CHEM-UA 652: Thermodynamics and Kinetics

Similar documents
Gases and the Virial Expansion

CHEM-UA 652: Thermodynamics and Kinetics

Thus, the volume element remains the same as required. With this transformation, the amiltonian becomes = p i m i + U(r 1 ; :::; r N ) = and the canon

2. Thermodynamics. Introduction. Understanding Molecular Simulation

G : Statistical Mechanics

Brief Review of Statistical Mechanics

V.E Mean Field Theory of Condensation

Midterm Examination April 2, 2013

Physics 112 The Classical Ideal Gas

Introduction Statistical Thermodynamics. Monday, January 6, 14

The Second Virial Coefficient & van der Waals Equation

HWK#7solution sets Chem 544, fall 2014/12/4

Supplemental Material for Temperature-sensitive colloidal phase behavior induced by critical Casimir forces

Quiz 3 for Physics 176: Answers. Professor Greenside

Contents. 1 Introduction and guide for this text 1. 2 Equilibrium and entropy 6. 3 Energy and how the microscopic world works 21

Physics 127b: Statistical Mechanics. Lecture 2: Dense Gas and the Liquid State. Mayer Cluster Expansion

Chem 4501 Introduction to Thermodynamics, 3 Credits Kinetics, and Statistical Mechanics

Chapter 2 Experimental sources of intermolecular potentials

V.C The Second Virial Coefficient & van der Waals Equation

Removing the mystery of entropy and thermodynamics. Part 3

Chapter 19 The First Law of Thermodynamics

i=1 n i, the canonical probabilities of the micro-states [ βǫ i=1 e βǫn 1 n 1 =0 +Nk B T Nǫ 1 + e ǫ/(k BT), (IV.75) E = F + TS =

A Brief Introduction to Statistical Mechanics

5.62 Physical Chemistry II Spring 2008

1 Particles in a room

Introduction. Statistical physics: microscopic foundation of thermodynamics degrees of freedom 2 3 state variables!

Chemistry 593: The Semi-Classical Limit David Ronis McGill University

although Boltzmann used W instead of Ω for the number of available states.

Intermolecular Potentials and The Second Virial Coefficient

(i) T, p, N Gibbs free energy G (ii) T, p, µ no thermodynamic potential, since T, p, µ are not independent of each other (iii) S, p, N Enthalpy H

Statistical Thermodynamics and Monte-Carlo Evgenii B. Rudnyi and Jan G. Korvink IMTEK Albert Ludwig University Freiburg, Germany

Derivation of Van der Waal s equation of state in microcanonical ensemble formulation

fiziks Institute for NET/JRF, GATE, IIT-JAM, JEST, TIFR and GRE in PHYSICAL SCIENCES

Monte Carlo Methods. Ensembles (Chapter 5) Biased Sampling (Chapter 14) Practical Aspects

Advanced Topics in Equilibrium Statistical Mechanics

An Extended van der Waals Equation of State Based on Molecular Dynamics Simulation

X α = E x α = E. Ω Y (E,x)

Nanoscale simulation lectures Statistical Mechanics

[S R (U 0 ɛ 1 ) S R (U 0 ɛ 2 ]. (0.1) k B

Ideal Gas Behavior. NC State University

CHAPTER 4. Cluster expansions

Chapter 5 - Systems under pressure 62

Introduction to Thermodynamic States Gases

8.044 Lecture Notes Chapter 8: Chemical Potential

SYDE 112, LECTURE 7: Integration by Parts

Mechanics, Heat, Oscillations and Waves Prof. V. Balakrishnan Department of Physics Indian Institute of Technology, Madras

ChE 524 A. Z. Panagiotopoulos 1

Imperfect Gases. NC State University

Computer simulation methods (1) Dr. Vania Calandrini

Understanding Molecular Simulation 2009 Monte Carlo and Molecular Dynamics in different ensembles. Srikanth Sastry

Tutorial on obtaining Taylor Series Approximations without differentiation

PHYS 352 Homework 2 Solutions

Lecture 7: Kinetic Theory of Gases, Part 2. ! = mn v x

1. Thermodynamics 1.1. A macroscopic view of matter

(# = %(& )(* +,(- Closed system, well-defined energy (or e.g. E± E/2): Microcanonical ensemble

where (E) is the partition function of the uniform ensemble. Recalling that we have (E) = E (E) (E) i = ij x (E) j E = ij ln (E) E = k ij ~ S E = kt i

5.9 Representations of Functions as a Power Series

Part II: Statistical Physics

d 1 µ 2 Θ = 0. (4.1) consider first the case of m = 0 where there is no azimuthal dependence on the angle φ.

MATH 250 TOPIC 13 INTEGRATION. 13B. Constant, Sum, and Difference Rules

Physics is time symmetric Nature is not

Lecture 14. Entropy relationship to heat

The non-interacting Bose gas

Thermodynamics 1 Lecture Note 2

Physics 505 Homework No. 12 Solutions S12-1

4.1 Constant (T, V, n) Experiments: The Helmholtz Free Energy

As in the book. Each of the other thermopotentials yield similar and useful relations. Let s just do the Legendre Transformations and list them all:

CE 530 Molecular Simulation

Grand Canonical Formalism

2m + U( q i), (IV.26) i=1

The Partition Function Statistical Thermodynamics. NC State University

Advanced Thermodynamics. Jussi Eloranta (Updated: January 22, 2018)

A. Temperature: What is temperature? Distinguish it from heat.

Experimental Soft Matter (M. Durand, G. Foffi)

1 Multiplicity of the ideal gas

More Thermodynamics. Specific Specific Heats of a Gas Equipartition of Energy Reversible and Irreversible Processes

Thermodynamics part III.

Quantum ideal gases: bosons

ChE 210B: Advanced Topics in Equilibrium Statistical Mechanics

MATHS 267 Answers to Stokes Practice Dr. Jones

Addison Ault, Department of Chemistry, Cornell College, Mount Vernon, IA. There are at least two ways to think about statistical thermodynamics.

e x3 dx dy. 0 y x 2, 0 x 1.

INDIAN INSTITUTE OF TECHNOLOGY GUWAHATI Department of Physics MID SEMESTER EXAMINATION Statistical Mechanics: PH704 Solution

Chapter 18 Thermal Properties of Matter

Solutions to Sample Questions for Final Exam

MA1131 Lecture 15 (2 & 3/12/2010) 77. dx dx v + udv dx. (uv) = v du dx dx + dx dx dx

Mathematics 205 Solutions for HWK 23. e x2 +y 2 dxdy

Bohr & Wheeler Fission Theory Calculation 4 March 2009

So far in talking about thermodynamics, we ve mostly limited ourselves to

Spontaneity: Second law of thermodynamics CH102 General Chemistry, Spring 2012, Boston University

Statistics 100A Homework 5 Solutions

Supporting Information

Critical Properties of Isobaric Processes of Lennard-Jones Gases

Molecular Interactions F14NMI. Lecture 4: worked answers to practice questions

Calculus II. Calculus II tends to be a very difficult course for many students. There are many reasons for this.

Part II: Statistical Physics

Lecture 6: Ideal gas ensembles

Monte Carlo. Lecture 15 4/9/18. Harvard SEAS AP 275 Atomistic Modeling of Materials Boris Kozinsky

Infinite series, improper integrals, and Taylor series

Physics 127b: Statistical Mechanics. Renormalization Group: 1d Ising Model. Perturbation expansion

Transcription:

1 CHEM-UA 652: Thermodynamics and Kinetics Notes for Lecture 4 I. THE ISOTHERMAL-ISOBARIC ENSEMBLE The isothermal-isobaric ensemble is the closest mimic to the conditions under which most experiments are performed, namely under conditions of constant N, P, and T. In order to fix the pressure and temperature, the system must allowed to exchange heat with its surrounds, and the surrounds must also act as a kind of isotropic piston on the system, allowing the system s volume V to fluctuate in response to an applied pressure P. If the volume fluctuates, then we can treat is as an additional mechanical variable in the system. If this is the case, then the energy E can change in response to the exchange of heat, while there is an additional energy term P V, which when added to the energy change E, gives the total energy change in response to both the piston and the heat exchange. This sum E +P V goes by the name enthalpy and is denoted H: H = E +P V 1) The mechanical version of this simply replaced E by the mechanical energy Ex), and we would integrate a Boltzmann distribution of Ex) + P V over all coordinates, momenta, and volumes in order to produce a partition function: N,P,T) = C N dv = = dv e βpv C N dxe βex)+pv) dx e βex) dv e βpv QN,V,T) 2) The last line of this shows that there is a very simple relation between the canonical partition function and that of the isothermal-isobaric ensemble. Once we know the canonical partition function, we just need to do one volume integral in orer to determine the partition function of the isothermal-isobaric ensemble. The thermodynamic relations are also completely analogous to those of the canonical ensemble. Thus, the average enthalpy H is given by and the constant-pressure heat capacity C P becomes H = ln N,P,T) 3) β C P = k B β 2 2 ln N,P,T) 4) β2 The only real difference from the canonical ensemble is that we determine an average volume V from the pressure derivative of N,P,T): V = k B T ln N,P,T) 5) P II. TREATING INTERACTIONS: VIRIAL COEFFICIENTS We have seen that if we can determine a canonical partition function, then we can also easily compute the properties of a system at constant pressure by simply applying Eq. 2). But how can we compute a canonical partition function for a system in which there are real interactions present. One way is to use computer simulation techniques. If we run a molecular dynamics or Monte Carlo calculation, then, in addition to approximating the actual microscopic motion of a system, we are, in effect, using the simulation to do

2 the many-dimensional integrals over coordinates and momenta! Of course, the downside of this is that computer simulations can be time consuming, and they only allow us to examine one thermodynamic state point at a time. Each new state point is a new simulation. In addition, the computer will spit out reams of data, and we have, somehow, to make sense of it all. As the Hungarian-American physicist and Nobel laureate, Eugene Wigner once said, It is nice to know that the computer understands the problem, but I would like to understand it too. For this reason, analytical techniques can play an extremely valuable role, as they provide insights that cannot be easily gleaned from computer simulations. Let us start, therefore, by considering a gas at low density and compute the canonical partition function for this system. We will assume that the density is sufficiently low that each particle interacts with just one other particle at a time which would certainly not the case in a liquid or solid). Because there interactions are pairwise, we can assume that the potential energy is a sum of pair-wise terms: Ur 1,...,r N ) = N 1 N i=1 j=i+1 u r i r j ) 6) A typical plot of such a potential and the force it produces between two particles is shown in Fig. 1. However, if each ur 12 ) F 12 ur 12 ), F 12 r 12 FIG. 1: Typical potential energy and force curves for two particles in a real gas. particle interacts with just one other particle at a time, then we can simplify the complicated sums above and write the potential as Ur 1,...,r N ) = u r 1 r 2 )+u r 3 r 4 )+ u r N 1 r N ) 7)

providedthatwerecognizethis asbeingjustonepossiblewaywecanpairup theparticles. Ifweaccountforallpossible ways we can do this as a combinatorial factor in the counting of microstates, then we have all of the information we need to construct the partition function. Solet sseehowwecandothis. Consider, first, thecaseofjust fourparticles. Inhowmanywayscanwepairthemupso that each particle interacts with just one other? There are, in fact, three possibilities. We have, for example, 1,2)3,4), but we also have 1,3)2,4), and 1,4)2,3). Note that the case 1,2)3,4) corresponds to what we have written in Eq. 6). That is, for this pairing, we would write the potential as Ur 1,r 2,r 3,r 4 ) = u r 1 r 2 )+u r 3 r 4 ). The three possibilities can be expressed as 3 1. Now suppose we had six particles. In how many ways can we pair these up. If you enumerate them all, you ll find that there are 15 possibilities: 1,2)3,4)5,6) 1,2)3,5)4,6) 1,2)3,6)4,5) 3 1,3)2,4)5,6) 1,3)2,5)4,6) 1,3)2,6)4,5) 1,4)2,3)4,6) 1,4)2,5)3,6) 1,4)2,6)3,5) 1,5)2,3)4,6) 1,5)2,4)3,6) 1,5)2,6)3,4) 1,6)2,3)4,5) 1,6)2,4)3,5) 1,6)2,5)3,4) Note that 15 = 5 3 1. Thus, in teneral, if we have N particles, where N must be even so that we can pair them up, then the number of ways we can pair them up is N 1)N 3)N 5) = N 1)!!. Recall that the canonical partition function is QN,V,T) = C N dx p e β N i=1 p2 i /2m dx r e βur1,...,rn) 8) Let us denote by ZN,V,T) ZN,V,T) = dr 1 dr N e βur1,...,rn) 9) We will be interested in computing the equation of state for our system, which we obtain by computing the pressure from ) lnq P = k B T 1) V However, ZN,V,T) contains all of the volume dependence of Q, so we really only need to worry about it, and the pressure would then be given by ) lnz P = k B T 11) V

In the limit that we can consider each particle as interacting with just one other particle at a time, for which there are N 1)!! possible combinations, we can write Z as ZN,V,T) = N 1)!! dr 1 dr N e βu r1 r2 ) βu rn 1 rn ) e βu r3 r4 ) e = N 1)!! dr 1 dr 2 e βu r1 r2 ) dr 3 dr 4 e βu r3 r4 ) 4 dr N 1 dr N e βu rn 1 rn ) 12) All of the integrals in the brackets are exactly the same. More over, for N particles, there are N/2 pairs we can make, so we can write the above expression as ZN,V,T) = N 1)!! dr 1 dr 2 e βu r1 r2 ) N/2 13) Next, let s look at the prefactor of N 1)!!. Recall that N! = NN 1)N 2) 1. This is a product of N terms, and if N is N! N N e N, which is known as Sterling s approximation. Similarly for N 1)!! = N 1)N 3) 3 1, the leading term is N 1 N, and there are N/2 terms in the product, so the approximation for a double factorial when N is large becomes N 1)!! N N/2 e N/2. Thus, we can write Z as N/2 ZN,V,T) e N N/2 dr 1 dr 2 e βu r1 r2 ) 14) In fact, the e N/2 we can simply absorb into the C N in the definition of Q, as it is volume independent and will not affect anything our calculations. Thus, it is sufficient for us to consider simply ZN,V,T) N dr 1 dr 2 e βu r1 r2 ) N/2 15) In order to simplify the integral, let us change variables to R = 1 2 r 1 +r 2 ), r = r 1 r 2 16) for which dr 1 dr 2 = drdr. Substituting this in, we obtain ZN,V,T) N drdr e βu r ) N/2 17) Since the integrand does not depend on R, we can easily integrate over R: a a a dr = dx dy dz = a 3 = V 18) Then N/2 ZN,V,T) = V N N/2 dr e βu r ) 19) The remaining integral still has an implied volume dependence in the sense that a a a dr e βu r ) = dx dy dz e βu r ) In order to make this volume dependence easier to work with, we go into a set of scaled coordinates s = r a = r V 1/3 = V 1/3 r 2)

The components of s all lie in the interval,1: s x,1,s y,1,s z,1. Moreover, dr = dxdydz = a 3 ds x ds y ds z = Vds. Putting this into the partition function, we have ZN,V,T) = V NV N/2 N/2 ds e βu v1/3 s ) 5 = V N N ds e βu V 1/3 s ) N/2 21) This is now an expression we can differentiate with respect to V to obtain the pressure. Using Eq. 11), we find { P = k B T } N/2 lnv N +ln N ds e βu V 1/3 s ) V Let = k B T V = Nk BT V {N lnv + N2 ln N +k B T V { N 2 ln N IV) = ds e βu V 1/3 s ) } ds e βu V 1/3 s ) } ds e βu V 1/3 s ) k If we assume that the interaction is a small perturbation to ideal-gas behavior, then we can take IV) to be a small quantity. In fact, we will see below that because di/dv 1/V 2, IV) 1/V, and since 1/V 1/N, IV) 1/N. In this case, NIV) 1, and we can use the approximation ln1+x) x to write lnniv) = ln1+niv) 1) NIV) 1) 24) 22) 23) so that P = Nk BT V + Nk BT 2 V NIV) 1) = Nk BT V + N2 k B T 2 Let s now look at the volume derivative of IV). Applying the chain rule, we have V = β u ds V 1/3 s) s1 3 V 2/3 e βu V 1/3 s ) V 25) 26) Now that we ve taken the volume derivative, we can go back to unscaled coordinates via s = V 1/3 r. Remember ds = 1/V)dr. If we do this, then the integral looks much less ugly: V = β 3V 2 dr u r r e βu r ) 27) The term appearing in the integrand r u r comes up frequently in classical mechanics and is called the virial. Recognizing that u r ) is a function of r = r, which is just the length of the vector r, it follows that so that u r = du r dr r = du r dr r r u r = u r r r r = rdu dr 28) 29)

Thus, V = β 3V 2 Now transform to spherical polar coordinates in r, which gives V = β 3V 2 2π dφ π dr r du dr e βur) 3) dθ sinθ dr r 3du dr e βur) 31) Note that we have made the approximation that the r integral can be taken from to. Strictly speaking, this is wrong because this integral should be limited by the maximum r allowed by the containing volume. However, as Fig. 1 indicates, u r) = u/ r vanishes very quickly as r increases, so that the integrand in the above integral goes to long before we hit the boundary of the box. Given that, extending the integral out to just adds a lot of zero to the integral, which will not change the answer! Integratinf over the angles, we obtain V = 4πβ 3V 2 We now simplify this expression by the integral as which we can integrate by parts to give V = 4π 3V 2 V = 4πβ 3V 2 1 β r 3 e βur) dr r 3du dr e βur) 32) dr r 3 d dr e βur) 33) dr d )e dr r3 βur) Unfortunately, when we try to evaluate the first term, we find that it is at r =. Even if we admit that this is a consequence of our approximation that r can be integrated out to, it does not really matter because this term would simply increase with increasing system size, i.e., increasing volume, which is unphysical, as the pressure cannot behave this way pressure is an intensive quantity. Fortunately, there is a trick that can save us. Recognizing that ur) as r, exp βur)) 1 as. Hence, let us back up and just subtract this one from the exponential where we introduce the derivative. That is, if we write V = 4πβ 3V 2 1 β dr r 3 d dr e βur) = 4πβ 3V 2 1 β dr r 3 d dr ) e βur) 1 we do not change anything since the derivative of 1 with respect to r is anyway. We can now integrate this by parts to give V = 4π 3V 2 r 3 ) ) e βur) 1 d ) dr dr r3 e βur) 1 36) Now the first term vanishes both at r = and at r =, and we are left with and the pressure becomes P = Nk BT V V = 4π V 2 N2 k B T V 2 dr r 2 ) e βur) 1 2π dr r 2 ) e βur) 1 Using the fact that R = N k B, and introducing the density ρ = n/v, we finally obtain P = ρrt +ρ 2 RT 2πN dr r 2 ) e βur) 1 P 2πN ρrt = 1+ρ dr r 2 e 1) βur) 39) 6 34) 35) 37) 38)

7 The term in brackets is denoted B 2 T): B 2 T) = 2πN dr r 2 ) e βur) 1 and is called a virial coefficient, as it arises from the classical mechanical virial mentioned above. 4) More realistic treatments of non-ideal gases generate an equation of state in the form of a power series in the density in which all of the coefficients are functions of temperature. This can be expressed as P ρrt = 1+ B k+1 T)ρ k 41) This equation of state is called the virial equation of state. The coefficients B k+1 T) are called virial coefficients. In the low density limit, the most important term in this power series is the k = 1 or linear term, which gives the approximate equation of state P ρrt 1+B 2T)ρ 42) where the dominant coefficient B 2 T) is called the second virial coefficient. These can be measured experimentally. The following plot shows the second virial coefficient for several gases. Although it is difficult to see, these curves pass FIG. 2: Experimental second virial coefficient plots for several real gases. through a shallow maximum that is not captured by the van der Waals equation. Thus, the van der Waals equation is only in qualitative agreement with these but is not fully quantitative. Let us look at some simple examples of B 2 T) calculations. 1. Hard-sphere potential: ur) = { r r > 43)

8 For this potential, the calculation of B 2 T) proceeds as follows: B 2 T) = 2πN e β 1 ) r dr+ 2 = 2πN 1)r 2 dr + = 2πN r 2 dr e β 1 ) r 2 dr 1 1)r 2 dr = 2 3 π3 N 44) which is also the paramter b in the van der Waals equation. In this case, there is no temperature dependence in the result for B 2 T). 2. Square-well potential plus hard-sphere potential: In this case, the potential takes the form { r ur) = ǫ r λ r > λ In this case, the hard-sphere potential is supplemented by a short-range constant attractive potential that has a square-well shape. The parameter λ > 1 determins the range of the square well. For this example, the calculation of B 2 T) proceeds as follows: B 2 T) = 2πN e β 1 ) λ r 2 dr + e βǫ 1 ) r dr+ 2 e β 1 ) r 2 dr = 2πN = 2πN 3 r 2 dr+ e βǫ 1 ) λ r 2 dr 3 + e βǫ 1 ) r 3 3 = 2πN 3 3 + e βǫ 1 ) 1 λ 3 3 3) 3 = 2πN 3 3 λ λ 45) 1 e βǫ 1 ) λ 3 1 ) 46) A plot of this second virial coefficient is given in the figure below. 3. Hard-sphere plus C 6 /r 6 potential: For this example, the potential is { r ur) = r > C6 r 6 This is a particular realization of a potential that leads to the van der Waals equation of state. The C 6 /r 6 attractive part of the potential comes directly from a full quantum mechanical treatment of the electron distribution around each atom in an interacting pair. Such an attractive potential attempts to model London dispersion or van der Waals forces. For this example, the calculation of B 2 T) proceeds as follows: B 2 T) = 2πN e β 1 ) r dr+ 2 = 2πN 3 3 + e βc6/r6 1 ) ) e βc6/r6 1 r 2 dr r 2 dr 47) 48)

9 2π 3 N /3 FIG. 3: Second virial coefficient for the hard-sphere plus square-well potential. he remaining integral cannot be evaluated in closed form, but let us suppose that βc 6 /r 6 << 1 for all values of r considered, i.e., r <. Then, we can expand the exponential using the power series e x 1+x+x 2 /2!+. If we retain just the first two terms, then we have the approximation e βc6/r6 1 1+ βc 6 r 6 1 = βc 6 r 6 49) With this approximation, the calculation of B 2 T) can proceed as follows: B 2 T) 2πN 3 3 +βc 1 6 r 6r2 dr = 2πN 3 3 +βc 6 = 2πN 3 3 +βc 6 = 2πN 3 3 + βc 6 3 3 = 2πN 3 3 1 C 6 3k B T 6 1 1 r 4dr ) 3r 3 A plot of this function will follow the curve presented in the Figure below. 5)

1 b B 2 T) T FIG. 4: Second virial coefficient for the hard-sphere plus C 6/r 6 potential. However, in this example, we can improve the approximation by retaining more terms of the exponential we expanded above. In fact, let us see what happens if we retain all of the terms in the power series. That is, we will use the fact that e x can be represented exactly as e x = k= x k k! = 1+ x k k! 51) so that e βc6/r6 1 = = 1+ 1 k! βc6 r 6 β k C k 6 k! ) k 1 r 6k 52) With the expansion, the calculation of B 2 T) proceeds as follows: B 2 T) = 2πN 3 3 + β k C6 k k! 1 r 6kr2 dr

= 2πN = 2πN = 2πN 3 3 3 3 + 3 3 + 1 3 β k C k 6 k! β k C k 6 k! 1 r 6k 2dr 1 6k 1) 6k 1 βc 6 ) k 1 k! 6k 1) 6k+2 The final result represents and exact expression for B 2 T) for this particular model of ur). How does it improve on the result obtained above retaining just two terms in the expansion of the exponential? Let us plot the final result as a function of temperature in the following way: Unfortunately, we cannot resum the final expression into a nice closed form, but we can evaluate the sum numerically on a computer. However, on a computer, we cannot sum an infinite number of terms. Rather, we replace the upper limit of with a maximum number of terms K max and then just increase K max until the result converges. So, let s do that for several values of K max and see how the curve changes as we increase K max. This will show us how the result converges with K max. The result is shown in the figure below. We see from the figure that the sum converges very rapidly with K max 11 53) 2πN 3 /3 B 2 T) K max = 2 K max = 3 K max = 4 K max = 5 K max = 6 T FIG. 5: Second virial coefficient for the potential in Eq.47. and only requres a few terms. Not unexpectedly, the most significant deviations from K max = 2 occur in the low-temperature part of the curves.

4. Perhaps the most realistic representation of the potential ur) shown in Fig. 2 of Lecture 2 is the so-called Lennard-Jones potential, for which 12 ) 6 ur) = 4ǫ 54) r) r A plot of this potential very closely resembles that shown in Fig. 2 of Lecture 2. In particular, there is a well-defined minimum, which we can compute by taking the derivative of ur), setting it to, and solving for r at that point: 12 U 12 r) = 4ǫ r 13 66 r 7 = 12 12 12 r 13 = 66 r 7 2 6 r 13 = 1 r 7 2 6 = r 6 r = 2 1/6 r m 55) where r m denotes the location of the minimum. The value of the Lennard-Jones potential at that point is ) 12 ) 6 ur m ) = u2 1/6 ) = 4ǫ 2 1/6 2 1/6 1 = 4ǫ 2 2 1 2 = 4ǫ 1 4 = ǫ 56) Thus, the parameter ǫ measures the depth of the minimum at r = r m. In addition, u) =, and for r <, the potential exhibits a very steep increase, very much like the hard-sphere potential. Hence, is a measure of the van der Waals radius of each particle. The calculation of the second virial coefficient can now be set up as { 12 B 2 T) = 2πN exp 4βǫ r) r Let us make a change of variables from r to ) )} 6 1 r 2 dr 57) x = r so that dr = dx. Let us also define β = βǫ. With this change, the integral for B 2 T) becomes B 2 T) = 2π 3 N e 4β x 12 x 6 ) 1 x 2 dx 58) Again, this integral cannot be evaluated analytically in closed form. We could expand the exponential in an infinite series as we did in the previous example. However, if we were to do this, the evaluation of the various terms in the integral would be considerably more complicated than in the previous example, as we would need to evaluated increasingly higher powers of a binomial form x 12 x 6 ) k, Instead, since this is just a one-dimensional integral, we can evaluate it numerically using a method such as Simpson s rule. Recall that

Simpson s rule for the integral of a function fx) using a set of n evenly spaced values for x x, x 1,...,x n ) is just b a fx)dx = h 3 fx )+4fx 1 )+2fx 2 )+4fx 3 )+2fx 4 )+ 2fx n 2 )+4fx n 1 )+fx n ) 59) where h = b a)/n, and the points x,...,x n are given by x j = a+jh, j =,...,n. 13 Applying this to the integral in Eq. 58), the resulting curve for B 2 T) as a function of T appears as in the figure below: In this figure, k B T = 1/β, and FIG. 6: Second virial coefficient for the Lennard-Jones potential evaluated using numerical integration. Experimental points are also plotted. Because of the use of reduced units, data for all gases land on a single master curve, as an example of the law of corresponding states. B 2T) = B 2T) 2π 3 N /3 Unfortunately, the curve in the book does not carry the curve out to high enough temperature to see the maximum. However, on page 233 of the book Statistical Mechanics by Donald A. McQuarrie, a more complete curve for the Lennard-Jones potential is shown, and in this curve the temperature axis is sufficiently long to see the maximum. This book is available on Google Books.