09.07.2015 Views

lecture notes on statistical mechanics - Scott Pratt - Michigan State ...

lecture notes on statistical mechanics - Scott Pratt - Michigan State ...

lecture notes on statistical mechanics - Scott Pratt - Michigan State ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

LECTURE NOTES ON STATISTICAL MECHANICS<strong>Scott</strong> <strong>Pratt</strong>Department of Physics and Astr<strong>on</strong>omy<strong>Michigan</strong> <strong>State</strong> UniversityPHY 831 – 2007-2013


4.7 The Boltzmann Equati<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 654.8 Phase Space Density and Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 694.9 Hubble Expansi<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 704.10 Evaporati<strong>on</strong>, Black-Body Emissi<strong>on</strong> and the Compound Nucleus . . . . . . . . . . . 734.11 Diffusi<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 754.12 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 785 Lattices and Spins 835.1 Statistical Mechanics of Ph<strong>on</strong><strong>on</strong>s . . . . . . . . . . . . . . . . . . . . . . . . . . . . 835.2 Feromagnetism and the Ising Model in the Mean Field Approximati<strong>on</strong> . . . . . . . . 855.3 One-Dimensi<strong>on</strong>al Ising Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 885.4 Lattice Gas, Binary Alloys, Percolati<strong>on</strong>... . . . . . . . . . . . . . . . . . . . . . . . . 905.5 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 916 Landau Field Theory 946.1 What are fields? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 946.2 Calculating surface energies in field theory . . . . . . . . . . . . . . . . . . . . . . . 966.3 Correlati<strong>on</strong>s and Susceptibilities in the Critical Regi<strong>on</strong> . . . . . . . . . . . . . . . . 986.4 Critical Exp<strong>on</strong>ents in Landau Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 1016.5 Validity of Landau Theory Near T c : The Ginzburg Criteria . . . . . . . . . . . . . . 1026.6 Critical Phenomena, Scaling and Exp<strong>on</strong>ents . . . . . . . . . . . . . . . . . . . . . . 1036.7 Symmetry Breaking and Universality Classes . . . . . . . . . . . . . . . . . . . . . . 1056.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107


1 Fundamentals of Statistical Physics“I know nothing ... nothing” - John Banner1.1 Ignorance, Entropy and the Ergodic TheoremC<strong>on</strong>sider a large number of systems N s → ∞, each of which can be in some state specific quantumstate. Let n i be the number of systems that are in the state i. We will define the ignorance I as ameasure of the number of ways to arrange the systems given n 0 , n 1 · · · .I =N s !n 0 !n 1 ! · · ·, (1.1)with the c<strong>on</strong>straint that n 0 + n 1 + · · · = N s . Our immediate goal is to find n i that maximizesignorance while satisfying the c<strong>on</strong>straint. If the observer knows nothing about the populati<strong>on</strong>probabilities, the values n i should be chosen to maximize I. However, before doing so, we willdefine S as:S ≡ 1 ln(I), (1.2)N swhich will be maximized when I is maximized. By defining it as the log of the ignorance, S willhave some c<strong>on</strong>venient properties which we will see below. The quantity S is the entropy, the mostfundamental quantity of <strong>statistical</strong> <strong>mechanics</strong>. It is divided by the number of systems so that <strong>on</strong>ecan speak of the entropy in an individual system. Using Stirling’s expansi<strong>on</strong>,lim ln N! = N ln N − N + (1/2) ln N + (1/2) ln(2π) + 1/(12N) + · · · , (1.3)N→∞we keep the first two terms to see thatS = 1 N s(N s ln N s − ∑ in i ln n i − N s + ∑ in i + · · ·)(1.4)= − ∑ ip i ln p i as N s → ∞,where p i ≡ n i /N s is the probability a given system is in state i. As N s → ∞, all terms bey<strong>on</strong>d thep i ln p i term in Eq. (1.4) vanish. Note that if all the probability is c<strong>on</strong>fined to <strong>on</strong>e state, the entropywill be zero. Furthermore, since for each probability, 0 < p i ≤ 1, the entropy is always positive.Our goal is to maximize S. Maximizing a multi-dimensi<strong>on</strong>al functi<strong>on</strong> (in this case a functi<strong>on</strong> ofn 0 , n 1 · · · ) with a c<strong>on</strong>straint is often d<strong>on</strong>e with Lagrange multipliers. In that case, <strong>on</strong>e maximizesthe quantity, S − λC(⃗n), with respect to all variables and with respect to λ. Here, the c<strong>on</strong>straintC must be some functi<strong>on</strong> of the variables c<strong>on</strong>strained to zero, in our case C = ∑ i p i − 1. Thecoefficient λ is called the Lagrange multiplier. Stating the minimizati<strong>on</strong>,(∂− ∑ [ ])∑p j ln p j − λ p j − 1 = 0, (1.5)∂p ijj∂∂λ(− ∑ jp j ln p j − λ[ ∑jp j − 1])= 0.1


The sec<strong>on</strong>d expressi<strong>on</strong> leads directly to the c<strong>on</strong>straint ∑ j p j = 1, while the first expressi<strong>on</strong> leads tothe following value for p i ,ln p i = −λ − 1, or p i = e −λ−1 . (1.6)The parameter λ is then chosen to normalize the distributi<strong>on</strong>, e −λ−1 multiplied by the number ofstates is unity. The important result here is that all states are equally probable. This is the resultof stating that you know nothing about which states are populated, i.e., maximizing ignorance isequivalent to stating that all states are equally populated. This can be c<strong>on</strong>sidered as a fundamentalprinciple – Disorder (or entropy) is maximized. All <strong>statistical</strong> <strong>mechanics</strong> derives from thisprinciple.ASIDE: REVIEW OF LAGRANGE MULTIPLIERSImagine a functi<strong>on</strong> F (x 1 · · · x n ) which <strong>on</strong>e minimizes w.r.t. a c<strong>on</strong>straint C(x 1 · · · x n ) = 0. Thegradient of F projected al<strong>on</strong>g the hyper-surface of the c<strong>on</strong>straint must vanish, or equivalently, ∇Fmust be parallel to ∇C. The c<strong>on</strong>stant of proporti<strong>on</strong>ality is the Lagrange multiplier λ,The two gradients are parallel if,∇F = λ∇C.∇(F − λC) = 0.However, this c<strong>on</strong>diti<strong>on</strong> <strong>on</strong> its own merely enforces that C(x 1 · · · x n ) is equal to some c<strong>on</strong>stant, notnecessarily zero. If <strong>on</strong>e fixes λ to an arbitrary value, then solves for ⃗x by solving the parallel-gradientsc<strong>on</strong>straint, <strong>on</strong>e will find a soluti<strong>on</strong> to the minimizati<strong>on</strong> c<strong>on</strong>straint with C(⃗x) = some c<strong>on</strong>stant, butnot zero. Fixing C = 0 can be accomplished by additi<strong>on</strong>ally requiring the c<strong>on</strong>diti<strong>on</strong>,∂(F − λC) = 0.∂λThus, the n-dimensi<strong>on</strong>al minimizati<strong>on</strong> problem with a c<strong>on</strong>straint is translated into an (n + 1)-dimensi<strong>on</strong>al minimizati<strong>on</strong> problem with no c<strong>on</strong>straint, where λ plays the role as the extra dimensi<strong>on</strong>.1.2 The Ergodic TheoremThe principle of maximizing entropy is related to the Ergodic theorem, which provides the way tounderstand why all states are equally populated from the perspective of dynamics. The Ergodictheorem is built <strong>on</strong> the symmetry of time-reversal, i.e., the rate at which <strong>on</strong>e changes from state i tostate j is the same as the rate at which <strong>on</strong>e changes from state j to state i. If a state is particularlydifficult to enter, it is equivalently difficult to exit. Thus, a time average of a given system will cyclethrough all states and, if <strong>on</strong>e waits l<strong>on</strong>g enough, the system will spend equal amounts of net timein each state.This can also be understood by c<strong>on</strong>sidering an infinite number of systems where each state ispopulated with equal probability. For every transiti<strong>on</strong> from state i to state j, if the probabilityof transiti<strong>on</strong>ing from j to i is equal to the probability of transiti<strong>on</strong>ing from i to j, it is clear theprobability distributi<strong>on</strong> will stay uniform. Thus, a uniform probability distributi<strong>on</strong> is stable, if <strong>on</strong>eassumes time reversal.2


Satisfacti<strong>on</strong> of time reversal is sometimes rather subtle. As an example, c<strong>on</strong>sider two largeidentical rooms, a left room and a right room, separated by a door manned by a security guard. Ifthe rooms are populated by 1000 randomly oscillating patr<strong>on</strong>s, and if the security guard grants anddenies access with equal probability when going right-to-left vs. left-to-right, the populati<strong>on</strong> of thetwo rooms will, <strong>on</strong> average, be equal. However, if the security guard denies access to the left roomwhile granting exit of the left room, the populati<strong>on</strong> will ultimately skew towards the right room.This explicit violati<strong>on</strong> of the principle of maximized entropy derives from the fact that movingleft-to-right and right-to-left, i.e. the time reversed moti<strong>on</strong>s, are not treated equivalently.The same security guard could, in principle, police the traversal of gas molecules between twopartiti<strong>on</strong>s of a box. Such paradoxes were discussed by Maxwell, and the security guard is referredto as Maxwell’s dem<strong>on</strong>. As described by Maxwell,... if we c<strong>on</strong>ceive of a being whose faculties are so sharpened that he can follow every moleculein its course, such a being, whose attributes are as essentially finite as our own, would be ableto do what is impossible to us. For we have seen that molecules in a vessel full of air at uniformtemperature are moving with velocities by no means uniform, though the mean velocity of anygreat number of them, arbitrarily selected, is almost exactly uniform. Now let us suppose thatsuch a vessel is divided into two porti<strong>on</strong>s, A and B, by a divisi<strong>on</strong> in which there is a smallhole, and that a being, who can see the individual molecules, opens and closes this hole, so asto allow <strong>on</strong>ly the swifter molecules to pass from A to B, and <strong>on</strong>ly the slower molecules to passfrom B to A. He will thus, without expenditure of work, raise the temperature of B and lowerthat of A, in c<strong>on</strong>tradicti<strong>on</strong> to the sec<strong>on</strong>d law of thermodynamics.This apparent violati<strong>on</strong> of the sec<strong>on</strong>d law of thermodynamics was explained by Leó Szilárd in 1929,who showed that the dem<strong>on</strong> would have to expend energy to measure the speed of the molecules,and thus increase entropy somewhere, perhaps in his brain, thus ensuring that the entropy of theentire system (gas + dem<strong>on</strong>) increased. Check out http://en.wikipedia.org/wiki/Maxwell’s dem<strong>on</strong>.We have defined the entropy with logarithms in such a way that it is additive for two uncorrelatedsystems. For instance, we c<strong>on</strong>sider a set of N a systems of type a which can be arranged I a ways, anda sec<strong>on</strong>d independent set of N b systems of type b which can be arranged I b ways. The combinedsystems can be arranged I = I a I b number of ways, and the entropy of the combined systems isS = ln I = S a + S b .1.3 Statistical EnsemblesThe previous secti<strong>on</strong> discussed the manifestati<strong>on</strong>s of maximizing ignorance, or equivalently entropy,without regard to any c<strong>on</strong>straints aside from the normalizati<strong>on</strong> c<strong>on</strong>straint. In this secti<strong>on</strong>, wediscuss the effects of fixing energy and/or particle number. These other c<strong>on</strong>straints can be easilyincorporated by applying additi<strong>on</strong>al Lagrange multipliers. For instance, c<strong>on</strong>serving the averageenergy can be enforced by adding an extra Lagrange multiplier β related to fixing the averageenergy per system. Minimizing the entropy per system with respect to the probability p i for beingin state i,(∂− ∑ p j ln p j − λ[ ∑ p j − 1] − β[ ∑ )p j ϵ j −∂p Ē] = 0, (1.7)ijjjgivesp i = exp(−1 − λ − βϵ i ). (1.8)3


Thus, the states are populated proporti<strong>on</strong>al to the factor e −βϵ i, which is the Boltzmann distributi<strong>on</strong>,with β being identified as the inverse temperature. Again, the parameter λ is chosen to normalizethe probability. However, again the Lagrange multipliers for a given β <strong>on</strong>ly enforce the c<strong>on</strong>straintthat the average energy is some c<strong>on</strong>stant, not the particular energy <strong>on</strong>e might wish. Thus, <strong>on</strong>e mustadjust β to find the desired energy, a sometimes time-c<strong>on</strong>suming process.For any quantity which is c<strong>on</strong>served <strong>on</strong> the average, <strong>on</strong>e need <strong>on</strong>ly add a corresp<strong>on</strong>ding Lagrangemultiplier. For instance, a multiplier α could be used to restrict the average particle number orcharge. The probability for being in state i would then be:p i = exp(−1 − λ − βϵ i − αQ i ). (1.9)Typically, the chemical potential µ is used to reference the multiplier,α = −µ/T. (1.10)The charge Q i could refer to the bary<strong>on</strong> number, electric charge, or any other c<strong>on</strong>served quantity.It could be either positive or negative. If there are many c<strong>on</strong>served charges, Q can be replaced by⃗Q and µ can be replaced by ⃗µ.Rather than enforcing the last Lagrange multiplier c<strong>on</strong>straint, that derivatives w.r.t. the multiplierare zero, we are often happy with knowing the soluti<strong>on</strong> for a given temperature and chemicalpotential. Inverting the relati<strong>on</strong> to find values of T and µ that yield specific values of the energyand particle number is often difficult, and as will be seen later <strong>on</strong>, it is often the temperature andchemical potentials that are effectively c<strong>on</strong>strained in many physical situati<strong>on</strong>s.In most textbooks, the charge is replaced by a number N. This is fine if the number of particlesis c<strong>on</strong>served such as a gas of Arg<strong>on</strong> atoms. However, the more general case includes cases of rathercomplicated chemical reacti<strong>on</strong>s, or multiple c<strong>on</strong>served charges. For instance, both electric chargeand bary<strong>on</strong> number are c<strong>on</strong>served in the hadr<strong>on</strong>ic medium in the interior of a star. Positr<strong>on</strong>s andelectr<strong>on</strong>s clearly c<strong>on</strong>tribute to the charge with opposite signs. Without apology, these <str<strong>on</strong>g>notes</str<strong>on</strong>g> willswitch between using N or Q depending <strong>on</strong> the c<strong>on</strong>text. As l<strong>on</strong>g as a sum is c<strong>on</strong>served, there is nodifference in referring to it as a number or as a charge.In <strong>statistical</strong> <strong>mechanics</strong> <strong>on</strong>e first c<strong>on</strong>siders which quantities <strong>on</strong>e wishes to fix, and which quantities<strong>on</strong>e wishes to allow to vary (but with the c<strong>on</strong>straint that the average is some value). This choicedefines the ensemble. The three most comm<strong>on</strong> ensembles are the micro-can<strong>on</strong>ical, can<strong>on</strong>ical andgrand-can<strong>on</strong>ical. The ensembles differ by which quantities vary, as seen in Table 1. If a quantity isallowed to vary, then a Lagrange multiplier defines the average quantity for that ensemble.Ensemble Energy Chargesmicro-can<strong>on</strong>ical fixed fixedcan<strong>on</strong>ical varies fixedgrand can<strong>on</strong>ical varies variesTable 1:Ensembles vary by what quantity is fixed and what varies.4


1.4 Partiti<strong>on</strong> Functi<strong>on</strong>sSince probabilities are proporti<strong>on</strong>al to e −βϵ i−αQ i, the normalizati<strong>on</strong> requirement can be written as:p i = 1 Z e−βϵ i−αQ i, (1.11)Z = ∑ ie −βϵ i−αQ i,where Z is referred to as the partiti<strong>on</strong> functi<strong>on</strong>. Partiti<strong>on</strong> functi<strong>on</strong>s are c<strong>on</strong>venient for calculatingthe average energy or charge,⟨E⟩ =∑i ϵ ie −βϵ i−αQ iZ(1.12)= − ∂ ln Z,∂β= T 2 ∂ ln Z (fixed α = −µ/T ),⟨Q⟩ =∑∂Ti Q ie −βϵ i−αQ iZ= − ∂ ln Z,∂α= T ∂ ln Z (fixed T ).∂µThe partiti<strong>on</strong> functi<strong>on</strong> can also be related to the entropy,(1.13)S = − ∑ ip i ln p i = ∑ ip i (ln Z + βϵ i + αQ i ) (1.14)= ln Z + β⟨E⟩ + α⟨Q⟩,which is derived using the normalizati<strong>on</strong> c<strong>on</strong>diti<strong>on</strong>, ∑ p i = 1, al<strong>on</strong>g with Eq. (1.11).EXAMPLE:C<strong>on</strong>sider a 3-level system with energies −ϵ, 0 and ϵ. As a functi<strong>on</strong> of T find:(a) the partiti<strong>on</strong> functi<strong>on</strong>, (b) the average energy, (c) the entropyThe partiti<strong>on</strong> functi<strong>on</strong> is:Z = e ϵ/T + 1 + e −ϵ/T , (1.15)and the average energy is:⟨E⟩ =∑i ϵ ie −ϵ i/TAt T = 0, ⟨E⟩ = −ϵ, and at T = ∞, ⟨E⟩ = 0.The entropy is given by:Z= ϵ −eϵ/T + e −ϵ/TZ. (1.16)S = ln Z + ⟨E⟩/T = ln(e ϵ/T + 1 + e −ϵ/T ) + ϵ T5−e ϵ/T + e −ϵ/T, (1.17)e ϵ/T + 1 + e−ϵ/T


and is zero for T = 0 and becomes ln 3 for T = ∞.ANOTHER EXAMPLE:C<strong>on</strong>sider two single-particle levels whose energies are −ϵ and ϵ. Into these levels, we place twoelectr<strong>on</strong>s (no more than <strong>on</strong>e electr<strong>on</strong> of the same spin per level). As a functi<strong>on</strong> of T find:(a) the partiti<strong>on</strong> functi<strong>on</strong>, (b) the average energy, (c) the entropyFirst, we enumerate the states. We can have <strong>on</strong>e state with both electr<strong>on</strong>s in the lower level,<strong>on</strong>e state where both electr<strong>on</strong>s are in the higher level, and four states with <strong>on</strong>e electr<strong>on</strong> in the lowerlevel and <strong>on</strong>e in the higher level. These two states differ based <strong>on</strong> whether the electr<strong>on</strong>s are in the↑↑, ↑↓, ↓↑, or ↓↓ c<strong>on</strong>figurati<strong>on</strong>s. The partiti<strong>on</strong> functi<strong>on</strong> is:and the average energy is:Z = e 2ϵ/T + 4 + e −2ϵ/T , (1.18)⟨E⟩ =∑i ϵ ie −ϵ i/TAt T = 0, ⟨E⟩ = −2ϵ, and at T = ∞, ⟨E⟩ = 0.The entropy is given by:Z= 2ϵ −e2ϵ/T + e −2ϵ/TZ. (1.19)S = ln Z + ⟨E⟩/T = ln(e 2ϵ/T + 4 + e −2ϵ/T ) + 2ϵT−e 2ϵ/T + e −2ϵ/T, (1.20)e 2ϵ/T + 4 + e−2ϵ/T and is zero for T = 0 and becomes ln 6 for T = ∞.With problems such as these, you need to very carefully differentiate between single-particlelevels and system levels, as <strong>on</strong>ly the latter appear in the sum for the partiti<strong>on</strong> functi<strong>on</strong>.1.5 Thermodynamic Potentials and Free EnergiesSince logs of the partiti<strong>on</strong> functi<strong>on</strong>s are useful quantities, they are often directly referred to asthermodynamic potentials or as free energies. For the case where both the charge and energy areallowed to vary, the potential is referred to as the grand can<strong>on</strong>ical potential,Ω GC ≡ T ln Z GC = T S − ⟨E⟩ + µ⟨Q⟩, (1.21)which is merely a restating of Eq. (1.14). Here, the subscript GC de<str<strong>on</strong>g>notes</str<strong>on</strong>g> that it is the grandcan<strong>on</strong>ical potential, sometimes called the “grand potential”. This differs from the can<strong>on</strong>ical case,where the energy varies but the charge is fixed (<strong>on</strong>ly states of the same charge are c<strong>on</strong>sidered).For that case there is no µQ/T term in the argument of the exp<strong>on</strong>entials used for calculating thepartiti<strong>on</strong> funcit<strong>on</strong>. For the can<strong>on</strong>ical case, the analog of the potential is the Helmholtz free energy,F ≡ −T ln Z C = ⟨E⟩ − T S. (1.22)For the grand potential and the Helmholts free energy, the signs and temperature factors are unfortunateas the definiti<strong>on</strong>s are historical in nature. Additi<strong>on</strong>ally, <strong>on</strong>e might c<strong>on</strong>sider the microcan<strong>on</strong>icalcase where energy is strictly c<strong>on</strong>served. In this case, <strong>on</strong>e <strong>on</strong>ly c<strong>on</strong>siders the entropy.6


The choice of can<strong>on</strong>ical vs. grand can<strong>on</strong>ical vs. micro-can<strong>on</strong>ical ensembles depends <strong>on</strong> thespecific problem. Micro-can<strong>on</strong>ical treatments tend to be more difficult and are <strong>on</strong>ly necessary forsmall systems where the fluctuati<strong>on</strong>s of the energy in a heat bath would be of the same order asthe average energy. Microcan<strong>on</strong>ical treatments are sometimes used for modestly excited atomicnuclei. Can<strong>on</strong>ical treatments are useful when the particle number is small, so that fluctuati<strong>on</strong>s ofthe particle number are similar to the number. Highly excited light nuclei are often treated withcan<strong>on</strong>ical approaches, allowing the energy to fluctuate, but enforcing the absolute c<strong>on</strong>servati<strong>on</strong> ofparticle number. Usually, calculati<strong>on</strong>s are simplest in the grand can<strong>on</strong>ical ensemble. In principle,the grand can<strong>on</strong>ical ensemble should <strong>on</strong>ly be used for a system in c<strong>on</strong>tact with both a heat bath anda particle bath. However, in practice it becomes well justified for a system of a hundred particles ormore, where the excitati<strong>on</strong> energy exceeds several times the temperature. Heavier nuclei can oftenbe justifiably treated from the perspective of the grand can<strong>on</strong>ical ensemble.The various ensembles can be related to <strong>on</strong>e another. For instance, if <strong>on</strong>e calculates the entropyin the microcan<strong>on</strong>ical ensemble, S(E, Q), <strong>on</strong>e can derive the Helmholtz free energy F (T, Q), byfirst calculating the density of states ρ(E). If the microcan<strong>on</strong>ical entropy was calculated for stateswithin a range δE, the density of states would found byρ(Q, E) = eS(Q,E)δE , (1.23)from which <strong>on</strong>e could calculate the Helmholtz free energy,∫F (Q, T ) = −T ln Z C = −T ln dE ρ(Q, E)e −E/T . (1.24)Finally, the grand can<strong>on</strong>ical potential can be generated from F ,{ }∑Ω GC (µ/T, T ) = T ln Z GC = T ln e −F (Q,T )/T e µQ/T . (1.25)QIt is tempting to relate the various ensembles through the entropy. For instance, <strong>on</strong>e couldcalculate ⟨E⟩ and ⟨Q⟩ in a grand can<strong>on</strong>ical ensemble from ln Z GC as a functi<strong>on</strong> of µ and T , thencalculate the entropy from Eq. (1.21). One could then use the value of S to generate the F in Eq.(1.22). However, the entropies in Eq.s (1.21) and (1.22) are not identical. They differ for smallsystems, but become identical for large systems. This difference derives from the fact that morestates are available if the charge or energy is allowed to fluctuate.In additi<strong>on</strong> to charge and energy, thermodynamic quantities can also depend <strong>on</strong> the volume Vof the system, which is assumed to be fixed for all three of the ensembles discussed above. Thedependence of the thermodynamic potentials <strong>on</strong> the volume defines the pressure. In the grandcan<strong>on</strong>ical ensemble, the pressure isP (µ, T, V ) ≡∂∂V Ω GC(µ/T, T, V ). (1.26)For large systems, those where the dimensi<strong>on</strong>s are much larger than the range of any interacti<strong>on</strong> orcorrelati<strong>on</strong>s, the potential Ω GC is proporti<strong>on</strong>al to the volume and P = Ω GC /V . In this case,P V = T ln Z GC = T S − E + µQ, (1.27)and the pressure plays the role of a thermodynamic potential for the grand can<strong>on</strong>ical ensemble.However, <strong>on</strong>e should remain cognizant that this is <strong>on</strong>ly true for the bulk limit with no l<strong>on</strong>g-rangeinteracti<strong>on</strong>s or correlati<strong>on</strong>s.7


1.6 Thermodynamic Relati<strong>on</strong>sIt is straightforward to derive thermodynamic relati<strong>on</strong>s involving expressi<strong>on</strong>s for the entropy in thegrand can<strong>on</strong>ical ensemble in Eq. (1.21) and the definiti<strong>on</strong> of the pressure, Eq. (1.27),S = ln Z GC (µ/T, T, V ) + E/T − (µ/T )Q, (1.28)( ∂ ln ZGCdS =− ⟨E⟩ ) ( )∂ ln ZGCdT +∂T T 2 ∂(µ/T ) − ⟨Q⟩ d(µ/T ) + P T dV + dE T − µdQT ,T dS = P dV + dE − µdQ.Here, Eq.s (1.12) and (1.13) were used to eliminate the first two terms in the sec<strong>on</strong>d line. Thisis probably the most famous thermodynamic relati<strong>on</strong>. For <strong>on</strong>e thing, it dem<strong>on</strong>strates that if twosystems a and b are in c<strong>on</strong>tact and can trade energy or charge, or if <strong>on</strong>e can expand at the other’sexpense, i.e.,dE a = −dE b , dQ a = −dQ b , dV a = −dV b , (1.29)that entropy can be increased by the transfer of energy, particle number or volume, if the temperatures,chemical potentials or pressures are not identical.(PadS a + dS b = − P ) (b 1dV a + − 1 ) (µadE a − − µ )bdQ a . (1.30)T a T b T a T b T a T bIf not blocked, such transfers will take place until the temperatures, pressures and chemical potentialsare equal in the two systems. At this point equilibrium is attained.From the expressi<strong>on</strong> for the energy in terms of partiti<strong>on</strong> functi<strong>on</strong>s,⟨E⟩ =∑i ϵ ie −ϵ i/T +µ i Q i /T∑i e−ϵ i/T +µ i Q i /T , (1.31)<strong>on</strong>e can see that energy rises m<strong>on</strong>ot<strong>on</strong>ically with temperature since higher temperatures give higherrelative weights to states with higher energy. Thus, as energy moves from a higher temperaturesystem to a lower temperature system, the temperatures will approach <strong>on</strong>e another. This is no lessthan the sec<strong>on</strong>d law of thermodynamics, i.e., stating that heat <strong>on</strong>ly moves sp<strong>on</strong>taneously from hotto cold is equivalent to stating that entropy must increase. Similar arguments can be made for thecharge and pressure, as charge will sp<strong>on</strong>taneously move to regi<strong>on</strong>s with lower chemical potentials,and higher pressure regi<strong>on</strong>s will expand into lower density regi<strong>on</strong>s.1.7 What to Minimize... What to MaximizeFor many problems in physics, there are parameters which adjust themselves according to thermodynamicc<strong>on</strong>siderati<strong>on</strong>s such as maximizing the entropy. For instance, in a neutr<strong>on</strong> star the nuclearmean field might adjust itself to minimize the free energy, or a liquid might adjust its chemicalc<strong>on</strong>centrati<strong>on</strong>s. Here, we will refer to such a parameter as x, and discuss how thermodynamics canbe used to determine x.For a system at fixed energy, <strong>on</strong>e would plot the entropy as a functi<strong>on</strong> of x, S(x, E, V, Q) andfind the value for x at which the entropy would be maximized. If the system is c<strong>on</strong>nected to aheat bath, where energy was readily exchanged, <strong>on</strong>e would also have to c<strong>on</strong>sider the change of theentropy in the heat bath. For a system a,dS tot = dS a + dS bath = dS a − dE a /T, (1.32)8


and the total entropy would be maximized when dS would be zero for changes dx. Stating thatdS tot = 0 is equivalent to stating that dF a = 0 at fixed T , where F a = E a − T S a is the Helmholtzfree energy of system a,dS tot | T,Q,V = − 1 T dF a| T,Q,V = 0. (1.33)Thus, if a system is in c<strong>on</strong>tact with a heat bath at temperature T , the value of x adjusts itselfto minimize (since the entropy comes in with the opposite sign) F (x, T, V, Q). If charges are alsoallowed to enter/exit the system with a bath at chemical potential µ, the change in entropy isHere, stating that dS tot = 0 is equivalent to stating,dS tot = dS a − dE a /T + µ a dQ a /T = 0. (1.34)dS tot | T,V,µ = d(S a − βE a + βµQ a )| T,V,µ = d P VT∣ = 0, (1.35)T,V,µThus, at fixed chemical potential and temperature a system adjusts itself to maximize the pressure.The calculati<strong>on</strong>s above assumed that the volume was fixed. In some instances, the pressure andparticle number might be fixed rather than the chemical potential and volume. In that case,This can then be stated as:dS tot = dS a − dE a /T − βP dV a = 0. (1.36)d(S a − βE a − βP V a ) | T,Q,P = −βd(µQ) = 0. (1.37)Here, the term µQ is referred to as the Gibb’s free energy,Summarizing the four cases discussed above:G ≡ µQ = E + P V − T S. (1.38)• If E, V and Q are fixed, x is chosen to maximize S(x, E, V, Q).• If Q and V are fixed, and the system is c<strong>on</strong>nected to a heat bath, x is chosen to minimizeF (x, T, V, Q).• If the system is allowed to exchange charge and energy with a bath at chemical potential µand temperature T , x is chosen to maximize P (x, T, µ, V ) (or Ω GC in the case that there arel<strong>on</strong>g range correlati<strong>on</strong>s or interacti<strong>on</strong>s).• If the charge is fixed, but the volume is allowed to adjust itself at fixed pressure, the systemwill attempt to minimize the Gibb’s free energy, G(x, T, P, Q).Perhaps the most famous calculati<strong>on</strong> of this type c<strong>on</strong>cerns the electroweak phase transiti<strong>on</strong> whichpurportedly took place in the very early universe, 10 −12 s after the big bang. In that transiti<strong>on</strong> theHigg’s c<strong>on</strong>densate ⟨ϕ⟩ became n<strong>on</strong>-zero in a process called sp<strong>on</strong>taneous symmetry breaking. Thefree energy can be calculated as a functi<strong>on</strong> of ⟨ϕ⟩ (Search literature of 1970s by Jackiw, Colemanand Linde). The resulting free energy depends <strong>on</strong> T and ⟨ϕ⟩, as illustrated qualitatively in Fig. 1.1.9


Figure 1.1: The free energy has a minimum at n<strong>on</strong>zero⟨ϕ⟩ at T = 0. In such circumstances a c<strong>on</strong>densedfield forms, as is the case for the Higg’s c<strong>on</strong>densate forthe electroweak transiti<strong>on</strong> or the scalar quark c<strong>on</strong>densatefor the QCD chiral transiti<strong>on</strong>. At high temperature, theminimum moves back to ⟨ϕ⟩ = 0. If ⟨ϕ⟩ is disc<strong>on</strong>tinuousas a functi<strong>on</strong> of T , a phase transiti<strong>on</strong> ensues. If ⟨ϕ⟩ is multidimensi<strong>on</strong>al, i.e., ϕ has two comp<strong>on</strong>ents with rotati<strong>on</strong>alsymmetry, the transiti<strong>on</strong> violates symmetry and isreferred to a sp<strong>on</strong>taneous symmetry breaking. Such potentialsare often referred to as “mexican hat” potentials.1.8 Free EnergiesPerusing the literature, <strong>on</strong>e is likely to find menti<strong>on</strong> of several free energies. For macroscopic systemswithout l<strong>on</strong>g-range interacti<strong>on</strong>s <strong>on</strong>e can write the grand potential as P V . The other comm<strong>on</strong> freeenergies can also be expressed in terms of the pressure as listed below:Grand Potential Ω = P V = T S − E + µQHelmholtz Free Energy F = E − T S = µQ − P VGibb’s Free Energy G = µQ = E + P V − T SEnthalpy ∗ H = E + P V = T S + µQTable 2: Various free energies∗ Enthalpy is c<strong>on</strong>served in what is called a Joule-Thoms<strong>on</strong> process, see http://en.wikipedia.org/wiki/Joule-Thoms<strong>on</strong> effect#Proof that enthalpy remains c<strong>on</strong>stant in a Joule-Thoms<strong>on</strong> process.In the grand can<strong>on</strong>ical ensemble, <strong>on</strong>e finds the quantities E and Q by taking derivatives of thegrand potential w.r.t. the Lagrange multipliers. Similarly, <strong>on</strong>e can find the Lagrange multipliers bytaking derivatives of some of the other free energies with respect to Q or E. To see this we beginwith the fundamental thermodynamic relati<strong>on</strong>,T dS = dE + P dV − µdQ. (1.39)Since the Helmholtz free energy is usually calculated in the can<strong>on</strong>ical ensemble,F = −T ln Z C (Q, T, V ) = E − T S, (1.40)it is typically given as a functi<strong>on</strong> of Q and T . N<strong>on</strong>etheless, <strong>on</strong>e can still find µ by c<strong>on</strong>sidering,dF = dE − T dS − SdT = −P dV − SdT + µdQ. (1.41)This shows that the chemical potential isµ = ∂F∂N ∣ . (1.42)V,T10


Similarly, in the microcan<strong>on</strong>ical ensemble <strong>on</strong>e typically calculates the entropy as a functi<strong>on</strong> ofE and Q. In this case the relati<strong>on</strong>dS = βdE + βP dV − βµdQ, (1.43)allows <strong>on</strong>e to read off the temperature asβ = ∂S∂E ∣ . (1.44)V,NIn many text books, thermodynamics begins from this perspective. We have instead begun with theentropy defined as − ∑ p i ln p i , as this approach does not require the c<strong>on</strong>tinuum limit, and allows thethermodynamic relati<strong>on</strong>s to be generated from general <strong>statistical</strong> c<strong>on</strong>cepts. The questi<strong>on</strong> “Whenis <strong>statistical</strong> <strong>mechanics</strong> thermodynamics?” is not easily answered. A more meaningful questi<strong>on</strong>c<strong>on</strong>cerns the validity of the fundamental thermodynamic relati<strong>on</strong>, T dS = dE + P dV − µdQ. Thisbecomes valid as l<strong>on</strong>g as <strong>on</strong>e can justify differentials, which is always the case for grand can<strong>on</strong>icalensemble, where E and Q are averages. However, for the other ensembles <strong>on</strong>e must be careful thatthe discreteness of particle number or energy levels invalidates the use of differentials, which is nevera problem for macroscopic systems.1.9 Forces at Finite TemperatureLet’s assume a particle feels a force dependent <strong>on</strong> it’s positi<strong>on</strong> r, but that the force depends <strong>on</strong> howthe thermal medium reacts to the particle’s positi<strong>on</strong>. Normally, <strong>on</strong>e would c<strong>on</strong>sider the force fromthe gradient of the energy, F i = −∂ i E(r). Indeed, this is the case in a thermal system,F i = − ∂E(r)∂r i∣ ∣∣∣S,where the partial derivative is taken at c<strong>on</strong>stant entropy. This is based <strong>on</strong> the idea that the particlemoves slowly, and that the particle remains in the a given, but shifting, energy level of the systemas the coordinate is slowly varied.However, thermal calculati<strong>on</strong>s are often performed more readily at fixed temperature, ratherthan fixed entropy. For that reas<strong>on</strong>, <strong>on</strong>e would like to know how to calculate the force usingderivatives where the temperature is fixed, and then correct for the fact that the temperature mustbe altered to maintain c<strong>on</strong>stant entropy. For simplicity, c<strong>on</strong>sider a <strong>on</strong>e-dimensi<strong>on</strong>al system.δE = ∂E∂x ∣ δx + ∂ET∂T ∣ δT, (1.45)xδS = 0 = ∂S∂x ∣ δx + ∂ST∂T ∣ δT.xNow, solve for δT in the lower expressi<strong>on</strong> and substitute into the upper expressi<strong>on</strong> to obtain,δE| S= ∂E∂x ∣ δx − ∂ T E∂ x Sδx (1.46)T∂ T S{ ∂E= δx∂x ∣ − T ∂S}T∂x ∣ ,T11


where the last step made use of the definiti<strong>on</strong> of temperature from Eq. (1.44),∂ T E∂ T S= dE/dS = T. (1.47)So, by inspecti<strong>on</strong> of Eq. (1.46) the force can now be state as∂E∂(E − T S)∂x ∣ = S∂x ∣ , (1.48)Tand the Helmholtz free energy, F = E − T S, plays the role of the potential. This result should notbe interpreted as the particle feels a force that pushes it towards minimizing the potential energyand maximizing the entropy. Particles d<strong>on</strong>’t feel entropy. The −T S term comes from having takengradients at fixed temperature, whereas the definiti<strong>on</strong> of the force is the derivative of the potentialenergy with respect to x at fixed entropy.1.10 Maxwell Relati<strong>on</strong>sMaxwell relati<strong>on</strong>s are a set of equivalences between derivatives of thermodynamic variables. Theycan prove useful in several cases, for example in the Clausius-Claeyr<strong>on</strong> equati<strong>on</strong>, which will bediscussed later. Derivati<strong>on</strong>s of Maxwell relati<strong>on</strong>s also tends to appear in written examinati<strong>on</strong>s.All Maxwell relati<strong>on</strong>s can be derived from the fundamental thermodynamic relati<strong>on</strong>,dS = βdE + (βP )dV − (βµ)dQ, (1.49)from which <strong>on</strong>e can readily identify β, P and µ with partial derivatives,β = ∂S∂E ∣ , (1.50)V,QβP = ∂S∂V ∣ ,E,Qβµ = − ∂S∂Q∣ .E,VHere, the subscripts refer to the quantities that are fixed. Since for any c<strong>on</strong>tinuous functi<strong>on</strong> f(x, y),∂ 2 f/(∂x∂y) = ∂ 2 f/(∂y∂x), <strong>on</strong>e can state the following identities which are known as Maxwellrelati<strong>on</strong>s,∂β∂V ∣ = ∂(βP )E,Q∂E ∣ , (1.51)V,Q∂β∂Q∣ = − ∂(βµ)E,V∂E ∣ ,V,Q∂(βP )∂Q ∣ = − ∂(βµ)β,V∂V ∣ .β,QAn entirely different set of relati<strong>on</strong>s can be derived by beginning with quantities other than dS.For instance, if <strong>on</strong>e begins with dF ,F = E − T S, (1.52)dF = dE − SdT − T dS,= −SdT − P dV − µdQ,12


three more Maxwell relati<strong>on</strong>s can be derived by c<strong>on</strong>sidering sec<strong>on</strong>d derivatives of F with respect toT, V and Q. Other sets can be derived by c<strong>on</strong>sidering the Gibbs free energy, G = E + P V − T S,or the pressure, P V = T S − E + µQ.EXAMPLE:Validate the following Maxwell relati<strong>on</strong>:Start with the standard relati<strong>on</strong>,∂V∂Q∣ = ∂µP,T∂P ∣ (1.53)T,QdS = βdE + (βP )dV − (βµ)dQ. (1.54)This will yield Maxwell’s relati<strong>on</strong>s where E, V and Q are held c<strong>on</strong>stant. Since we want a relati<strong>on</strong>where T, N and P/T are held c<strong>on</strong>stant, we should c<strong>on</strong>sider:This allows <strong>on</strong>e to realize:V = −d(S − (βP )V − βE) = −Edβ − V d(βP ) − (βµ)dQ. (1.55)∂(S − βP V − βE)∂(βP )∣ ,β,Qβµ = −∂(S − βP V − βE)∂Q∣ . (1.56)β,βPUsing the fact that ∂ 2 f/(∂x∂y) = ∂ 2 f/(∂y∂x),∂V∂Q∣ = ∂βµβP,T∂βP ∣ . (1.57)β,QOn the l.h.s., fixing βP is equivalent to fixing P since T is also fixed. On the r.h.s., the factors ofβ in ∂(βµ)/∂(βP ) cancel because T is fixed. This then leads to the desired result, Eq. (1.53).ANOTHER EXAMPLE:Show that∂E∂Q∣ = µ − P ∂µ∣S,P∂POne way to start is to look at the relati<strong>on</strong> and realize that V does not appear. Thus, let’s rewritethe fundamental thermodynamic relati<strong>on</strong> with dV = · · · . This ensures that we w<strong>on</strong>’t have any V sappearing <strong>on</strong> the r.h.s. of the equati<strong>on</strong>s going forward.∣S,QdV = T P dS − 1 P dE + µ P dQ.Now, since the relati<strong>on</strong> will involve S, P and Q being c<strong>on</strong>stant add d(E/P ) from each side to getd(V + E/P ) = T P dS + Ed(1/P ) + µ P dQ∂E∂Q∣ = ∂(µ/P )S,P∂(1/P ) ∣S,Q= µ − P ∂µ∣∂P13∣S,Q


Some students find a nmem<strong>on</strong>ic useful for deriving some of the Maxwell relati<strong>on</strong>s. Seehttp://en.wikipedia.org/wiki/Thermodynamic square. However, this is more cute than useful. Asyou will see with some of the more difficult problems at the end of this chapter.1.11 Fluctuati<strong>on</strong>sStatistical averages of higher order terms in the energy or the density can also be extracted from thepartiti<strong>on</strong> functi<strong>on</strong>. For instance, in the grand can<strong>on</strong>ical ensemble (using β = 1/T and α = −µ/Tas the variables),∂ 2 β ln Z(β, α) = ∂ β∂ β ZZ= 1 Z ∂2 βZ −( 1Z ∂ βZ) 2(1.58)= ⟨ E 2 ⟩ − ⟨E ⟩ 2= ⟨ (E − ⟨E⟩) 2⟩ .Thus, the sec<strong>on</strong>d derivative of ln Z gives the fluctuati<strong>on</strong> of the energy from its average value.Similarly, <strong>on</strong>e can see that∂ β ∂ α ln Z = ⟨(E − ⟨E⟩)(Q − ⟨Q⟩)⟩ (1.59)∂ β ∂ α ln Z = ⟨ (Q − ⟨Q⟩) 2⟩ . (1.60)Fluctuati<strong>on</strong>s play a critical role in numerous areas of physics, most notably in critical phenomena.1.12 Problems1. Using the methods of Lagrange multipliers, find x and y that minimize the following functi<strong>on</strong>,f(x, y) = 3x 2 − 4xy + y 2 ,subject to the c<strong>on</strong>straint,3x + y = 0.2. C<strong>on</strong>sider 2 identical bos<strong>on</strong>s (A given level can have an arbitrary number of particles) in a2-level system, where the energies are 0 and ϵ. In terms of ϵ and the temperature T , calculate:(a) The partiti<strong>on</strong> functi<strong>on</strong> Z C(b) The average energy ⟨E⟩. Also, give ⟨E⟩ in the T = 0, ∞ limits.(c) The entropy S. Also give S in the T = 0, ∞ limits.(d) Now, c<strong>on</strong>nect the system to a particle bath with chemical potential µ. Calculate Z GC (µ, T ).Find the average number of particles, ⟨N⟩ as a functi<strong>on</strong> of µ and T . Also, give theT = 0, ∞ limits.Hint: For a grand-can<strong>on</strong>ical partiti<strong>on</strong> functi<strong>on</strong> of n<strong>on</strong>-interacting particles, <strong>on</strong>e can state14


that Z GC = Z 1 Z 2 · · · Z n , where Z i is the partiti<strong>on</strong> functi<strong>on</strong> for <strong>on</strong>e single-particle level,Z i = 1+e −β(ϵ i−µ) +e −2β(ϵ i−µ) +e −3β(ϵ i−µ) · · · , where each term refers to a specific numberof bos<strong>on</strong>s in that level.3. Repeat the problem above assuming the particles are identical Fermi<strong>on</strong>s (No level can havemore than <strong>on</strong>e particle, e.g., both are spin-up electr<strong>on</strong>s).4. Beginning with the expressi<strong>on</strong>,T dS = dE + P dV − µdQ,show that the pressure can be derived from the Helmholtz free energy, F = E − T S, withP = − ∂F∂V ∣ .Q,T5. Beginning with:derive the Maxwell relati<strong>on</strong>,dE = T dS − P dV + µdQ,∂V∂µ ∣ = − ∂NS,P∂P ∣ .S,µ6. Assuming that the pressure P is independent of V when written as a functi<strong>on</strong> of µ and T ,i.e., ln Z GC = P V/T (true if the system is much larger than the range of interacti<strong>on</strong>),(a) Find expressi<strong>on</strong>s for E/V and Q/V in terms of P/T , and partial derivatives of P/Tw.r.t. α ≡ −µ/T and β ≡ 1/T . Here, assume the chemical potential is associated withthe c<strong>on</strong>served number Q.(b) Find an expressi<strong>on</strong> for C V = dE/dT | N,V in terms of P/T , E, Q, V and the derivativesof P/T , E and Q w.r.t. β and α.7. Beginning with the definiti<strong>on</strong>,Show that∂C p∂PC P = T ∂S∂T∣ ,N,P∣ ∣ = −T ∂2 V ∣∣∣P,NT,N∂T 2Hint: Find a quantity Y for which both sides of the equati<strong>on</strong> become ∂ 3 Y/∂ 2 T ∂P .8. Beginning withT dS = dE + P dV − µdQ, and G ≡ E + P V − T S,(a) Show thatS = − ∂G∂T∣ ,N,PV = ∂G∂P∣ .N,T15


(b) Beginning withδS(P, N, T ) = ∂S ∂S ∂SδP + δN +∂P ∂N ∂T δT,Show that the specific heats,C P ≡ T ∂S∂T ∣ , C V ≡ T ∂SN,P∂T ∣ ,N,Vsatisfy the relati<strong>on</strong>:C P = C V − T(∂V∂T) 2 (∣P,N∂V∂P) −1∣T,NNote that the compressibility, ≡ −∂V/∂P , is positive (unless the system is unstable),therefore C P > C V .9. From Sec. 1.11, it was shown how to derive fluctuati<strong>on</strong>s in the grand can<strong>on</strong>ical ensemble.Thus, it is straightforward to find expressi<strong>on</strong>s for the following fluctuati<strong>on</strong>s, ϕ EE ≡⟨δEδE⟩/V, ϕ QQ ≡ ⟨δQδQ⟩/V and ϕ QE ≡ ⟨δEδQ⟩/V . In terms of the 3 fluctuati<strong>on</strong>s calculatedin the grand can<strong>on</strong>ical ensemble, and in terms of the volume and the temperature T ,express the specific heat at c<strong>on</strong>stant volume,C V = dEdT∣ .Q,VNote that the fluctuati<strong>on</strong> observables, ϕ ij , are intrinsic quantities, assuming that the correlati<strong>on</strong>sin energy density occur within a finite range and that the overall volume is much largerthan that range.16


2 Statistical Mechanics of N<strong>on</strong>-Interacting Particles“Its a gas! gas! gas!” - M. Jagger & K. RichardsWhen students think of gases, they usually think back to high school physics or chemistry, andthink of the ideal gas law. This simple physics applies for gases which are sufficiently dilute thatinteracti<strong>on</strong>s between particles can be neglected and that identical-particle statistics can be ignored.For this chapter, we will indeed c<strong>on</strong>fine ourselves to the assumpti<strong>on</strong> that interacti<strong>on</strong>s can be neglected,but will c<strong>on</strong>sider in detail the effects of quantum statistics. Quantum statistics play adominant role in situati<strong>on</strong>s across all fields of physics. For instance, Fermi gas c<strong>on</strong>cepts play theessential role in understanding nuclear structure, stellar structure and dynamics, and the propertiesof metals and superc<strong>on</strong>ductors. For Bose systems, quantum degeneracy plays the central role inliquid Helium and in many atom/molecule trap systems. Even in the presence of str<strong>on</strong>g interacti<strong>on</strong>sbetween c<strong>on</strong>stituents, the role of quantum degeneracy still plays the dominant role in determiningthe properties and behaviors of numerous systems.2.1 N<strong>on</strong>-Interacting GasesA n<strong>on</strong>-interacting gas can be c<strong>on</strong>sidered as a set of independent momentum modes, labeled by themomentum p. For particles in a box defined by 0 < x < L x , 0 < y < L y , 0 < z < L z , the wavefuncti<strong>on</strong>s have the form,ψ px ,p y ,p z(x, y, z) ∼ sin(p x x/¯h) sin(p y y/¯h) sin(p z z/¯h) (2.1)with the quantized momentum ⃗p satisfying the boundary c<strong>on</strong>diti<strong>on</strong>s,p x L x /¯h = n x π, n x = 1, 2, · · · , (2.2)p y L y /¯h = n y π, n y = 1, 2, · · · ,p z L z /¯h = n z π, n z = 1, 2, · · ·The number of states in d 3 p is dN = V d 3 p/(π¯h) 3 . Here p x , p y and p z are positive numbers asthe momenta are labels of the static wavefuncti<strong>on</strong>s, each of which can be c<strong>on</strong>sidered as a coherentcombinati<strong>on</strong> of a forward and a backward moving plane wave. The states above are not eigenstatesof the momentum operator, as sin px/¯h is a mixture of left-going and right-going states. If <strong>on</strong>e wishesto c<strong>on</strong>sider plane waves, the signs of the momenta can be either positive or negative. Since <strong>on</strong>edoubles the range of momentum for each dimensi<strong>on</strong> when <strong>on</strong>e includes both positive and negativemomenta, <strong>on</strong>e needs to introduce a factor of 1/2 to the density of states for each dimensi<strong>on</strong> whenincluding the negative momentum states. The density of states in three dimensi<strong>on</strong>s then becomes:dN = Vd3 p(2π¯h) 3 . (2.3)For <strong>on</strong>e or two dimensi<strong>on</strong>s, d 3 p is replaced by d D p and (2π¯h) 3 is replaced by (2π¯h) D , where D isthe number of dimensi<strong>on</strong>s. For two dimensi<strong>on</strong>s, the volume is replaced by the area, and in <strong>on</strong>edimensi<strong>on</strong>, the length plays that role.In the grand can<strong>on</strong>ical ensemble, each momentum mode can be c<strong>on</strong>sidered as being an independentsystem, and being described by its own partiti<strong>on</strong> functi<strong>on</strong>, z p . Thus, the partiti<strong>on</strong> functi<strong>on</strong> of17


the entire system is the product of partiti<strong>on</strong> functi<strong>on</strong>s of each mode,Z GC = ∏ pz p , z p = ∑ n pe −n pβ(ϵ p −µq) . (2.4)Here n p is the number of particles of charge q in the mode, and n p can be 0 or 1 for Fermi<strong>on</strong>s, or canbe 0, 1, 2, · · · for bos<strong>on</strong>s. The classical limit corresp<strong>on</strong>ds to the case where the probability of havingmore than 1 particle would be so small that the Fermi and Bose cases become indistinguishable.Later <strong>on</strong> it will be shown that this limit requires that µ is much smaller (in units of T ) than theenergy of the lowest mode.For thermodynamic quantities, <strong>on</strong>ly ln Z GC plays a role, which allows the product in Eq. (2.4)to be replaced by a sum,ln Z GC = ∑ ln z p (2.5)p∫ V d 3 p= (2s + 1)(2π¯h) ln z 3 p∫ V d 3 p= (2s + 1)(2π¯h) ln(1 + 3 e−β(ϵ−µq) ) (Fermi<strong>on</strong>s),∫ ()V d 3 p= (2s + 1)(2π¯h) ln 1(Bos<strong>on</strong>s).3 1 − e −β(ϵ−µq)Here, the intrinsic spin of the particles is s. For the Fermi expressi<strong>on</strong>, where z p = 1 + e −β(ϵp−µ) ,the first term represents the c<strong>on</strong>tributi<strong>on</strong> for having zero particles in the state and the sec<strong>on</strong>d termaccounts for having <strong>on</strong>e particle in the level. For bos<strong>on</strong>s, which need not obey the Pauli exclusi<strong>on</strong>principle, <strong>on</strong>e would include additi<strong>on</strong>al terms for having 2, 3 · · · particles in the mode, and theBos<strong>on</strong>ic relati<strong>on</strong> has made use of the fact that 1 + x + x 2 + x 3 · · · = 1/(1 − x).EXAMPLE:For a classical two-dimensi<strong>on</strong>al, n<strong>on</strong>-interacting, n<strong>on</strong>-relativistic gas of fermi<strong>on</strong>s of mass m, andcharge q = 1 and spin s, find the charge density and energy density of a gas in terms of µ and T .The charge can be found by taking derivatives of ln Z GC with respect to µ,Q =∂∂(βµ) ln Z GC, (2.6)= (2s + 1)A∫ ∞02πp dp(2π¯h) · e −β(p2 /2m−µ)2 1 + e −β(p2 /2m−µ)= (2s + 1)A mT ∫ ∞2π¯h 2 dxe−(x−βµ)1 + e . −(x−βµ)The above expressi<strong>on</strong> applies to Fermi<strong>on</strong>s. To apply the classical limit, <strong>on</strong>e takes the limit that theβµ → −∞,QA0∫mT ∞= (2s + 1)2π¯h 2 dx e −(x−βµ) (2.7)0= (2s + 1) mT2π¯h 2 eβµ .18


In calculating the energy per unit area, <strong>on</strong>e follows a similar procedure beginning with,E = − ∂∂β ln Z GC, (2.8)∫E∞A = (2s + 1)02πp dp p 2(2π¯h) 2 2me −β(p2 /2m−µ)1 + e −β(p2 /2m−µq)= (2s + 1) mT 2 ∫ ∞2π¯h 2 dx xe−(x−βµ)1 + e −(x−βµ)≈ (2s + 1) mT 202π¯h 2 eβµ , (2.9)where in the last step we have taken the classical limit. If we had begun with the Bos<strong>on</strong>ic expressi<strong>on</strong>,we would have came up with the same answer after assuming βµ → −∞.As can be seen in working the last example, the energy and charge can be determined by thefollowing integrals,∫= (2s + 1)QVEV∫= (2s + 1)d D pf(ϵ), (2.10)(2π¯h)Dd D p(2π¯h) D ϵ(p)f(ϵ),where D is again the number of dimensi<strong>on</strong>s and f(p) is the occupati<strong>on</strong> probability, a.k.a. thephase-space occupancy or the phase-space filling factor,⎧⎨ e −β(ϵ−µ) /(1 + e −β(ϵ−µ) ), Fermi<strong>on</strong>s, Fermi − Dirac distributi<strong>on</strong>f(ϵ) = e −β(ϵ−µ) /(1 − e −β(ϵ−µ) ), Bos<strong>on</strong>s, Bose − Einstein distributi<strong>on</strong> (2.11)⎩e −β(ϵ−µ) , Classical, Boltzmann distributi<strong>on</strong>.Here f(ϵ) can be identified as the average number of particles in mode p. These expressi<strong>on</strong>s arevalid for both relativistic or n<strong>on</strong>-relativisitc gases, the <strong>on</strong>ly difference being that for relativistic gasesϵ(p) = √ m 2 + p 2 .The classical limit is valid whenever the occupancies are always much less than unity, i.e.,e β(µ−ϵ0) ≪ 1, or equivalently β(µ−ϵ 0 ) → −∞. Here, ϵ 0 refers to the energy of the lowest momentummode, which is zero for a massless gas, or for a n<strong>on</strong>-relativistic gas where the rest-mass energy isignored, ϵ = p 2 /2m. If <strong>on</strong>e preserves the rest-mass energy in ϵ(p), then ϵ 0 = mc 2 . The decisi<strong>on</strong> toignore the rest-mass in ϵ(p) is accompanied by translating µ by the same amount. For relativisticsystems, <strong>on</strong>e usually keeps the rest mass energy because <strong>on</strong>e can not ignore the c<strong>on</strong>tributi<strong>on</strong> fromanti-particles. For instance, if <strong>on</strong>e has a gas of electr<strong>on</strong>s and positr<strong>on</strong>s, the chemical potentialmodifies the phase space occupancy through a factor e βµ , while the anti-particle is affected by e −βµ ,and absorbing the rest-mass into µ would destroy the symmetry between the expressi<strong>on</strong>s for thedensities of particles and anti-particles.2.2 Equi-partiti<strong>on</strong> and Virial TheoremsParticles undergoing interacti<strong>on</strong> with an external field, as opposed to with each other, also behaveindependently. The phase space expressi<strong>on</strong>s from the previous secti<strong>on</strong> are again important here,19


as l<strong>on</strong>g as the external potential does not c<strong>on</strong>fine the particles to such small volumes that theuncertainty principle plays a role. The structure of the individual quantum levels will becomeimportant <strong>on</strong>ly if the energy spacing is not much smaller than the temperature.The equi-partiti<strong>on</strong> theorem applies for the specific case where the quantum statistics can beneglected, i.e. f(ϵ) = e −β(ϵ−µ) , and <strong>on</strong>e of the variables, p or q, c<strong>on</strong>tributes to the energy quadratically.The equi-partiti<strong>on</strong> theory states that the average energy associated with that variable is then(1/2)T . This is the case for the momentum of n<strong>on</strong>-relativistic particles being acted <strong>on</strong> by potentialsdepending <strong>on</strong>ly <strong>on</strong> the positi<strong>on</strong>. In that case,E = p2 x + p 2 y + p 2 z2m+ V (r). (2.12)When calculating the average of p 2 x/2m (in the classical limit),⟨ ⟩ ∫ p2x= (2π¯h)−3 dp x dp y dp z d 3 r (p 2 x/2m)e −(p2 x +p2 y +p2 z )/2mT −V (r)/T2m (2π¯h) ∫ , (2.13)−3 dp x dp y dp z d 3 r e −(p2 x+p 2 y+p 2 z)/2mT −V (r)/Tthe integrals over p y , p z and r factorize and then cancel leaving <strong>on</strong>ly,⟨ ⟩ p2x=2m∫dpx (p 2 x/2m)e −p2 x/2mT∫dpx e −p2 x /2mT = T 2 . (2.14)The average energy associated with p x is then independent of m.The same would hold for any other degree of freedom that appears in the Hamilt<strong>on</strong>ian quadratically.For instance, if a particle moves in a three-dimensi<strong>on</strong>al harm<strong>on</strong>ic oscillator,the average energy isH = p2 x + p 2 y + p 2 z2m+ 1 2 mω2 xx 2 + 1 2 mω2 yy 2 + 1 2 mω2 zz 2 , (2.15)⟨H⟩ = 3T, (2.16)with each of the six degrees of freedom c<strong>on</strong>tributing T/2.The generalized equi-partiti<strong>on</strong> theorem works for any variable that is c<strong>on</strong>fined to a finite regi<strong>on</strong>,and states ⟨q ∂H ⟩= T. (2.17)∂qFor the case where the potential is quadratic, this merely states the ungeneralized form of theequi-partiti<strong>on</strong> theorem. To prove the generalized form, <strong>on</strong>e need <strong>on</strong>ly integrate by parts,⟨q ∂H ⟩ ∫dq q(∂H/∂q) e−βH(q)= ∫ (2.18)∂qdq e−βH(q)= −T ∫ dq q(∂/∂q) e∫ −βH(q)dq e−βH(q)T[qe −βH(q)∣ ∣ ∞ + ∫ ]dq e −βH(q) −∞=∫dq e−βH(q)= T if H(q → ±∞) = ∞.20


Note that disposing of the term with the limits ±∞ requires that H(q → ±∞) → ∞, or equivalently,that the particle is c<strong>on</strong>fined to a finite regi<strong>on</strong> of phase space. The rule also works if p takes theplace of q.The virial theorem has an analog in classical <strong>mechanics</strong>, and is often referred to as the “ClausiusVirial Theorem”. It states, ⟨ ⟩ ⟨ ⟩∂H ∂Hp i = q i . (2.19)∂p i ∂q iTo prove the theorem we use Hamilt<strong>on</strong>’s equati<strong>on</strong>s of moti<strong>on</strong>,⟨ ddt (q ip i )⟩= ⟨p i ˙q i ⟩ + ⟨q i ṗ i ⟩ =⟨ ⟩∂Hp i −∂p i⟨ ⟩∂Hq i . (2.20)∂q iThe theorem will be proved if the time average of d/dt(p i q i ) = 0. Defining the time average as theaverage over an infinite interval τ,⟨ ddt (p iq i )⟩≡ 1 ∫ τdt d τ dt (p iq i ) = 1 τ (p iq i )| τ 0 , (2.21)0which will equal zero since τ → ∞ and p i q i is finite. Again, this finite-ness requires that theHamilt<strong>on</strong>ian c<strong>on</strong>fines both p i and q i to a finite regi<strong>on</strong> in phase space. In the first line of theexpressi<strong>on</strong> above, the equivalence between the <strong>statistical</strong> average and the time-weighted averagerequired the the Ergodic theorem.EXAMPLE:Using the virial theorem and the n<strong>on</strong>-generalized equi-partiti<strong>on</strong> theorem, show that for a n<strong>on</strong>relativisticparticle moving in a <strong>on</strong>e-dimensi<strong>on</strong>al potential,that the average potential energy isV (x) = κx 6 ,⟨V (x)⟩ = T/6.To proceed, we first apply the virial theorem to relate an average of the potential to an averageof the kinetic energy which is quadratic in p and can thus be calculated with the equi-partiti<strong>on</strong>theorem, ⟨x ∂V ⟩ ⟨= p ∂H ⟩=∂x ∂p⟨ ⟩ p2Next, <strong>on</strong>e can identify ⟨x ∂V ⟩= 6 ⟨V (x)⟩ ,∂xto see thatm⟨V (x)⟩ = T 6 .⟨ ⟩ p2= 2 = T. (2.22)2mOne could have obtained this result in <strong>on</strong>e step with the generalized equi-partiti<strong>on</strong> theorem.21


2.3 Degenerate Bose GasesThe term degenerate refers to the fact there are some modes for which the occupati<strong>on</strong> probabilityf(p) is not much smaller than unity. In such cases, Fermi<strong>on</strong>s and Bos<strong>on</strong>s behave very differently.For bos<strong>on</strong>s, the occupati<strong>on</strong> probability,f(ϵ) =e−β(ϵ−µ), (2.23)1 − e−β(ϵ−µ) will diverge if µ reaches ϵ. For the n<strong>on</strong>-relativistic case (ϵ = p 2 /2m ignoring the rest-mass energy)this means that a divergence will ensue if µ ≥ 0. For the relativistic case (ϵ = √ p 2 + m 2 includesthe rest-mass in the energy) a divergence will ensue when µ ≥ m. This divergence leads to thephenomena of Bose c<strong>on</strong>densati<strong>on</strong>.To find the density required for Bose c<strong>on</strong>densati<strong>on</strong>, <strong>on</strong>e needs to calculate the density for µ → 0 − .For lower densities, µ will be negative and there are no singularities in the phase space density. Thecalculati<strong>on</strong> is a bit tedious but straight-forward,∫ρ(µ = 0 − , T ) = (2s + 1)∫= (2s + 1)d D p(2π¯h) Dd D p(2π¯h) De −p2 /2mT1 − e −p2 /2mT∞∑n=1(∫(mT )D/2= (2s + 1)(2π¯h) De −np2 /2mT ,)∑ ∞d D xe −x2 /2n=1n −D/2(2.24)For D = 1, 2, the sum over n diverges. Thus for <strong>on</strong>e or two dimensi<strong>on</strong>s, Bose c<strong>on</strong>densati<strong>on</strong> neveroccurs (at least for massive particles where ϵ p ∼ p 2 ) as <strong>on</strong>e can attain an arbitrarily high densitywithout having µ reach zero. This can also be seen by expanding f(ϵ) for ϵ p and µ near zero,f(p → 0) ≈1βp 2 /m − βµ . (2.25)For D = 1, 2, the phase space weight p D−1 dp will not kill the divergence of f(p → 0) in the limitµ → 0. Thus, <strong>on</strong>e can obtain arbitrarily high densities without µ actually reaching 0. Since µ neverreaches zero, even as the density approaches infinite, Bose c<strong>on</strong>densati<strong>on</strong> never occurs.For D = 3, the c<strong>on</strong>densati<strong>on</strong> density is finite,(∫(mT )3/2ρ(µ = 0) = (2s + 1)(2π¯h) 3where ζ(n) is the Riemann-Zeta functi<strong>on</strong>,(mT )3/2= (2s + 1)3ζ(3/2),(2π)3/2¯h)∑ ∞d 3 xe −x2 /2n=1n −3/2 (2.26)ζ(n) ≡∞∑j=11j n (2.27)22


The Riemann-Zeta functi<strong>on</strong> appears often in <strong>statistical</strong> <strong>mechanics</strong>, and especially often in graduatewritten exams. Some values are:ζ(3/2) = 2.612375348685... ζ(2) = π 2 /6, ζ(3) = 1.202056903150..., ζ(4) = π 4 /90 (2.28)Once the density exceeds ρ c = ρ(µ = 0), µ no l<strong>on</strong>ger c<strong>on</strong>tinues to rise. Instead, the occupati<strong>on</strong>probability of the ground state,f =e−β(ϵ 0−µ)1 − e , (2.29)−β(ϵ 0−µ)becomes undefined, and any additi<strong>on</strong>al density above ρ c is carried by particles in the ground state.The gas has two comp<strong>on</strong>ents. For the normal comp<strong>on</strong>ent, the momentum distributi<strong>on</strong> is describedby the normal Bose-Einstein distributi<strong>on</strong> with µ = 0 and has a density of ρ c , while the c<strong>on</strong>densati<strong>on</strong>comp<strong>on</strong>ent has density ρ − ρ c . Similarly, <strong>on</strong>e can fix the density and find the critical temperatureT c , below which the system c<strong>on</strong>denses.Having a finite fracti<strong>on</strong> of the particles in the ground state does not explain superfluidity byitself. However, particles at very small relative momentum (in this case zero) will naturally formordered systems if given some kind of interacti<strong>on</strong> (For dilute gases the interacti<strong>on</strong> must be repulsive,as otherwise the particles would form droplets). An ordered c<strong>on</strong>glomerati<strong>on</strong> of bos<strong>on</strong>s will also movewith remarkably little fricti<strong>on</strong>. This is due to the fact that a “gap” energy is required to removeany of the particles from the ordered system, thus allowing the macroscopic c<strong>on</strong>densate to movecoherently as a single object carrying a macroscopic amount of current even though it moves veryslowly. Since particles bounce elastically off the c<strong>on</strong>densate, the drag force is proporti<strong>on</strong>al to thesquare of the velocity (just like the drag force <strong>on</strong> a car moving through air) and is negligible fora slow moving c<strong>on</strong>densate. Before the innovati<strong>on</strong>s in atom traps during the last 15 years, the<strong>on</strong>ly known Bose c<strong>on</strong>densate was liquid Helium-4, which c<strong>on</strong>denses at atmospheric pressure at atemperature of 2.17 ◦ K, called the lambda point. Even though liquid Helium-4 is made of tightlypacked, and therefore str<strong>on</strong>gly interacting, atoms, the predicti<strong>on</strong> for the critical temperature usingthe calculati<strong>on</strong>s above is <strong>on</strong>ly off by a degree or so.EXAMPLE:Phot<strong>on</strong>s are bos<strong>on</strong>s and have no charge, thus no chemical potential. Find the energy density andpressure of a two-dimensi<strong>on</strong>al gas of phot<strong>on</strong>s (ϵ = p) as a functi<strong>on</strong> of the temperature T . Note:whereas in three dimensi<strong>on</strong>s phot<strong>on</strong>s have two polarizati<strong>on</strong>s, they would have <strong>on</strong>ly <strong>on</strong>e polarizati<strong>on</strong>in two dimensi<strong>on</strong>s. Also, in two dimensi<strong>on</strong>s the pressure is a force per unity length, not a force perunit area as it is in three dimensi<strong>on</strong>s.First, write down an expressi<strong>on</strong> for the pressure (See HW problem),∫1P = 2πpdp p2 e −ϵ/T, (2.30)(2π¯h) 2 2ϵ 1 − e−ϵ/T which after replacing ϵ = p,P =∫14π¯h 2= T 34π¯h 2 ∫e−p/Tp 2 dp . (2.31)1 − e−p/T x 2 dx { e −x + e −2x + e −3x + · · · } ,= T {32π¯h 2 1 + 1 2 + 1 }3 3 + · · · = T 33 2π¯h 2 ζ(3),23


where the Riemann-Zeta functi<strong>on</strong>, ζ(3), is 1.202056903150... Finally, the energy density, by inspecti<strong>on</strong>of the expressi<strong>on</strong> for the pressure, is E/V = 2P .2.4 Degenerate Fermi GasesAt zero temperature, the Fermi-Dirac distributi<strong>on</strong> becomes,f(T → 0) =e−β(ϵ−µ)1 + e −β(ϵ−µ) ∣∣∣∣β→∞= Θ(µ − ϵ), (2.32)a step-functi<strong>on</strong>, which is unity for energies below the chemical potential and zero for ϵ > µ. Thedensity is then,ρ(µ, T = 0) =For three dimensi<strong>on</strong>s, the density of states is:D(ϵ)/V =∫ µ0dϵ D(ϵ)/V (2.33)(2s + 1)(2π¯h) 3 (4πp2 ) dpdϵ . (2.34)For small temperatures (much smaller than µ) <strong>on</strong>e can expand f in the neighborhood of µ,f(ϵ) = f 0 + δf, (2.35){ e−ϵ ′δf =/T /(1 + e −ϵ′ /T ) − 1, ϵ ′ ≡ ϵ − µ < 0e −ϵ′ /T /(1 + e −ϵ′ /T ), ϵ ′ > 0{ −eϵ ′=/T /(1 + e ϵ′ /T ), ϵ ′ < 0e −ϵ′ /T /(1 + e −ϵ′ /T ), ϵ ′ .> 0The functi<strong>on</strong> δf is illustrated in Fig. 2.1. The functi<strong>on</strong> δf(ϵ ′ ) is ∓1/2 at ϵ ′ = 0 and returns to zerofor |ϵ ′ | >> T . Also, δf is an odd functi<strong>on</strong> in ϵ ′ , which will be important below.One can expand D(µ + ϵ ′ ) in powers of ϵ ′ , and if the temperature is small, <strong>on</strong>e need <strong>on</strong>ly keepthe first term in the Taylor expansi<strong>on</strong> if <strong>on</strong>e is c<strong>on</strong>sidering the limit of small temperature.ρ(µ, T ) ≈ ρ(µ, T = 0) + 1 ∫ ∞[dϵ ′ D(µ) + ϵ ′ dD ]δf(ϵ ′ ), (2.36)V −∞dϵE≈E V V (µ, T = 0) + 1 ∫ ∞[dϵ ′ (µ + ϵ ′ ) D(µ) + ϵ ′ dD ]δf(ϵ ′ ).Vdϵ−∞Since <strong>on</strong>e integrates from −∞ → ∞, <strong>on</strong>e need <strong>on</strong>ly keep terms in the integrand which are even inϵ ′ . Given that δf is odd in ϵ ′ , the following terms remain,δρ(µ, T ) = 2 dDI(T ), (2.37)( ) V dϵEδ = µδρ + 2 D(µ)I(T ), (2.38)V VI(T )∫ ∞≡ dϵ ′ ϵ ′ e −ϵ′ /T0 1 + e −ϵ′ /T∫ ∞= T 2 dx xe−x1 + e . −x024


10.5f(ε)0.5δ f(ε)0ε'00µε-0.50εFigure 2.1: The phase space density at small temperature is shown in red, while the phase space densityat zero temperature is shown in green.With some trickery, I(T ) can be written in terms of ζ(2),∫ ∞I(T ) = T 2 dx x ( e −x − e −2x + e 3x − e −4x + · · · ) , (2.39)0∫ ∞= T 2 dx x {( e −x + e −2x + e 3x + · · · ) − 2 ( e −2x + e −4x + e 6x + e −8x + · · · )} ,0= T 2 ζ(2) − 2(T/2) 2 ζ(2) = T 2 ζ(2)/2,where ζ(2) = π 2 /6. This gives the final expressi<strong>on</strong>s:δρ(µ, T ) = π2 T 2( ) EδV6VdDdϵ∣ , (2.40)ϵ=µ= µδρ + π2 T 2D(µ). (2.41)6VThese expressi<strong>on</strong>s are true for any number of dimensi<strong>on</strong>s, as the dimensi<strong>on</strong>ality affects the answerby changing the values of D and dD/dϵ. Both expressi<strong>on</strong>s assume that µ is fixed. If ρ if fixedinstead, the µδρ term in the expressi<strong>on</strong> for δE is neglected. It is remarkable to note that theexcitati<strong>on</strong> energy rises as the square of the temperature for any number of dimensi<strong>on</strong>s. One powerof the temperature can be thought of as characterizing the range of energies δϵ, over which theFermi distributi<strong>on</strong> is affected, while the sec<strong>on</strong>d power comes from the change of energy associatedwith moving a particle from µ − δϵ to µ + δϵ.EXAMPLE:Find the specific heat for a relativistic three-dimensi<strong>on</strong>al gas of Fermi<strong>on</strong>s of mass m, spin s anddensity ρ at low temperature T .Even though the temperature is small, the system can still be relativistic if µ is not muchsmaller than the rest mass. Such is the case for the electr<strong>on</strong> gas inside a neutr<strong>on</strong> star. In that case,25


ϵ = √ m 2 + p 2 which leads to dϵ/dp = p/ϵ, and the density of states and the derivative from Eq.(2.34) becomes(2s + 1)D(ϵ) = V 4πpϵ . (2.42)(2π¯h)3The excitati<strong>on</strong> energy is then:δE fixed N = (2S + 1) p fϵ f V12¯h 3 T 2 , (2.43)where p f and ϵ f are the Fermi momentum and Fermi energy, i.e., the momentum where ϵ(p f ) =ϵ f = µ at zero temperature. For the specific heat this yields:C V = dE/dTρV= (2S + 1) 2p fϵ f12¯h 3 T. (2.44)ρTo express the answer in terms of the density, <strong>on</strong>e can use the fact that at zero temperature,ρ =(2s + 1) 4πp 3 f(2π¯h) 3 3 , (2.45)to substitute for p f in the expressi<strong>on</strong> for C V .One final note: Rather than using the chemical potential or the density, author’s often referto the Fermi energy ϵ f , the Fermi momentum p f , or the Fermi wave-number k f = p f /¯h. Thesequantities usually refer to the density, rather than to the chemical potential, though not necessarilyalways. That is, since the chemical potential changes as the temperature rises from zero to maintaina fixed density, it is somewhat arbitrary as to whether ϵ f refers to the chemical potential, or whetherit refers to the chemical potential <strong>on</strong>e would have used if T were zero. Usually, the latter criteriumis used for defining ϵ f , p f and k f . Furthermore, ϵ f is always measured relative to the energy forwhich p = 0, whereas µ is measured relative to the total energy which would include the potential.For instance, if there was a gas of density ρ, the Fermi momentum would be defined by:1 4πρ = (2s + 1)(2π¯h) 3 3 p3 f, (2.46)and the Fermi energy (for the n<strong>on</strong>-relativistic case) would be given by:ϵ f = p2 f2m . (2.47)If there were an attractive potential of strength V , the chemical potential at zero temperature wouldthen be ϵ f −V . If the temperature were raised while keeping ρ fixed, p f and ϵ f would keep c<strong>on</strong>stant,but µ would change to keep ρ fixed, c<strong>on</strong>sistent with Eq. (2.40).ANOTHER EXAMPLE:C<strong>on</strong>sider a two-dimensi<strong>on</strong>al n<strong>on</strong>-relativistic gas of spin-1/2 Fermi<strong>on</strong>s of mass m at fixed chemicalpotential µ.(A) For a small temperature T , find the change in density to order T 2 .26


(B) In terms of ρ and m, what is the energy per particle at T = 0(C) In terms of ρ, m and T , what is the change in energy per particle to order T 2 .First, calculate the density of states for two dimensi<strong>on</strong>s,AD(ϵ) = 2(2π¯h) 2πpdp 2 dϵ =A2m, (2.48)π¯hwhich is independent of p f . Thus, dD/dϵ = 0, and as can be seen from Eq. (2.40), the density doesnot change to order T 2 .The number of particles in area A is:N = 2A 1(2π¯h) 2 πp2 f =and the total energy of those particles at zero temperature is:E = 2A 1(2π¯h) 2 ∫ pf02πp f dp fand the energy per particle becomesEN ∣ = p2 fT =04m = ϵ f2 , p f =From Eq. (2.41), the change in the energy is:and the energy per particle changes by an amount,A2π¯h 2 p2 f, (2.49)p 2 f2m =A8mπ¯h 2 p4 f, (2.50)√2π¯h 2 ρ. (2.51)δE = D(µ) π26 T 2 = A πm6¯h 2 T 2 , (2.52)δEN = πm6¯h 2 ρ T 2 (2.53)2.5 Grand Can<strong>on</strong>ical vs. Can<strong>on</strong>icalIn the macroscopic limit, the grand can<strong>on</strong>ical ensemble is usually the easiest to work with. However,when working with <strong>on</strong>ly a few particles, there are differences between the restricti<strong>on</strong> that <strong>on</strong>ly thosestates with exactly N particles are c<strong>on</strong>sidered, as opposed to the restricti<strong>on</strong> that the average numberis N. This difference can matter when the fluctuati<strong>on</strong> of the number, which usually goes as √ N,is not much less than N itself. To understand how this comes about, c<strong>on</strong>sider the grand can<strong>on</strong>icalpartiti<strong>on</strong> functi<strong>on</strong>,Z GC (µ, T ) = ∑ e µN/T Z C (N, T ). (2.54)NThermodynamic quantities depend <strong>on</strong> the log of the partiti<strong>on</strong> functi<strong>on</strong>. If the c<strong>on</strong>tributi<strong>on</strong>s to Z GCcome from a range of ¯N ± δN, we can approximate ZGC by,Z GC (µ, T ) ≈ δNe µ ¯N/T Z C ( ¯N, T ) (2.55)ln Z GC ≈ µ ¯N/T + ln Z C ( ¯N, T ) + ln δN,27


and if <strong>on</strong>e calculates the entropy from the two ensembles, ln Z GC +βĒ−µ ¯N/T in the grand can<strong>on</strong>icalensemble and ln Z + βĒ in the can<strong>on</strong>ical ensemble, <strong>on</strong>e will see that the entropy is greater in thegrand can<strong>on</strong>ical case by an amount,S GC ≈ S C + ln δN. (2.56)The entropy is greater in the grand can<strong>on</strong>ical ensemble because more states are being c<strong>on</strong>sidered.However, in the limit that the system becomes macroscopic, ln Z C will rise proporti<strong>on</strong>al to N,and if <strong>on</strong>e calculates the entropy per particle, S/N, the can<strong>on</strong>ical and grand can<strong>on</strong>ical results willbecome identical. For 100 particles, δN ≈ √ N = 10, and the entropy per particle differs <strong>on</strong>ly byln 10/100 ≈ 0.023. As a comparis<strong>on</strong>, the entropy per particle of a gas at room temperature is <strong>on</strong>the order of 20 and the entropy of a quantum gas of quarks and glu<strong>on</strong>s at high temperature isapproximately four units per particle.2.6 Gibb’s ParadoxNext, we c<strong>on</strong>sider the manifestati<strong>on</strong>s of particles being identical in the classical limit, i.e., the limitwhen there is little chance two particles are in the same single-particle level. We c<strong>on</strong>sider thepartiti<strong>on</strong> functi<strong>on</strong> Z C (N, T ) for N identical spinless particles,Z C (N, T ) =z(T )NN!where z(T ) is the partiti<strong>on</strong> functi<strong>on</strong> of a single particle,, (2.57)z(T ) = ∑ pe −βϵ p. (2.58)In the macroscopic limit, the sum over momentum states can be replaced by an integral, ∑ p →[V/(2π¯h) 3 ] ∫ d 3 p.If the particles were not identical, there would be no 1/N! factor in Eq. (2.57). This factororiginates from the fact that if particle a is in level i and particle b is in level j, there is no differencefrom the state where a is in j and b is in i. This is related to Gibb’s paradox, which c<strong>on</strong>cerns theentropy of two gases of n<strong>on</strong>-identical particles separated by a partiti<strong>on</strong>. If the partiti<strong>on</strong> is lifted, theentropy of each particle, which has a c<strong>on</strong>tributi<strong>on</strong> proporti<strong>on</strong>al to ln V , will increase by an amountln 2. The total entropy will then increase by N ln V . However, if <strong>on</strong>e states that the particles areidentical and thus includes the 1/N! factor in Eq. (2.57), the change in the total entropy due tothe 1/N! factor is, using Stirling’s approximati<strong>on</strong>,∆S (from 1/N! term) = − ln N! + 2 ln(N/2)! ≈ −N ln 2. (2.59)Here, the first term comes from the 1/N! term in the denominator of Eq. (2.57), and the sec<strong>on</strong>dterm comes from the c<strong>on</strong>siderati<strong>on</strong> of the same terms for each of the two sub-volumes. Thisexactly cancels the N ln 2 factors inherent to doubling the volume, and the total entropy is notaffected by the removal of the partiti<strong>on</strong>. This underscores the importance of taking into accountthe indistinguishability of particles even when there is no significant quantum degeneracy.28


To again dem<strong>on</strong>strate the effect of the particles being identical, we calculate the entropy for athree-dimensi<strong>on</strong>al classical gas of N particles in the can<strong>on</strong>ical ensemble.S = ln Z C + βE = N ln z(T ) − N log(N) + N + βE, (2.60)∫( ) 3/2VmTz(T ) =d 3 p e −p2 /2mT = V(2π¯h) 3 2π¯h 2 , (2.61)[ ( ) ] 3/2V mTS/N = lnN 2π¯h 2 + 5 2 , (2.62)where the equipartiti<strong>on</strong> theorem was applied, E/N = 3T/2). The beauty of the expressi<strong>on</strong> is thatthe volume of the system does not matter, <strong>on</strong>ly V/N, the volume per identical particle. This meansthat when calculating the entropy per particle in a dilute gas, <strong>on</strong>e can calculate the entropy of <strong>on</strong>eparticle in a volume equal to the subvolume required so that the average number of particles in thevolume is unity. Note that if the comp<strong>on</strong>ents of a gas have spin s, the volume per identical particlewould be (2s + 1)V/N.2.7 Iterative Techniques for the Can<strong>on</strong>ical and Microcan<strong>on</strong>ical EnsemblesThe 1/N! factor must be changed if there is a n<strong>on</strong>-negligible probability for two particles to be inthe same state p. For instance, the factor z N in Eq. (2.57) c<strong>on</strong>tains <strong>on</strong>ly <strong>on</strong>e c<strong>on</strong>tributi<strong>on</strong> for allN particles being in the ground state, so there should be no need to divide that c<strong>on</strong>tributi<strong>on</strong> byN! (if the particles are bos<strong>on</strong>s, if they are Fermi<strong>on</strong>s <strong>on</strong>e must disregard any c<strong>on</strong>tributi<strong>on</strong> with morethan 1 particle per state). To include the degeneracy for a gas of bos<strong>on</strong>s, <strong>on</strong>e can use an iterativeprocedure. First, assume you have correctly calculated the partiti<strong>on</strong> functi<strong>on</strong> for N − 1 particles.The partiti<strong>on</strong> functi<strong>on</strong> for N particles can be written as:Z(N) = 1 Nz (n) ≡ ∑ p{z (1) Z(N − 1) ± z (2) Z(N − 2) + z (3) Z(N − 3) ± z (4) Z(N − 4) · · · } , (2.63)e −nβϵ p.Here, the ± signs refer to Bos<strong>on</strong>s/Fermi<strong>on</strong>s. If <strong>on</strong>ly the first term is c<strong>on</strong>sidered, <strong>on</strong>e recreates Eq.(2.57) after beginning with Z(N = 0) = 1 (<strong>on</strong>ly <strong>on</strong>e way to arrange zero particles). Proof of thisiterative relati<strong>on</strong> can be accomplished by c<strong>on</strong>sidering the trace of e −βH using symmetrized/antisymmetrizedwave functi<strong>on</strong>s (See <strong>Pratt</strong>, PRL84, p.4255, 2000).If energy and particle number are c<strong>on</strong>served, <strong>on</strong>e then works with the microcan<strong>on</strong>ical ensemble.Since energy is a c<strong>on</strong>tinuous variable, c<strong>on</strong>straining the energy is usually somewhat awkward. However,for the harm<strong>on</strong>ic oscillator energy levels are evenly spaced, which allows iterative treatments tobe applied. If the energy levels are 0, ϵ, 2ϵ · · · , the number of ways, N(A, E) to arrange A identicalparticles so that the total energy is E = nϵ can be found by an iterative relati<strong>on</strong>,N(A, E) = 1 AA∑ ∑(±1) a−1 N(A − a, E − aϵ i ), (2.64)a=1 i29


where i de<str<strong>on</strong>g>notes</str<strong>on</strong>g> a specific quantum state. For the Fermi<strong>on</strong>ic case this relati<strong>on</strong> can be used to solvea problem c<strong>on</strong>sidered by Euler, “How many ways can a integers be arranged to sum to E withoutusing the same number twice?”Iterative techniques similar to the <strong>on</strong>es menti<strong>on</strong>ed here can also incorporate c<strong>on</strong>servati<strong>on</strong> ofmultiple charges, angular momentum, and in the case of QCD, can even include the c<strong>on</strong>straint thatthe overall state is a coherent color singlet (<strong>Pratt</strong> and Ruppert, PRC68, 024904, 2003).2.8 Enforcing Can<strong>on</strong>ical and Microcan<strong>on</strong>ical C<strong>on</strong>straints through Integratingover Complex Chemical PotentialsFor systems that are interactive, or for systems where quantum degeneracy plays an important role,grand can<strong>on</strong>ical partiti<strong>on</strong> functi<strong>on</strong>s are usually easier to calculate. However, knowing the grandcan<strong>on</strong>ical partiti<strong>on</strong> functi<strong>on</strong> for all imaginary chemical potentials can lead <strong>on</strong>e to the can<strong>on</strong>icalpartiti<strong>on</strong> functi<strong>on</strong>. First, c<strong>on</strong>sider the grand can<strong>on</strong>ical partiti<strong>on</strong> functi<strong>on</strong> with an imaginary chemicalpotential, µ/T = iθ,Z GC (µ = iT θ, T ) = ∑ e −βE i+iθQ i. (2.65)iUsing the fact that,<strong>on</strong>e can see that,δ Q,Qi = 1 ∫ 2πdθ e i(Q−Qi)θ , (2.66)2π 0∫1 2πdθ Z GC (µ = iT θ, T )e −iQθ2π 0= ∑ e −βE iδ Q,Qii(2.67)= Z C (Q, T ).A similar expressi<strong>on</strong> can be derived for the microcan<strong>on</strong>ical ensemble,12π∫ ∞−∞dz Z C (T = i/z)e −izE = 12π= ∑ i∫ ∞−∞dz ∑ ie iz(ϵ i−E)δ(E − ϵ i ) = ρ(E).(2.68)As stated previously, the density of states plays the role of the partiti<strong>on</strong> functi<strong>on</strong> in the microcan<strong>on</strong>icalensemble.2.9 Problems1. Beginning with the expressi<strong>on</strong> for the pressure for a n<strong>on</strong>-interacting gas of bos<strong>on</strong>s,P VT= ln Z GC = ∑ ln ( 1 + e −β(ϵp−µ) + e −2β(ϵp−µ) + · · · ) , ∑ ∫V→ (2s + 1)(2π¯h) 3 ppd 3 p,show thatP =(2s + 1)(2π¯h) 3Here, the energy is relativistic, ϵ = √ p 2 + m 2 .∫d 3 p p2f(p), where f =e−β(ϵ−µ)3ϵ301 − e −β(ϵ−µ) .


2. Derive the corresp<strong>on</strong>ding expressi<strong>on</strong> for Fermi<strong>on</strong>s in the previous problem.3. Derive the corresp<strong>on</strong>ding expressi<strong>on</strong> for Bos<strong>on</strong>s/Fermi<strong>on</strong>s in two dimensi<strong>on</strong>s in the previousproblem. Note that in two dimensi<strong>on</strong>, P describes the work d<strong>on</strong>e per expanding by a unitarea, dW = P dA.4. For the two-dimensi<strong>on</strong>al problem above, show that P gives the rate at which momentum istransfered per unit length of the boundary of a 2-d c<strong>on</strong>fining box. Use simple kinematicc<strong>on</strong>siderati<strong>on</strong>s.5. C<strong>on</strong>sider classical n<strong>on</strong>-relativistic particles acting through a radially symmetric potential,V (r) = V 0 exp(r/λ).Using the equi-partiti<strong>on</strong> and virial theorems, show that⟨ rλ V (r) ⟩= 3T.6. C<strong>on</strong>sider a relativistic (ϵ = √ m 2 + p 2 ) particle moving in <strong>on</strong>e dimensi<strong>on</strong>,(a) Using the generalized equi-partiti<strong>on</strong> theorem, show that⟨ ⟩ p2= T.ϵ(b) Show the same result by explicitly performing the integrals in⟨ ⟩ ∫p2 dp (p 2 /ϵ)e= ∫ −ϵ/T.ϵ dp e−ϵ/THINT: Integrate the numerator by parts.7. C<strong>on</strong>sider a massless three-dimensi<strong>on</strong>al gas of bos<strong>on</strong>s with spin degeneracy N s . Assuming zerochemical potential, find the coefficients A and B for the expressi<strong>on</strong>s for the pressure andenergy density,( ) EP = AN s T 4 , = BN s T 4V8. Show that if the previous problem is repeated for Fermi<strong>on</strong>s that:A F ermi<strong>on</strong>s = 7 8 A Bos<strong>on</strong>s,B F ermi<strong>on</strong>s = 7 8 B Bos<strong>on</strong>s.9. C<strong>on</strong>sider a three-dimensi<strong>on</strong>al solid at low temperature where both the l<strong>on</strong>gitudinal and transversesound speeds are given by c s = 3000 m/s. Calculate the ratio of specific heats,C V (due to ph<strong>on</strong><strong>on</strong>s)C V (due to phot<strong>on</strong>s) ,31


where the phot<strong>on</strong> calculati<strong>on</strong> assumes the phot<strong>on</strong>s move in a vacuum of the same volume.Note that for sound waves the energy is ϵ = ¯hω = c s¯hk = c s p. For ph<strong>on</strong><strong>on</strong>s, there are threepolarizati<strong>on</strong>s (two transverse and <strong>on</strong>e l<strong>on</strong>gitudinal), and since the temperature is low, <strong>on</strong>e canignore the Debye cutoff which excludes high momentum nodes, as their wavelengths are belowthe spacing of the crystal.Data: c = 3.0 × 10 8 m/s, ¯h = 1.05457266 × 10 −34 Js.10. For a <strong>on</strong>e-dimensi<strong>on</strong>al n<strong>on</strong>-relativistic gas of spin-1/2 Fermi<strong>on</strong>s of mass m, find the change ofthe chemical potential δµ(T, ρ) necessary to maintain a c<strong>on</strong>stant density per unity length, ρ,while the temperature is raised from zero to T . Give answer to order T 2 .11. For a two-dimensi<strong>on</strong>al gas of spin-1/2 n<strong>on</strong>-relativistic Fermi<strong>on</strong>s of mass m at low temperature,find both the quantities below:d(E/A)d(E/A)dT ∣ ,∣µ,VdTGive both answers to the lowest n<strong>on</strong>-zero order in T , providing the c<strong>on</strong>stants of proporti<strong>on</strong>alityin terms of the chemical potential at zero temperature and m.12. C<strong>on</strong>sider harm<strong>on</strong>ic oscillator levels, 0, ϵ, 2ϵ · · · , populated by A Fermi<strong>on</strong>s of the same spin.(a) Using the iterative relati<strong>on</strong> in Eq. (2.64) solve for the number of ways, N(A, E), toarrange A particles to a total energy E for all A ≤ 3 and all E ≤ 6ϵ.(b) For N(A = 3, E = 6ϵ), list all the individual ways to arrange the particles.13. C<strong>on</strong>sider a two-dimensi<strong>on</strong>al gas of spin-less n<strong>on</strong>-degenerate n<strong>on</strong>-relativistic particles of massm.∣N,V(a) Show that the grand can<strong>on</strong>ical partiti<strong>on</strong> functi<strong>on</strong> Z GC (µ, T ) is{ }µ/T AmTZ GC (µ, T ) = exp e2π¯h 2(b) Using Eq. (2.67) in Sec. 2.8 and the expressi<strong>on</strong> above, show that the can<strong>on</strong>ical partiti<strong>on</strong>functi<strong>on</strong> Z C (N, T ) is( AmT) N2π¯hZ C (N, T ) =2 N!32


3 Interacting Gases and the Liquid-Gas Phase Transiti<strong>on</strong>“yeah my distracti<strong>on</strong> is my defenseagainst this lack of inspirati<strong>on</strong>against this slowly deflati<strong>on</strong>yeah the further the horiz<strong>on</strong> you knowthe more it warps my gazethe foreground’s out of focusbut you know I kinda hope it’sjust a phase, just a phasejust a phase, just a phasejust a phase, just a phasejust a phase” - A. Difranco3.1 Virial Expansi<strong>on</strong>The virial expansi<strong>on</strong> is defined as an expansi<strong>on</strong> of the pressure in powers of the density,[∞∑( ) ] n−1 ∫ρ (2j + 1)P = ρT A 1 + A n , ρ 0 ≡ d 3 p e −ϵp/T . (3.1)ρ 0 (2π¯h) 3n=2The leading term is always A 1 = 1, as low density matter always behaves as an ideal gas. Thec<strong>on</strong>tributi<strong>on</strong> from interacti<strong>on</strong>s to the sec<strong>on</strong>d order term, A 2 , is negative for attractive interacti<strong>on</strong>sand positive for repulsive interacti<strong>on</strong>s. Quantum statistics also affects A 2 . For bos<strong>on</strong>s, the c<strong>on</strong>tributi<strong>on</strong>is negative, while for fermi<strong>on</strong>s it is positive. The negative c<strong>on</strong>tributi<strong>on</strong> for bos<strong>on</strong>s can beunderstood by c<strong>on</strong>sidering the Bose-Einstein form for the phase-space occupancy, which provides arelatively str<strong>on</strong>ger enhancement to low-momentum particles. This lowers the average momentumof particles colliding with the boundaries, which lowers the pressure relative to P = ρT . Similarly,the c<strong>on</strong>tributi<strong>on</strong> is positive for fermi<strong>on</strong>s, as the average momentum of a particle in a Fermi gas israised by quantum statistics.Like perturbati<strong>on</strong> theory, virial expansi<strong>on</strong>s are <strong>on</strong>ly justifiable in the limit that they are unimportant,i.e., either <strong>on</strong>ly the first few terms matter. For that reas<strong>on</strong>, the sec<strong>on</strong>d-order coefficientis often calculated carefully from first principles, but subsequent coefficients are often insertedphenomenologically. The virial expansi<strong>on</strong> will be c<strong>on</strong>sidered more rigorously in Chapter 6, wherecoefficients will be expressed in terms of phase shifts (<strong>on</strong>ly sec<strong>on</strong>d order) and perturbati<strong>on</strong> theory.3.2 The Van der Waals Eq. of <strong>State</strong>All equati<strong>on</strong>s of state, P (ρ, T ), should behave as P = ρT at low density. The sec<strong>on</strong>d-order virialcoefficient is usually negative, as l<strong>on</strong>g-range interacti<strong>on</strong>s are usually attractive and dominate at lowdensity. For molecules, the l<strong>on</strong>g range interacti<strong>on</strong> usually originates from induced dipoles attracting<strong>on</strong>e another and falls off as 1/r 6 at large r. At high densities, particles tend to run out of space andrepel <strong>on</strong>e another. At some density, the c<strong>on</strong>tributi<strong>on</strong> to the pressure from interacti<strong>on</strong>s tends to rise<strong>on</strong>ce again.33


The Van der Waals parameterizati<strong>on</strong> of the equati<strong>on</strong> of state incorporates both an attractivecomp<strong>on</strong>ent for low density and repulsi<strong>on</strong> at high density,P =Here, the sec<strong>on</strong>d order virial coefficient isρT1 − ρ/ρ s− aρ 2 . (3.2)A 2 = ρ 0ρ s− aρ 0T , (3.3)and as l<strong>on</strong>g as a is large enough, the correcti<strong>on</strong> to the pressure to order ρ 2 is negative. However, athigh density, the first term in Eq. (3.2) will dominate, and in fact P → ∞ as ρ → ρ s . The quantityρ s is the saturati<strong>on</strong> density, as it is the highest density <strong>on</strong>e can obtain before the pressure jumpsto infinity. As the temperature goes to zero, a system will approach the saturati<strong>on</strong> density if thereis no pressure. The inverse of ρ s is often referred to as the excluded volume. The Van der Waalsequati<strong>on</strong> of state thus implies a hard-core interacti<strong>on</strong>.EXAMPLE:Using the Maxwell relati<strong>on</strong>,∂(P/T )∂β∣ = − ∂EN,V∂V ∣ ,N,Tderive the energy per particle E/N for a Van der Waals equati<strong>on</strong> of state, as a functi<strong>on</strong> of ρ and T .First,( )∂ P= ∂ ()ρ− βaρ 2 = −aρ 2 .∂β T ∂β 1 − ρ/ρ sBeggining with E/N = 3/2T at ρ = 0,EN = 3 2 T + 1 ∫ VdV aρ 2 = 3 ∫ ρN ∞ 2 T − 10 ρ dρ 2 aρ2 = 3 2 T − aρ.Remarkably, the energy per particle is independent of ρ s .If <strong>on</strong>e plots P vs. V , the behavior is no l<strong>on</strong>ger m<strong>on</strong>ot<strong>on</strong>ic for low T as seen in Fig. 3.1. Forthese temperatures below the critical temperature,T c = 8aρ s27 , (3.4)multiple densities can provide the same pressure. We refer to the inflecti<strong>on</strong> point as the criticalpoint, at which V c = 3/ρ s , and P c = aρ 2 s/27. We will derive this expressi<strong>on</strong> shortly.A regi<strong>on</strong> of densities for T < T c is unstable to phase separati<strong>on</strong>. To find the regi<strong>on</strong> we c<strong>on</strong>siderthe c<strong>on</strong>diti<strong>on</strong>s for phase co-existence.T gas = T liq , P gas = P liq , µ gas = µ liq . (3.5)34


Figure 3.1: The lower panel illustratesP vs V for several isotherms for the Vander Waals equati<strong>on</strong> of state. For T < T c ,the behavior is no l<strong>on</strong>ger m<strong>on</strong>ot<strong>on</strong>ic andphase separati<strong>on</strong> can occur. For T c , thereis an inflecti<strong>on</strong> point, where d 2 P/dV 2 =dP/dV = 0. This is the critical point. Inthe upper panel, the T = 0.7·T c isotherm isshown al<strong>on</strong>g with the two points denotingthe liquid and gas densities. Graphically,these points are found by requiring that thearea between the isotherm and a horiz<strong>on</strong>talline c<strong>on</strong>necting between two points <strong>on</strong> theisotherm integrates to zero.The first two c<strong>on</strong>diti<strong>on</strong>s can be satisfied by taking any two points <strong>on</strong> the same isotherm in Fig. 3.1that are at the same pressure. The third c<strong>on</strong>diti<strong>on</strong>, that the chemical potentials are equal, can alsobe illustrated with the isotherm. Beginning with<strong>on</strong>e can solve for Ndµ,T dS = dE + P dV − µdN, T S = P V + E − µN, (3.6)d(E − T S + µN) = −P dV − SdT − Ndµ (3.7)= −d(P V )Ndµ = −SdT + V dP.Following al<strong>on</strong>g an isotherm, dT = 0, and between two points illustrated in the upper panel of Fig.3.1 the change in the chemical potential is∫ ∫ ∫∫ vgasµ gas − µ liq = vdP = d(P v) − P dv = P gas v gas − P liq v liq − P dv = 0. (3.8)v liqHere, v = V/N is the volume per particle. The c<strong>on</strong>diti<strong>on</strong> that µ liq = µ gas can then be statedgraphically by stating that the integrated area between the P vs. V isotherm in the upper panelof Fig. 3.1 and the line between the two points is zero, i.e., there are equal amounts of area aboveand below the line.35


As an example of coexistence, Fig. 3.2 shows how the coexistence regi<strong>on</strong> is remarkably similarfor a variety of transiti<strong>on</strong>s when viewed after scaling the variables. Here, we make <strong>on</strong>ly a quickremark that the actual form of the coexistence curve, and the behavior of the matter, in the regi<strong>on</strong>near the critical point defies the simple Van der Waals picture. In that regi<strong>on</strong>, there are no l<strong>on</strong>gertwo distinct phases and the density fluctuates wildly. Critical phenomena represents a field in itself,and will be discussed later in the course.Figure 3.2: Phase coexistence for avariety of substances shows a remarkableuniversality after being scaled. TheVan der Waals al<strong>on</strong>g with Maxwell c<strong>on</strong>structi<strong>on</strong>sfails when working very closeto T c , but explains the behavior well forT < T c . This figure was compiled byGuggenheim in 1945 (J.Chem).EXAMPLE:Solve for T c , P c and V c in the Van der Waals equati<strong>on</strong> of state using the scaled variables, p = P/aρ 2 s,t = T/aρ s , v = ρ s V/N = ρ s /ρ.In terms of these variables,p =tv − 1 − 1 v . 2For T < T c there are two points where dP/dV = 0. At <strong>on</strong>e of these points, d 2 P/dV 2 < 0. At thecritical point,dpdv = d2 pdv = 0. 2If the sec<strong>on</strong>d derivative were not zero, P vs. V would not be m<strong>on</strong>ot<strong>on</strong>ic. These relati<strong>on</strong>s become:t c(v c − 1) = 2 1 , 22 vc3 (v c − 1) = 6 1 .3 vc4This gives, v c = 3 and t c = 8/27. Plugging this into the expressi<strong>on</strong> for p, p c = 1/27.t c3.3 Clausius-Clapeyr<strong>on</strong> Equati<strong>on</strong>The Clausius-Clapeyr<strong>on</strong> equati<strong>on</strong> allows us to plot the coexistence line in a P-V diagram if giventhe latent heat and the change of the volume per particle as <strong>on</strong>e goes across the phase line <strong>on</strong> a P36


vs. T representati<strong>on</strong> of the phase boundary.dPdT =L/T(V/N) gas − (V/N) liq(3.9)The phase line can be found by inspecting the coexistence curve, e.g. Fig. 3.1, for the value of thecoexistence pressure for a given isotherm.Figure 3.3: For phase coexistence, P , T and µ are c<strong>on</strong>tinuousacross the coexistence line. The requirement thatµ (B) − µ (A) is the same <strong>on</strong> both sides of the coexistence lineis the starting point for proving the Clausius-Clapeyr<strong>on</strong> equati<strong>on</strong>.To prove the relati<strong>on</strong> c<strong>on</strong>sider the coexistence line in the P-V diagram sketched in Fig. 3.3. SinceP , T and µ are c<strong>on</strong>tinuous across the line,µ (B)liq− µ(A) liqSince Ndµ = −SdT + V dP ,( ) ( )S V− dT + dPNliqNliq= −= µ(B) gas − µ (A)gas . (3.10)( SN)dT +gas( VN)dPdT = (S/N) gas − (S/N) liq(V/N) gas − (V/N) liq.dP, (3.11)gasGiven that the latent heat is L = T ∆S, this proves the relati<strong>on</strong>. Since L refers T ∆S between thetwo adjacent points across the phase boundary <strong>on</strong> the P vs. T coexistence line, it can be identifiedwith the heat added to change the phase if d<strong>on</strong>e either at c<strong>on</strong>stant T or at c<strong>on</strong>stant P .3.4 Perturbati<strong>on</strong> TheoryThe pressure P (µ, T ), density ρ(µ, T ), and energy density E/V (µ, T ), are all affected by interacti<strong>on</strong>s.But since, all such quantities are derived from the partiti<strong>on</strong> functi<strong>on</strong>, we will c<strong>on</strong>centrate <strong>on</strong>calculating the alterati<strong>on</strong> of the partiti<strong>on</strong> functi<strong>on</strong>. The relati<strong>on</strong>sE = −(∂/∂β) ln Z GC , ρ = (T/V )(∂/∂µ) ln Z GC , P/T = (∂/∂V ) ln Z GC , (3.12)hold for any system, interacting or not.The modern way to include interacti<strong>on</strong>s is through perturbati<strong>on</strong> theory. An alternative methodis based <strong>on</strong> “Mayer cluster diagrams”, but is applicable <strong>on</strong>ly for classical systems with interacti<strong>on</strong>sof the type,∫H int = d 3 r a d 3 r b V (r a − r b )ρ(r a )ρ(r b ). (3.13)37


The cluster technique is well described in Pathria (Statistical Mechanics, R.K. Pathria, Pergamm<strong>on</strong>,or in Huang, Sec. 10.1). More general interacti<strong>on</strong>s might involve exchanging off-shell phot<strong>on</strong>s orglu<strong>on</strong>s. Since the standard methods of quantum perturbati<strong>on</strong> theory apply to all cases, and naturallyincorporate relativity and quantum <strong>mechanics</strong>, we will focus <strong>on</strong> these methods. Our discussi<strong>on</strong> willstay brief. Perturbati<strong>on</strong> theory at finite temperature can easily be a subject for an entire course.Perturbati<strong>on</strong> theory at finite temperature c<strong>on</strong>cerns calculati<strong>on</strong> of Z GC ,Z GC = ∑ i⟨i|e −β(H−µQ) |i⟩, (3.14)where i refers to specific quantum states. Even if the system has 10 23 particles, and is interacting,eigen-states for the bulk system still exist. We will separate H − µQ into two parts, H 0 and V ,where H 0 is usually c<strong>on</strong>sidered to be the part for which the state is expressed simply by listingwhich single-particle states are occupied, i.e., the many-body wave functi<strong>on</strong> is a product of singleparticlewave functi<strong>on</strong>s. We note that in our notati<strong>on</strong> H 0 includes the c<strong>on</strong>tributi<strong>on</strong> from −µQ. Inthe language of creati<strong>on</strong> and destructi<strong>on</strong> operators, H 0 is usually the part of the field which canbe written as ∑ i (ϵ i − µ)a † i a i, while the interacti<strong>on</strong> V represents all other parts of the Hamilt<strong>on</strong>ianwhere the creati<strong>on</strong> and destructi<strong>on</strong> operators appear in higher powers. The eigen-states are oftenmomentum eigen-states, though they might also be eigen-states of a classical field. In the nuclearshell model, <strong>on</strong>e uses a basis of harm<strong>on</strong>ic-oscillator eigen-states, while in some other variants ofnuclear many-body theory <strong>on</strong>e might use eigen-states of a potential v(r), chosen to be c<strong>on</strong>sistentwith the density. In nuclear physics the interacti<strong>on</strong> V is also often referred to as the “residual”interacti<strong>on</strong>, referring to the fact that it is what is left over after having absorbed as much as possibleinto the effective potential v(r).After identifying V , <strong>on</strong>e moves to the interacti<strong>on</strong> representati<strong>on</strong>, with the definiti<strong>on</strong>s,U int (β) ≡ e βH 0e −βH , (3.15)V (β) ≡ e βH 0V e −βH 0,which leads to the relati<strong>on</strong> for the evoluti<strong>on</strong> operator U,∂∂β U int(β) = e βH 0H 0 e −βH − e βH 0(H 0 + V )e −βH = −e βH 0V e −βH (3.16)= −V (β)U int (β).The goal is to calculate the partiti<strong>on</strong> functi<strong>on</strong>,Z GC = Tr e −βH 0U int (β). (3.17)This is the standard technique for quantum perturbati<strong>on</strong> theory in the interacti<strong>on</strong> representati<strong>on</strong>,<strong>on</strong>ly in those cases <strong>on</strong>e is calculating the evoluti<strong>on</strong> operator e −iHt . The derivati<strong>on</strong>s are equivalent,except that it/¯h is replaced with β. Thus, the Boltzmann factor is identified with the evoluti<strong>on</strong>operator in imaginary time. Furthermore, <strong>statistical</strong> calculati<strong>on</strong>s c<strong>on</strong>sider the trace rather than theexpectati<strong>on</strong> between specific initial and final states.The differential eqauti<strong>on</strong> for U(β) in Eq. (3.16) can be equivalently stated as an integral equati<strong>on</strong>,U int (β) = 1 −∫ β038dβ ′ V (β ′ )U int (β ′ ). (3.18)


One can now replace U in the r.h.s.expressi<strong>on</strong>,with the complete expressi<strong>on</strong> for U to see the iterativeU int (β) =1 −−∫ β0∫ βdβ ′ V (β ′ ) +0 0 0∫ β0∫ β ′dβ ′ dβ ′′ V (β ′ )V (β ′′ ) (3.19)0∫ β ′ ∫ β ′′dβ ′ dβ ′′ dβ ′′′ V (β ′ )V (β ′′ )V (β ′′′ ) + · · · .Here, U int (β) and V (β) are matrices, with the indices referring to specific states.complicati<strong>on</strong>s associated with the limits for each variable,To overcomeU int (β) =1 − T∫ β0dβ ′ V (β ′ ) + 1 ∫ β ∫ β2! T dβ ′ dβ ′′ V (β ′ )V (β ′′ ) (3.20)− 1 ∫ β ∫ β ∫ β3! T dβ ′ dβ ′′ dβ ′′′ V (β ′ )V (β ′′ )V (β ′′′ ) + · · · ,00where the limits of the various integrati<strong>on</strong>s are now independent from <strong>on</strong>e another, with the factthat the factors V (β) at lower values of β are always pushed to the right by the “time-ordering”operator T . Since there are n! ways to order such operators, a factor 1/n! is added to correct forthe over-counting. If V (β) were to commute with V (β ′ ) (which would be true if [V, H 0 ] = 0), thetime-ordering operator would be irrelevant, and U int could be written as exp{− ∫ dβ ′ V (β ′ )}. Thiswould be the case if V were diag<strong>on</strong>al in the same basis that diag<strong>on</strong>alized H 0 . This is the case foruniform fields, which might change the energy of a particle of momentum p, but does not change thefact that p is an eigenstate. The more general case of scattering does not satisfy the commutati<strong>on</strong>c<strong>on</strong>diti<strong>on</strong> as particles with momentum p 1 and p 2 are scattered into p ′ 1 and p ′ 2 by V . Some classes ofinteracti<strong>on</strong>s represent exchanges of charge or can even change the number of quanta. For instance,in the three-point interacti<strong>on</strong> in electro-weak theory, V ii ′, i might refer to an electr<strong>on</strong> and i ′ mightrefer to a neutrino and W − bos<strong>on</strong>.The states i and i ′ in the matrices V ii ′ used in perturbati<strong>on</strong> theory are many-body states, with ilabeling the c<strong>on</strong>figurati<strong>on</strong> of perhaps 10 23 particles. However, the interacti<strong>on</strong> V rarely involves all thequanta. Instead, it might involve changing the state of two particles, say those with momentum p 1and p 2 to p ′ 1 and p ′ 2. For this two-particle-to-two-particle case, we c<strong>on</strong>sider two particles interactingn<strong>on</strong>-relativistically through a potential U(r 1 −r 2 ). This interacti<strong>on</strong> is diag<strong>on</strong>al in the r 1 , r 2 basis(butclearly not diag<strong>on</strong>al in the p 1 , p 2 basis) because the particles do not move instantaneously and00⟨r 1 , r 2 |V |r ′ 1, r ′ 1⟩ = U(r 1 − r 2 )δ 3 (r 1 − r ′ 1)δ 3 (r 2 − r ′ 2). (3.21)The same interacti<strong>on</strong> sandwiched between momentum states describes a change in the relativemomentum,q ≡ p 1 − p 2 → q ′ ≡ p ′ 1 − p ′ 2, (3.22)with the total momentum not changing. This is expected given the translati<strong>on</strong> invariance of the039


interacti<strong>on</strong>. Performing the machinati<strong>on</strong>s,∫⟨p 1 , p 2 |V |p ′ 1, p ′ 2⟩ = d 3 r 1 d 3 r 2 ⟨p 1 , p 2 |r 1 , r 2 ⟩⟨r 1 , r 2 |V |r 1 , r 2 ⟩⟨r 1 , r 2 |p ′ 1, p ′ 2⟩ (3.23)= 1 ∫d 3 rV 2 1 d 3 r 2 e ip 1·r 1 /¯h+ip 2·r 2 /¯h U(r 1 − r 2 )e −ip′ 1·r 1/¯h−ip ′ 2·r 2/¯h= 1 ∫d 3 R e i(P−P′ )·R/¯h d 3 r e i(q−q′ )·r/¯h U(r)V{2 = (2π¯h)3 δ=3 (P − P ′)Ũ(k ≡ (q − V 2 q′ )/¯h), if P = P ′ not yet c<strong>on</strong>strained.= 1 Ũ(k ≡ (q − V q′ )/¯h), if P = P ′ already c<strong>on</strong>strained.r = r 1 − r 2 , P = p 1 + p 2 , R = m 1r 1 + m 2 r 2m 1 + m 2, q = m 2p 1 − m 1 p 2m 1 + m 2.The inverse powers of the volume, which originated from having normalized wave functi<strong>on</strong>s e ikx / √ V ,will disappear when summing over the density of states, and are often omitted in many text books.Thus, the matrix element between two-particle states typically includes a δ functi<strong>on</strong> representingthe c<strong>on</strong>servati<strong>on</strong> of momentum and the Fourier transform of the potential Ũ(k), where k is thedifference in the relative momenta (divided by ¯h) between the incoming and outgoing states.For many-body theory, it is more physical to think of the interacti<strong>on</strong> in terms of creati<strong>on</strong> anddestructi<strong>on</strong> operators. In that case,V = 1 V∑⟨p 1 , p 2 |V |p ′ 1, p ′ 2⟩a † pa † ′p ′ 1 p 2a ′ p1 a p2 , (3.24)1 ,p′ 2 ,p 1,p 2where a † p/a p are creati<strong>on</strong>/destructi<strong>on</strong> operators that create/remove a particle with momentum p.In the interacti<strong>on</strong> representati<strong>on</strong>, a † p(β) = e −βϵ pa † p, andV (β) =∑⟨p 1 , p 2 |V |p ′ 1, p ′ 2⟩a † pa † ′p ′ 1 p 2a ′ p1 a p2 e −β(ϵ′ 1 +ϵ′ 2 −ϵ 1−ϵ 2 ) . (3.25)1 ,p′ 2 ,p 1,p 2For such a two-particle-to-two-particle matrix element, V (β) would destroy particles with momentump 1 and p 2 , then create two new particles with p ′ 1 and p ′ 2. The initial many-particle state i mightundergo several combinati<strong>on</strong>s of V at the various “times” before being returned to its original c<strong>on</strong>figurati<strong>on</strong>,as demanded by the trace. For some applicati<strong>on</strong>s, V (β) might refer to the annihilati<strong>on</strong>of <strong>on</strong>e particle and the creati<strong>on</strong> of two particles, such as in the decay of a W bos<strong>on</strong> into an electr<strong>on</strong>and a neutrino. Similarly, the merging of the electr<strong>on</strong> and neutrino into a W bos<strong>on</strong> would berepresented by a term with two annihilati<strong>on</strong> operators and <strong>on</strong>e creati<strong>on</strong> operator.Our immediate goal is to show how perturbati<strong>on</strong> theory, whose formalism refers to many-bodystates, can be rewritten in terms of matrix elements involving <strong>on</strong>ly a few particles. We will see thatwhen <strong>on</strong>e calculates the log of the partiti<strong>on</strong> functi<strong>on</strong>, that indeed, the correcti<strong>on</strong>s due to interacti<strong>on</strong>sinvolve the c<strong>on</strong>siderati<strong>on</strong> of matrix elements such as those in Eq. (3.23) that involve <strong>on</strong>ly a fewparticles at a time.To accomplish the reducti<strong>on</strong> menti<strong>on</strong>ed above, we will first show how the n th -order c<strong>on</strong>tributi<strong>on</strong>in the perturbative expansi<strong>on</strong> of the grand can<strong>on</strong>ical partiti<strong>on</strong> functi<strong>on</strong> Z (The GC subscript willbe suppressed) can be factorized into a “c<strong>on</strong>nected” piece and a lower-order term for the expansi<strong>on</strong>40


β β β β β β β Figure 3.4: This illustrates how a termT ⟨V (β 1 ) · · · V (β n )⟩ might look for n = 11.The term V (β 1 ) is c<strong>on</strong>nected in an m = 5subdiagram which can be factored out of the matrix element. The number of waysthat the V (β 1 ) term can be c<strong>on</strong>nected toa group of m out of n is (n − 1)!/[(n −m)!(m − 1)!].β ββ β β β of Z. The n th order term is:Z n = (−1)nn!∫ β0dβ 1 · · · dβ n∑T ⟨ i|e −βH 0V (β 1 ) · · · V (β n )|i ⟩ (3.26)iWe now c<strong>on</strong>sider the specific term V (β 1 ) as illustrated a specific path illustrated in Fig. 3.4. Eachvertex represents a two-body-to-two-body matrix element, with the two incoming lines (from theleft) representing the annihilated momenta, and the two outgoing lines (to the right) representingthe new momenta.The term V (β 1 ) will be “c<strong>on</strong>nected” to those other potential terms V (β i ) for which <strong>on</strong>e can forman independent chain of initial momenta which are transformed then return to the same independentmomenta. In the figure the number of such terms is m = 5. By careful inspecti<strong>on</strong> <strong>on</strong>e can see thatthe remainder of the matrix element is independent and factorizes away from those terms c<strong>on</strong>nectedto V (β 1 ). One can re-express Z n as a sum over all m,Z n = (−1)n ∑[∫](n − 1)!dβ 1 · · · dβ m T ⟨⟨V (β 1 ) · · · V (β m )⟩⟩n! (m − 1)!(n − m)!c<strong>on</strong>nected(3.27)m≤n[∫· dβ 1 ′ · · · dβ n−m ′ T ⟨⟨ V (β 1) ′ · · · V (β n−m) ⟩⟩]′= 1 ∑∫(−1) m 1Z n−m dβ 1 · · · dβ m T ⟨⟨V (β 1 ) · · · V (β m )⟩⟩n(m − 1)!c<strong>on</strong>nected.m≤nHere, the double brackets denote the thermal trace, ⟨⟨A⟩⟩ = ∑ i ⟨i|e−βH 0A|i⟩. The preceeding 1/nfactor prevents <strong>on</strong>e from immediately factorizing Z into Z multiplied by a piece depending <strong>on</strong>ly <strong>on</strong>41


c<strong>on</strong>nected diagrams. To overcome this limitati<strong>on</strong>, we replace V with λV , where λ = 1. Then, sinceZ n will depend <strong>on</strong> λ as λ n , we can writeλ ddλ Z = ∑ λ ddλ Z n (3.28)n= ∑ ∑ (−1) m ∫Z n−mdβ 1 · · · dβ m T ⟨⟨V (β 1 ) · · · V (β m )⟩⟩(m − 1)!c<strong>on</strong>nectedm n>m= Z · ∑ (−1) m ∫dβ 1 · · · dβ m T ⟨⟨V (β 1 ) · · · V (β m )⟩⟩(m − 1)!c<strong>on</strong>nected.mAfter dividing both sides by Z <strong>on</strong>e finds an expressi<strong>on</strong> for ln Z,ddλ ln Z = 1 ∑ (−1) m ∫dβ 1 · · · dβ m T ⟨⟨V (β 1 ) · · · V (β m )⟩⟩λ (m − 1)!c<strong>on</strong>nected. (3.29)mThe final expressi<strong>on</strong> then comes from noting the the m th -order c<strong>on</strong>nected piece behaves as λ m .Thus, <strong>on</strong>e can integrate both sides of the equati<strong>on</strong> from λ = 0 to λ = 1 to obtain our final result:ln Z = ln Z (V=0) + ∑ (−1) m ∫dβ 1 · · · dβ m T ⟨⟨V (β 1 ) · · · V (β m )⟩⟩m!c<strong>on</strong>nected. (3.30)mThis profound expressi<strong>on</strong> permits <strong>on</strong>e to calculate correcti<strong>on</strong>s to the partiti<strong>on</strong> functi<strong>on</strong> for 10 23particles in terms of calculable terms, each of which represents <strong>on</strong>ly a few particles.EXAMPLE:C<strong>on</strong>sider n<strong>on</strong>-relativistic n<strong>on</strong>-degenerate particles of mass m interacting through a potentialU(r) = U 0 e −r/a .Find the correcti<strong>on</strong> to the pressure to first-order in perturbati<strong>on</strong> theory for fixed chemical potential.Give answer in terms of α = (2j + 1)e µ/T .First, calculate the matrix element from Eq. (3.23). Since the first order expressi<strong>on</strong> has thesame total momentum in the bra and ket,⟨p 1 , p 2 |V |p ′ 1, p ′ 2⟩ = 1 Ũ(k = 0).VSince k = 0, the integral over r especially easy.⟨p 1 , p 2 |V |p 1 , p 2 ⟩ = 1 V∫d 3 r U(r). (3.31)Now, <strong>on</strong>e can calculate the matrix element necessary for perturbati<strong>on</strong> theory,⟨p 1 , p 2 |e −βH 0V (β 1 )|p 1 , p 2 ⟩ = ⟨p 1 , p 2 |e −βH 0e −β 1H 0V e β 1H 0|p 1 , p 2 ⟩ (3.32)= e −β(P 2 /2M+q 2 /2m red )+2βµ 1 ∫d 3 r U(r).V42


The exp<strong>on</strong>entials with β 1 canceled because the state was the same <strong>on</strong> the left and right of V . Thiscancellati<strong>on</strong> <strong>on</strong>ly happens in first-order.Approaching the home stretch, <strong>on</strong>e now uses Eq. (3.30) to work toward the answer,ln Z = ln Z 0 − 1 ∑(∫ βdβ 1 e −β(P 2 /2M+q 2 /2m red )+2βµ 1 ) ∫d 3 r U(r)2P,q0V= ln Z 0 − e 2βµ β V (2j + 1) 2 ∫ ∫∫d 3 P d 3 q e −β(P 2 /2M+q 2 /2m red )d 3 r U(r)2 (2π¯h) 6= ln Z 0 − β 2 V (MT )3/2 (m red T ) 3/2 ∫α2(2π) 3¯h 6 d 3 r U(r)= ln Z 0 − 1 α 2 V T 2 m 3 ∫2 (2π) 3¯h 6 d 3 r U(r)= ln Z 0 − 4πU 0 a 3 α2 V T 2 m 3(2π) 3¯h 6 .Here, M = 2m is the total mass, and m red = m/2 is the reduced mass, and α was defined as(2j + 1)e µ/T . The preceeding factors of 1/2 correct for double-counting. It is good practice tocheck the dimensi<strong>on</strong>s after a calculati<strong>on</strong>s such as this. In this case, all the units cancel as ln Z isdimensi<strong>on</strong>less.Although this example may have seen somewhat lengthy, it is the simplest example of a calculati<strong>on</strong>that <strong>on</strong>e can do in perturbati<strong>on</strong> theory at finite T . When performing higher-order calculati<strong>on</strong>s,or when including Bose and Fermi effects, calculati<strong>on</strong>s are inherently more difficult. Such calculati<strong>on</strong>swill not be c<strong>on</strong>sidered here, and are more appropriate for a course dedicated to many-bodytheory. However, even if <strong>on</strong>e does not intend to ever perform such calculati<strong>on</strong>s, it is very worthwhileto have an understanding of the fundamentals of perturbati<strong>on</strong> theory shown here:• Perturbati<strong>on</strong> theory is an expansi<strong>on</strong> in the order of interacti<strong>on</strong>s.• Calculati<strong>on</strong>s involve matrix elements of the type ⟨p 1 , p 2 |V |p 1 , p 2 ⟩ (or perhaps with differentnumbers of particles if interacti<strong>on</strong>s create or destroy quanta), which can be calculated easily.• Diagrams, such as the <strong>on</strong>e shown in Fig. 3.4 allow <strong>on</strong>e to separate c<strong>on</strong>tributi<strong>on</strong>s to ln Z bywhether they involve the interacti<strong>on</strong> of 1 particle, 2 particles · · · n particles in a “c<strong>on</strong>nected”manner.3.5 Virial Expansi<strong>on</strong>The perturbative expansi<strong>on</strong> of ln Z can be c<strong>on</strong>sidered as a sum of c<strong>on</strong>nected diagrams, each ofwhich involves a specific number n of incoming particles. The example illustrated in Fig. 3.4has four incoming particles. The number of incoming particles should not be c<strong>on</strong>fused with thethe perturbative order, i.e., the number of vertices in the diagram, which in this case was five. Forinteracti<strong>on</strong>s that c<strong>on</strong>serve the number of quanta, each diagram will depend <strong>on</strong> the chemical potentialas e nβµ , and since βµ → −∞ in the low density limit, expanding in the density is equivalent to43


expanding in n. The partiti<strong>on</strong> functi<strong>on</strong>, or equivalently the pressure, can then be calculated orderby-orderas:P/T = B 1 e βµ + B 2 e 2βµ + · · · B n e nβµ + · · · (3.33)Furthermore, since the density is ∂P/∂µ, <strong>on</strong>e can write a similar expansi<strong>on</strong> for the density,ρ = B 1 e βµ + 2B 2 e 2βµ + · · · nB n e nβµ + · · · (3.34)The virial expansi<strong>on</strong> is defined as an expansi<strong>on</strong> of the pressure in powers of the density,[∞∑( ) ] n−1 ∫ρ (2j + 1)P = ρT A 1 + A n , ρ 0 ≡ B 1 = d 3 p e −ϵp/T . (3.35)ρ 0 (2π¯h) 3n=2By inspecti<strong>on</strong>, <strong>on</strong>e can determine the first two virial coefficients, A 1 and A 2 , in terms of the coefficientsB i ,A 1 = 1, A 2 = −B 2 /ρ 0 . (3.36)The fact that A 2 and B 2 have opposite signs is easy to understand. An attractive interacti<strong>on</strong> willincrease both the density and the pressure by increasing the Boltzmann factor which behaves ase −V/T . Since the density increase more due to the extra powers of n in Eq. (3.34), the ratio of P/ρwill fall for attractive interacti<strong>on</strong>s. Subsequent coefficients are more difficult to calculate. The nexttwo are,A 3 = −2B 3 /ρ 0 + 4B 2 2/ρ 2 0, A 4 = −3B 4 /ρ 0 + 18B 2 B 3 /ρ 2 0 − 20B 3 2/ρ 3 0. (3.37)In Mayer’s cluster theory, <strong>on</strong>e can make a further reducti<strong>on</strong> and show that the complex termsin A 3 , A 4 · · · that include products of several powers of B can all be neglected if <strong>on</strong>e <strong>on</strong>ly uses“multiply c<strong>on</strong>nected” graphs (See Pathria).Like perturbati<strong>on</strong> theory, virial expansi<strong>on</strong>s are <strong>on</strong>ly justifiable in the limit that they are unimportant,i.e., either <strong>on</strong>ly the first few terms matter, or the expansi<strong>on</strong> does not c<strong>on</strong>verge. For thatreas<strong>on</strong>, the sec<strong>on</strong>d-order coefficient is often calculated carefully from first principles, but subsequentcoefficients are often inserted phenomenologically.3.6 Virial Coefficients from Phase ShiftsQuantum scattering theory can provide a simple means for calculating the sec<strong>on</strong>d virial coefficientin terms of phase shifts. In scattering theory, the relative wave functi<strong>on</strong> is a distorted plane wave,where the plane wave e iq·r is written as an expansi<strong>on</strong> in partial waves,e iq·r = 1qr∑i l (2l + 1)ϕ l (r)P l (cos θ), (3.38)leach of which is an eigenstate of angular momentum. Each partial wave can then be written as anincoming piece plus an outgoing piece. The effect of a spherically symmetric potential is to shiftthe phase of the outgoing part, leading to the following form at large r,ϕ l (r → ∞) ∼ eiδ lqr sin[qr/¯h − lπ + δ l(q)]. (3.39)44


If <strong>on</strong>e c<strong>on</strong>siders the relative wave functi<strong>on</strong> to be c<strong>on</strong>fined to a spherical volume of radius R by aninfinite potential, the boundary c<strong>on</strong>diti<strong>on</strong> is:The density of states is then,dndq=(2l + 1)πqR¯h − lπ + δ l = nπ, n = 1, 2, · · · (3.40)(R/¯h + dδ )l,dqdndϵ=(2l + 1)π(Rm/q¯h + dδ )ldϵAs compared to no interacti<strong>on</strong>, the density of state is changed by an amount∆ dndϵ(2l + 1) dδ l=π dϵ(3.41)(3.42)The most transparent way to calculate the correcti<strong>on</strong> to the partiti<strong>on</strong> functi<strong>on</strong> is to c<strong>on</strong>siderthe additi<strong>on</strong> to the density of states as a new mass state. If the phase shift jumped by π in asmall energy level with relative energy ϵ ∗ (which is what happens with a narrow res<strong>on</strong>ance), the<strong>statistical</strong> effect would be identical to having a free particle with spin l and mass M = m a + m band an additi<strong>on</strong>al Boltzmann weight e −ϵ∗ /T . From that perspective, <strong>on</strong>e can write the correcti<strong>on</strong> toln Z asln Z = ln Z 0 + V ∫d 3 P e ∑ −P 2 /2MT +2µ/Te −ϵ∗ /T(3.43)(2π¯h) 3 ϵ ∗ln Z = ln Z 0 + V(2π¯h) 3 ∫2µ/T V (MT )3/2= ln Z 0 + e(2π) 3/2¯h 3d 3 P e ∑ −P 2 /2MT +2µ/Tl∑∫dϵl∫dϵ(2l + 1) dδ lπ dϵ e−ϵ/T .(2l + 1) dδ lπ dϵ e−ϵ/TBy comparis<strong>on</strong> with the definiti<strong>on</strong> of the virial expansi<strong>on</strong> coefficients,A 2 = −2 ∑ ∫3/2 (2l + 1) dδ ldϵπ dϵ e−ϵ/T . (3.44)lFor bos<strong>on</strong>s, the wave functi<strong>on</strong> is symmetrized, which eliminates any odd values for l in thepartial wave expansi<strong>on</strong>. It also doubles the c<strong>on</strong>tributi<strong>on</strong> for even values of l. This adds an extrapower of 2 to the above expressi<strong>on</strong>, and the sum would cover <strong>on</strong>ly even l. For fermi<strong>on</strong>s, a factor of2 is also added, but the sum covers <strong>on</strong>ly odd l. In the real world, particles have spins, and since<strong>on</strong>ly pairs of the same spin projecti<strong>on</strong> are identical, both odd and even values of l c<strong>on</strong>tribute. Anexample is nn scattering. The spin half neutr<strong>on</strong>s can have their intrinsic spins combined as either aS = 1 triplet or a S = 0 singlet for which the spin comp<strong>on</strong>ent of the wave functi<strong>on</strong>s are symmetricor antisymmetric respectively. The spatial wave functi<strong>on</strong>s are then antisymmetric and symmetricrespectively. Phase shifts are measured for both the triplet (<strong>on</strong>ly odd l) and for the singlet (<strong>on</strong>lyeven l). In this case the virial coefficient becomes:A 2 = −2 5/2 { ∑l=0,2,4...∫dϵ(2l + 1)πdδ l,S=1dϵe −ϵ/T + 345∑l=1,3,5...∫dϵ(2l + 1)π}dδ l,S=0e −ϵ/T . (3.45)dϵ


The real situati<strong>on</strong> is a bit more complicated if <strong>on</strong>e wants to accurately address the S = 1 terms.First, due to spin-orbit coupling, m l and m S are not good quantum numbers, and <strong>on</strong>e needs toadd the quantum number J to label the phase shifts. Sec<strong>on</strong>dly, the tensor interacti<strong>on</strong> mixes stateswith different l if they have the same J (l must differ by even number to c<strong>on</strong>serve parity). Thesecouplings are typical causes for headaches am<strong>on</strong>g nuclear physicists.EXAMPLE:The scattering length a parameterizes the phase shift at low energy as:δ l=0 = −ap/¯h.For neutr<strong>on</strong>s, a ≈ −2 × 10 −14 m. C<strong>on</strong>sider a neutr<strong>on</strong> gas at a temperature of T = 4 × 10 −21 Joules(room temperature), at what density will the correcti<strong>on</strong> to the pressure in a sec<strong>on</strong>d-order virialexpansi<strong>on</strong> be of the same strength as the ρT . DATA: ¯h = 1.05 × 10 −34 J·s, m = 1.67 × 10 −27 kg.Ignore the spin when making the estimate.First lets calculate the virial coefficent using Eq. (3.44),∫A 2 = −2 3/2= 2 3/2 a ∫π¯hdϵ 1 π∫dδdϵ e−ϵ/T = −2 3/2dp e −p2 /2mT = 2 3/2 ā hNow, the first and sec<strong>on</strong>d terms will be of the same order when,Solving for ρ,−A 2 ρ = ρ 0 =(mT )3/2(2π) 3/2¯h 3 ≈ 1030 m −3 .ρ ≈ 2 × 10 33 .dδdp e−ϵ/Tdp 1 π√mT2π ≈ −6 × 10−4 .The corresp<strong>on</strong>ding mass density, 3000 g/cm 3 , is hundreds of times higher than normal matterdensities, but is very low compared to the density of the nucleus, ≈ 2 × 10 15 g/cm 3 . A realisticcalculati<strong>on</strong> would have to account for the fact the neutr<strong>on</strong>s are spin-half objects, and also includethe effects of quantum degeneracy.The phase shift expressi<strong>on</strong>s discussed here are not applicable <strong>on</strong>ce Coulomb is introduced intothe problem. Coulomb forces are l<strong>on</strong>g range, hence the scattered waves never fully emerge from theinfluence of the force.3.7 An Aside: A Brief Review of Creati<strong>on</strong> and Destructi<strong>on</strong> OperatorsWhen phase space densities are low, the <strong>on</strong>ly matrix elements that matter are those c<strong>on</strong>necting anempty single-particle level to <strong>on</strong>e with a singly-occupied level. For higher occupati<strong>on</strong>s, <strong>on</strong>e mustworry about the more general case, ⟨m|V |n⟩, where m and n refer to the number of particles in thelevel. Although the m and n dependence could be nearly any functi<strong>on</strong>, in nearly all cases it willfollow the systematics of creati<strong>on</strong> and destructi<strong>on</strong> operators. The systematics plays a central rolein the study of c<strong>on</strong>densates and lasers, or in lattice-gauge calculati<strong>on</strong>s.46


Creati<strong>on</strong> and destructi<strong>on</strong> operators are built <strong>on</strong> the physics of a simple <strong>on</strong>e-dimensi<strong>on</strong>al harm<strong>on</strong>icoscillator:H = p22m + 1 2 mω2 x 2 . (3.46)Here, the spring c<strong>on</strong>stant is k = mω 2 , where ω = √ k/m is the angular frequency of the oscillator.Creati<strong>on</strong> and destructi<strong>on</strong> operators are defined as:a † =√ mω2¯h x + i √12¯hmω p,√ √mω 1a = 2¯h x − i p. (3.47)2¯hmωWith this definiti<strong>on</strong>,[a, a † ] = 1, H = ¯hω(a † a + 1/2). (3.48)Here, a is the destructi<strong>on</strong> operator while a † is the creati<strong>on</strong> operator.Now, we c<strong>on</strong>sider some eigenstate |m⟩ which has an energy ϵ m ,The state a|m⟩ must then also be an eigenstate,H|m⟩ = ϵ m |m⟩. (3.49)H(a|m⟩) = ¯hω [ (a † a + 1/2) ] a|m⟩ = ¯hωa [ a † a − 1/2 ] |m⟩ = (ϵ m − ¯hω)(a|m⟩). (3.50)Thus, for any eigenstate, successive operati<strong>on</strong>s of the destructi<strong>on</strong> operator will yield more eigenstateswith energies lowered by ¯hω – unless there is some state |0⟩ for whichIf we assume there is such a state, its energy must bea|0⟩ = 0. (3.51)H|0⟩ = ¯hω(a † a + 1/2)|0⟩ = (¯hω/2)|0⟩. (3.52)One could have also c<strong>on</strong>sidered the state a † |m⟩. Similarly, applicati<strong>on</strong> of the commutati<strong>on</strong>relati<strong>on</strong>s leads to fact thatH(a † |m⟩) = (ϵ m + ¯hω)(a † |m⟩). (3.53)Thus, the eigen-states of the Hamilt<strong>on</strong>ian are:ϵ = (1/2, 3/2, 5/2 · · · )¯hω. (3.54)Thus, we have shown that the basis of eigenstates are the vacuum, plus some states (a † ) m |0⟩.However, these states are not yet normalized. To normalize them, we calculate the norm by applyingthe commutati<strong>on</strong> operatorsThe normalized eigenstates are thus:⟨0|a m (a † ) m |0⟩ = m⟨0|a m−1 (a † ) m−1 |0⟩ = m! . (3.55)|n⟩ = 1 √n!(a † ) n |0⟩ (3.56)47


In quantum field theory, there are field operators for each point in space x. Each of these pointsmight c<strong>on</strong>tain 0, 1, 2 · · · particles. Whereas the Schrödinger equati<strong>on</strong> involves wave functi<strong>on</strong>s, ψ(x)and ψ ∗ (x), field operators, Ψ(x) and Ψ † (x), are actually creati<strong>on</strong> and destructi<strong>on</strong> operators,[Ψ(x), Ψ † (x ′ )] = δ(x − x ′ ). (3.57)The fact that quantum field theories behave in this manner is <strong>on</strong>e of the most profound aspectsof modern quantum theory. It is as if each point in space is its own harm<strong>on</strong>ic oscillator. Thekinetic energy term then acts like a term coupling adjacent oscillators, so that the eigenstates ofthe Hamilt<strong>on</strong>ian are also eigenstates of momentum. One can then associate a harm<strong>on</strong>ic oscillatorof frequency ¯hω = ϵ p for each value of the momentum p.EXAMPLES:• Calculate ⟨0|aaa † aa † a † |0⟩Commute all the creati<strong>on</strong> operators to the right:⟨0|aaa † aa † a † |0⟩ = ⟨0|a 3 (a † ) 3 |0⟩ − ⟨0|a 2 (a † ) 2 |0⟩ = 3! − 2! = 4.• Calculate ⟨n|a † a † a † a|m⟩.Commute the destructi<strong>on</strong> operator to the left,⟨n|a † a † a † a|m⟩ = ⟨n|a(a † ) 3 |m⟩ − 3⟨n|(a † ) 2 |m⟩=1 {√ ⟨0|a (n+1) (a † ) m+3 |0⟩ − 3⟨0|a n (a † ) m+2 |0⟩ }n!m!= δ n,m+21√n!m!{(n + 1)! − 3n!}= δ n,m+2 (n − 2) √ n(n − 1).• The wave functi<strong>on</strong> for the ground state of the harm<strong>on</strong>ic oscillator is:ψ 0 (x) =1(2πR 2 ) 1/2 e−x2 /4R 2 ,R = ¯h/(2mω).Find ψ 1 (x), the wave functi<strong>on</strong> of the first excited state by operating a † <strong>on</strong> ψ 0 (x).Here, simply use the fact that the creati<strong>on</strong> operator is:a † =√ mω2¯h x − i √12¯hmω (−i∂ x).Differentiating,ψ 1 (x) =√2mω¯h xψ 0(x) = (x/R)ψ 0 .48


3.8 The Partiti<strong>on</strong> Functi<strong>on</strong> as a Path IntegralFor interacti<strong>on</strong>s that defy perturbati<strong>on</strong> theory, several n<strong>on</strong>-perturbative methods are based <strong>on</strong> thepath-integral picture of the partiti<strong>on</strong> functi<strong>on</strong>:Z = ∑ ⟨i|e −βH |i⟩ =∑⟨i 1 |e −δβH |i 2 ⟩⟨i 2 |e −δβH |i 3 ⟩⟨i 3 | · · · |i n ⟩⟨i n |e −δβH |i 1 ⟩. (3.58)ii 1 ,i 2 ,··· ,i nHere, δβ = β/n, and completeness was used to insert the intermediate states. If n → ∞, δβ → 0and the exp<strong>on</strong>ential becomes 1 − δβH, which is linear in H. The quantity is c<strong>on</strong>sidered to be apath integral because each set i 1 · i n can be c<strong>on</strong>sidered as a quantum path.In various forms of the nuclear shell model, a subset of all states is c<strong>on</strong>sidered in the sum overi. The quantity M = (1 − δβH) is treated as a matrix, and the path integral is simple the trace ofthis matrix to the n th power. Since states in the trace are many-body states, even the three-particle- three-hole excitati<strong>on</strong>s of a few single-particle levels can amount to milli<strong>on</strong>s of many-body states.N<strong>on</strong>etheless, such brute-force methods are often applied. The T = 0 limit is reached by havingβ → ∞ which requires many applicati<strong>on</strong>s of M. In fact, this is related to as the Lanczos methodfor finding the ground state of a matrix Hamilt<strong>on</strong>ian.An even more brute-force method for calculating the path integral is lattice gauge theory. Inlattice calculati<strong>on</strong>s, the states |i⟩ refer to c<strong>on</strong>figurati<strong>on</strong>s of the relativistic quantum field operatorsat each point in space. In order to make the calculati<strong>on</strong>s tenable, the c<strong>on</strong>tinuum of space pointsis divided up into a lattice. The finite steps in imaginary time, δβ, represent a discretizati<strong>on</strong> in afourth dimensi<strong>on</strong>.One possibility for describing the states |i⟩ would be to c<strong>on</strong>sider each point <strong>on</strong> the lattice as it’sown harm<strong>on</strong>ic oscillator basis, then the state i would have n(i x , i y , i z ) excitati<strong>on</strong>s at each latticepoint. However, this is not how lattice calculati<strong>on</strong>s are performed. Instead, the state i refers to aset of complex field values at each point, η ix,i y,i z, where η is a complex number describing the state:|⃗η⟩ = exp ( ⃗η · ⃗a † − ⃗η ∗ · ⃗a ) |0⟩ (3.59)= e −|η|2 /2 exp ( ⃗η · ⃗a †) |0⟩Here, ⃗a is short-hand for a 1 , a 2 · · · where the comp<strong>on</strong>ents refer to different single-particle levels.The last step above exploited the Campbell-Baker-Hausdorff lemma,e A+B = e A e B e −[A,B]/2 (3.60)The states |⃗η⟩ are c<strong>on</strong>venient because they are normalized eigenstates of the destructi<strong>on</strong> operators.For instance,⃗a|⃗η⟩ = ⃗η|⃗η⟩. (3.61)In order to use this basis in the path-integral formulati<strong>on</strong>, the states must satisfy completeness,which in this case is (homework problem),∫1dη r dη i |η⟩⟨η| = 1, (3.62)πwhere η r and η i are the real and imaginary parts of η. Thus, each step in imaginary time can bec<strong>on</strong>sidered a set of values of η, and the path can be c<strong>on</strong>sidered as a path through these sets ofnumbers.49


EXAMPLE:The state |η⟩ is not orthog<strong>on</strong>al to |η ′ ⟩. Find the overlap, ⟨η ′ |η⟩.Expanding the exp<strong>on</strong>entials in the bra,⟨η ′ |η⟩ = e −|η′ | 2 /2 ⟨0|e −η′∗a |η⟩= e −|η′ | 2 /2 ⟨0|e η′∗η |η⟩.Expanding the ket and keeping <strong>on</strong>ly the first term, since the others d<strong>on</strong>’t overlap with ⟨0|,⟨η ′ |η⟩ = exp{−|η ′ | 2 /2 − |η| 2 /2 + η ′∗ η}Since the kinetic terms in the Hamilt<strong>on</strong>ians have gradients, the dominant paths are those where ηis smooth. Rather than randomly sampling all paths, lattice calculati<strong>on</strong>s use Metropolis algorithmswhich c<strong>on</strong>sider small changes of the path, then keep or reject the new path based <strong>on</strong> the relative<strong>statistical</strong> weight of the new and old paths.In order to perform the summati<strong>on</strong>s in Eq. (3.58) <strong>on</strong>e c<strong>on</strong>siders element of the type,⟨⃗η n+1 |e −δβH(⃗a† ,⃗a) |⃗η n ⟩ = ⟨⃗η n+1 |e −δβH(⃗η∗ n+1 ,⃗ηn) |⃗η n ⟩, (3.63)which assume that the Hamilt<strong>on</strong>ian is normal ordered with all the creati<strong>on</strong> operators pushed to theleft. One difficulty with these elements is that the creati<strong>on</strong> and destructi<strong>on</strong> operators are evaluatedfor different η. To overcome this, <strong>on</strong>e can use Eq. (3.59) to rewrite,into the the matrix element to obtain|⃗η n ⟩ = e −[δ⃗η∗ ⃗η+⃗η ∗ δ⃗η]/2 e δ⃗η·⃗a† |⃗η n+1 ⟩ (where δη ≡ η n − η n+1 ), (3.64)⟨⃗η n+1 |e −δβH |⃗η n ⟩ = ⟨⃗η n+1 |e −δβH e [δ⃗η∗·⃗η−⃗η ∗·δ⃗η]/2 |⃗η n+1 ⟩. (3.65)Here, we have assumed δβ and δη are small. After referring to δ⃗η = ˙⃗ηδβ, <strong>on</strong>e obtains the relati<strong>on</strong>,⟨⃗η n+1 |e −δβH |⃗η n ⟩ = ⟨⃗η n+1 |e δβ[ ˙⃗η ∗ η/2−⃗η ∗ ˙⃗η/2−H] |⃗η n ⟩. (3.66)One can now rewrite the path integral in Eq. (3.58) asZ = 1 ∏{∫}d⃗pπ N i d⃗q i ⟨⃗p i , ⃗q i |e δβ[i⃗p i· ˙⃗q i /2−i ˙⃗p i·⃗q i /2−H(⃗p i ,⃗q i )] |⃗p i , ⃗q i ⟩ , (3.67)iwhere N is the number of points (Dimensi<strong>on</strong> of p times number of time steps) over which fields areintegrated. Here, ⃗p and ⃗q are the real and imaginary part of ⃗η (aside from a factor of √ 2),η ≡ p + iq √2. (3.68)Calculati<strong>on</strong> of Z thus entails dividing β into an infinite number of steps δβ, then integrating overall ⃗p and ⃗q for each step, with the boundary c<strong>on</strong>diti<strong>on</strong>s, ⃗p(β) = ⃗p(0), ⃗q(β) = ⃗q(0). This is a pathintegral and is often written as:Z = 1 ∫ { }∏{∫ β ( ) }d⃗p(βπ N i )d⃗q(β i ) exp dτL ˙⃗p(τ), ˙⃗q(τ), ⃗p(τ), ⃗q(τ) . (3.69)0i50


Here, L = i⃗p · ˙⃗q − H is the Lagrangian, with the factor of i coming from the fact that the time stepis in imaginary time. We also note that integrati<strong>on</strong> by parts and the boundary c<strong>on</strong>diti<strong>on</strong>s that thepath return to its original starting point made terms of the form (p ˙q − ṗq)/2 equivalent to p ˙q.Lattice gauge theory is useful for relativistic theories where the particle number is not a goodquantum number, and the η basis is as reas<strong>on</strong>able as any, and is especially useful for situati<strong>on</strong>swhere c<strong>on</strong>densed fields play a dominant role, such as QCD. Whereas the method described aboveapplies <strong>on</strong>ly for bos<strong>on</strong>ic fields (since the creati<strong>on</strong> operators follow bos<strong>on</strong>ic rules), the method, aftersome c<strong>on</strong>torti<strong>on</strong>s, can also be applied for fermi<strong>on</strong>s.Quantum chromodynamics is a n<strong>on</strong>-abelian relativistic field theory, which is very str<strong>on</strong>gly interactingand cannot be treated in perturbati<strong>on</strong> theory. It is also characterized by c<strong>on</strong>densedquark-antiquark pairs and by a c<strong>on</strong>densate of glu<strong>on</strong>ic fields. At temperatures in the neighborhoodof 170 MeV, the state of the system changes rather abruptly. Quarks no l<strong>on</strong>ger bel<strong>on</strong>g towell defined color singlets (color dec<strong>on</strong>finement) and the background c<strong>on</strong>densate of quark-antiquarkpairs melts (restorati<strong>on</strong> of chiral symmetry). Lattice gauge calculati<strong>on</strong> provide the <strong>on</strong>ly means to“solve” such a str<strong>on</strong>gly interacting theory. A large community is devoted to precisely such calculati<strong>on</strong>s.These calculati<strong>on</strong>s involve large and sometimes very specialized computati<strong>on</strong>al facilities.Calculati<strong>on</strong>s typically exceed 10 20 floating point operati<strong>on</strong>s.3.9 Problems1. C<strong>on</strong>sider a low density three-dimensi<strong>on</strong>al gas of n<strong>on</strong>-relativistic spin-zero bos<strong>on</strong>s of mass mat temperature T = 1/β and chemical potential µ.(a) Find ρ 0 as defined in Eq. (3.1) in terms of m and T .(b) Expand the density ρ to sec<strong>on</strong>d order in e βµ , i.e., to e 2βµ . Express your answers for thispart and the next two parts in terms of ρ 0 .(c) Expand ρ 2 to sec<strong>on</strong>d order in e βµ .(d) Expand δP ≡ P − ρT to sec<strong>on</strong>d order in e βµ . (Hint: it is easier if you use the expressi<strong>on</strong>for P expanded in f 0 = e βµ−βϵ , i.e., ln(1 + f 0 + f 2 0 + f 3 0 · · · ) = − ln(1 − f 0 ), then expandthe logarithm in powers of f 0 )(e) Determine the sec<strong>on</strong>d virial coefficient defined by Eq. (3.1).2. C<strong>on</strong>sider the Van der Waals equati<strong>on</strong> of state in scaled variables,where p = P/aρ 2 s, v = V/V s , t = T/aρ s .p =tv − 1 − 1 v 2 ,(a) Derive the Maxwell relati<strong>on</strong>,∂(P/T )∂β∣ = − ∂EN,V∂V ∣ .N,T(b) Find the scaled energy per particle e ≡ E/(aρ s N) as a functi<strong>on</strong> of v and t using theMaxwell relati<strong>on</strong> above. Begin with the fact that e = (3/2)t as v → ∞.51


(c) Show that the change of entropy/particle s = S/N between two values of v at a fixedtemperature t is:s b − s a = ln[(v b − 1)/(v a − 1)].(d) Using the fact that ts = e + pv − µ, show thatµ b − µ a = − 2 + 2 [vb+ tv b v a v b − 1 −v ]a− t lnv a − 1( )vb − 1.v a − 1(e) Show that as t → 0, p b will equal p a if v b → ∞ and v a = 1 + t. Then, show that in thesame limit, µ a will equal µ b if v b = te 1/t .(f) Find the latent heat L = t(s b − s a ) for the small t limit. How does it compare with theminimum of e at t = 0?(g) At t = 0, the system will have p = 0 in order to minimize the energy.Clausius-Clapeyr<strong>on</strong> equati<strong>on</strong>, find dp/dt al<strong>on</strong>g the coexistence line at t = 0.Using the3. Calculate the sec<strong>on</strong>d-order virial coefficient to first order in perturbati<strong>on</strong> theory for distinguishableparticles of mass m at temperature T interacting through the two-body potential,{ 0 , r < aV (r) =V 0 /r 6 , r > a4. C<strong>on</strong>sider a gas of distinguishable particles of mass m at temperature T and chemical potentialµ in a volume V interacting through a mean field,V = V 0 .(a) Using Eq. (3.30) calculate the correcti<strong>on</strong> to ln Z to first order in perturbati<strong>on</strong> theory.(b) Calculate the m th order correcti<strong>on</strong>.(c) Sum all the correcti<strong>on</strong>s to see thatln Z =V ∫(2π¯h) 3d 3 p e −(ϵ p+V 0 −µ)/T .5. Using Eq. (3.44), calculate the sec<strong>on</strong>d-order virial coefficient for a gas of distinguishable n<strong>on</strong>relativisticparticles of mass m at temperature T that interact through a hard core potential,{ ∞ , r < aV (r) =0 , r > aC<strong>on</strong>sider <strong>on</strong>ly the s-wave c<strong>on</strong>tributi<strong>on</strong> (valid at low T ).6. C<strong>on</strong>sider the state|η⟩ ≡ exp ( ηa † − η ∗ a ) |0⟩,where a † and a are creati<strong>on</strong>/destructi<strong>on</strong> operators, [a, a † ] = 1.(a) Show that |η⟩ is an eigenstate of the destructi<strong>on</strong> operator,a|η⟩ = η|η⟩.52


(b) Show that |η⟩ obey the relati<strong>on</strong>s,⟨η ′ |η⟩ = exp{η ′∗ η − [|η| 2 + |η ′ | 2 ]/2}.(c) Show that the completeness relati<strong>on</strong> is satisfied,∫1dη r dη i ⟨n 1 |η⟩⟨η|n 2 ⟩ = δ n1 ,nπ2,where |n⟩ is a normalized state with n particles, |n⟩ = (1/ √ n!)(a † ) n |0⟩.7. C<strong>on</strong>sider a Hamilt<strong>on</strong>ian,where j(t) is a real functi<strong>on</strong>.H = H 0 + j(t)(a + a † ), H 0 = ¯hωa † a,(a) Show thate −iH 0t/¯h a = ae −iH 0t/¯h e iωt .(b) Show that the state{ ∫ t]|ψ(t)⟩ ≡ e −iH0t/¯h exp −(i/¯h) dt ′ j(t ′ )[e }−iωt′ a + e iωt′ a †−∞is an eigenstate of the Hamilt<strong>on</strong>ian, i.e., showi¯h ∂ |ψ(t)⟩ = H|ψ⟩.∂t(c) Assuming that j(t) = 0 for t > t ′ , find an expressi<strong>on</strong> for η in terms of j such that|ψ(t > t ′ )⟩ is a coherent state,|ψ(t)⟩ = e −iH 0t/¯h exp { ηa † − η ∗ a } |0⟩.53


4 Dynamics of Liquids and GasesVisi<strong>on</strong> without acti<strong>on</strong> is a daydream. Acti<strong>on</strong> without visi<strong>on</strong> is a nightmare. - Japanese proverb4.1 “Adiabatic” Expansi<strong>on</strong>sThe science of thermodynamics was at <strong>on</strong>e time centered around engines. In particular, the c<strong>on</strong>straints<strong>on</strong> the efficiency of engines related to entropy are of both intellectual and practical interest.One begins withdE = T dS − P dV + µdN, (4.1)where the term T dS is referred to as the heat. Heat is not an energy, but is a change in energy. Oneshould not refer to the amount of heat in a gas or liquid – <strong>on</strong>ly the amount of heat added or removedfrom a gas or liquid. Adding energy without changing the particle number or the volume will raisethe entropy. The most comm<strong>on</strong> method for adding heat is through thermal c<strong>on</strong>tact such as themicroscopic transfer of energy between the molecules in a cylinder and the sides of the cylinder.An insulated gas is <strong>on</strong>e for which there is no change in the energy aside from the work d<strong>on</strong>e <strong>on</strong>the gas. Thus, for an insulated gas, Eq. (4.1) implies that an insulated gas is equivalent to fixing theentropy and particle number. However, this is clearly not the case if the volume changes suddenly.If a pist<strong>on</strong> is instantly moved so that the volume of a cylinder is doubled, the gas molecules will nothave time to change their energy, and dE would be zero. In this case, the entropy of each molecule,which is proporti<strong>on</strong>al to ln[(mT ) 3/2 V ], would increase by a factor of ln 2 due to the increase involume. Thus, an insulated system could have a change in entropy if the boundaries are suddenlychanged. In fact, Eq. (4.1) assumes that the system changes slowly. Otherwise, the gas will notdo the work P dV . Here, “slowly” requires that the boundaries move much more slowly than themolecular velocity of the gas.One must remain cognizant of the meaning of several terms, all of which are often thrown aboutrather casually, with complete disregard for the c<strong>on</strong>fusi<strong>on</strong> that might ensue from the carelessness.Some of the terms are:1. isentropic: Entropy is c<strong>on</strong>served, ∆S = 0.2. insulated: No energy is transferred from thermal c<strong>on</strong>tact with the outside.3. adiabatic: Different meanings to different people, and sometimes different meanings to thesame pers<strong>on</strong> depending <strong>on</strong> the c<strong>on</strong>text. Sometimes means “slowly”, sometimes means ∆S = 0.4. reversible: Usually means ∆S = 0.The slowness c<strong>on</strong>siderati<strong>on</strong>s of the previous paragraph are the source of much c<strong>on</strong>fusi<strong>on</strong> relatedto the definiti<strong>on</strong> of the word “adiabatic”. In many instances (especially to engineers), adiabaticmeans that no heat enters a gas, ∆S = 0. Whereas for many problems, adiabatic means thatthings change slowly. For instance, if reading a quantum <strong>mechanics</strong> book, <strong>on</strong>e might encounterthe example of a particle in the ground state of a square well that is expanded slowly. The textmight simply say that the well expands adiabatically. This might be c<strong>on</strong>trasted with the suddenexpansi<strong>on</strong> of the well, and a questi<strong>on</strong> might be posed regarding the probability that the particleremain in the ground state after the well changed in both circumstances. For the particle in thewell, it will remain in the ground state in a slow (adiabatic) expansi<strong>on</strong>, but will find its probabilitysplit between many states in the sudden expansi<strong>on</strong>. In the adiabatic case, the entropy remains54


zero since the system is in <strong>on</strong>e state, whereas the entropy becomes finite for the sudden expansi<strong>on</strong>.Thus, in this case slowness of the expansi<strong>on</strong> and c<strong>on</strong>serving entropy are syn<strong>on</strong>ymous and the term“adiabatic” is not ambiguous. However, this is not always the case, as sometimes the reader mustpay careful attenti<strong>on</strong> to discern whether an author is referring to zero change in entropy vs. a slowchange of parameters. To make the situati<strong>on</strong> worse, some introductory physics books describe anadiabatic expansi<strong>on</strong> as being rapid. The intent of such a statement is that in a rapid expansi<strong>on</strong>,heat has no time to leak out of the c<strong>on</strong>tainer, so that entropy is c<strong>on</strong>served. However, it must notbe rapid relative to the thermal velocity of the gas, or <strong>on</strong>e would create entropy because of the lackof slowness.EXAMPLE:A gas of N a indistinguishable, n<strong>on</strong>-relativistic, spinless, n<strong>on</strong>-degenerate, weakly interacting ‘a’particles of mass m a are c<strong>on</strong>fined to half of an insulated box by a barrier. Each half has a volumeV . The other half of the box is a vacuum. The particles are thermalized at a temperature T whenthe barrier is removed suddenly.• What is the average energy and entropy per particle before the barrier is removed?The energy per particle for a n<strong>on</strong>-degenerate n<strong>on</strong>-relativistic gas is (3/2)T . The entropy perparticle is:S/N = (ln Z)/N + βE/NZ = 1 ( ∫) N Vd 3 p e −p2 /2mT= 1 ( ) 3N/2 mTN! (2π¯h) 3 N! V N 2π¯h 2 ,where Z, E and S are the partiti<strong>on</strong> functi<strong>on</strong>, average energy and entropy. Using E/N =(3/2)T and ln(N!) ≈ N ln N − N,S/N = 5 ( ) ] 3/2 mT[(V/N)2 + ln 2π¯h 2 .• What is the average energy and entropy per particle after the barrier is removed and theparticles fill the entire volume?The energy is the same and the entropy per particle increases by an amount ln 2 due to theincrease in volume.• The barrier is instead slowly pushed to the right like a pist<strong>on</strong>. What is the average energyand entropy per particle be after the particles fill the entire volume?In this case the entropy is fixed. If V/N doubles, T 3/2 must decrease by a factor of 2, whichreduces T and E/N by a factor of 2 −2/3 .• The experiment is changed so that a separate type of ‘b’ particles populate the empty halfof the box before the barrier is introduced. The particles are at the same initial density andtemperature, and the masses and spins of the two types of particles are identical. What isthe average energy and entropy per particle after the partiti<strong>on</strong> is suddenly removed and thegases have thoroughly mixed?The volume that plays a role in the entropy is the volume per identical particle. Though thedensity is the same, the volume per identical particle has doubled, thus increasing the entropyper particle by an amount ln 2. The energy per particle is unchanged.55


4.2 Molecular Excitati<strong>on</strong>s and ThermodynamicsAtoms can be excited by moving electr<strong>on</strong>s into higher-lying states. Molecules also have intrinsicexcitati<strong>on</strong>s due to vibrati<strong>on</strong>, rotati<strong>on</strong> or any other internal degrees of freedom. Nuclei also havea rich excitati<strong>on</strong> spectrum, including single-particle-like excitati<strong>on</strong>s as well as collective rotati<strong>on</strong>aland vibrati<strong>on</strong>al states. These states greatly affect the thermodynamic properties of the matter. Fora single atom, the average internal energy per particle will be:∑l⟨E int /N⟩ =d lϵ l e −ϵ l/Tz intz int ≡ ∑ d l e −ϵl/T ,l= − ∂∂β ln z int, (4.2)where d l is the degeneracy of the energy level l.For a spectrum of <strong>on</strong>e-dimensi<strong>on</strong>al harm<strong>on</strong>ic oscillator levels separated by ¯hω,The average internal energy is then,z int = 1 + e −β¯hω + e −2β¯hω + e −3β¯hω · · · (4.3)=11 − e . −β¯hωe−β¯hω⟨E int /N⟩ = ¯hω , (4.4)1 − e−β¯hω which looks exactly like what we had for bos<strong>on</strong>s in specific energy level since those energies werealso evenly spaced, 0, ϵ, 2ϵ · · · .Taking the high and low T limits in Eq. (4.4),⟨E int /N⟩ ={¯hωe −¯hω/T , T > ¯hω(4.5)Thus, the excitati<strong>on</strong>s can be neglected if the temperature is much less than the energy required toexcite the level. If the temperature is much higher, the harm<strong>on</strong>ic excitati<strong>on</strong>s c<strong>on</strong>tribute just as <strong>on</strong>ewould expect from c<strong>on</strong>sidering the equipartiti<strong>on</strong> theorem result for a classical harm<strong>on</strong>ic oscillator.The equipartiti<strong>on</strong> theorem states that for each degree of freedom that appears quadraticallyin the Hamilt<strong>on</strong>ian, an amount of energy T/2 is added to the system. For a <strong>on</strong>e-dimensi<strong>on</strong>alharm<strong>on</strong>ic oscillator, both x and p x c<strong>on</strong>tribute T/2. Vibrati<strong>on</strong>al excitati<strong>on</strong>s of diatomic moleculestend to behave like a <strong>on</strong>e-dimensi<strong>on</strong>al harm<strong>on</strong>ic oscillator, c<strong>on</strong>tributing <strong>on</strong>e unit of T per particlefor temperatures much higher than ¯hω. For most diatomic molecules, ¯hω is thousands of degreesK, which means that vibrati<strong>on</strong>al excitati<strong>on</strong>s can be largely ignored at room temperature.Any rotating object whose positi<strong>on</strong> can be denoted by <strong>on</strong>e unit vector (or directi<strong>on</strong> θ and ϕ),and whose energy is independent of such energies, will have a rotati<strong>on</strong>al spectra,ϵ rot = ¯h2 l(l + 1), (4.6)2I56


where I is the moment of inertia of the molecule. Since the orientati<strong>on</strong> of a diatomic molecule isdetermined by a single unit vector, they will have such a spectra. Al<strong>on</strong>g with the degeneracy of(2l + 1), the internal partiti<strong>on</strong> functi<strong>on</strong> is given by the sum,z rot =∞∑(2l + 1) exp { −β¯h 2 l(l + 1)/2I } . (4.7)l=0For T > ¯h 2 /I, many terms in the sumc<strong>on</strong>tribute, which allows the sum to be approximated by an integral,∫z rot (T >> ¯h 2 /I) = 2 dl le −l2 /(2IT/¯h 2 )(4.8)The energy (= −∂ ln z/∂β) then becomes= 2IT/¯h 2 .⟨E rot /N⟩| T >>¯h 2 /I = T. (4.9)The characteristic energy ¯h 2 /2I is below room temperature for most molecules. Thus, for a dilutegas of diatomic molecules, the energy per particle is:⎧⟨ ⟩ E⎨ 3T/2, T


For the three limits above, C P = 5/2, 7/2, 9/2. If C P and C V are defined as T dS/dT , not perparticle, the relati<strong>on</strong> becomesC P = C V + N. (4.18)EXAMPLE:C<strong>on</strong>sider an isentropic expansi<strong>on</strong> for a dilute gas of O 2 molecules near room temperature, wherethe volume changes from V a to V b .• Find the final temperature T b in terms of the volumes and the original temperature T a .The fundamental thermodynamic relati<strong>on</strong> is:T dS = 0 = dE + P dV = N (C V dT + T dV/V ) ,with C V = 5/2. This becomes∫ Tb∫ Vb(5/2) dT/T = − dV/V,T a V aT b = T a(VbV a) −2/5.• Find the final pressure P b in terms of the original pressure P a and V a and V b .Using the ideal gas law, P = NT/V ,P b= T ( ) −7/5b V a Vb= .P a T a V b V aOften the ratio C P /C V is referred to as γ, which in this case is 7/5. The above equati<strong>on</strong> canthen be equivalently stated asP V γ | N,S= c<strong>on</strong>stant.4.3 Efficiency of Engines: The Third Law of ThermodynamicsThe third law of thermodynamics is nothing more than a statement that the entropy of the entiresystem must increase. For a closed amount of gas that is heated, where the gas does work as itexpands, there is a limit <strong>on</strong> the efficiency of the engine. Since the cycle returns to its original volumeand temperature, the net work d<strong>on</strong>e by the gas must equal the net heat added during the heatingpart of the cycle, Q H = W . Similarly, since the entropy must return to its original value, entropymust leave the system during part of the cycle to compensate for the entropy gained during thepart of the cycle where heat is added. The exhaust heat,Q X = T X ∆S, (4.19)where ∆S is determined by the heat added during the heating porti<strong>on</strong> of the cycle,T H ∆S = Q H . (4.20)58


Since the net ∆S in a closed cycle will be zero, if <strong>on</strong>e wishes to minimize the exhaust heat andmaximize the heat flowing in, <strong>on</strong>e should design a cycle so that the heat flows in while the cycle hasits highest temperature and is exhausted at the coldest porti<strong>on</strong> of the cycle. By having the entropyleave when the temperature is as low as possible Q X will be as small as possible, and by havingthe heat enter when T H is at its highest value, ∆S will be as small as possible, which also helpsminimize Q H . All other parts of the cycle should follow isentropic paths. Such a cycle is known asthe Carnot cycle, and has four parts as illustrated in Fig. 4.1.PcVFigure 4.1: Illustrati<strong>on</strong> of Carnot Cycle.(a-b) From a hot and compressed stage characterizedby T H , P a and V a , the system expands isothermallyto P b , and V b . Since the temperature is c<strong>on</strong>stant,P b V b = P a V a . Heat is added during this stage.(b-c) At fixed entropy, the system is cooled to a lowertemperature T X . For this segment, P c Vcγ = P b V γb .(c-d) The gas is compressed isothermally at temperatureT X until it returns to the original entropy, at whichpoint the pressure and volume are P d and V d . Heat isexhausted during this stage.(d-a) At fixed entropy, the gas is compressed back toits original temperature T H , pressure and volume. Toc<strong>on</strong>serve entropy, P d V γd= P aV γa .The efficiency of a Carnot engine is defined as the work, W = Q H − Q X , divided by the inputenergy Q H ,e = W Q H= Q H − Q XQ H. (4.21)Since entropy is <strong>on</strong>ly added at the points where heat is added, and since the entropy added at T Hequals the entropy lost at T X ,e = (T H − T X )∆ST H ∆S= 1 − T XT H. (4.22)The maximum efficiency is thus determined by the temperatures at which heat enters and is exhaustedduring the cycle. For that reas<strong>on</strong>, diesel engines, which burn hotter than gasoline engines,are more efficient.EXAMPLE:A Carnot engine uses a gas whose ratio of specific heats is γ = C P /C V . Express the efficiency interms of γ and the ratio of volumes V c /V b from Fig. 4.1.Since the efficiency is given in terms of the ratio of the temperatures, <strong>on</strong>e must express T H /T Xin terms of the volumes. For the isentropic expansi<strong>on</strong> from b − c,P b V γb= P c V γc ,which when combined with the ideal gas law, P V = NT ,T H V γ−1b= T X V γ−1c , or T HT X=59(VcV b) γ−1.


This givese = 1 − T ( ) γ−1X Vb= 1 − .T H V cThe efficiency thus improves for a higher compressi<strong>on</strong> engine. Whereas the compressi<strong>on</strong> in a gasolineengine is typically between 7 and 11, diesel engines have compressi<strong>on</strong> ratios from 14 to 25.4.4 HydrodynamicsHydrodynamics explains the dynamics of gases and fluids in the limit that the fluid remains locallyequilibrated during the moti<strong>on</strong>. Mathematically, this c<strong>on</strong>diti<strong>on</strong> can be stated in terms of thecharacteristic times associated with the fluid compared to the characteristic times associated withmicroscopic equilibrati<strong>on</strong>. For instance, the atoms in a gas or fluid will kinetically equilibrate aftera few collisi<strong>on</strong>s, thus the relevant microscopic time scale is the collisi<strong>on</strong> time τ coll . For an isotropicexpansi<strong>on</strong>, the expansi<strong>on</strong> rate can be defined as1∼˙ρτ exp∣ρ∣ , (4.23)where ρ represents any c<strong>on</strong>served charge or number density. Hydrodynamics is valid in the limitτ exp >> τ coll . The expansi<strong>on</strong> rate can also be related to the divergence of velocity through theequati<strong>on</strong> of c<strong>on</strong>tinuity,DDρ = −ρ∇ · v,Dt Dt ≡ ∂ + v · ∇, (4.24)∂tHere D/Dt = ∂ t + v · ∇ is the rate of change as observed by an co-mover, i.e., <strong>on</strong>e whose velocityis the same as that of the fluid. The expansi<strong>on</strong> rate is then equal to the divergence of the velocity,1/τ exp = ∇ · v. (4.25)For anisotropic expansi<strong>on</strong>s, <strong>on</strong>e might have a c<strong>on</strong>diti<strong>on</strong> where ∇ · v = 0 even though the variouscomp<strong>on</strong>ents ∂ i v j ≠ 0, e.g. collapsing al<strong>on</strong>g the x axis while expanding al<strong>on</strong>g the y axis. Thecharacteristic expansi<strong>on</strong> time would then be defined by the largest comp<strong>on</strong>ents of the velocitygradient, |∂ i v j |. For systems where τexp −1 is less than, but not much much less than, the collisi<strong>on</strong>rate τ −1coll, <strong>on</strong>e can apply viscous hydrodynamics. However, ideal hydrodynamics is based <strong>on</strong> theassumpti<strong>on</strong> that velocity gradients are always sufficiently small that the matter is always locallyequilibrated.Ideal hydrodynamics involves solving three equati<strong>on</strong>s simultaneously. The first equati<strong>on</strong> describesthe accelerati<strong>on</strong> of the matter,DvDt = − 1 ∇P, (4.26)ρ mwhere ρ m is the mass density. The upper-case D de<str<strong>on</strong>g>notes</str<strong>on</strong>g> the derivative being taken in the frame ofthe fluid. If <strong>on</strong>e c<strong>on</strong>siders a slab of volume δV = Aδx, and multiplies both sides of the equati<strong>on</strong> forthe x th comp<strong>on</strong>ent by ρδV , the l.h.s. becomes ma x and the r.h.s. becomes P (x)A − P (x + δx)A,which is the net force pushing <strong>on</strong> the volume element in the x directi<strong>on</strong>. Thus, this expressi<strong>on</strong> issimply the differential expressi<strong>on</strong> of F = ma, where the force is the pressure multiplied by the area.60


The sec<strong>on</strong>d equati<strong>on</strong> expresses the fact that entropy is c<strong>on</strong>served, or equivalently that dU =−P dV where U is the thermal energy,DUDt= −PDVDt . (4.27)If <strong>on</strong>e c<strong>on</strong>siders a cubic volume element, V = L x L y L z , <strong>on</strong>e can derive thatThis givesdV = dL x L y L z + L x dL y L z + L x L y dL z (4.28)dL x = dt[v x (x = L x ) − v i (x = 0)] = dtL x ∂ x v x ,DVDt= V ∇ · v.DUDt= −P V ∇ · v (4.29)D(U/V )dt= −(P + U/V )∇ · vDϵDt= −(P + ϵ)∇ · v,where the final equati<strong>on</strong> simply uses ϵ to refer to the energy density.The third equati<strong>on</strong> is simply the equati<strong>on</strong> of c<strong>on</strong>tinuity for c<strong>on</strong>served charges. Summarizing,hydrodynamics is driven by the equati<strong>on</strong>s:DvDtDϵDtDρDt= − 1ρ m∇P (4.30)= −(P + ϵ)∇ · v,= −ρ∇ · v,(4.31)Here, the energy density in the frame of the matter is U/V = ϵ. Respectively, the three equati<strong>on</strong>sexpress: the fact that the net force <strong>on</strong> an element is the integrated pressure pushing <strong>on</strong> its surface,entropy c<strong>on</strong>servati<strong>on</strong> and particle number c<strong>on</strong>servati<strong>on</strong>. The co-moving derivatives, D/Dt = ∂ t +v · ∇, represent the time derivative as viewed by an observer in the frame where v = 0. Whensolving the hydrodynamic equati<strong>on</strong>s of moti<strong>on</strong>, there are three variables, ρ, ϵ and v, which <strong>on</strong>eintegrates forward with the three equati<strong>on</strong>s of moti<strong>on</strong>. The pressure and mass densities are not freevariables as P is uniquely determined by ρ and ϵ, and ρ m is determined by ρ. In many cases, thereare multiple c<strong>on</strong>served currents. In that case <strong>on</strong>e writes c<strong>on</strong>servati<strong>on</strong> laws for a vector of chargedensities ρ.Viscous hydrodynamics incorporates correcti<strong>on</strong>s proporti<strong>on</strong>al to the expansi<strong>on</strong> rate divided bythe equilibrati<strong>on</strong> rate. Since the expansi<strong>on</strong> rate is proporti<strong>on</strong>al to the derivative of the velocity,the accelerati<strong>on</strong> from Eq. (4.26) is expanded to include terms which are linear in derivatives of thevelocity such as:Dv i= − 1 {∂ i P − ∂ j [ηω ij ] − ∂ i [ζ∇ · v] + D }Dt ρ m Dt [κ∂ iT ] , (4.32)ω ij ≡ ∂ j v i + ∂ i v j − (2/3)δ ij (∇ · v).61


The equati<strong>on</strong> expressing entropy c<strong>on</strong>servati<strong>on</strong> is also altered:DϵDt = −(P + ϵ)∇ · v + η ∑ ijω 2 ij + ζ(∇ · v) 2 + κ∇ 2 T. (4.33)These equati<strong>on</strong>s are known as Navier-Stokes hydrodynamics. The coefficient ζ is the referred to asthe bulk viscosity. In an expanding system it effectively lowers the pressure proporti<strong>on</strong>al to ∇ · v.The shear viscosity η <strong>on</strong>ly manifests itself when the velocity gradients are anisotropic, as ω = 0for a rotati<strong>on</strong>ally symmetric velocity gradient, e.g., v = v 0 + ar. The coefficient κ is the heatc<strong>on</strong>ductivity and comes from the fact that <strong>on</strong>ce the system is not equilibrated, the energy mightflow with a different velocity than the particle number. This term appears in variety of forms inthe literature depending <strong>on</strong> whether the velocity v is that of the particles or that of the energy.For radiatively dominated systems it is more practical to use the energy and momentum density todefine the velocity, i.e., the velocity is that required to make the momentum density zero. In thelimit where temperatures are very high, this last term can be neglected.EXAMPLE:C<strong>on</strong>sider matter which initially has a uniform temperature T 0 and a three-dimensi<strong>on</strong>al densityprofile,ρ(r, t = 0) = ρ 0 e −r2 /2R 2 0 ,and expands according to an equati<strong>on</strong> of state P = ρT according to ideal hydrodynamics.• Show that if the subsequent evoluti<strong>on</strong> of the density and temperature are parameterized bythat entropy will be c<strong>on</strong>served ifρ(r, t) = ρ 0R 3 0R 3 (t) e−r2 /2R(t) 2 ,R 2 (t)T (t) = R 2 0T 0 .T (r, t) = T (t)The entropy per particle, σ = S/N, can be calculated by c<strong>on</strong>sidering a single particle in avolume V/N in the can<strong>on</strong>ical ensemble (see discussi<strong>on</strong> of Gibb’s paradox in chapter 2),σ = (βE + ln Z [ ∫ ] [ ( ) ] 3/2C)(V/N)= (3/2) + lnd 3 1 mTpe −βϵp = 5/2 + lnN(2π¯h) 3 ρ 2π¯h 2 ,The total entropy is then∫∫S(t) = d 3 rρ(r)σ(r) = c<strong>on</strong>stant + d 3 r ρ(r) ln [ T 3/2 /ρ ]∫= c<strong>on</strong>stant ′ + d 3 r ρ(r) { ln(T 3/2 R 3 ) + r 2 /2R 2}= c<strong>on</strong>stant ′ + (N/2)⟨r 2 /R 2 ⟩ + N ln(T 3/2 R 3 )= c<strong>on</strong>stant ′′ + N ln(T 3/2 R 3 ).In the sec<strong>on</strong>d to the last step, the average ⟨r 2 /R 2 ⟩ = 3, and was absorbed into the c<strong>on</strong>stant.The remaining time dependence comes from the fact that T and R are functi<strong>on</strong>s of time. Thusthe total entropy will be c<strong>on</strong>stant if R 2 T is c<strong>on</strong>stant.62


• Assuming that the velocity profile is linear,v(r, t) = A(t)r,find A(t) and R(t) that satisfy the hydrodynamic equati<strong>on</strong>s of moti<strong>on</strong> and current c<strong>on</strong>servati<strong>on</strong>.Current c<strong>on</strong>servati<strong>on</strong> leads to the expressi<strong>on</strong>,0 = ∂ρ∂t + v ∂ρ ( ) ∂v∂r + ρ ∂r + 2v r{}= ρ − 3ṘR + r2R Ṙ − Ar r 3 R + 3A ,2which fortunately can be satisfied for all r ifNext we c<strong>on</strong>sider the equati<strong>on</strong>s of moti<strong>on</strong>,∂v∂t + v ∂v∂rṘ(t) = A(t)R(t).= −1 ∂Pmρ ∂r = −T ∂ρmρ ∂rAr ˙ + A 2 r = T rmR . 2Again, the equati<strong>on</strong>s of moti<strong>on</strong> can satisfied for all r if˙ A(t) + A 2 (t) =T (t)mR 2 (t)After substituting for T using the fact that R 2 T = R 2 0T 0 , we have two equati<strong>on</strong>s for theevoluti<strong>on</strong> of A and R,Ṙ(t) = A(t)R(t),A(t) ˙ + A 2 (t) = T 0 R02m R 4 (t) .Using the first equati<strong>on</strong> to eliminate A and ˙ A in the sec<strong>on</strong>d equati<strong>on</strong>,m ¨R(t) = R2 0T 0R 3 (t) .This can be solved by noting that the term <strong>on</strong> the right looks like a force affecting a particleof mass m at positi<strong>on</strong> R where the potential is V (R) = R0T 2 0 /2R 2 . Given the potential <strong>on</strong>ecan solve for the trajectory witht =∫ RThe subsequent integrati<strong>on</strong> yields:R 0R(t) 2 = R 2 0 +∫dx Rv = dx√ .2(E − V (x))/mR 0( )T0t 2 , A(t) =mThe temperature can then be found by T = T 0 R 2 0/R 2 .( )T0 tm R 2 (t) .63


4.5 Relativistic HydrodynamicsSome of the most important applicati<strong>on</strong>s of hydrodynamics involve relativistic moti<strong>on</strong>. These includeastr<strong>on</strong>omical applicati<strong>on</strong>s as well as some laboratory applicati<strong>on</strong>s such as relativistic heavyi<strong>on</strong> collisi<strong>on</strong>s. In some cases, such as black holes, general relativity comes into play as well. Hydrodynamicscan incorporate special relativity in a most elegant fashi<strong>on</strong> by c<strong>on</strong>sidering the stressenergy tensor,T αβ = (P + ϵ)u α u β − P g αβ , (4.34)where ϵ is the energy density in the frame of the fluid ϵ = U/V , and u α is the relativistic fourvelocity. The equati<strong>on</strong>s of moti<strong>on</strong> for hydrodynamics and for entropy c<strong>on</strong>servati<strong>on</strong> are expressed as∂ α T αβ = 0. (4.35)This is effectively four separate c<strong>on</strong>servati<strong>on</strong> equati<strong>on</strong>s, <strong>on</strong>e for each value of β. To see that theserepresent the equati<strong>on</strong>s of moti<strong>on</strong> for hydrodynamics, <strong>on</strong>e can view the system in a frame wherethe velocities are small, and if they reproduce the n<strong>on</strong>-relativistic equati<strong>on</strong>s in that frame, they willbe correct in all frames since they have the correct Lorentz form. To see this, we write the 4 × 4stress-energy tensor for the limit where u = (1, v x , v y , v z ),T αβ =⎛⎜⎝ϵ (P + ϵ)v x (P + ϵ)v y (P + ϵ)v z(P + ϵ)v x −P 0 0(P + ϵ)v y 0 −P 0(P + ϵ)v z 0 0 −P⎞⎟⎠ (4.36)By inspecti<strong>on</strong> <strong>on</strong>e can see that the ∂ α T α0 = 0 leads to the equati<strong>on</strong> for entropy c<strong>on</strong>servati<strong>on</strong>, Eq.(4.29). The other three equati<strong>on</strong>s, ∂ α T αi = 0, become,∂v∂t = − 1 ∇P. (4.37)P + ϵIn the n<strong>on</strong>-relativistic limit P + ϵ ∼ ρ m , and these equati<strong>on</strong>s reproduce the n<strong>on</strong>-relativisitic expressi<strong>on</strong>for accelerati<strong>on</strong>. Finally, after writing the equati<strong>on</strong> for current c<strong>on</strong>servati<strong>on</strong> relativistically,∂ α (ρu α ) = 0, all the n<strong>on</strong>-relativistic equati<strong>on</strong>s are satisfied, and if they are satisfied in <strong>on</strong>e frame,they must be satisfied in all frames given that they have a manifestly c<strong>on</strong>sistent Lorentz structure.One might ask about whether the energy density includes the vacuum energy. In that case thereis also a vacuum energy. By “vacuum”, <strong>on</strong>e refers to the energy density at zero temperature andparticle density. In that case, the pressure isP = − dE vacdV = −d(ϵ vacV )dVwhere ϵ vac and P vac denote the vacuum energy density and pressure.relativistic hydrodynamics is insensitive to the vacuum energy density.4.6 Hydrodynamics and Sound= −ϵ vac , (4.38)Thus, the P + ϵ term inTo see how a sound wave is a hydrodynamic wave, we linearize the equati<strong>on</strong>s of moti<strong>on</strong> and theequati<strong>on</strong> of c<strong>on</strong>tinuity, keeping terms <strong>on</strong>ly of first order in the velocity ⃗v, and first order in the64


deviati<strong>on</strong> of the mass density, δρ m (ρ m = ρ (0)m + δρ m ),∂⃗v∂t∂δρ m∂t= − (dP/dρ m)| S∇δρρ (0)m , (4.39)m= −ρ (0)m ∇ · ⃗v.Through substituti<strong>on</strong>, <strong>on</strong>e can eliminate the ⃗v in the equati<strong>on</strong>s by taking the divergence of theupper equati<strong>on</strong> and comparing to the ∂ t of the sec<strong>on</strong>d equati<strong>on</strong>.The soluti<strong>on</strong> to this equati<strong>on</strong> is a cosine wave,∂ 2∂t δρ 2 m = dP ∣ ∣∣∣S∇ 2 δρ m . (4.40)dρ mδρ m (r, t) = A cos (k · r − ωt + ϕ), (4.41)where ϕ is an arbitrary phase and the speed of the wave is√ ∣dP ∣∣∣Sc s = ω/k = . (4.42)dρ mOne could have performed the substituti<strong>on</strong> to eliminate δρ m rather than v, to obtain the compani<strong>on</strong>expressi<strong>on</strong>,∂ 2∂t 2 δ⃗v = c2 s∇ 2 ⃗v, (4.43)for which the corresp<strong>on</strong>ding soluti<strong>on</strong> for the velocity would be,⃗v(r, t) = −A c scos (k · r − ωt + ϕ), (4.44)ρ (0)mwhere again ω and k can be anything as l<strong>on</strong>g as ω/k = c s .If matter is in a regi<strong>on</strong> where c 2 s = dP/dρ m | S < 0, the speed of sound is imaginary. This meansthat for a real wavelength λ = 2π/k, the frequency ω is imaginary. Physically, this signals unstablemodes which grow in time. This can occur for matter which suddenly finds itself deep inside adensity-temperature regi<strong>on</strong> for which it would prefer to separate into two separate phases. Seeproblem 5 of this chapter.4.7 The Boltzmann Equati<strong>on</strong>When local kinetic equilibrium is far from being maintained, even viscous hydrodynamics cannot bejustified. Fortunately, for many of the examples where hydrodynamic descripti<strong>on</strong> is unwarranted,Boltzmann descripti<strong>on</strong>s are applicable. The Boltzmann approach involves following the trajectoriesof individual particles moving through a mean field punctuated by random collisi<strong>on</strong>s. The probabilityof having collisi<strong>on</strong>s is assumed to be independent of the particle’s past history and correlati<strong>on</strong>swith other particles are explicitly ignored. I.e., the chance of encountering a particle depends <strong>on</strong>ly<strong>on</strong> the particles momentum and positi<strong>on</strong>, and the average phase space density of the system. Thus,the Boltzmann descripti<strong>on</strong> is effectively a <strong>on</strong>e-body theory. A Boltzmann descripti<strong>on</strong> is justified iftwo criteria are met:65


• The particles are sufficiently dilute that classical trajectories do not violate the uncertaintyprinciple. Such violati<strong>on</strong>s occur when the collisi<strong>on</strong> time falls below the ¯h/T . Many examplessatisfy this criteria, including most liquids.• Correlati<strong>on</strong>s between particles in space and time are negligible. By definiti<strong>on</strong>, liquids aretightly correlated. Applicati<strong>on</strong>s of the Boltzmann equati<strong>on</strong>s is then questi<strong>on</strong>able, however,str<strong>on</strong>gly correlated systems are often ideal candidates for hydrodynamics.The Boltzmann equati<strong>on</strong> represents a tool for solving for the evoluti<strong>on</strong> of the phase space densityf(p, r, t) a.k.a. the phase space occupancy, which is defined byf(p, r, t) =1 dN(2S + 1) d 3 pd 3 r/(2π¯h) . (4.45)3The definiti<strong>on</strong> is such that f = 1 corresp<strong>on</strong>ds to <strong>on</strong>e particle per quantum state. In a kineticallythermalized system, f would be given by the Bose or Fermi distributi<strong>on</strong>s, and at each point in coordinatespace, a few parameters, T (x, t), µ(x, t) and the collective velocity v(x, t), would determinethe entire momentum dependence. In c<strong>on</strong>trast, a Boltzmann approach requires storing the entirestructure of f(p, r, t).In the absence of interacti<strong>on</strong> particles move in uniform trajectories and the Boltzmann equati<strong>on</strong>is simple,∂f∂t + v p · ∇f = 0. (4.46)This simply states that the co-moving derivative, D/Dt = ∂ t + v∂ x , of the phase space density ofparticles with zero velocity is zero. Adding external forces,0 = ∂f∂t + v p · ∇f + dpdt · ∇ pf (4.47)= ∂f∂t + v p · ∇f + F · ∇ p f.Here, the force F acts like an external field. It could originate from an external source or it couldbe the mean field driven by the phase space density itself. After adding this term, the equati<strong>on</strong> nowstates that after adding an external force, the phase space density will be unchanged if you moveand accelerate with the particles. Thus, in the absence of collisi<strong>on</strong>s, you can choose any particlewith momentum p 0 at positi<strong>on</strong> r 0 , and as you follow the trajectory through time, the phase spacedensity evaluated at that particle’s new momentum and positi<strong>on</strong> will be that same as it was initially.This can be understood physically by stating that if there are n particles in a phase space cell, thatthe same n particles will be in that phase space cell if you follow its trajectory.The fact that the number of particles in a phase space cell is unchanged is not a trivial statement.If <strong>on</strong>e were to choose a phase space cell of dimensi<strong>on</strong>s dx, dy, dz, dp x , dp y , dp z , then follow thetrajectories of the points within the cell, the new phase space cell would be distorted in shape.Despite the fact that dp ′ x ≠ dp x , · · · , the product d 3 p ′ d 3 r ′ would equal the original phase spacevolume d 3 pd 3 r. This result is known as the Liouville theorem. If <strong>on</strong>e c<strong>on</strong>siders the number ofeffective states within a volume, violating the Liouville theorem would involve either inserting orremoving quantum states into the system. As l<strong>on</strong>g as the momenta and coordinates satisfy thecommutati<strong>on</strong> relati<strong>on</strong> [p, x] = i¯h, the number of states must be fixed. The Liouville theorem can66


also be applied to other combinati<strong>on</strong>s of coordinates and their can<strong>on</strong>ical momentum, such as L zand ϕ.To prove the Liouville theorem we c<strong>on</strong>sider a rectangular cell of dimensi<strong>on</strong>s ∆p i and ∆x i . If theboundaries of the cell move with the velocities corresp<strong>on</strong>ding to the coordinates of the boundaries,the number of particles in the cell will remain fixed. The phase space density will then remain fixedif the volume of the cell,1 ∏∆Ω = ∆p(2π¯h) 3 i ∆x i , (4.48)remains fixed. In a time step dt the widths ∆p i and ∆x i change by amounts,id∆p i = dt∆p i∂F i (p, x)∂p i, (4.49)d∆x i = dt∆x i∂v i (p, x)∂x i.If v is <strong>on</strong>ly a functi<strong>on</strong> of v, i.e., v = p/m and if F i is <strong>on</strong>ly a funcit<strong>on</strong> of x, it is clear that the phasespace volume is fixed. However, it remains fixed even if there are momentum-dependent forces orpositi<strong>on</strong> dependent velocities, as can be seen by c<strong>on</strong>sidering Hamilt<strong>on</strong>ians Eq.s of moti<strong>on</strong>,F i = − ∂H∂x i, (4.50)v i = ∂H∂p i, (4.51)<strong>on</strong>e finds()d(∆x i ∆p i ) = ∆x i d∆p i + ∆p i d∆x i = ∆x i ∆p i dt − ∂2 H+∂2 H= 0. (4.52)∂x i ∂p i ∂p i ∂x iOne of the most notable c<strong>on</strong>sequences of the Liouville theorem comes in beam dynamics. Magnetscan focus charged particle beams, thus narrowing them down to a small regi<strong>on</strong>. However, thisfocusing is always accompanied by broadening the distributi<strong>on</strong> in momentum space. C<strong>on</strong>versely,increasing the momentum resoluti<strong>on</strong> of a beam always broadens the beam spot in coordinate space.Collisi<strong>on</strong>s ruin the c<strong>on</strong>tinuity of trajectories, and thus violate Liouville’s theorem. They alsorepresent the <strong>on</strong>ly means towards achieving thermalizati<strong>on</strong>. Collisi<strong>on</strong>s are incorporated into theBoltzmann equati<strong>on</strong> through the expressi<strong>on</strong>,∂f(p, r, t)+ v p · ∇f(p, r, t) + F · ∇ p f(p, r, t) (4.53)∂t∫d 3 p ′ b=(2π¯h) dΩ v dσ(p a , p b → p, p ′ b )3 relf(p a , r, t)f(p b , r, t)dΩ∫d 3 p b−(2π¯h) dΩ v dσ(p, p b → p ′ a, p ′ b )3 relf(p, r, t)f(p b , r, t)dΩHere, v rel is the magnitude of the relative velocity and Ω refers to the angle between the incomingand outgoing relative momentum. The first term <strong>on</strong> the r.h.s. of Eq. (4.53) is the gain term due toparticles with momentum p a and p b scattering into the state p. The sec<strong>on</strong>d term is the loss term67


and represents the scattering of a particle with momentum p with momentum p b . Given eitherboth outgoing momentum (first term <strong>on</strong> r.h.s.) or incoming (sec<strong>on</strong>d term <strong>on</strong> r.h.s.), momentumand energy c<strong>on</strong>servati<strong>on</strong> al<strong>on</strong>g with the angle Ω determine the other two momenta.The expressi<strong>on</strong> in Eq. (4.53) is what is usually referred to as the Boltzmann equati<strong>on</strong>, though<strong>on</strong>e might still call it the Boltzmann equati<strong>on</strong> after adding a few opti<strong>on</strong>s. One such opti<strong>on</strong> involvesmaking the scattering kernel n<strong>on</strong>-local. In the kernel above, the a and b particles are assumed to beat the same locati<strong>on</strong>. It is straight-forward, though a bit tedious, to add a dependence <strong>on</strong> r a − r bto the cross secti<strong>on</strong> so that the a and b particles scatter at r a and r b respectively. This would alsoadd another layer of integrati<strong>on</strong>.A more comm<strong>on</strong> extensi<strong>on</strong> to the Boltzmann equati<strong>on</strong> involves incorporating quantum statisticsby replacing the products of phase space densities of the outgoing particles,f(p ′ a, r)f(p ′ b, r) → f(p ′ a, r)f(p ′ b, r)[1 ± f(p ′ a, r][1 ± f(p ′ b, r)]. (4.54)With this extensi<strong>on</strong>, the equati<strong>on</strong> is sometimes referred to as the Vlasov equati<strong>on</strong>, or the Boltzmann-Uehling-Uhlenbeck equati<strong>on</strong>. The enhancement (Bos<strong>on</strong>) or suppressi<strong>on</strong> (Fermi<strong>on</strong>) factors will pushthe system towards an equilibrated Bose or Fermi distributi<strong>on</strong>. To see how this is accomplished wec<strong>on</strong>sider a single momentum mode p, and assume that the rate at which the populati<strong>on</strong> changes is:∂f p∂t = A p − B p f p . (4.55)Here, A p is the rate at which scatterings fill the level p and B is the rate at which a particle willleave the level if it is already present. If the populati<strong>on</strong> equilibrates at f p = e −β(ϵp−µ) , the ratio ofthe coefficients must be:A p /B p = e −β(ϵ p−µ) . (4.56)The additi<strong>on</strong> of the enhancement/suppressi<strong>on</strong> factors alters the rate equati<strong>on</strong>,Solving for the new equilibrated phase space density,f p =∂f p∂t = A p(1 ± f p ) − B p f p . (4.57)A pB p ∓ A p=e−β(ϵp−µ). (4.58)1 ∓ e−β(ϵp−µ) EXAMPLE:C<strong>on</strong>sider a box of thermalized gas with f(p, r, t < 0) = e −β(ϵp−µ) . The box is infinitely l<strong>on</strong>g in they and z directi<strong>on</strong>s but is c<strong>on</strong>fined in the x directi<strong>on</strong> by two walls, <strong>on</strong>e at x = L and the other atx = −L. At time t = 0 the walls magically dissolve. Assuming the gas particles do not interactwith <strong>on</strong>e another, solve for f(p, r, t > 0).Since f is fixed al<strong>on</strong>g a trajectory, The initial phase space density is:f(p, r, 0) = e −β(ϵ p−µ) Θ(L − x)Θ(x + L).Using the Liouville theorem, with the particle’s velocity being v = p/m,f(p, r, t) = f(p, r − vt, 0) = e −β(ϵ p−µ) Θ(L − x + v x t)Θ(x − v x t + L).68


Writing v x = p x /m, and using the fact that Θ(x) = Θ(Cx),f(p, r, t) = e −β(ϵp−µ) Θ(p x − m(x − L)/t)Θ(m(x + L)/t − p x ).Thus, the distributi<strong>on</strong> at r is a thermal <strong>on</strong>e, except for cutting off the tails of the momentumdistributi<strong>on</strong> for p x above m(x + L)/t and below m(x − L)/t.Numerical soluti<strong>on</strong>s of the Boltzmann equati<strong>on</strong> tend to fall into two categories. One class oftreatments involves storing f(p, r, t) <strong>on</strong> a mesh, then explicitly solving for the evoluti<strong>on</strong> using theBoltzmann equati<strong>on</strong>. This can be difficult since a three-dimensi<strong>on</strong>al Boltzmann equati<strong>on</strong> involvesstoring six-dimensi<strong>on</strong>al phase space informati<strong>on</strong>. If each dimensi<strong>on</strong> is represented by 40 points, <strong>on</strong>ealready begins to challenge the typical amount of memory in a desktop workstati<strong>on</strong> system.A sec<strong>on</strong>d approach is to represent the particles as sample particles. For instance, if a nuclearcollisi<strong>on</strong> has a few hundred nucle<strong>on</strong>s, <strong>on</strong>e might model the scatterings of 100× more particles. Ifthe cross secti<strong>on</strong> were scaled down by a factor of 100, the collisi<strong>on</strong> rates would be the same as if thesampling rati<strong>on</strong> was 1-to-1. In the limit of a large sample factor, the collisi<strong>on</strong>s become local andcorrelati<strong>on</strong>s are washed out, and the simulati<strong>on</strong> approaches the Boltzmann equati<strong>on</strong>.4.8 Phase Space Density and EntropyIn the first chapter, we derived the fact that entropy is S = ∑ i −p i ln p i , where p i is the probabilitythat a state is occupied. If <strong>on</strong>e looks at a single unit of phase space, the probability of havingthe various states n = 0, 1 · · · is determined by the phase space density (at least in a thermalsystem). Thus, the entropy in a given cell should be solely determined by the phase space density,and summing over the cells should give the total entropy. Since the phase space density is a givencell is fixed in the absence of collisi<strong>on</strong>s, it follows that the entropy is also fixed in the absence ofcollisi<strong>on</strong>s.If the phase space density is low, there are <strong>on</strong>ly two probabilities for a given mode, n = 0 andn = 1, with probabilities (1 − f) and f respectively. For Fermi<strong>on</strong>s, these two probabilities are allthat are allowed even if f approaches unity. The entropy for fermi<strong>on</strong>s is thus:S = −(1 − f) ln(1 − f) − f ln(f) (for fermi<strong>on</strong>s), ∼ f[1 − ln(f)] as f → 0. (4.59)Thus, given f(p, r, t), <strong>on</strong>e can obtain an expressi<strong>on</strong> for the entropy that is independent of anyassumpti<strong>on</strong> about thermalizati<strong>on</strong>, and the total entropy is found by summing over phase spacecells:∫1S total = (2J + 1) d 3 p d 3 r [−(1 − f) ln(1 − f) − f ln(f)] . (4.60)(2π¯h) 3To see that the total entropy is fixed in a collisi<strong>on</strong>less system, <strong>on</strong>e can c<strong>on</strong>sider Eq. (4.60) with thereplacement that f(p, r, t) = f(p, r − v p t, 0),S(t) =(2J + 1)(2π¯h) 3∫d 3 pd 3 r {−(1 − f(p, r − v p t, 0) ln(1 − f(p, r − v p t, 0)) (4.61)−f(p, r − v p t, 0) ln(f(p, r − v p t, 0))}If <strong>on</strong>e makes the substituti<strong>on</strong> r ′ ≡ r − v p t, the r.h.s. immediately becomes S(t = 0).69


EXAMPLE:Assuming that the probability of n bos<strong>on</strong>s being in a specific phase space cell is p n ∝ e −nβ(ϵ−µ) , findan expressi<strong>on</strong> for the entropy in terms of the phase space density f.First calculate the normalized value of p n in terms of x ≡ e −β(ϵ−µ) .The entropy is thenp n =x n1 + x + x 2 + x 3 · · · = xn (1 − x).S = − ∑ np n ln(p n ) = −(1 − x) ∑ nx n ln[x n (1 − x)] = −(1 − x) ∑ nx n [n ln(x) + ln(1 − x)]Using the fact that ∑ nx n = x∂ x∑ x n ,S = −x ln(x) − ln(1 − x).1 − xFinally, to get an expressi<strong>on</strong> in terms of f, <strong>on</strong>e inverts the expressi<strong>on</strong>, f = x/(1 − x) to obtainx =Inserting this into the expressi<strong>on</strong> for the entropy,f1, (1 − x) =1 + f 1 + f .S = −f ln(f) + (1 + f) ln(1 + f).Using the Bos<strong>on</strong> result in the example above, and the Fermi<strong>on</strong> result, Eq. (4.60), the entropyfor Bos<strong>on</strong>s/Fermi<strong>on</strong>s can be simultaneously expressed,∫1S total = (2J + 1) d 3 p d 3 r [±(1 ± f) ln(1 ± f) − f ln(f)] . (4.62)(2π¯h) 34.9 Hubble Expansi<strong>on</strong>A Hubble expansi<strong>on</strong> involves the expansi<strong>on</strong> of a system in a translati<strong>on</strong>ally invariant manner.Translati<strong>on</strong>al invariance implies no pressure gradient (relativistically you must be in the referenceframe of the matter) and thus no accelerati<strong>on</strong>. Gravitati<strong>on</strong>al effects can still provide accelerati<strong>on</strong>,but we leave that for a course in general relativity. If the matter originates from a point at x = y =z = t = 0, and if the accelerati<strong>on</strong> is zero, the collective velocity of the matter at later times mustbe:v = r t . (4.63)This simply states that an observer moving with a fluid element of velocity v will be at positi<strong>on</strong>r = vt if there is no accelerati<strong>on</strong>.In the absence of collisi<strong>on</strong>s, a particle with momentum p will slowly traverse matter until itultimately reaches matter whose velocity is that of its own, v p = p/E. Thus, if an observer movingwith the matter measures a particle’s momentum at some time, and if the same particle doesnot collide and reaches a sec<strong>on</strong>d observer, who is also co-moving with the local matter, the sec<strong>on</strong>d70


observer should see a smaller momentum. If a particle of momentum p 0 and energy E 0 = √ p 2 + m 2at time t 0 at the origin, moves without colliding, at time t, its momentum as measured by a localobserver (that moves with the matter) will be:Given that the positi<strong>on</strong> isp = γp 0 − γvE 0 =1√ p r/t0 − √ E 0. (4.64)1 − (r/t)2 1 − (r/t)2r = v p (t − t 0 ) = (p 0 /E 0 )(t − t 0 ), (4.65)the sec<strong>on</strong>d term that is proporti<strong>on</strong>al to E 0 can also be written as piece proporti<strong>on</strong>al to momentumas measured by a co-mover at the later time isp =1√ p 0 − √ (t − t 0)/tp 0 = √ p 0t 01 − (r/t)2 1 − (r/t)2 t2 − r . (4.66)2The denominator can be identified with the proper time, i.e., the time a co-moving observer wouldhave measured since the matter was initially at a point:τ ≡Rewriting the expressi<strong>on</strong> for the locally-measured momentum,t√1 − v2 = √ t 2 − r 2 . (4.67)p = p 0τ 0τ . (4.68)Here, p 0 and p are the momenta of the particle as measured by observers who are at the samepositi<strong>on</strong> as the particle, but moving with the local matter. Thus, as l<strong>on</strong>g as the particle does notsuffer a collisi<strong>on</strong>, its locally measured momentum falls inversely with the proper time. If it nevercollides, it eventually ends up with matter whose velocity matches that of its own, and the locallymeasured momentum approaches zero.If n<strong>on</strong>e of the particles collide between τ 0 and τ, the distributi<strong>on</strong> of momenta at any point, asmeasured by a co-mover, must be the same as the distributi<strong>on</strong> at τ 0 after all the momenta are scaledby the factor τ 0 /τ. This is true for both the relativistic and n<strong>on</strong>-relativistic (T


For hydrodynamic expansi<strong>on</strong> (or collisi<strong>on</strong>less expansi<strong>on</strong>s) entropy is c<strong>on</strong>served. If <strong>on</strong>e is atmoving with the fluid the collective velocity is zero and the equati<strong>on</strong> of c<strong>on</strong>tinuity gives:∂s∂τSince the collective velocity is v = r/τ, this gives:This gives= −s∇ · v. (4.69)∂s∂τ = −3 s τ . (4.70)s(τ) = s(τ 0 ) τ 3 0τ 3 . (4.71)For a massless gas, the entropy density is proproti<strong>on</strong>al to T 3 ,s(τ)s(τ 0 ) = T 3T 3 0= τ 3 0τ 3 , (4.72)and T = T (τ 0 )τ 0 /τ, which is the same expressi<strong>on</strong> given in the example before for a collisi<strong>on</strong>lessexpansi<strong>on</strong>. This equivalence owes itself to Eq. (4.9), which shows that the collisi<strong>on</strong>less systemappears locally equilibrated. Since the system is already locally equilibrated, collisi<strong>on</strong>s have noeffect. For that reas<strong>on</strong>, phot<strong>on</strong>-phot<strong>on</strong> collisi<strong>on</strong>s would have no effect <strong>on</strong> the thermal backgroundradiati<strong>on</strong> spectrum.For a n<strong>on</strong>-relativistic gas of particles with c<strong>on</strong>served particle number, the entropy per particleis determined by the factor ln[(mT ) 3/2 /ρ]. The same equati<strong>on</strong> for c<strong>on</strong>tinuity used for the entropyabove could be applied to ρ, and would give,Fixing the entropy per particle,orρ(τ) = ρ(τ 0 ) τ 3 0τ 3 . (4.73)T (τ) 3/2ρ(τ)= T (τ 0) 3/2ρ(τ 0 ) , (4.74)T (τ) = T (τ 0 ) τ 02τ . (4.75)2Thus, a massive gas cools much more quickly than a gas of massless particles. In cosmology,this difference in cooling inspired the terminology “radiati<strong>on</strong> dominated” for the massless case vs“matter dominated” for the n<strong>on</strong>-relativistic case. When the universe was approximately a hundredthousand years old, the temperature of the universe left the radiati<strong>on</strong>-dominated era and movedinto the matter-dominated behavior era. It should be emphasized that the simple 1/τ 2 coolingfor the matter (mostly hydrodgen) is not valid due to the heating associated with galaxy and starformati<strong>on</strong>, but that for the phot<strong>on</strong>s, the 1/τ cooling is observed as they decoupled from the matterwhen atoms formed and the neutral matter became largely transparent to l<strong>on</strong>g-wavelength phot<strong>on</strong>s.72


4.10 Evaporati<strong>on</strong>, Black-Body Emissi<strong>on</strong> and the Compound NucleusEvaporati<strong>on</strong> plays a crucial role in the cooling of hot stars, hot nuclei, and hot people. For largesystems, where the outer surface is cooler than the inside, accurate estimates of the evaporati<strong>on</strong> rateinvolves integrating over the emissi<strong>on</strong> probability over the volume of the star, c<strong>on</strong>voluting with theescape probability. Detailed microscopic approaches are needed for problems of neutrino emissi<strong>on</strong>from supernovas, or for calculating details of spectra from stars where the outer gas might absorbspecific wave lengths.Such emissi<strong>on</strong> is often approximated by assuming the particles are thermalized at a distance R,and that all outgoing particles at this distance escape. The number emitted per unit area per unittime is:dNdA d 3 p dt = v p · ˆn dN thermd 3 p d 3 r Θ(v p · ˆn), (4.76)where ˆn is the unit normal vector of the surface. For an equilibrated gas,dN thermd 3 p d 3 rf(p) (2J + 1)= (2J + 1) =(2π¯h)3(2π¯h) 3e −(ϵp−µ)/T1 ∓ e −(ϵ p−µ)/T . (4.77)EXAMPLE:The surface of a star has a temperature T , which is much less than the mass of an electr<strong>on</strong> m. Ifthe electr<strong>on</strong>s have chemical potential µ, what is the ratio of emitted phot<strong>on</strong>s to emitted electr<strong>on</strong>s.(Assume the electr<strong>on</strong>s are n<strong>on</strong>-degenerate.)Since both particles have two polarizati<strong>on</strong>s, the ratio will be determined by the ratio of velocitiesand Boltzmann factors:N eN γ=∫d 3 p (p/m)e −β(p2 /2m−µ)∫d3 p e −βp /(1 − e −βp )= (mT )2 8πe βµ ∫ du ue −u8πmT 3 ζ(3)= e βµ mT ζ(3)For small systems, that have an equilibrated temperature throughout the volume, such as acompound nucleus, time-reversal (or Weisskopf) arguments can be applied to the problem. For thiscase, <strong>on</strong>e c<strong>on</strong>siders the equilibrati<strong>on</strong> between a drop at temperature T and a large box at the sametemperature in which the drop lives. The rate at which particles are absorbed is∫Γ abs = ρ⟨σ abs v⟩ =d 3 p(2π¯h) 3 f(p)σ abs(p)v p . (4.78)If the drop is a black, σ abs = πR 2 , where R is the radius of the drop. One can now compare to theemissi<strong>on</strong> rate from Eq.s (4.76) and (4.77),∫Γ emm = 4πR 2 d 3 p(2π¯h) v 3 p · ˆnf(p)Θ(v p · ˆn) (4.79)⟨cos θΘ(cos θ)⟩ =∫= 4πR 2 ⟨cos θΘ(cos θ)⟩∫ 10d 3 p(2π¯h) 3|v p| f(p),d cos θ cos θ∫ 1−1 d cos θ = 1 4 . (4.80)73


Here, the average over emissi<strong>on</strong> directi<strong>on</strong>s yields a factor of 1/4, and if the absorpti<strong>on</strong> cross secti<strong>on</strong>is that of a black body, σ abs = πR 2 , the absorpti<strong>on</strong> and emissi<strong>on</strong> rates become equal. Thus, <strong>on</strong>ecan relate the emissi<strong>on</strong> rate to the absorpti<strong>on</strong> rate if <strong>on</strong>e assumes that at equilibrium the two ratesmust be equal.For small systems with modest excitati<strong>on</strong> energy, the emitted particle might carry a significantfracti<strong>on</strong> of the total excitati<strong>on</strong> energy and the assumpti<strong>on</strong> of an equilibrated temperature andchemical potential becomes unwarranted. For such systems, the temperature before and after theemissi<strong>on</strong> is no l<strong>on</strong>ger identical. To address such systems, <strong>on</strong>e can c<strong>on</strong>sider two sets of states. In thefirst set, the drop has energy between E and E + δE. In the sec<strong>on</strong>d, the drop has energy betweenE − ϵ p and E − ϵ p + δE, and a particle in an individual quantum state of momentum p runs freely.Again, the rates from <strong>on</strong>e set to the other should be equal, with the absorpti<strong>on</strong> rate being,Γ abs = σ abs(p)v p, (4.81)Vwhere V is the volume of the box. Since the emissi<strong>on</strong> from a given state in set 1 to a given statein set 2 equals the rate from the same state in set 2 to the specific state in set 1, the ratio of therates for all states emitted from set 1 to set 2 equals the net rate of the inverse process multipliedby the ratio of states.N(E − ϵ p )Γ emm (p) = Γ absN(E)= σ abs(p)v pV(4.82)N(E − ϵ p ), (4.83)N(E)where N is the number of states of the compound nucleus per unit of excitati<strong>on</strong> energy, usuallycalled the level density. This is the rate for emitting a particle into a specific state. The rate into arange of states in d 3 p is found by multiplying by the number of states per d 3 p, (2J + 1)V/(2π¯h 3 ),which givesdΓ emm= (2J + 1) σ abs(p)v p N(E − ϵ p ). (4.84)d 3 p(2π¯h) 3 N(E)Using the definiti<strong>on</strong> of the entropy, ∆S = ln[N(E)/N(E − ϵ p )], <strong>on</strong>e findsdΓ emmd 3 p= (2J + 1) σ abs(p)v p(2π¯h) 3 eS(E−ϵp)−S(E) . (4.85)If the temperature changes little for such emissi<strong>on</strong>, ∆S = (ϵ p − µ)/T , and <strong>on</strong>e obtains the sameexpressi<strong>on</strong> as before (aside from the degeneracy correcti<strong>on</strong>s for f(p), which would be accounted forif <strong>on</strong>e correctly accounted for the probability that the large box might already have particles so thatthe emissi<strong>on</strong> would increase the number of particles in a given momentum state from n to n + 1).EXAMPLE:C<strong>on</strong>sider a neutr<strong>on</strong> star from which spin-1/2 neutrinos are emitted from within a radius R, whichis at a uniform temperature T . Assume that the probability of an incoming neutrino of momentump being absorbed is η


Simply use the black-body rate from Eq.(4.85). Since the star is macroscopic, <strong>on</strong>e can assumethat ∆S = (ϵ p − µ)/T .dΓd 3 p= 2ηπR2(2π¯h) 3 e−βϵ p.Here, we have ignored the chemical potential, as we have assumed the neutrino is captured by aninverse beta-decay process, n + ν → p + e, with chemical potentials being equal, µ n = µ p + µ e sincethe reacti<strong>on</strong>s equilibrate. This is reas<strong>on</strong>able if the neutrinos leave easily.To find the rate at which energy leaves the star, <strong>on</strong>e must integrate over d 3 p with the Boltzmannfactor calculated using ϵ = p.dEdt= 2ηπR2(2π¯h) 3 ∫= 6ηR2π 2¯h 3 T 4 .d 3 p pe −βϵ pOne nice feature of this result is that the emissi<strong>on</strong> rate equals the black-body rate multiplied bythe absorpti<strong>on</strong> probability, η.4.11 Diffusi<strong>on</strong>Diffusi<strong>on</strong> is the c<strong>on</strong>tinuous limit of a <strong>statistical</strong> random walk. The prototypical example is animpurity in a gas, but there are innumerable examples, both within physical processes or anyrandom process. For instance, the growth and survival of wrapped characters <strong>on</strong> l<strong>on</strong>g paragraphscan be understood with the diffusi<strong>on</strong> equati<strong>on</strong>.The diffusi<strong>on</strong> equati<strong>on</strong> is based <strong>on</strong> the drift equati<strong>on</strong>,j(x) = −D∇ρ. (4.86)This states that a current j will develop in the presence of a density gradient. If <strong>on</strong>e combines thisexpressi<strong>on</strong> with the equati<strong>on</strong> of c<strong>on</strong>tinuity,∂ρ∂t= −∇ · j, (4.87)<strong>on</strong>e finds the diffusi<strong>on</strong> equati<strong>on</strong>∂ρ(x, t)= D∇ 2 ρ. (4.88)∂tHere D is the diffusi<strong>on</strong> c<strong>on</strong>stant.To see the equivalence with a random walk, we write the equivalent equati<strong>on</strong> for a random walk,δN i = Γδt {(N i+1 − N i ) + (N i−1 − N i )} , (4.89)where Γ is the probability of moving in a specific directi<strong>on</strong> per unit time. Thus, in a small time stepδt, the number of particles leaving a locati<strong>on</strong> is 2Γδt. To compare to the diffusi<strong>on</strong> equati<strong>on</strong>, it isinsightful to c<strong>on</strong>sider the change in ρ for a small time step δt and to rewrite the diffusi<strong>on</strong> equati<strong>on</strong>with the derivatives being expressed in terms of changes per ∆x,δρ(x, t) =D δt {[ρ(x + ∆x, t) − ρ(x, t)] + [ρ(x − ∆x, t) − ρ(x, t)]} . (4.90)∆x2 75


Comparing the diffusi<strong>on</strong> equati<strong>on</strong> in this form to the analogous equati<strong>on</strong> for a random walk in Eq.(4.89), then making the statement that ρ is proporti<strong>on</strong>al to N i , allows <strong>on</strong>e to identify the diffusi<strong>on</strong>c<strong>on</strong>stant:D = Γ(∆x) 2 . (4.91)The diffusi<strong>on</strong> c<strong>on</strong>stant D can now be identified with microscopic quantities. For instance, if thecollisi<strong>on</strong> time is τ coll and the mean free path is λ, the size of the random step is ∆x ∼ λ and the rateat which <strong>on</strong>e takes some random step in a specific directi<strong>on</strong> is Γ ∼ 1/2τ coll . The diffusi<strong>on</strong> c<strong>on</strong>stantis:D ∼ λ 2 /τ = λv therm . (4.92)Here, we neglect factors of two, since the collisi<strong>on</strong> rate depends <strong>on</strong> both the velocity of the diffusingparticles, as well as the velocity of the colliding partners, as well as the fracti<strong>on</strong> of momentum lostper collisi<strong>on</strong>. More exact expressi<strong>on</strong>s can be evaluated with a detailed microscopic analysis.A simple soluti<strong>on</strong> to the diffusi<strong>on</strong> equati<strong>on</strong> is a Gaussian,ρ(x, t) =1√4πDtexp{ −x24DtThe width of the Gaussian increases as t 1/2 , beginning at zero.}. (4.93)EXAMPLE:C<strong>on</strong>sider molecules that diffuse in a c<strong>on</strong>tainer whose walls are positi<strong>on</strong>ed at x = 0 and x = L. Ifthe molecules collide with the walls, they stick and disappear from the distributi<strong>on</strong>, and the densityρ(x) must be zero at x = 0 and x = L. C<strong>on</strong>sider a soluti<strong>on</strong> of the form,Using the diffusi<strong>on</strong> equati<strong>on</strong> find A(t).Inserting this expressi<strong>on</strong> into Eq. (4.88),ρ(x, t) = A(t) sin(πx/L).(A ˙π) 2= −D A,L{( } π 2A(t) = A 0 exp −D t .L)One can also check the soluti<strong>on</strong> to see that the rate at which the integrated number falls equals thedrift into the surfaces. Integrating the density,∫N(t) = dx ρ(x, t) = 2A(t)L/π,Ṅ = 4 ˙ ALπ= −πDA LSince the currents into each wall is D∂ρ/∂x, <strong>on</strong>e should get the same result from the drift equati<strong>on</strong>,Indeed, this agrees.Ṅ = −2D ∂ρ∂x = −ADπ L .76


EXAMPLE:C<strong>on</strong>sider the same problem, but with an initially uniform density ρ 0 . Express ρ(t) as a Fourierexpansi<strong>on</strong> in terms of sin(mπx/L).First, recall the properties of a Fourier transform in a finite interval, 0 < x < L,∑F (x) = F m sin(mπx/L),F m = 2 Lm=1,2,3···∫ LNow, use the following expressi<strong>on</strong> for the density,0dx F (x) sin(mπx/L).ρ(x, t) = ∑ mA m (t) sin(mπx/L).Before proceeding further, find A m (t = 0),∫ LA m (t = 0) = 2 dx ρ 0 sin(mπx/L) (4.94)L 0{ 4ρ0 /mπ, odd m,=(4.95)0, even m.Each c<strong>on</strong>tributi<strong>on</strong> behaves independently, and can be solved exactly as the previous example, whichyields,{ ( mπ) } 2A m (t) = A m (t = 0) exp −D t .LThus, the higher order c<strong>on</strong>tributi<strong>on</strong>s are damped with rapidly increasing exp<strong>on</strong>ential dampingfactors. For large times soluti<strong>on</strong>s are independent of the initial c<strong>on</strong>diti<strong>on</strong>s and approach the formof the previous example, i.e., <strong>on</strong>ly the m = 1 term survives.For grins, <strong>on</strong>e can use the properties of Fourier transforms to show that ζ(2) = π 2 /6. If <strong>on</strong>eintegrates,∫ LSome algebra shows that0dxρ(x, t = 0) = ρ 0 L,∑m=1,3···The previous expressi<strong>on</strong> thus becomes= ∑m=1,3···= ∑m=1,3···1m 2 = 3 4A m (t = 0)8ρ 0 L(mπ) 2 .∑m=1,2,3···ρ 0 L = 8ρ 0Lπ 2 34 ζ(2),ζ(2) = π26 .77∫ L01m 2 = 3 4 ζ(2).dx sin(mπx/L)


For some problems, <strong>on</strong>e may wish to use the method of images, similar to what is used for someboundary-value problems in electromagnetism. A simple example c<strong>on</strong>siders placing a charge at x 0 ,where there is an absorbing plate at the x = 0 plane. In the absence of the plate, the density wouldevolve as1ρ(x, t) = √ e −(x−x 0) 2 /(4Dt) . (4.96)4πDtHowever, this would not satisfy the boundary c<strong>on</strong>diti<strong>on</strong> that ρ(x = 0, t) = 0. By c<strong>on</strong>sidering asec<strong>on</strong>d soluti<strong>on</strong> centered at −x 0 , <strong>on</strong>e can subtract the two soluti<strong>on</strong>s to find a soluti<strong>on</strong> that satisfiesboth the boundary c<strong>on</strong>diti<strong>on</strong> at x = 0 and the initial c<strong>on</strong>diti<strong>on</strong>, that for x > 0 the charge is <strong>on</strong>lyat <strong>on</strong>e point at t = 0. The overall soluti<strong>on</strong> is thenρ(x, t) ={1 √4πDt(e −(x−x 0) 2 /(4Dt) − e −(x+x 0) 2 /(4Dt)), x > 00, x < 0When x 0 is small, compared to √ Dt, <strong>on</strong>e can expand the soluti<strong>on</strong> in powers of x 0 , and find(4.97)∣dρ ∣∣∣x0ρ(x, t) ≈ x 0dx 0 =0xx 0= √ /(4Dt) 4π(Dt)3 e−x2 .(4.98)For this limiting soluti<strong>on</strong>, <strong>on</strong>e can solve for the remaining fracti<strong>on</strong> of original particles,N(t) ==∫ ∞0x 0√πDt.Further as a check, <strong>on</strong>e can see that dN/dt = − x 0j = −D∂ x ρ at x = 0.4.12 Problems4 √ πDtdx ρ(x, t) (4.99)3, which indeed matches the flux into the wall,1. A molecule of mass m has two internal states, a spin-zero ground state and a spin-1 excitedstate which is at energy X above the ground state. Initially, a gas of such molecules is attemperature T i before expanding and cooling isentropically to a temperature T f . Neglectquantum degeneracy of the momentum states for the following questi<strong>on</strong>s.(a) What is the initial energy per particle? Give answer in terms of m, T i , X and the initialdensity ρ i .(b) Derive an expressi<strong>on</strong> for the initial entropy per particle in terms of the same variables.(c) After insentropically cooling to T f , find the density ρ f . Give answer in terms of ρ i , T i ,T f and X.2. Repeat problem #1 above, but assume the molecule has the excitati<strong>on</strong> spectrum of a 3-dimensi<strong>on</strong>al harm<strong>on</strong>ic oscillator, where the energy levels are separated by amounts ¯hω = X,with X


3. C<strong>on</strong>sider an ideal gas with C p /C v = γ going through the Carnot cycle illustrated in Fig. 4.1.The initial volume for N molecules at temperature T H expands from V a to V b = 2V a and thento V c = 2V b .(a) In terms of NT H , find the work d<strong>on</strong>e while expanding from a − b.(b) Again, in terms of NT H , how much heat was added to the gas while expanding froma − b.(c) In terms of NT H , find the work d<strong>on</strong>e while expanding from b − c.(d) What is the efficiency of the cycle (abcd)?4. C<strong>on</strong>sider a hydrodynamic slab which has a Gaussian profile al<strong>on</strong>g the x directi<strong>on</strong> but istranslati<strong>on</strong>ally invariant in the y and z directi<strong>on</strong>s. Assume the matter behaves as an ideal gasof n<strong>on</strong>-relativistic particles. Initially, the matter is at rest and has a profile,ρ(x, t = 0) = ρ 0 exp(−x 2 /2R 2 0),with an initial uniform temperature T 0 . Assume that as it expands it maintains a Gaussianprofile with a Gaussian radius R(t).(a) Show that entropy c<strong>on</strong>servati<strong>on</strong> requires( ) 2/3 R0T (t) = T 0 .R(b) Assuming the velocity has the form v = A(t)x, show that c<strong>on</strong>servati<strong>on</strong> of particle currentgivesA = ṘR .(c) Show that the hydrodynamic expressi<strong>on</strong> for accelerati<strong>on</strong> givesA ˙ + A 2 = R2/30 T 0mR 8/3Putting these two expressi<strong>on</strong>s together show that R(t) can be found by solving andinverting the integral,√ ∫ m Rdxt = √3T 0 1 − (R0 /x) .2/35. C<strong>on</strong>sider a gas obeying the Van der Waals equati<strong>on</strong> of state,P =R 0ρT1 − ρ/ρ s− aρ 2 .(a) First, using Maxwell relati<strong>on</strong>s, show that the speed of sound for P (ρ, T ) is given bymc 2 s = ∂P∂ρ∣ +T(∂P∂T) 2∣ρTρ 2 C V.79


(b) C<strong>on</strong>sider a gas described by the Van der Waals equati<strong>on</strong> of state, which has a densityequal to that of the critical point, ρ s /3. What is the range of temperatures for whichthe matter has unstable s<strong>on</strong>ic modes? Express your answer in terms of the criticaltemperature T c .6. C<strong>on</strong>sider an initially thermalized three-dimensi<strong>on</strong>al Gaussian distributi<strong>on</strong> for the phase spacedensity of n<strong>on</strong>-relativistic particles of mass m,}f(p, r, t = 0) = f 0 exp{− r2−p2.2mT 0Assume the particles move freely for t > 0.(a) At a given r and t > 0, show that f(p, r, t) can be expressed in terms of a locallythermalized distributi<strong>on</strong> of the form,}f(p, r, t) = C(t)e −r2 /2R 2 (t) (p − mv(r, t))2exp{− .2mT (r, t)Then, find C(t), R(t), v(r, t) and T (r, t). In additi<strong>on</strong> to r and t, these parameters shoulddepend <strong>on</strong> the initial Gaussian size R 0 , the initial temperature T 0 and the mass m.(b) Find the density as a functi<strong>on</strong> of r and t, then compare your result for the density andtemperature to that for a hydrodynamic expansi<strong>on</strong> of the same initial distributi<strong>on</strong> asdescribed in the example given in the <str<strong>on</strong>g>lecture</str<strong>on</strong>g> <str<strong>on</strong>g>notes</str<strong>on</strong>g>.(c) Using the fact that a hydrodynamic expansi<strong>on</strong> assumes infinitely high collisi<strong>on</strong> rate, andthat the Boltzmann soluti<strong>on</strong> was for zero collisi<strong>on</strong> rate, make profound remarks abouthow the hydrodynamic and free-streaming evoluti<strong>on</strong>s compare with <strong>on</strong>e another.(d) Calculate the total entropy as a functi<strong>on</strong> of time for the previous problem assuming f 0is small.7. Assume that there exists a massive species of neutrinos, m ν = 10 eV. Further assume thatit froze out at the same time as the example of the text, when the Hubble time as 10 7 yearsand the temperature was 4000 K. If the Hubble expansi<strong>on</strong> was without accelerati<strong>on</strong> after thatpoint, and if the current Hubble time is 14 × 10 9 years, find:2R 2 0(a) the current effective temperature of the massive neutrino.(b) If the neutrino has two polarizati<strong>on</strong>s (just like phot<strong>on</strong>s) what would be the relativepopulati<strong>on</strong>, N ν /N γ , at freezeout? Assume the chemical potential for the neutrino is zero(otherwise there would be more neutrinos than anti-neutrinos) and treat the neutrin<strong>on</strong><strong>on</strong>-relativistically,f ν (p) =e−βϵ1 + e −βϵ ≈ e−β(m+p2 /2m) .8. C<strong>on</strong>sider a hot nucleus of radius 5 fm at a temperature of 1 MeV. The chemical potentialfor a cold (or warm) nucleus is approximately the binding energy per particle, µ ≈ −7 MeV.Estimate the mean time between emitted neutr<strong>on</strong>s. (Treat the nucleus as if it has fixedtemperature and assume a Boltzmann distributi<strong>on</strong> (not Fermi) for f(p)).80


9. C<strong>on</strong>sider a hot nucleus of radius R with an electric charge of Z at temperature T . Assumingthat prot<strong>on</strong>s and neutr<strong>on</strong>s have the same chemical potential, find the ratio of prot<strong>on</strong> spectrato neutr<strong>on</strong> spectra,dN prot /d 3 pdtdN neut /d 3 pdtas a functi<strong>on</strong> of the momentum p. Approximate the two masses as being equal, m n = m p ,and neglect quantum degeneracy. Assume that all incoming nucle<strong>on</strong>s would be captured andthermalized if they reach the positi<strong>on</strong> R. HINT: The emissi<strong>on</strong> ratio equals the ratio of capturecross secti<strong>on</strong>s.10. A drift detector works by moving the electr<strong>on</strong>s i<strong>on</strong>ized by a track through a gas towardsreadout plates.(a) Assuming the mean free path of the electr<strong>on</strong>s is λ = 300 nm, and assuming their velocityis thermal at room temperature (v therm ≈ √ T/m), estimate the size R (Gaussian radius)that the diffusi<strong>on</strong> cloud imprints <strong>on</strong>to the plates after drifting for 200 µs. Use theapproximati<strong>on</strong> that D ≈ λv therm .(b) Assume an electric field E is resp<strong>on</strong>sible for the drift velocity. If the drift velocity isapproximately a · τ/2, where a is the accelerati<strong>on</strong> and τ = λ/v therm is the collisi<strong>on</strong> time,find an analytic expressi<strong>on</strong> for the Gaussian size R of the cloud if it travels a distanceL. Give R in terms of the temperature T , the electr<strong>on</strong> mass m and charge e, the electricfield E, the mean free path λ and L.11. A cloud of radioactive mosquitos is being blown by the wind parallel to a high-voltagemosquito-zapping plate. At at time t = 0, the cloud is at the first edge of the plate (x = 0).The probability they start out at a distance y from the plate is:ρ(y, t = 0) = A 0 ye −y2 /(2R 2 0 )The speed of the breeze is v x , and the length of the plate is L. The mosquito’s moti<strong>on</strong> in they directi<strong>on</strong> can be c<strong>on</strong>sidered a diffusi<strong>on</strong> process with diffusi<strong>on</strong> c<strong>on</strong>stant D.(a) What is the distributi<strong>on</strong> ρ(y, t)? Assume that the form for ρ(t) is the same as for ρ 0 <strong>on</strong>lywith A and R becoming functi<strong>on</strong>s of t. Solve <strong>on</strong>ly for t < L/v x .(b) What fracti<strong>on</strong> of mosquitos survive?12. For an expanding system, the Diffusi<strong>on</strong> equati<strong>on</strong> is modified by adding an extra term proporti<strong>on</strong>alto ∇ · v,∂ρ∂τ + ρ(∇ · v) = D∇2 ρ.The sec<strong>on</strong>d term accounts for the fact that the density would fall due to the fact that thematter is expanding, even if the the particles were not diffusing. For a Hubble expansi<strong>on</strong>,v = r/τ, and the ∇ · v = 3/τ (would be 2/τ or 1/τ in 2-D or 1-D). Thus in a Hubbleexpansi<strong>on</strong>, the diffusi<strong>on</strong> equati<strong>on</strong> becomes∂ρ∂t + 3 τ ρ = D∇2 ρ.81


Instead of the positi<strong>on</strong> r, <strong>on</strong>e can use the variable⃗η ≡ ⃗r τ .The advantage of using η is that since the velocity gradient is 1/τ, the velocity differencebetween two points separated by dr = τdη is dv = dη. Thus if two particles move with thevelocity of the local matter, their separati<strong>on</strong> η 1 − η 2 will remain fixed. Next, <strong>on</strong>e can replacethe density ρ = dN/d 3 r withρ η ≡ dNd 3 η = τ 3 ρ.Here, we have been a rather sloppy with relativistic effects, but for |η| much smaller than thespeed of light, they can be ignored. One can now rewrite the diffusi<strong>on</strong> equati<strong>on</strong> for ρ η ,∂ρ η∂τ = D∇2 ρ η .This looks like the simple diffusi<strong>on</strong> equati<strong>on</strong> without expansi<strong>on</strong>, however since the densityis changing D is no l<strong>on</strong>ger a c<strong>on</strong>stant which invalidates using the simple Gaussian soluti<strong>on</strong>sdiscussed in the chapter. For an ultrarelativistic gas, with perturbative interacti<strong>on</strong>s, thescattering cross secti<strong>on</strong>s are roughly proporti<strong>on</strong>al to 1/T 2 , and the density falls as 1/T 3 ,which after c<strong>on</strong>sidering the fact that the temperature then falls as 1/τ, the diffusi<strong>on</strong> c<strong>on</strong>stantwould roughly rise inversely with the time,D(τ) = D 0ττ 0.This time dependence would be different if the particles had fixed cross secti<strong>on</strong>s, or if the gaswas not ultra-relativistic. However, we will assume this form for the questi<strong>on</strong>s below.(a) Transform the three-dimensi<strong>on</strong>al diffusi<strong>on</strong> equati<strong>on</strong>,∂ρ η∂τ = D(τ)∇2 ρ ηinto an equati<strong>on</strong> where all derivatives w.r.t. r are replaced with derivatives w.r.t. η.(b) Rewrite the expressi<strong>on</strong> so that all menti<strong>on</strong> of τ is replaced by s ≡ ln(τ/τ 0 ).(c) If a particle is at ⃗η = 0 at τ 0 , find ρ η (η, s).82


5 Lattices and SpinsOh, what tangled webs we weave when first we practice to believe - L.J. Peter5.1 Statistical Mechanics of Ph<strong>on</strong><strong>on</strong>sPh<strong>on</strong><strong>on</strong>s are sound waves through a crystal, with the name“ph<strong>on</strong><strong>on</strong>” coming from the Greek ph<strong>on</strong>efor sound, and made to sound like phot<strong>on</strong>, with which ph<strong>on</strong><strong>on</strong>s have much in comm<strong>on</strong>. Ph<strong>on</strong><strong>on</strong>s playa critical role in determining the specific heat and heat diffusi<strong>on</strong> properties of solids. Additi<strong>on</strong>ally, itis the exchange of ph<strong>on</strong><strong>on</strong>s that provide the attractive interacti<strong>on</strong> necessary for superc<strong>on</strong>ductivity.S<strong>on</strong>ic excitati<strong>on</strong>s in solids have discrete eigen-modes, just as they do for air in organ pipes.Using k = (k x , k y , k z ) to denote the wave number, the density waves in a rectangular solid behaveasπ π πδρ = A sin(k · r), k x = n x , k y = n y , k z = n z . (5.1)L x L y L zAs quantum excitati<strong>on</strong>s, the momentum and energy of the excitati<strong>on</strong>s are determined by the simplerelati<strong>on</strong>s,p = ¯hk, ϵ = ¯hω, (5.2)which looks exactly the same as that for phot<strong>on</strong>s. The difference being that for ph<strong>on</strong><strong>on</strong>s, ω = c s k,with c s being the speed of sound rather than the speed of light. In the macroscopic limit, we statethe density of such modes as:dN = 3V d3 k(2π) 3= 3V d3 p(2π¯h) 3 , (5.3)where the factor 3 <strong>on</strong> the r.h.s. arises because there are 2 transverse modes and <strong>on</strong>e l<strong>on</strong>gitudinalmode for any wave number. This is exactly the same as the expressi<strong>on</strong> for free particles with J = 1.A sec<strong>on</strong>d difference between ph<strong>on</strong><strong>on</strong>s and phot<strong>on</strong>s is that wavelengths of s<strong>on</strong>ic excitati<strong>on</strong>s cannotbe arbitrarily short, and are cut off at the scale of the lattice spacing. To determine the wavelengthand frequency of the cutoff, the usual argument is based <strong>on</strong> the number of modes of the s<strong>on</strong>icexcitati<strong>on</strong>s as compared to the number of vibrati<strong>on</strong>al modes in the individual atoms in the lattice.If <strong>on</strong>e treats the lattice as a collecti<strong>on</strong> of coupled three-dimensi<strong>on</strong>al harm<strong>on</strong>ic oscillators, each sitec<strong>on</strong>tributes three modes if the coupling is ignored. Once the coupling between sites is introduced,the new soluti<strong>on</strong>s become linear combinati<strong>on</strong>s of the previous 3N modes. The cutoff in wave numbercan be determined by equating the number of modes,which gives the cutoff wave number as3N =3V(2π) 3 ∫ kD0d 3 k, (5.4)( ) 6π 2 1/3Nk D =. (5.5)VThe subscript D refers to Debye, who first c<strong>on</strong>sidered the cutoff in 1912. It is more comm<strong>on</strong> torefer to the frequency of the cutoff rather than the wave number. The Debye frequency is( ) 6π 2 1/3Nω D = k D c s = c s . (5.6)V83


In the real world, the transverse and l<strong>on</strong>gitudinal modes have different speeds of sound. One canmake correcti<strong>on</strong>s for this and assign different values of ω D and k D for l<strong>on</strong>gitudinal and transversemodes, but as the physics of the high-frequency modes is described by the simple cutoff prescripti<strong>on</strong>at the ∼ 10% level, we will simply assume that both modes have the same cutoff in frequency.Figure 5.1: The specific heat from ph<strong>on</strong><strong>on</strong>s rises asT 3 at low temperature, then approaches a c<strong>on</strong>stantof 3N at large T , where the thermal properties revertto those of the harm<strong>on</strong>ic oscillator states up<strong>on</strong> whichthe s<strong>on</strong>ic modes are built.The occupancy of a given mode is given bye−¯hc sk/Tf(k) =1 − e , (5.7)−¯hc sk/Twhich is the expressi<strong>on</strong> for Bose modes. The modes are bos<strong>on</strong>ic because they are built <strong>on</strong> harm<strong>on</strong>icoscillator states which allow multiple excitati<strong>on</strong>s. We will ignore the added complicati<strong>on</strong> of assigningthe l<strong>on</strong>gitudinal and transverse modes different speeds of sound. The energy and specific heat are:3VE =d(2π¯h)∫p


This is the expected result for N three-dimensi<strong>on</strong>al harm<strong>on</strong>ic oscillators. We could have predictedthis result by simply knowing that at high temperature the coupling between oscillators shouldbecome irrelevant and the specific heat should be determined by the number of degrees of freedom.To study the behavior of C V at low temperature we first integrate by parts, using the fact thate −x /(1 − e −x ) 2 = −(d/dx)e −x /(1 − e −x ). This yields,e −x DD(x D ) = −3x D + 12(1 − e −x D )x 3 D∫ xD0e −xdx x 3 . (5.12)1 − e−x At low temperatures, x D → ∞, which allows <strong>on</strong>e to ignore the first term while setting the limit <strong>on</strong>the sec<strong>on</strong>d term to ∞.D(x D → ∞) = 72 ζ(4), (5.13)x 3 D( ) 3 TC V (T → 0) = 216Nζ(4)¯hω D= 36 V T 3π 2 (¯hc s ) 3 ζ(4),where ζ(4) = π 4 /90 is the Riemann-Zeta functi<strong>on</strong>. Aside from the different spin factor, this is theidentical result <strong>on</strong>e would obtain for phot<strong>on</strong>s if c s were replaced by the speed of light. FYI: Thesec<strong>on</strong>d term in Eq. (5.12) is sometimes expressed in terms of the functi<strong>on</strong>s D n (x), where D n withthe subscript n is the more general definiti<strong>on</strong> of a Debye functi<strong>on</strong>,D n (x) ≡ 1 ∫ xdtx 30t n e −t. (5.14)1 − e−t 5.2 Feromagnetism and the Ising Model in the Mean Field Approximati<strong>on</strong>In the next secti<strong>on</strong> we will c<strong>on</strong>sider the Ising model, which is based <strong>on</strong> a simple Hamilt<strong>on</strong>ian forspins <strong>on</strong> a lattice,H = −J ∑ σ i σ j − µB ∑ σ i . (5.15)n.n.iHere, the notati<strong>on</strong> <strong>on</strong> the sum, n.n., refers to the fact that <strong>on</strong>e sums over all pairs of nearestneighbors, and the values of the spins σ i are limited to ±1 (not ±1/2). One can include a highernumber of spins, but those models go by a different name, Pott’s model. In a two-dimensi<strong>on</strong>al Pott’smodel the spins can point in a finite number n of directi<strong>on</strong>s which are separated by ∆θ = 2π/n,and the interacti<strong>on</strong>s between neighboring spins behave as −J cos(θ i − θ j ). In the limit of n → ∞, itis called the XY model, and for n = 2 it becomes the Ising model. Despite the apparent simplicityof the Ising model interacti<strong>on</strong>, the soluti<strong>on</strong> is remarkably rich and difficult to solve.In this secti<strong>on</strong>, we c<strong>on</strong>sider the mean-field limit of the Ising model. In that limit, the spins aretreated independently, with neighboring spins being treated as if they possessed the average spin,unrelated to the spin of the immediate neighbor. Thus, the interacti<strong>on</strong> for a single spin isH i = −qJ⟨σ⟩σ i − µBσ i , (5.16)85


where q is the number of nearest neighbors and ⟨σ⟩ is the average spin. First, we c<strong>on</strong>sider the casewith zero external field B. One can then calculate the average of σ,⟨σ i ⟩ = eβqJ⟨σ⟩ − e −βqJ⟨σ⟩= tanh(βqJ⟨σ⟩). (5.17)e βqJ⟨σ⟩ + e−βqJ⟨σ⟩ Since the average of a particular spin equals the average of all spins, ⟨σ i ⟩ = ⟨σ⟩, and <strong>on</strong>e has asimple equati<strong>on</strong> to solve for ⟨σ⟩,⟨σ⟩ = tanh(βqJ⟨σ⟩). (5.18)The soluti<strong>on</strong> can be seen graphically in Fig. 5.2 by plotting the l.h.s. and the r.h.s. of Eq. (5.18)as a functi<strong>on</strong> of ⟨σ⟩ and finding where the two graphs intersect. The ⟨σ⟩ = 0 soluti<strong>on</strong> alwaysexists. A sec<strong>on</strong>d n<strong>on</strong>-trivial soluti<strong>on</strong> exists <strong>on</strong>ly for βqJ > 1. Otherwise the tanh functi<strong>on</strong> bendsover and the two functi<strong>on</strong>s do not intersect. This criteria gives the critical temperature for thetransiti<strong>on</strong>, βqJ = 1 or T c = qJ. More neighbors or a higher coupling increases T c . For T < T c , the⟨σ⟩ = 0 soluti<strong>on</strong> also satisfies Eq. (5.18) but yield a higher free energy. Fig. 5.3 shows the meanmagnetizati<strong>on</strong> as a functi<strong>on</strong> of temperature from solving Eq. (5.18).Figure 5.2: Graphical soluti<strong>on</strong> to Eq. (5.18). ForβqJ = 0.75, there is no soluti<strong>on</strong> aside from ⟨σ⟩ = 0,whereas for βqJ = 1.25, there is an additi<strong>on</strong>al soluti<strong>on</strong>as marked where the lines intersect. The soluti<strong>on</strong>exists for βqJ > 1, or for T < T c = qJ.Figure 5.3: The mean magnetizati<strong>on</strong> as a functi<strong>on</strong>of temperature in the mean field approximati<strong>on</strong>.Sp<strong>on</strong>taneousmagnetizati<strong>on</strong> ensues for T < T c =qJ for zero field, whereas for finite the transiti<strong>on</strong> issmoothed over and some magnetizati<strong>on</strong> ensues forT > T c .86


EXAMPLE:Find an expressi<strong>on</strong> for ⟨σ⟩ in the limit of small deviati<strong>on</strong> δT = T c − T .In this limit ⟨σ⟩ will be small, so the tanh functi<strong>on</strong> in Eq. (5.18) can be expanded for small σ,⟨σ⟩ = βqJ⟨σ⟩ − 1 3 (βqJ⟨σ⟩)3 + · · · .Solving for ⟨σ⟩,⟨σ⟩ ≈√3(βqJ − 1)(βqJ) 3≈√3(T c − T )T c. (5.19)One can add an external magnetic field to the mean field equati<strong>on</strong>. This adds a term µσB to theFigure 5.4: Graphical soluti<strong>on</strong> to Eq. (5.21), magnetizati<strong>on</strong>with external field. The magnetic fieldraises the tanh curves from the origin so that soluti<strong>on</strong>sexist for any temperature, not just T < T c .energy for each spin, which leads to the following expressi<strong>on</strong> for the average spin,or⟨σ i ⟩ = eβ(qJ⟨σ⟩+µB) − e −β(qJ⟨σ⟩+µB), (5.20)e β(qJ⟨σ⟩+µB) + e−β(qJ⟨σ⟩+µB) ⟨σ⟩ = tanh β (qJ⟨σ⟩ + µB) . (5.21)Unlike the soluti<strong>on</strong> with B = 0, the ⟨σ⟩ ̸= 0 soluti<strong>on</strong> exists for all temperatures. The singularity inthe plot of ⟨σ⟩ vs. T is then smoothed out as shown in Fig. 5.3.The results described above can alternatively be found by c<strong>on</strong>sidering the free-energy. Thisalternative method is a bit l<strong>on</strong>ger, but has the advantage of showing how the ⟨σ⟩ = 0 soluti<strong>on</strong> isless favorable. Since the free energy is given by F = E − T S, we first find an expressi<strong>on</strong> for theentropy in terms of ⟨σ⟩, the temperature, the field and the coupling qJ. Assuming the probabilitythat a spin in the spin-up state is p,SN = − ∑ ip i ln p i = −p ln p − (1 − p) ln(1 − p). (5.22)87


To express this in terms of ⟨σ⟩, <strong>on</strong>e inverts the expressi<strong>on</strong>⟨σ⟩ = p − (1 − p) = 2p − 1, (5.23)to findThe entropy is thenSN= −1+ ⟨σ⟩2p = 1 + ⟨σ⟩2, (1 − p) = 1 − ⟨σ⟩ . (5.24)2( ) 1 + ⟨σ⟩ln2− 1 − ⟨σ⟩2( ) 1 − ⟨σ⟩ln2(5.25)The energy per spin isEN = −qJ 2 ⟨σ⟩2 − µB⟨σ⟩, (5.26)where the factor of 1/2 ensures that the interacti<strong>on</strong> between neighboring spins is not double counted.Now, <strong>on</strong>e must simply minimize F = E − T S with respect to ⟨σ⟩,∂ F∂⟨σ⟩ N = 0 (5.27){∂= − qJ ∂⟨σ⟩ 2 ⟨σ⟩2 − µB⟨σ⟩ + T 1 + ⟨σ⟩ ( ) 1 + ⟨σ⟩ln+ T 1 − ⟨σ⟩ ( )} 1 − ⟨σ⟩ln2 22 20 = −qJ⟨σ⟩ − µB + T ( ) 1 + ⟨σ⟩2 ln .1 − ⟨σ⟩Rearranging the terms so that the logarithm is isolated <strong>on</strong> <strong>on</strong>e side, then taking the exp<strong>on</strong>ential ofboth sides,This is precisely the earlier result, Eq. (5.21)5.3 One-Dimensi<strong>on</strong>al Ising Modele 2βqJ⟨σ⟩+2βµB = 1 + ⟨σ⟩1 − ⟨σ⟩ , (5.28)⟨σ⟩ = e2βqJ⟨σ⟩+2βµB − 1e 2βqJ⟨σ⟩+2βµB + 1= tanh(βqJ⟨σ⟩ + βµB).The Ising model can be written in any number of dimensi<strong>on</strong>s, but the <strong>on</strong>e-dimensi<strong>on</strong>al situati<strong>on</strong> isof particular interest because it can be solved analytically. The method for solving the model is ofinterest in its own right, but it also represents a striking less<strong>on</strong> of the folly of applying the meanfield approximati<strong>on</strong> in certain situati<strong>on</strong>s.It is intuitively clear that a <strong>on</strong>e-dimensi<strong>on</strong>al system cannot undergo sp<strong>on</strong>taneous magnetizati<strong>on</strong>.For any temperature above zero, there is always the possibility that any two c<strong>on</strong>secutive spins mightnot point in the same directi<strong>on</strong>. The penalty for a l<strong>on</strong>g chain to break into two equal regi<strong>on</strong>s, <strong>on</strong>espin-up the other spin-down, is thus finite. The same is not true for a two- or three-dimensi<strong>on</strong>alsystem. In those cases, <strong>on</strong>e must flip spins al<strong>on</strong>g a boundary whose size grows to infinity as the size88


of the system goes to infinity. This rather manifest fact is sometimes stated as a principle: “Thereis no l<strong>on</strong>g-range order in <strong>on</strong>e-dimensi<strong>on</strong>” (unless there is a l<strong>on</strong>g-range interacti<strong>on</strong>). This principalis manifestly violated in the mean-field approximati<strong>on</strong> which results in sp<strong>on</strong>taneous magnetizati<strong>on</strong>,independent of the dimensi<strong>on</strong>ality. This underscores the importance of understanding the role ofcorrelati<strong>on</strong>s, rather than averaged behavior, in driving the behavior of the system.Even though the average magnetizati<strong>on</strong> for the B = 0 case is trivially zero in a <strong>on</strong>e-dimensi<strong>on</strong>alsystem, the B ≠ 0 behavior seems n<strong>on</strong>-trivial at first glance. However, it can be solved analytically.To see this we c<strong>on</strong>sider the partiti<strong>on</strong> functi<strong>on</strong>,Z = ∑σ 1···σ nexp β {J(⃗σ 1 · ⃗σ 2 + ⃗σ 2 · ⃗σ 3 + · · · ⃗σ n · ⃗σ 1 ) + µB(σ 1 + σ 2 · · · + σ n )} , (5.29)where each spin is coupled through and energy −J⃗σ i · ⃗σ i+1 to the adjacent spin, and the last spinis coupled back to the first spin, as if the lattice is bent into a ring. The energy can be expressedas a sum over the spin values,H = ∑ i−Jσ i σ i+1 − 1 2 µB(σ i + σ i+1 ), (5.30)and the partiti<strong>on</strong> functi<strong>on</strong> can then be written as a product of matrix elements,Z = ∑⟨σ 1 |P|σ 2 ⟩⟨σ 2 |P|σ 3 ⟩ · · · ⟨σ n |P|σ 1 ⟩, (5.31)σ 1···σ n{P i,i+1 = exp β Jσ i σ i+1 + 1 }2 µB(σ i + σ i+1 ) ,( )eβ(J+µB)e=−βJe −βJ e β(J−µB) .Here, P is known as the transfer matrix, where the states with σ i = ±1 are represented by a twocomp<strong>on</strong>entvector with elements either up or down respectively. The partiti<strong>on</strong> functi<strong>on</strong> can thenbe expressed asZ = Tr P n , (5.32)which due to the cyclic property of the trace can also be written asZ = Tr (UPU † ) n , (5.33)where U is a unitary transformati<strong>on</strong>. We will use the transformati<strong>on</strong> that diag<strong>on</strong>alizes P, UPU † =P λ , soZ = Tr (P λ ) n . (5.34)The elements of P λ are the eigenvalues of P,( )λ+ 0P λ =, (5.35)0 λ −√λ ± = e βJ cosh(βµB) ± e 2βJ sinh 2 (βµB) + e −2βJ .One can now write the partiti<strong>on</strong> functi<strong>on</strong> as(Z = Tr Pλ n (λ+ )= Tr0)0(λ − ) n = (λ + ) n + (λ − ) n . (5.36)89


As n → ∞, the sec<strong>on</strong>d term becomes unimportant as can be seen from expanding ln Z,[ ( ) n ]}[ ( ) n ]ln Z = ln{(λ + ) n λ−λ−1 += n ln(λ + ) + ln 1 + . (5.37)λ + λ +As n → ∞, the sec<strong>on</strong>d term vanishes since λ − < λ + . This gives the final result for the partiti<strong>on</strong>functi<strong>on</strong>,√ln Z = n ln(λ), λ = e βJ cosh(βµB) + e 2βJ sinh 2 (βµB) + e −2βJ . (5.38)EXAMPLE:In terms of J, µB and T , find the average magnetizati<strong>on</strong> (per spin) for the 1-d Ising model.Since the magnetic field enters the Hamilt<strong>on</strong>ian as −µ ⃗ B · ⃗σ, the average spin isThis gives,⟨σ⟩ ==⟨σ⟩ = 1 d ln Zn d(βµB) .cosh(βµB) sinh(βµB)sinh(βµB) + √sinh 2 (βµB)+e√−4βJcosh(βµB) + sinh 2 (βµB) + e −4βJsinh(βµB)√sinh 2 (βµB) + e −4βJ .Although the exact analytic soluti<strong>on</strong> <strong>on</strong>ly exists for <strong>on</strong>e-dimensi<strong>on</strong>, the technique of applyinga transfer matrix can be applied to more dimensi<strong>on</strong>s. For a finite two-dimensi<strong>on</strong>al lattice of sizeL x × L y , <strong>on</strong>e can treat the c<strong>on</strong>figurati<strong>on</strong> of each row as a state, with 2 Lx possible values. One canthen c<strong>on</strong>struct the (2 L x× 2 L x) transfer matrix between neighboring columns. This is discussed inChapter 15 of Huang.5.4 Lattice Gas, Binary Alloys, Percolati<strong>on</strong>...Aside from the multitude of theoretical treatments of spins <strong>on</strong> lattices, lattices are also used to modelliquid-gas phase transiti<strong>on</strong>s, percolati<strong>on</strong> and fragmentati<strong>on</strong>. Mainly to provide an introducti<strong>on</strong> tothe vocabulary, we present a list of several such models here. The first model we describe is thelattice gas. Here, a lattice of sites are assigned either zero (no particle) or 1 (occupied) with a setprobability. The c<strong>on</strong>figurati<strong>on</strong> is weighted according to the probability exp{−N nn ϵ 0 /T }, where N nnis the number of nearest neighbor sites. Effectively, this models a gas with a short-range attractiveattracti<strong>on</strong>, represented by the positive weight (assuming ϵ 0 < 0). Since the occupancy is nevergreater than unity, it also mimics a repulsive interacti<strong>on</strong> at short distances. This can be used tomodel the shapes of drops, which can become very irregular near the critical point.Binary alloys are more physically represented by lattices. Each site is assigned a specific species.The lattice is first c<strong>on</strong>structed according to the physical structure of the modeled system, e.g., itmight be body-centered cubic, before N a + N b = N atoms are assigned to the N lattice sites. Theentire c<strong>on</strong>figurati<strong>on</strong> is weighted according to e −βH whereH = ϵ aa N aa + ϵ bb N bb + ϵ ab N ab , (5.39)90


where N ij are the number of neighboring combinati<strong>on</strong>s where <strong>on</strong>e species is i and the other is j. Forthe case of two equally populated species <strong>on</strong> a BCC lattice, the lowest energy state is chemicallydivided into alternate planes. Above some critical temperature, the l<strong>on</strong>g-range order disappears.Like the lattice gas, the binary model and the Ising model all involve assigning <strong>on</strong>e of two numbers(0 or 1 for a lattice gas, ’1’ or ’2’ for a binary alloy, and ’-1’ or ’1’ for the Ising model) to lattice sites,then adding a weight which depends <strong>on</strong>ly nearest neighbor occupancies. Thus, all three approachescan at times be mapped <strong>on</strong>to <strong>on</strong>e another.Another class of calculati<strong>on</strong>s are called “percolati<strong>on</strong> models”. These models usually addressthe questi<strong>on</strong> of how l<strong>on</strong>g a given chain might extend across a lattice, given that some of the linksmight be broken. Percolati<strong>on</strong> models can be divided into b<strong>on</strong>d percolati<strong>on</strong> or site percolati<strong>on</strong>. Inb<strong>on</strong>d percolati<strong>on</strong>, a lattice is c<strong>on</strong>structed with potential b<strong>on</strong>ds between all nearest neighbors. Onlya fracti<strong>on</strong> p of b<strong>on</strong>ds are filled, with the b<strong>on</strong>ds chosen randomly. For a two-dimensi<strong>on</strong>al (L × L)or three-dimensi<strong>on</strong>al (L × L × L) lattice, <strong>on</strong>e can ask the questi<strong>on</strong> of whether an unbroken chaincan be found that extends completely across the lattice. As L → ∞, an infinite chain will exist ifand <strong>on</strong>ly if p > p c . The critical value p c depends <strong>on</strong> dimensi<strong>on</strong>ality and the type of lattice, e.g.,cubic or BCC, and whether <strong>on</strong>e is doing b<strong>on</strong>d percolati<strong>on</strong> of site percolati<strong>on</strong>. For site percolati<strong>on</strong>,the sites are populated with probability p, and b<strong>on</strong>ds are assumed to exist if neighboring sitesare filled. For b<strong>on</strong>d percolati<strong>on</strong> of a simple three-dimensi<strong>on</strong>al cubic lattice, p c ≈ 0.2488 (Seehttp://mathworld.wolfram.com/Percolati<strong>on</strong>Threshold.html).Percolati<strong>on</strong> is used to study electric properties and fragmentati<strong>on</strong>. Percolati<strong>on</strong> models havebeen applied to nuclear fragmentati<strong>on</strong>, and compared to data taken here at the NSCL. The sameapproaches have been compared to experiments where a glass is dropped and shattered and analyzedfor the probability of a given shard having a specific mass.5.5 Problems1. The speed of sound in copper is 3400 m/s, and the number density is ρ Cu = 8.34 × 10 28 m −3 .(a) Assuming there are two free electr<strong>on</strong>s per atom, find an expressi<strong>on</strong> for C V /N a (whereN a is the number of atoms) from the free electr<strong>on</strong>s in copper by assuming a free gas ofelectr<strong>on</strong>s. Use the expressi<strong>on</strong> from chapter 2 for a low-T Fermi gas,δE = T 2 π26 D(ϵ),where D(ϵ) is the density of single particle electr<strong>on</strong> states at the Fermi surface. Giveanswer in terms of T and the Fermi energy ϵ F .(b) In terms of ϵ F and ¯hω D , find an expressi<strong>on</strong> for the temperature at which the specific heatfrom electr<strong>on</strong>ic excitati<strong>on</strong>s equals that from ph<strong>on</strong><strong>on</strong>s. Use the low T expressi<strong>on</strong> for thespecific heat of ph<strong>on</strong><strong>on</strong>s.(c) What is ¯hω D in eV? in K?(d) What is ϵ F in eV? in K?(e) What is the numerical value for the answer in (b) in K?2. C<strong>on</strong>sider the mean-field soluti<strong>on</strong> to the Ising model of Eq. (5.18).91


(a) For small temperatures, show that the variati<strong>on</strong> δ⟨σ⟩ = 1 − ⟨σ⟩ is:δ⟨σ⟩ ≈ 2e −2qJ/T .(b) C<strong>on</strong>sider a functi<strong>on</strong> of the form, y(T ) = e −1/T . Find dy/dT , d 2 y/dT 2 , and d n y/dT nevaluated at T = 0 + . What does this tell you about doing a Taylor expansi<strong>on</strong> of y(T )about T = 0 + ?3. Show that in the mean field approximati<strong>on</strong> of the Ising model the susceptibility,becomesχ ≡ d⟨σ⟩dB ,χ = (1 − ⟨σ⟩2 )µT − T c + ⟨σ⟩ 2 T c.4. The total energy for the Ising model in the mean field approximati<strong>on</strong> from summing over allthe sites in Eq. (5.16) isH = − N 2 qJ⟨σ⟩2 − NµB⟨σ⟩,where N is the number of sites, and the factor of 1/2 is a correcti<strong>on</strong> for double counting. Interms of T, T c , µB and ⟨σ⟩, find an expressi<strong>on</strong> forC v = 1 N5. C<strong>on</strong>sider the <strong>on</strong>e-dimensi<strong>on</strong>al Ising model. For the following, give analytic answers in termsof T , µB and J.dEdT .(a) What are the high B and low B limits for ⟨σ⟩?(b) What are the high T and low T limits for ⟨σ⟩?(c) Find an exact expressi<strong>on</strong> for the specific heat (per spin) in terms of T , µB and J.(d) What are the high B and low B = 0 limits for the specific heat?(e) What are the high T and low T limits for the specific heat?92


6. C<strong>on</strong>sider two-dimensi<strong>on</strong>al b<strong>on</strong>d percolati<strong>on</strong> of a large L × L simple square lattice. Firstc<strong>on</strong>sider the red lattice, where we will remove a percentage p of the red b<strong>on</strong>ds. Next, we willremove all blue b<strong>on</strong>ds that intersect a surviving red b<strong>on</strong>d. Thus, in the limit of a large lattice,the fracti<strong>on</strong> of blue b<strong>on</strong>ds broken will be (1 − p).(a) List which combinati<strong>on</strong>s of the following are possible.a) A c<strong>on</strong>nected string of blue b<strong>on</strong>ds extends all the way from the bottom blue row tothe top of the lattice.b) There is no c<strong>on</strong>nected string of blue b<strong>on</strong>ds that extends all the way from the bottomto the top.c) A c<strong>on</strong>nected string of red b<strong>on</strong>ds extends all the way from the left side to the rightside of the red lattice.d) There is no c<strong>on</strong>nected string of red b<strong>on</strong>ds that extends all the way from the left sideto the right side of the red lattice.(b) What is p c for a simple square lattice in b<strong>on</strong>d percolati<strong>on</strong>?7. C<strong>on</strong>sider a <strong>on</strong>e-dimensi<strong>on</strong>al b<strong>on</strong>d percolati<strong>on</strong> model of an infinitely l<strong>on</strong>g string, where theprobability of any set of neighbors being c<strong>on</strong>nected is p.(a) In terms of p, find the probability that a fragment will have size A.(b) What is the average size of a fragment?93


6 Landau Field TheoryNothing is real and nothing to get hung about - J. Lenn<strong>on</strong>, P. McCartney6.1 What are fields?Field theory represents a means by which the state of a system is expressed in terms of a quantityϕ which is a functi<strong>on</strong> of the positi<strong>on</strong> ⃗x. The quantity ϕ represents a measure of some quantityaveraged over a small volume δV . One example of such a quantity is the average magnetizati<strong>on</strong>per unity volume, ⃗m(⃗x), while another is the number density ρ(⃗x). In each case, when viewedmicroscopically, the magnetizati<strong>on</strong> density and the number density are not smooth quantities, butare carried by point-like objects. Thus, field theories are <strong>on</strong>ly applicable when looking at averagesor correlati<strong>on</strong>s <strong>on</strong> scales larger than the microscopic scale of the matter. As will be seen furtherahead, correlati<strong>on</strong> length scales tend to go to infinity near a critical point, a regi<strong>on</strong> where fieldtheories are especially valuable.Finite-temperature field theories are usually expressed in terms of the free-energy density, f(⃗x),∫F ≡ d 3 x f(ϕ, ∇ϕ), (6.1)f = κ 2 (∇ϕ)2 + V(ϕ).The sec<strong>on</strong>d statement, that derivatives appear <strong>on</strong>ly quadratically, is certainly an ansatz, but as wewill see below, the assumpti<strong>on</strong> is motivated from physical arguments. The choice of the symbol Ffor the free energy should not suggest that <strong>on</strong>e is always interested in the Helmholtz free energy, asfor some applicati<strong>on</strong>s it might represent the Gibb’s free energy or the entropy. If the parameter ϕchanges slowly compared to the time required to come to thermal equilibrium (c<strong>on</strong>stant temperature),the use of the Helmholtz free energy is justified, and if the pressure also equilibrates quicklycompared to ϕ, the Gibb’s free energy would be the appropriate choice.The gradient term is warranted whenever there is an attractive short-range interacti<strong>on</strong>. As anexample, <strong>on</strong>e can look at the two-dimensi<strong>on</strong>al X − Y model, where spins <strong>on</strong> a lattice are denotedby an angle θ, with the following interacti<strong>on</strong> between nearest neighbors,For small differences in the angle this can be approximated as:V = V 02= V 02= V 02∑V ij = V 0 [1 − cos(θ i − θ j )]. (6.2)(θ ix − θ ix−1) 2 + (θ iy − θ iy−1) 2 (6.3)i x ,i y∑ [(θix − θ ix−1) 2 + (θ iy − θ 2] (∆L) 2iy−1)(∆L) 2i x ,i y∫dx dy (∇θ) 2 ,where ∆L is the lattice spacing. Thus, in this example, it is easy to identify κ with the microscopicquantity V 0 . Since the free energy is F = E −TS, terms that c<strong>on</strong>tribute to the energy density carrythrough to the free energy density.94


A sec<strong>on</strong>d example is that of a liquid, or gas, where ϕ refers to the density ρ(⃗x). In that case theattractive interacti<strong>on</strong> is mainly the two-particle potential v(⃗r),V = 1 ∫d 3 x d 3 x ′ ⟨ρ(⃗x)ρ(⃗x ′ )⟩v(⃗x − ⃗x ′ ). (6.4)2For a field theory, we are interested in an average density at a given point ¯ρ(⃗x). In terms of theaverages, the two-body potential energy isV = 1 ∫d 3 x d 3 x ′ ¯ρ(⃗x)¯ρ(⃗x ′ )C(⃗x, ⃗x ′ )v(⃗x − ⃗x ′ ), (6.5)2C(⃗x, ⃗x ′ ) ≡ ⟨ρ(⃗x)ρ(⃗x′ )⟩¯ρ(⃗x)¯ρ(⃗x ′ ) .Here, C(⃗r) is the correlati<strong>on</strong> (unity if uncorrelated) for two particles to be separate by ⃗r. In thelow density limit (particles interacting <strong>on</strong>ly two-at-a-time) the correlati<strong>on</strong> is approximatelyNext, <strong>on</strong>e can make a gradient expansi<strong>on</strong> for ¯ρ,C(⃗x, ⃗x ′ ) ≈ e −v(⃗r)/T . (6.6)¯ρ(⃗x ′ ) ≈ ¯ρ(⃗x) + (x ′ i − x i )∂ i ¯ρ(⃗x) + 1 ∑(x ′ i − x i )(x ′ j − x j )∂ i ∂ j ¯ρ(⃗x). (6.7)2Since V (⃗r) is an even functi<strong>on</strong> of ⃗r, the odd terms in the expansi<strong>on</strong> can be discarded leaving∫ {V = d 3 x V local (¯ρ(⃗x)) − κ }2 ¯ρ(⃗x)∇2 ¯ρ(⃗x)(6.8)κ = − 1 ∫d 3 r r 2 e −v(r)/T v(r), (6.9)6∫V local (¯ρ) = ¯ρ2 d 3 r e −v(r)/T v(r),2where the approximati<strong>on</strong> for the correlati<strong>on</strong>, Eq. (6.6), has been used for the correlati<strong>on</strong> functi<strong>on</strong>.The factor of 1/3 in the expressi<strong>on</strong> for κ comes from assuming that v(⃗r) is rotati<strong>on</strong>ally symmetric,so that the integral with rx 2 is <strong>on</strong>e third the integral with r 2 . The term V local is the c<strong>on</strong>tributi<strong>on</strong> tothe potential energy when density gradients are neglected. We will drop the subscript “local” andsimply c<strong>on</strong>sider V(¯ρ) to be the free energy density of a system at uniform density. Assuming thatother c<strong>on</strong>tributi<strong>on</strong>s to the energy do not c<strong>on</strong>tribute to the ∇ 2 term, the free energy becomes:F ==∫∫d 3 xd 3 xi,j[V(¯ρ(⃗x)) − κ ]2 ¯ρ(⃗x)∇2 ¯ρ(⃗x)(6.10)[V(¯ρ(⃗x)) + κ 2 (∇¯ρ)2 ]. (6.11)where the last step involved integrating by parts. Since the definiti<strong>on</strong> of density usually implies anaverage the “bar” over ¯ρ can be dropped .An obvious questi<strong>on</strong> c<strong>on</strong>cerns how <strong>on</strong>e might justify dropping higher order gradients in theTaylor expansi<strong>on</strong> of ρ(⃗x ′ ). Certainly, such terms should exist. Calculating such higher-order terms95


would involve finding expressi<strong>on</strong>s for coefficients like were found for κ in Eq. (6.9) above. However,rather than <strong>on</strong>e extra power of r 2 as in Eq. (6.9), <strong>on</strong>e would find powers of r 4 , r 6 · · · . If the potentialv(r) falls off exp<strong>on</strong>entially with a sufficiently short range, these terms can be neglected, and evenif they fall off with a power-law (such as 1/r 6 for the Lennard-J<strong>on</strong>es potential) screening effectsoften give an exp<strong>on</strong>ential fall-off which results in exp<strong>on</strong>ential behavior for large r, which keeps thecoefficients finite. If <strong>on</strong>e follows through with the algebra for the extra terms, they will fall off asa ratio of the interacti<strong>on</strong> range to the characteristic length for the density changing, (R int /L char ) n .Thus, the higher-order terms can <strong>on</strong>ly be neglected if <strong>on</strong>e is c<strong>on</strong>sidering variati<strong>on</strong>s of the density<strong>on</strong> length scales much greater than the range of the interacti<strong>on</strong>.One comm<strong>on</strong> mistake students make when seeing a mean-field theory is to assume that thegradient term originated from kinetic terms in the microscopic equati<strong>on</strong>s of moti<strong>on</strong>. Instead, suchterms originate from attractive short-range interacti<strong>on</strong>s. In fact, if the interacti<strong>on</strong>s were repulsiveκ would have the wr<strong>on</strong>g sign. The terms do, however, represent kinetic-energy type terms in thefree energy. For instance, if V is quadratic, <strong>on</strong>e can express the free energy as a sum of independentc<strong>on</strong>tributi<strong>on</strong>s from “momentum” modes where δρ ∼ sin(kx). In that case, the ∇ 2 term in the freeenergy behaves as k 2 , just as it would for free particles or sound modes.We re-emphasize that the applicati<strong>on</strong> of field theories is based <strong>on</strong> the assumpti<strong>on</strong> that thecharacteristic length of the fluctuati<strong>on</strong>s is much larger than the range of interacti<strong>on</strong>. It is alsoassumed that the scale is greater than the characteristic thermal wavelength. If <strong>on</strong>e is c<strong>on</strong>sideringscales of the order of the thermal wavelength, which is inherently quantum mechanical, λ therm ∼¯h/ √ mT , <strong>on</strong>e should use a quantum approach such as density-functi<strong>on</strong>al theory.Quantum field theories are also often required to study dynamical excitati<strong>on</strong>s, such as ph<strong>on</strong><strong>on</strong>s.Quantum field theories are really extremely different from the classical, Landau, theories describedhere. Here, the fields are real quantities like density, whereas quantum fields typically need to besquared to represent an observable quantity. Also, the gradient terms in quantum field theories aretypically related to kinetic energy.6.2 Calculating surface energies in field theoryAs discussed in Chapter 3, two phases can co-exist in equilibrium. However, there is a penaltyassociated with the boundary. This is often referred to as a surface energy or a surface free energy.To calculate the surface free energy, we c<strong>on</strong>sider a large box where the left side (x > 0) is in a liquid state. If A is the area of the y − z plane,the total Helmholtz free energy (F = E − T S) is∫F/A = dx[V(ρ(x)) + κ ]2 (∂ xρ) 2 . (6.12)Our goal is to find the density profile that minimizes the free energy given the c<strong>on</strong>straint that ithas the density of the gas at x = −∞ and the density of the liquid at x = ∞. We will assume thatthe temperature is a c<strong>on</strong>stant over the profile, which is reas<strong>on</strong>able if energy moves more quicklythan particles. By having the system able to c<strong>on</strong>nect to particle baths at ±∞, <strong>on</strong>e can statethat the chemical potential should be chosen so that µ(x → ∞) = µ(x → −∞) = µ 0 , with µ 0being chosen so that the two phases are at equilibrium, i.e., they also have the same pressure,P (x → ∞) = P (x → −∞) = P 0 .The Helmholtz free energy is the quantity <strong>on</strong>e should minimize if a system is c<strong>on</strong>nected to aheat bath and the number of particles in the box is fixed. If the system can also exchange particles96


with a bath whose chemical potential is µ 0 , <strong>on</strong>e should include the free energy associated with theparticles that have been added or subtracted into the bath. In that case, the number is no l<strong>on</strong>gerc<strong>on</strong>served, and the quantity to be minimized is∫F − µ 0 N = A dx[V(ρ) − µ 0 ρ + κ ]2 (∂ xρ) 2 (6.13)∫= Adx[−P (ρ) + (µ − µ 0 )ρ + κ 2 (∂ xρ) 2 ]. (6.14)The last step used the thermodynamic identity that the free energy density for a uniform gas isV = µρ − P . Figure 6.1 illustrates a plot of −P + (µ − µ 0 )ρ as a functi<strong>on</strong> of the density forfixed temperature. By picking the appropriate value of µ 0 for the specified temperature T , the twominima are at the same level and phase coexistence can occur. In fact, at the minima V −µ 0 ρ = −P ,and the statement that both minima are equal is equivalent to saying that for the specified T , µ 0was chosen so that two densities have the same pressure.Figure 6.1: Here, P (ρ, T ) and µ(ρ, T ) are thepressure and chemical potential of a gas at uniformdensity ρ and temperature T . If the systemis c<strong>on</strong>nected to a heat bath with chemical potentialµ 0 , the system will choose a density tominimize −P + (µ − µ 0 )ρ. If µ 0 equals the chemicalpotential for phase coexistence, the two minimaare degenerate and are located at the coexistencedensities, ρ gas and ρ liq for the given T . Aboundary between the two regi<strong>on</strong>s must traversethe regi<strong>on</strong> of less favorable free-energies, hencethere is a surface free energy as described in Eq.(6.18). In this expressi<strong>on</strong>, the surface energy isfound by integrating over the densities betweenthe two coexistence values with the integrand beingproporti<strong>on</strong>al to the square root of the height,illustrated by the dashed lines.is,The cost of the net free energy for a profile ρ(x) that differs from the two coexistence densities∆F/A ==∫∫dxdx[−P + (µ − µ 0 )ρ + κ 2 (∂ xρ) 2] ∫−[P 0 − P + (µ − µ 0 )ρ + κ 2 (∂ xρ) 2 ],dx (−P 0 ) (6.15)where P 0 is the coexistence pressure for the temperature T . The last term, ∫ dx P 0 , is a c<strong>on</strong>stant forfixed limits, and is subtracted so that the integrand goes to zero at large x, allowing the integrati<strong>on</strong>regi<strong>on</strong> to be stated as −∞ < x < ∞. The quantity (µ − µ 0 )ρ + P 0 − P is represented by the heightof the lines in the shaded regi<strong>on</strong> of Fig. 6.1. After replacing dx(∂ x ρ) 2 = dρ(∂ x ρ), the integrati<strong>on</strong>can be restated in terms of the densities,∫ ρliq[P0 − P + (µ − µ 0 )ρ∆F/A = dρ+ κ ]∂ x ρ 2 ∂ xρ . (6.16)ρ gas97


Next, <strong>on</strong>e chooses ∂ x ρ to minimize the free energy. Whereas the first term would be minimizedby choosing a step functi<strong>on</strong>, ∂ x ρ = ∞, the sec<strong>on</strong>d term would be minimized by a smooth profile.Minimizing the functi<strong>on</strong>al integral yields,dd(∂ x ρ)[P0 − P + (µ − µ 0 )ρ+ κ ]∂ x ρ 2 ∂ xρ = 0, (6.17)√∂ x ρ = 2 P 0 − P + (µ − µ 0 )ρ,κ∆F/A = √ 2κ∫ ρliqρ gasdρ √ P 0 − P + (µ − µ 0 )ρ.Thus, the coefficient κ is resp<strong>on</strong>sible for the surface free energy. If there were no gradient termin the functi<strong>on</strong>al for the free energy, the density profile between equilibrated phases would becomea step functi<strong>on</strong>. Looking back at the expressi<strong>on</strong> for κ in the previous secti<strong>on</strong> shows that surfaceenergies will be larger for interacti<strong>on</strong>s with l<strong>on</strong>ger ranges.EXAMPLE:Let the functi<strong>on</strong> ∆Ψ ≡ P 0 − P + (µ − µ 0 )ρ be approximated with the following form,∆Ψ = B 2[(ρ − ρc ) 2 − α 2] 2.Find the surface energy in terms of B, α and κ.From Eq. (6.18), the surface energy isF/A = √ κB= √ κB∫ ρ−ρc =αρ−ρ c =−α∫ α−α= 4 3 α3√ κBdρdx (α 2 − x 2 )√[(ρ − ρ c ) 2 − α 2 ] 2Note that B, κ and α 2 would depend <strong>on</strong> the temperature. Since two independent minima no l<strong>on</strong>gerexist for T > T c , the coefficient α 2 goes to zero at T c , which means that the surface energy alsobecomes zero.6.3 Correlati<strong>on</strong>s and Susceptibilities in the Critical Regi<strong>on</strong>Correlati<strong>on</strong>s, e.g.,C(⃗r − ⃗r ′ ) ≡ ⟨(ϕ ∗ (⃗r) − ¯ϕ ∗ )(ϕ(⃗r ′ ) − ¯ϕ)⟩, (6.18)¯ϕ ≡ ⟨ϕ⟩,are central to the study of critical phenomena. Note that here we have assumed a complex field,which has two independent comp<strong>on</strong>ents ϕ r and ϕ i . For this case,⟨ϕ ∗ (⃗r)ϕ(⃗r ′ )⟩ = ⟨ϕ r (⃗r)ϕ r (⃗r ′ )⟩ + ⟨ϕ i (⃗r)ϕ i (⃗r ′ )⟩. (6.19)98


Thus, if <strong>on</strong>e were to solve for correlati<strong>on</strong>s of a <strong>on</strong>e-comp<strong>on</strong>ent (real) field, <strong>on</strong>e would divide theexpressi<strong>on</strong> derived here for complex fields by a factor of two. We work with complex fields <strong>on</strong>lybecause the Fourier transforms are somewhat less painful.Near the critical point the characteristic length of such correlati<strong>on</strong>s approaches infinity. Todem<strong>on</strong>strate this, we c<strong>on</strong>sider a purely quadratic functi<strong>on</strong>al, as the correlati<strong>on</strong>s can be found analytically.The effect of higher-order terms alters the behavior quantitatively (e.g., different valuesfor the critical exp<strong>on</strong>ents), but does not alter the qualitative behavior shown here. For quadraticpotentials,V(ϕ) = A 2 ϕ2 , (6.20)the problem can be divided into individual modes:∫ ∫ [ AF = d 3 r f = d 3 r2 |ϕ|2 + κ ]2 |∇ϕ|2= 1 2∑(6.21)⃗ k(A + κk2 ) ∣ ∣ ∣ ˜ϕ⃗k∣ ∣∣2, (6.22)where the Fourier and inverse-Fourier transforms are defined by˜ϕ ⃗k ≡ √ 1 ∫d 3 r e i⃗k·⃗r ϕ(⃗r), (6.23)Vϕ(⃗r) = 1 √V∑⃗ ke −i⃗ k·⃗r ˜ϕ⃗k .Since the modes are independent, and therefore uncorrelated, the correlati<strong>on</strong> in terms of momentumcomp<strong>on</strong>ents is straight-forward,⟨ ˜ϕ ⃗ ∗ ˜ϕ⃗k k′⟩ = δ ⃗k, ⃗ k ′⟨ ∣ ˜ϕ∣ ∣∣2⃗k ⟩ (6.24)2T= δ ⃗k, ⃗ k ′(A + κk 2 ) ,where the equipartiti<strong>on</strong> theory was used for the last step, i.e., if the free energy behaves as α|ϕ| 2 ,then α⟨|ϕ| 2 ⟩ = T . The reas<strong>on</strong> it is not T/2 comes from the fact ϕ is complex and |ϕ| 2 = ϕ 2 r + ϕ 2 i .One can now calculate the correlati<strong>on</strong> in coordinate space,⟨ϕ ∗ (⃗r)ϕ(⃗r ′ )⟩ = 1 V= 1 V∑e i⃗ k·⃗r−i ⃗ k ′·⃗r ′ ⟨ ˜ϕ ⃗ ∗ ˜ϕ⃗k k′⟩ (6.25)⃗ k, ⃗ k ′∑⃗ ke i⃗ k·(⃗r−⃗r ′ ) 2T(A + κk 2 ) .Here, the last term employed the equipartiti<strong>on</strong> theorem. Now, the sum over modes can be transformedto an integral,∫⟨ϕ ∗ (⃗r)ϕ(⃗r ′ 1)⟩ = d 3 k e i⃗ k·(⃗r−⃗r ′ ) 2T(2π) 3 (A + κk 2 ) . (6.26)99


This integral can be performed by first integrating over the angle between ⃗ k and (⃗r − ⃗r ′ ), thenperforming a c<strong>on</strong>tour integrati<strong>on</strong>,⟨ϕ ∗ (⃗r)ϕ(⃗r ′ = 0)⟩ ==∫ ∞1Tkdk sin(kr)2π 2 r −∞(A + κk 2 )T2πκr e−r/ξ , ξ = √ κ/A.(6.27)Here, ξ is referred to as the correlati<strong>on</strong> length. Near the critical point, the curvature of the freeenergyw.r.t. the order parameter ϕ switches sign, and as A goes to zero, ξ → ∞. The fact that thislength becomes much larger than any driving microscopic length legitimizes the scaling argumentsused in the discussi<strong>on</strong> of critical phenomena in the next secti<strong>on</strong>.In the case where ϕ refers to the magnetizati<strong>on</strong> density m(⃗r), the correlati<strong>on</strong> functi<strong>on</strong> alsoprovides the susceptibility,χ ≡ d ⟨m(⃗r = 0)⟩ (6.28)dB= d Tr e −β[H 0+ ∫ d 3rµBm(⃗r)] m(⃗r = 0)dB Tr e −β[H 0+ ∫ d 3 rµBm(⃗r)]∫ [d 3 r Tr e −βH m(⃗r)m(⃗r = 0) Tre −βH m(⃗r = 0) ] [∫ d 3 r Tr e −βH m(⃗r) ]= µ T= µ ∫T= µ ∫T= µ ∫κTre −βH− µ T[Tr e −βH ] 2d 3 r { ⟨m(⃗r)m(⃗r = 0)⟩ − ¯m 2} (6.29)d 3 r {⟨(m(⃗r) − ¯m)(m(⃗r = 0) − ¯m)⟩}rdr e −r/ξ= µ ξ2κ ,where Eq. (6.27) was inserted into the expressi<strong>on</strong> for the correlati<strong>on</strong> in the sec<strong>on</strong>d-to-last step, anda factor of 1/2 was added to account for the fact that this is not a complex field. Relati<strong>on</strong>s suchas this are comm<strong>on</strong> in linear resp<strong>on</strong>se theory, and often come under the catch-all phrase of thefluctuati<strong>on</strong>-dissipati<strong>on</strong> theorem. Linear resp<strong>on</strong>se theory c<strong>on</strong>siders the change of an observable ⟨δA⟩due to a small field that couples to a sec<strong>on</strong>d observable B, H int = BδF . This dependence can beexpressed in terms of a generalized susceptibility, χ, where δA = χδF defines χ. The central resultof linear resp<strong>on</strong>se theory is that χ can be expressed in terms of correlati<strong>on</strong>s of the type ⟨A(0)B(r)⟩.Although we will not have the time to cover linear resp<strong>on</strong>se theory, I recommend reading the firsttwo chapters of Forster 1 at some point in your life. Equati<strong>on</strong> (6.29) will be important in the nextsecti<strong>on</strong>s as it will be used to c<strong>on</strong>nect the divergent behavior of the correlati<strong>on</strong> functi<strong>on</strong> at the criticalpoint to other quantities such as the susceptibilities and specific heat.1 D. Forster, Hydrodynamic Fluctuati<strong>on</strong>s, Broken Symmetry, and Correlati<strong>on</strong> Functi<strong>on</strong>s100


6.4 Critical Exp<strong>on</strong>ents in Landau TheoryThe curvature used in the free energy A changes sign at the critical point. One can parameterizethis as:A(T ) = at, t ≡ (T − T c )/T c . (6.30)Thus, the correlati<strong>on</strong> length diverges as t −1/2 ,ξ =√ κA = √ κ/a t −1/2 , (6.31)χ = µ a t−1The exp<strong>on</strong>ents describing the divergences here are known as critical exp<strong>on</strong>ents. In Landau theory,all three-dimensi<strong>on</strong>al systems would have the same critical exp<strong>on</strong>ents,ξ ∼ |t| −ν , ν = 1/2, (6.32)χ ∼ |t| −γ , γ = 1.The specific heat is also divergent, but to see that we must first calculate the free energy.Equati<strong>on</strong> (6.21) gives the free energy for a particular set of amplitudes ϕ 1 , ϕ 2 · · · . To calculate thefree energy of the entire system, <strong>on</strong>e must sum over all the possible values of ϕ i ,∑F tot = −T lne −F (ϕ 1,ϕ 2··· )/T , (6.33)all c<strong>on</strong>figurati<strong>on</strong>s ϕ 1 ,ϕ 2···where F can be written as a sum over individual modes ⃗ k,( )∑e −F k(ϕ k )/Tϕ kF tot = −T ∑ ⃗ kln. (6.34)As was shown in Chapter 3, the sum over states ϕ k can be written as an integral over the real andimaginary parts of ϕ k , so that the partiti<strong>on</strong> functi<strong>on</strong> becomese −F k/T= 1 ∫dϕ r dϕ i e −(1/2)(A+κk2 )|ϕ k | 2 , (6.35)2π∼ (A + κk 2 ) −1 ,where the last expressi<strong>on</strong> ignores additive c<strong>on</strong>stants. Finally, the total free energy of the systemwill beF tot = T ∑ ⃗ kln(A + κk 2 ). (6.36)After inserting A = at, <strong>on</strong>e can solve for the c<strong>on</strong>tributi<strong>on</strong> to the specific heat from the temperaturedependence in A. Since ln Z = −F/T , <strong>on</strong>e can first take a derivative of F/T w.r.t. t to get anexpressi<strong>on</strong> for the energy density, then take a sec<strong>on</strong>d derivative to find the specific heat. The resultsisC V /V =∫a2(2π) 3d 3 k1011(at + κk 2 ) 2 ∼ t −1/2 . (6.37)


This last term represents another critical exp<strong>on</strong>ent,C V ∼ t −α , α = 1/2. (6.38)Another critical comp<strong>on</strong>ent describes the rise of the mean magnetizati<strong>on</strong> for temperatures justbelow T c . In our approximati<strong>on</strong>, we c<strong>on</strong>sider <strong>on</strong>ly small fluctuati<strong>on</strong>s about the free energy minimum,and our choice of the mean magnetizati<strong>on</strong> comes from the mean-field approximati<strong>on</strong> from Eq. (5.19)in Chapter 5,⟨m⟩ ∼ |t| β , β = 1/2. (6.39)Finally, for an external field with T = T c , mean field theory gives (see the homework problem inthis chapter)⟨m⟩ ∼ (µB) 1/δ , δ = 3. (6.40)The divergence of the specific heat is related to the divergence of the integrand in Eq. (6.37)when t is set to zero. Note that for d dimensi<strong>on</strong>s, the integrand goes as dk k (d−1) /k 4 , which will bedivergent for d ≤ 3. Thus, in four dimensi<strong>on</strong>s or higher, the specific heat is well defined at t → 0,whereas the behavior is more complicated for the two- and three-dimensi<strong>on</strong>al cases.Summarizing the critical exp<strong>on</strong>ents in Landau theory,⟨ϕ(0)ϕ(r)⟩ ∼ ξ 2−d (r/ξ) (1−d)/2 e −r/ξ , ξ ∼ t −ν , ν = 1/2, (6.41)C v ∼ t −α , α = 2 − d/2,< m > (t = 0) ∼ H 1/δ , δ = 3,< m > (H = 0, t < 0) ∼ t β , β = 1/2,χ ∼ t −γ , γ = 1,where d is the dimensi<strong>on</strong>ality. We will see in the next two secti<strong>on</strong>s that Landau theory is notvalid near T c for d ≤ 3. N<strong>on</strong>etheless, the features of Landau theory are qualitatively correct, andmost importantly, the correct method for visualizing critical behavior requires Landau theory forperspective.6.5 Validity of Landau Theory Near T c : The Ginzburg CriteriaThe exp<strong>on</strong>ents above were based <strong>on</strong> an assumpti<strong>on</strong> that <strong>on</strong>e could ignore the quartic piece of thefree energy density, which would <strong>on</strong>ly be justified if fluctuati<strong>on</strong>s of the order parameter were smallcompared to the mean value. To see whether that is the case, we c<strong>on</strong>sider the quartic casef(x) = A 2 |ϕ|2 + B 4 |ϕ|4 , (6.42)where A = at. For t < 0, A is negative and two distinct minima develop⟨|ϕ|⟩ 2 = −A/B = −at/B. (6.43)We will c<strong>on</strong>sider the case for t > 0, which means that the minimum is located at ϕ = 0, butthat as t → 0, the fluctuati<strong>on</strong>s will become large. The correlati<strong>on</strong> takes <strong>on</strong> the form,⟨ϕ(0)ϕ(r)⟩ ∼ ξ 2−d (r/ξ) (1−d)/2 e −r/ξ . (6.44)102


This expressi<strong>on</strong> was derived exactly for three dimensi<strong>on</strong>s previously, see Eq. (6.27). The integralsfor the three-dimensi<strong>on</strong>al integrati<strong>on</strong> can be generalized for an arbitrary number of dimensi<strong>on</strong>s din the limit that r is large, resulting in the expressi<strong>on</strong> above. If <strong>on</strong>e uses r = ξ as a characteristiclength scale for comparis<strong>on</strong>, the typical size for fluctuati<strong>on</strong>s are:⟨δϕδϕ⟩ typical ∼ T κ ξ2−d , (6.45)where dimensi<strong>on</strong>less c<strong>on</strong>stants such as (2π) d are ignored. For mean-field theory to be valid, <strong>on</strong>emust justify ignoring the quartic piece in the free energy above, orA2 ⟨δϕδϕ⟩ typical >> B 4 ⟨δϕδϕ⟩2 typical, (6.46)A2>> B 4 ⟨δϕδϕ⟩ typical,at >> B T κ ξ2−d .Sincethe criteria for validity becomesξ ∼ √ κ/at −1/2 , (6.47)t 2−d/2 >> BT c κ −d/2 a −2+d/2 . (6.48)As the critical point is approached t → 0, the criteria will be satisfied for d > 4. However, meanfield theory is unjustified in the critical regi<strong>on</strong> for three dimensi<strong>on</strong>s or fewer.Despite the fact that the theory is unjustified, Landau theory is still enormously useful. First,the theory is still justified away from the critical point, depending <strong>on</strong> the parameters. Sec<strong>on</strong>dly,it is qualitatively correct in many aspects and represents a crucial perspective from which <strong>on</strong>e canunderstand the more correct approaches based <strong>on</strong> scaling and renormalizati<strong>on</strong>, which are also fieldtheories, but differ in that they c<strong>on</strong>sider the effects of the ϕ 4 term.6.6 Critical Phenomena, Scaling and Exp<strong>on</strong>entsPhase transiti<strong>on</strong>s can be classified as either first or sec<strong>on</strong>d order. First-order transiti<strong>on</strong>s are characterizedby co-existing phases with the order parameter being different in the two phases. Insec<strong>on</strong>d-order transiti<strong>on</strong>s the derivative (or perhaps n th -order derivative) of the order parameter isdisc<strong>on</strong>tinuous. The term “order parameter” refers to any measure, not even necessarily observable,used to describe a system in a given state. Examples are the magnetizati<strong>on</strong> density (ferromagnetictransiti<strong>on</strong>), number density (liquid-gas transiti<strong>on</strong>), quark-antiquark c<strong>on</strong>densate (chiral transiti<strong>on</strong>),Higg’s c<strong>on</strong>densate (electro-weak transiti<strong>on</strong>), or field phase (superc<strong>on</strong>ductivity). Quantities thatinvolve comparing systems with different c<strong>on</strong>diti<strong>on</strong>s, such as the specific heat, which involves comparingthe entropy at two different temperatures, or the susceptibility, which involves comparingthe magnetizati<strong>on</strong> at two different external fields, are not order parameters.Transiti<strong>on</strong>s can bel<strong>on</strong>g to any of several universality classes, a grouping determined by thesymmetry associated with the phase transiti<strong>on</strong>. For instance, in the X − Y model the systembreaks two-dimensi<strong>on</strong>al rotati<strong>on</strong>al symmetry by choosing a specific directi<strong>on</strong> for aligning spins. Thethree-dimensi<strong>on</strong>al analog, the Heisenberg model, has a higher degree of symmetry, hence bel<strong>on</strong>gs to103


a different universality class and has different critical exp<strong>on</strong>ents, which will be introduced furtherbelow.At T c Landau’s field theory gave the following form for the correlati<strong>on</strong> functi<strong>on</strong> for magnetizati<strong>on</strong>density in the previous secti<strong>on</strong>,Γ(r) ≡ ⟨m(⃗r)m(r = 0)⟩ ∼ 1 r e−r/ξ , ξ = √ κ/A, (6.49)where the free energy density was treated in the Gaussian approximati<strong>on</strong>,f ≈ A 2 m2 + κ 2 (∇m)2 . (6.50)Here, A is a functi<strong>on</strong> of temperature. At T c , the sign of A will switch, and the correlati<strong>on</strong> lengthdiverges as√ κξ =dA/dT (T − T c) −1/2 ∼ t −1/2 , t ≡ T − T c. (6.51)T cIn interacting theories, interacti<strong>on</strong>s change the behavior of ξ near T c quantitatively, but not qualitatively.Quantum aspects of the interacti<strong>on</strong>s lead to screening effects that can be understood inrenormalizati<strong>on</strong> calculati<strong>on</strong>s, but are bey<strong>on</strong>d the scope of this class (See Chapter 18 in Huang).The resulting form for the correlati<strong>on</strong> functi<strong>on</strong> near T c is of the form,Γ(r) ∼ r 2−d+η e −r/ξ , ξ ∼ t −ν . (6.52)Here, d is the number of dimensi<strong>on</strong>s. In the three-dimensi<strong>on</strong>al Landau calculati<strong>on</strong>, d = 3, the twocritical exp<strong>on</strong>ents areη = 0, ν = 1/2. (6.53)However, in the real world they vary from this amount, depending <strong>on</strong> the universality class.The other four critical exp<strong>on</strong>ents used in the last secti<strong>on</strong> (α, β, γ, δ) are all determined by η andν. The relati<strong>on</strong>s can best be understood from dimensi<strong>on</strong>al analysis. These arguments are based <strong>on</strong>two assumpti<strong>on</strong>s:1. The n<strong>on</strong>-analytic part of the free energy density scales proporti<strong>on</strong>al to ∼ 1/ξ 3 .From dimensi<strong>on</strong>al grounds, if ξ is the <strong>on</strong>ly relevant length scale, and since the free-energydensity has dimensi<strong>on</strong>s of Energy/L d , it must behave as ξ −d , or equivalently as t dν . This isbased <strong>on</strong> a scaling argument that if <strong>on</strong>e looks at a fluctuati<strong>on</strong> of size X = Cξ at temperature t,it will have the same free energy as a fluctuati<strong>on</strong> of size X ′ = Cξ ′ at temperature t ′ . Assumingξ behaves as t −ν , the free energy scales as t dν , and the specific heat, which <strong>on</strong>e obtains bytaking two derivatives w.r.t. t, behaves as t −(2−dν) .2. The magnetizati<strong>on</strong> and fields all behave as if they have effective dimensi<strong>on</strong>alities, such thatchanging t, which causes a scale change of ξ, will scale all quantities according to their effectivedimensi<strong>on</strong>ality. Thus if ⟨m⟩ behaves as t β , Γ must behave as t 2β if evaluated at the same valueof r/ξ, which given the form of the correlati<strong>on</strong> functi<strong>on</strong> in Eq. (6.52), requires thatΓ(r/ξ = 1, t) ∼ t 2β , (6.54)ξ d+η−2 ∼ t 2β ,t ν(d+η−2) ∼ t 2β ,β = 1 ν(d + η − 2).2104


One would have obtained the same result for any choice of r as l<strong>on</strong>g as the comparis<strong>on</strong> (beforeand after the temperature change) was d<strong>on</strong>e at the same value of r/ξ. The field H must scaleso that the free energy in a fluctuati<strong>on</strong> of characteristic size, mHξ 3 , must be independent oft. This means that if H ∼ m δ , and m ∼ t β , thenm 1+δ ξ 3 ∼ t 0 , (6.55)t β(1+δ) ξ dν ∼ t 0 ,β(1 + δ) − dν = 0,δ = d − η + 2d + η − 2Finally, the susceptibility must have the same scaling as m/H, orχ ∼ m 1−δ ∼ t β(1−δ) (6.56)−γ = β(1 − δ) = ν(2 − η).Summarizing these arguments, <strong>on</strong>e can express critical exp<strong>on</strong>ents as follows:⟨ϕ(0)ϕ(r)⟩ ∼ r 2−d−η e −r/ξ , ξ ∼ t −ν , (6.57)C v ∼ t −α , α = 2 − νd,< m > (t = 0) ∼ H 1/δ , δ = d − η + 2d + η − 2 ,< m > (H = 0, t < 0) ∼ t β , β = ν(d + η − 2)/2,χ ∼ t −γ , γ = ν(2 − η),where <strong>on</strong>ly two of the six exp<strong>on</strong>ents are independent. Here, we have used the two exp<strong>on</strong>ents usedto describe the behavior of the correlati<strong>on</strong> functi<strong>on</strong>, η and ν, to define the remaining four. Since thecorrelati<strong>on</strong> functi<strong>on</strong> Γ(r) would nominally have dimensi<strong>on</strong>s of L 2−d , the exp<strong>on</strong>ent η is often referredto as the anomalous dimensi<strong>on</strong>. It is noteworthy that the scaling relati<strong>on</strong>s above are violated inLandau theory, as can be seen by comparing the mean-field value, β = 1/2, to the expressi<strong>on</strong> forβ above. This stems from the fact that Landau theory is not based <strong>on</strong> scaling arguments, and asemphasized by the Ginzburg criteria, should not be applicable for systems of dimensi<strong>on</strong>s less thanfour near T c .The relati<strong>on</strong>s above were finally rigorously justified with renormalizati<strong>on</strong> theory. For this KenWils<strong>on</strong> was awarded the Nobel prize. Renormalizati<strong>on</strong> theory is bey<strong>on</strong>d what we will cover in thisclass, but the interested student should read Chapter 18 of Huang.6.7 Symmetry Breaking and Universality ClassesIn Landau theory, the critical exp<strong>on</strong>ents were all determined by the dimensi<strong>on</strong>ality al<strong>on</strong>e, andseveral were even independent of the dimensi<strong>on</strong>s, e.g. β in Eq. (6.41). However, in nature they aredetermined by the symmetry breaking involved in the problem. For instance, if there is a singlereal field ϕ whose free energy density is of the “mexican hat” form,f = A 2 ϕ2 + B 4 ϕ4 , (6.58)105


the system must choose whether to create a field with ⟨ϕ⟩ > 0 or < 0 when A becomes negative.For a complex field with a free energy density of the same form, ϕ 2 → ϕ ∗ ϕ, the system must alsochoose the complex phase. For a system where the spins are allowed to align in any directi<strong>on</strong>, thesystem must choose a directi<strong>on</strong> (θ, ϕ) <strong>on</strong> which to align. Each of these three examples represents adifferent symmetry and leads to different exp<strong>on</strong>ents. The critical exp<strong>on</strong>ents are determined by thesymmetry, and problems of the same symmetry are referred to as being in the same universalityclass. The breaking of the symmetry at low temperature is called sp<strong>on</strong>taneous symmetry breaking,as opposed to the breaking of the symmetry by adding an external field, which is referred to asexplicit symmetry breaking.For the broken reflecti<strong>on</strong> symmetry above, the system chooses between two discrete minima,whereas for the example of the complex field or for the spin alignment, the set of possible minimaform an infinite c<strong>on</strong>tinuous set. These symmetries are called c<strong>on</strong>tinuous symmetries and have theproperty that, after being broken, <strong>on</strong>e can make incremental changes in the field while maintaininga minimum in the free energy. In c<strong>on</strong>trast, the real-field case <strong>on</strong>ly has a reflecti<strong>on</strong> symmetry, whichis referred to as a discrete symmetry. For instance, c<strong>on</strong>sider the case of the complex field. Writingthe free energy density in terms of the real and imaginary parts, ϕ x and ϕ y ,f = A 2 (ϕ2 x + ϕ 2 y) + B 4where we c<strong>on</strong>sider the case where α < 0 and minima appear for(ϕ2x + ϕ 2 y) 2+κ2 (∇ϕ)2 , (6.59)(ϕ 2 x + ϕ 2 y) = ϕ 2 0 = − A B . (6.60)If <strong>on</strong>e expands around the minima in the x directi<strong>on</strong>, <strong>on</strong>e can rewrite the fields asRewriting the free energy in terms of δϕ x and δϕ y ,δϕ x = ϕ x − ϕ 0 , δϕ y = ϕ y . (6.61)f ≈ f min + 1 ∂ 2 f2 δϕ2 x + 1 ∂ 2 f∂ 2 ϕ x 2 δϕ2 y + κ ∂ 2 ϕ y 2 (∇ϕ)2 (6.62)= f min + 1 2 A xδϕ 2 x + 1 2 A yδϕ 2 y + κ 2 (∇δϕ x) 2 + κ 2 (∇δϕ y) 2 ,A x = −2A, A y = 0.Thus, the y mode c<strong>on</strong>tributes to the free energy <strong>on</strong>ly through the term proporti<strong>on</strong>al to (∇δϕ y ) 2 , orin momentum space, the modes c<strong>on</strong>tribute to the free energy as k 2 . For relativistic field theories,modes for which the c<strong>on</strong>tributi<strong>on</strong> vanishes as k → 0 are called massless modes as they have norest energy. These modes are also referred to as Goldst<strong>on</strong>e bos<strong>on</strong>s (Goldst<strong>on</strong>e derived the generalidea that massless modes should be associated with breaking a c<strong>on</strong>tinuous symmetry) or Nambu-Goldst<strong>on</strong>e bos<strong>on</strong>s. Although the most famous example of such symmetry breaking is the electroweaktransiti<strong>on</strong> in particle physics, Higgs, Goldst<strong>on</strong>e and Nambu were also c<strong>on</strong>cerned with thenuclear physics problem of chiral symmetry breaking, and of course Landau was most c<strong>on</strong>cernedwith problems in c<strong>on</strong>densed matter physics, the most famous of which is Ginzburg-Landau theoryfor superc<strong>on</strong>ductivity, which employs precisely this model of complex scalar fields.106


In the linear sigma model for nuclear physics, <strong>on</strong>e c<strong>on</strong>siders a four-comp<strong>on</strong>ent field, ⃗ ϕ, where ϕ 0refers to the σ field, which in QCD is related to the quark-antiquark c<strong>on</strong>densate, and ϕ 1,2,3 are thethree pi<strong>on</strong>ic fields. The Hamilt<strong>on</strong>ian c<strong>on</strong>tains terms of the form,V (ϕ) = − m2 02(ϕ20 + ϕ 2 1 + ϕ 2 2 + ϕ3) 2 λ ( )+ ϕ24 0 + ϕ 2 1 + ϕ 2 2 + ϕ 2 23 . (6.63)Since the first term is negative, sp<strong>on</strong>taneous symmetry breaking ensues. Defining the directi<strong>on</strong> ofthe c<strong>on</strong>densate as the “zero” directi<strong>on</strong>, the c<strong>on</strong>densate generated is√m0⟨ϕ 0 ⟩ =λ . (6.64)This would result in <strong>on</strong>e massive mode, from the curvature of the potential in the “zero” directi<strong>on</strong>and three massless modes, since at the minimum,d 2 Vdϕ 2 ∣ √ = 0, i = 1, 2, 3. (6.65)i ϕ0 = m 0 /λThis would predict three Goldst<strong>on</strong>e bos<strong>on</strong>s, which would be massless pi<strong>on</strong>s. In nature, pi<strong>on</strong>s areindeed very light, but not massless. This is accounted for by adding a term proporti<strong>on</strong>al to ϕ 0in V , which explicitly breaks the symmetry, and makes the original symmetry approximate. Thesymmetries of the linear sigma model derive from underlying chiral symmetries in QCD, which isagain <strong>on</strong>ly an approximate symmetry. The symmetry is broken by the small current quark masses,≈ 5 MeV/c 2 .In electroweak physics, i.e., the Weinberg-Salam model, the situati<strong>on</strong> is more complicated dueto the gauge nature of the theory. Again, there are four scalar fields, with similar quadratic andquartic couplings. The symmetry breaking is the same, but the three accompanying scalar fieldsdo not become massless due to their gauge coupling to other particles. Remarkably, the three fieldsplay the part of the l<strong>on</strong>gitudinal comp<strong>on</strong>ent of the W ± and Z massive fields. Whereas the W andZ are massless in the bare Lagrangian, and have <strong>on</strong>ly two polarizati<strong>on</strong>s, they pick up mass throughthe symmetry breaking, which requires three polarizati<strong>on</strong>. That extra degree of freedom is providedby the three scalar fields.It is no exaggerati<strong>on</strong> to state that the most important problems addressed in the last 40 yearsin nuclear, particle, and c<strong>on</strong>densed-matter physics c<strong>on</strong>cern sp<strong>on</strong>taneously broken symmetries. Furthermore,<strong>on</strong>e of the most exciting astr<strong>on</strong>omical measurements, that of fluctuati<strong>on</strong>s of the 3-degreebackground radiati<strong>on</strong> is related to inflati<strong>on</strong>, which might be associated with the breaking of theelectro-weak symmetries or perhaps grand unified theories, under which quarks and lept<strong>on</strong>s appearin a single fermi<strong>on</strong>ic multiplet and the gauge bos<strong>on</strong>s all corresp<strong>on</strong>d to generators of a single group.6.8 Problems1. C<strong>on</strong>sider the example for which the surface energy was calculated, where∆Ψ ≡ P 0 − P + (µ − µ 0 )ρ = A 2[(ρ − ρc ) 2 − α 2] 2.Using Eq. (6.17), solve for the density profile ρ(x) between the two phases.107


2. C<strong>on</strong>sider the <strong>on</strong>e-dimensi<strong>on</strong>al Ising model, with the total energy in the mean field approximati<strong>on</strong>being,E = − ∑ 12 qJ⟨σ⟩σ i.iNote the factor of 1/2 being added relative to the effective energy for <strong>on</strong>e mode to accountfor double counting.(a) Using the definiti<strong>on</strong> of entropy, − ∑ i p i ln(p i ), show that the entropy per spin is:S/N = − 1 + ⟨σ⟩ ( ) 1 + ⟨σ⟩ln− 1 − ⟨σ⟩ ( ) 1 − ⟨σ⟩ln.2 2 2 2(b) The free-energy, F = E − T S, per spin is thenF/N = − 1 ( )(1 + ⟨σ⟩) 1 + ⟨σ⟩2 qJ⟨σ⟩2 + T ln+ T22Show that minimizing the free energy w.r.t. ⟨σ⟩ gives:( ) (1 + ⟨σ⟩)2βqJ⟨σ⟩ = ln.(1 − ⟨σ⟩)(1 − ⟨σ⟩)2(c) Compare the expressi<strong>on</strong> above to that from the previous expressi<strong>on</strong>,⟨σ⟩ = tanh(βqJ⟨σ⟩).( 1 − ⟨σ⟩ln2(d) If the density of spin sites per unit volume is ρ 0 , the free energy density is{f(σ, T ) = ρ 0 V(σ, T ) + κ }2 (∇σ)2 ,V(σ, T ) = − 1 ( )( )(1 + σ) 1 + σ (1 − σ) 1 − σ2 qJσ2 + T ln + T ln .2 22 2Derive an expressi<strong>on</strong> for the surface free energy that is an integral over the density withlimits from −σ eq and +σ eq . Write the integrand in terms of ρ 0 , κ, T and R(σ, T ) ≡V(σ, T ) − V(σ eq , T ), where σ eq is the soluti<strong>on</strong> to the transcendental expressi<strong>on</strong> for theequilibrated value for ⟨σ⟩.(e) Find the surface energy in the limit that T → 0 in terms of κ, ρ 0 and T c = qJ.3. Repeat the calculati<strong>on</strong> for the correlati<strong>on</strong> functi<strong>on</strong> from Sec. 6.3 for the <strong>on</strong>e dimensi<strong>on</strong>al case.In this case, the Fourier transforms are defined as:˜ϕ k ≡ √ 1 ∫dx e ikx ϕ(x),Lϕ(x) = √ 1 ∑e −ikx ϕ k .Lk).108


4. Beginning with the definiti<strong>on</strong> of the average density,⟨ρ⟩ = Tr ∫ e−βH+βµ d 3rρ(⃗r) ρ(⃗r = 0)Tr e −βH+βµ ∫ ,d 3 rρ(⃗r)prove the following relati<strong>on</strong> between the density-density correlati<strong>on</strong> functi<strong>on</strong> and dρ/dµ,T d⟨ρ⟩ ∫dµ = d 3 r { ⟨ρ(⃗r)ρ(⃗r = 0)⟩− < ρ > 2} .5. Returning to the expressi<strong>on</strong> for the mean spin in the Ising model in the mean field approximati<strong>on</strong>,⟨σ⟩ = tanh(βqJ⟨σ⟩ + βµB),solve for ⟨σ⟩ in the limit that µB is small and T = T c . Note this provides the exp<strong>on</strong>ent δ inmean field theory. Hint: Expand tanh and keep <strong>on</strong>ly the lowest order n<strong>on</strong>-zero terms for both⟨σ⟩ and for µB.6. Assume the following form for the correlati<strong>on</strong> functi<strong>on</strong>,⟨m(0)m(r)⟩ − ⟨m⟩ 2 = Cr 2−d−η e −r/ξ ,ξ ∼ t −νnear T c .(a) Find the susceptibility using Eq. (6.29) in terms of C, ξ, η and the dimensi<strong>on</strong>ality d.When expressing your answer, you can use the shorthand Ω d for ∫ dΩ in d dimensi<strong>on</strong>s,i.e.,∫ ∫d d r... = dΩ r d−1 dr...For your entertainment, Ω d = 2π d/2 /Γ(d/2).(b) If χ diverges as t −γ near T c , find γ in terms of η and ν.7. Assume that the free energy obeys the following form,∫ { AF = d 3 r2 ϕ2 + C }2 ϕ6 .Assuming that near T c , A ∼ at, find the critical exp<strong>on</strong>ent in mean field theory β where,below T c .⟨ϕ⟩ ∼ t β109

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!