical equation of state, which cannot be used to specify all thermodynamic functions, while the statistical mechanical
DOS IN MATERIALS SCIENCE, M.Tech 34

EQUATIONS OF STATES IN THERMODYNAMICS

derivation yields the partition function for the system, which does allow all thermodynamic functions to be specified, including the mechanical equation of state. Conventional derivation Consider first one mole of gas which is composed of non-interacting point particles that satisfy the ideal gas law

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. 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 that volume. To see this, we must realize that a particle is surrounded by a sphere of radius r = 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 per particle (of average diameter d or radius r) is , which was divided by 2 to prevent overcounting. 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. Next, we introduce a pairwise attractive force between the particles. Van der Waal assumed that, notwithstanding the existence of this force, the density of the fluid is homogeneous. Further 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. That is, the bulk of them have more
DOS IN MATERIALS SCIENCE, M.Tech 35

EQUATIONS OF STATES IN THERMODYNAMICS

attracting particles to their right than to their left when they are relatively close to the lefthand wall of the container. The same statement holds with left and right interchanged. 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 . 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.

DOS IN MATERIALS SCIENCE, M.Tech

36

EQUATIONS OF STATES IN THERMODYNAMICS

Statistical thermodynamics derivation The canonical partition function Q of an ideal gas consisting of N = nNA identical particles, is

where

is the thermal de Broglie wavelength,

with the usual definitions: h is Planck's constant, m the mass of a particle, k Boltzmann's constant and T the absolute temperature. In an ideal gas q 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 independent, the total partition function still factorizes, , but the intermolecular potential necessitates two modifications to q. 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) .

DOS IN MATERIALS SCIENCE, M.Tech

37

EQUATIONS OF STATES IN THERMODYNAMICS

Second, we insert a Boltzmann factor exp[ - /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 4 r2dr 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 V. We get

DOS IN MATERIALS SCIENCE, M.Tech

38

EQUATIONS OF STATES IN THERMODYNAMICS

Other thermodynamic parameters We reiterate that the extensive volume V is related to the volume per particle v=V/N where N = nNA is the number of particles in the system. The equation of state does not give us all the thermodynamic parameters of the system. We can take the equation for the Helmholtz energy A

From the equation derived above for lnQ, we find

Where is an undetermined constant, which may be taken from the Sackur-Tetrode equation for an ideal gas to be:

This equation expresses A in terms of its natural variables V and T , and therefore gives us all thermodynamic information about the system. The mechanical equation of state was already derived above

The entropy equation of state yields the entropy (S )

from which we can calculate the internal energy

DOS IN MATERIALS SCIENCE, M.Tech

39

EQUATIONS OF STATES IN THERMODYNAMICS

Similar equations can be written for the other thermodynamic potential and the chemical potential, but expressing any potential as a function of pressure p will require the solution of a third-order polynomial, which yields a complicated expression. Therefore, expressing the enthalpy and the Gibbs energy as functions of their natural variables will be complicated. Reduced form Although the material constants a and b in the usual form of the van der Waals equation differs for every single fluid considered, the equation can be recast into an invariant form applicable to all fluids. Defining the following reduced variables (fR, fC is the reduced and critical variables version of f, respectively),

, where

as shown by Salzman. The first form of the van der Waals equation of state given above can be recast in the following reduced form:

This equation is invariant for all fluids; that is, the same reduced form equation of state applies, no matter what a and b may be for the particular fluid. This invariance may also be understood in terms of the principle of corresponding states. If two fluids have the same reduced pressure, reduced volume, and reduced temperature, we say that their states are corresponding. The states of two fluids may be corresponding even if their measured pressure, volume, and temperature are very different. If the two fluids' states are corresponding, they exist in the same regime of the reduced form equation of state. Therefore, they will respond to changes in roughly the same way, even though their measurable physical characteristics may differ significantly.
DOS IN MATERIALS SCIENCE, M.Tech 40

EQUATIONS OF STATES IN THERMODYNAMICS

Cubic equation The van der Waals equation is a cubic equation of state. That is we can write the equation into a cubic form of the volume. In the reduced formulation the cubic equation is:

At the critical temperature, where

we get as expected

For TR < 1, there are 3 values for vR. For TR > 1, there is 1 real value for vR. 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. Maxwell equal area rule Below the critical temperature TR < 1 an isotherm of the Van der Waals equation oscillates as shown.

DOS IN MATERIALS SCIENCE, M.Tech

41

EQUATIONS OF STATES IN THERMODYNAMICS

Maxwell's rule eliminates the oscillating behavior of the isotherm in the phase transition zone by defining it as a certain isobar in that zone. The above isotherm is for a reduced temperature of . The Maxwell correction is at a vapor pressure of between the reduced volume of the pure liquid pure gas at the vapor pressure. which is unstable; the Van der Waals and the

Along the red portion of the isotherm

equation fails to describe real substances in this region because the equation always assumes that the fluid is uniform while between a and c on the isotherm it becomes more stable to be a coexistence of two different phases, a denser phase which we normally call liquid and a sparser phase which we normally call gas. To fix this problem James Clerk Maxwell (1875) replaced the isotherm between a and c with a horizontal line positioned so that the areas of the two hatched regions are equal. The flat line portion of the isotherm now corresponds to liquid-vapor equilibrium. The portions ad and ce are interpreted as metastable states of super-heated liquid and super-cooled vapor respectively.[3] The equal area rule can be expressed as:

where PV is the vapor pressure (flat portion of the curve), VL is the volume of the pure liquid phase at point a on the diagram, and VG is the volume of the pure gas phase at point c on the diagram. The sum of these two volumes will equal the total volume V. Maxwell justified the rule by saying that work done on the system in going from c to b should equal work released on going from a to b. (The area on a PV diagram corresponds to mechanical work). Thats because the change in the free energy function A(T,V) equals the work done during a reversible process, and the free energy function - being a state variable should take on a unique value regardless of path. In particular, the value of A at point b should calculate the same regardless of whether the path came from left or right, or went straight across the horizontal isotherm or around the original Van der Waals isotherm. Maxwells argument is not totally convincing since it requires a reversible path through a region of thermodynamic instability. Nevertheless, more subtle arguments based on modern theories of phase equilibrium seem to confirm the Maxwell Equal Area construction and it remains a valid modification of the Van der Waals equation of state.

DOS IN MATERIALS SCIENCE, M.Tech

42

EQUATIONS OF STATES IN THERMODYNAMICS

The Maxwell equal area rule can also be derived from an assumption of equal chemical potential of coexisting liquid and vapour phases. On the isotherm shown in the above plot, points a and c are the only pair of points which fulfill the equilibrium condition of having equal pressure, temperature and chemical potential. It follows that systems with volumes intermediate between these two points will consist of a mixture of the pure liquid and gas with specific volumes equal to the pure liquid and gas phases at points a and c. The Van der Waals equation may be solved for VG and VL as functions of the temperature and the vapor pressure PV. Since:

where A is the Helmholtz free energy, it follows that the equal area rule can be expressed as:

Since the gas and liquid volumes are functions of PV and T only, this equation is then solved numerically to obtain PV as a function of temperature (and number of particles N), which may then be used to determine the gas and liquid volumes.

Theorem of corresponding states According to van der Waals, the theorem of corresponding states (or principle of corresponding states) indicates that all fluids, when compared at the same reduced temperature and reduced pressure, have approximately the same compressibility factor and all deviate from ideal gas behavior to about the same degree. Material constants that vary for each type of material are eliminated, in a recast reduced form of a constitutive equation. The reduced variables are defined in terms of critical variables. It originated with the work of Johannes Diderik van der Waals in about 1873 when he used the critical temperature and critical pressure to characterize a fluid. The most prominent example is the van der Waals equation of state, the reduced form of which applies to all fluids.

DOS IN MATERIALS SCIENCE, M.Tech

43

EQUATIONS OF STATES IN THERMODYNAMICS

Compressibility factor at the critical point

The compressibility factor at the critical point, which is defined as the subscript

, where

indicates the critical point is predicted to be a constant independent of

substance by many equations of state; the Van der Waals equation e.g. predicts a value of . Substance H2O
4

Value 0.23[4] 0.31[4] 0.30[5] 0.30[5] 0.29[5] 0.29[5] 0.29[5]

He

He H2 Ne N2 Ar

DOS IN MATERIALS SCIENCE, M.Tech

44

EQUATIONS OF STATES IN THERMODYNAMICS

BIBILOGRAPHY: Thermodynamics by Khurmi Application of Thermodynamics by R K Guptha Basics of Thermodynamics by B A Srinivas Google Wikipedia Encyclopaedia

DOS IN MATERIALS SCIENCE, M.Tech

45

You might also like