9 Statistical physics

9.1 Foundations: From Microstates to Thermodynamics

Statistical physics is the art of predicting macroscopic laws from microscopic uncertainty. We never know the exact 102310^{23}-body state; instead we organize ignorance into ensembles whose averages reproduce temperature, pressure, magnetization, and friends. This section builds the core: microstates vs macrostates, the fundamental postulate, the three standard ensembles, and the thermodynamic limit. We also show how the information-theoretic view turns “maximum entropy” into the unique engine behind the familiar distributions.


9.1.1 Microstates, macrostates, and phase space

A microstate is a complete specification of the system’s degrees of freedom. For NN classical point particles, it is a point in 6N6N-dimensional phase space with coordinates

Γ(r1,,rN;p1,,pN)\Gamma \equiv (\boldsymbol r_{1},\dots,\boldsymbol r_{N};\boldsymbol p_{1},\dots,\boldsymbol p_{N})

A macrostate is defined by a few bulk variables, e.g., (E,V,N)(E,V,N) for an isolated gas. Many microstates map to the same macrostate. Counting how many leads to entropy.

Liouville’s theorem states that Hamiltonian flow is incompressible in phase space, so volume elements are preserved along trajectories. This makes uniform distributions over energy shells dynamically consistent.


9.1.2 Fundamental postulate and Boltzmann’s principle

Fundamental postulate. For an isolated system with fixed (E,V,N)(E,V,N) at equilibrium, all microstates compatible with these constraints are equally likely.

Let Ω(E,V,N)\Omega(E,V,N) be the number of microstates with energies in [E,E+δE][E,E+\delta E] for a small but macroscopic δE\delta E. Then the entropy is

S(E,V,N)=kBlnΩ(E,V,N)S(E,V,N) = k_{B}\,\ln \Omega(E,V,N)

Thermodynamic temperature follows from

1T=(SE)V,N\frac{1}{T} = \left(\frac{\partial S}{\partial E}\right)_{V,N}

and pressure and chemical potential are

pT=(SV)E,N,μT=(SN)E,V\frac{p}{T} = \left(\frac{\partial S}{\partial V}\right)_{E,N},\qquad -\,\frac{\mu}{T} = \left(\frac{\partial S}{\partial N}\right)_{E,V}

For weakly interacting subsystems AA and BB, additivity of SS at fixed total EE yields the equilibrium condition TA=TBT_{A}=T_{B} by maximizing SA(EA)+SB(EB)S_{A}(E_{A})+S_{B}(E_{B}) subject to EA+EB=EE_{A}+E_{B}=E.


9.1.3 Thermodynamic limit, extensivity, and concavity

Define the thermodynamic limit as NN\to\infty, VV\to\infty with n=N/Vn=N/V fixed. In this limit:

  • Entropy becomes extensive: S(λE,λV,λN)λS(E,V,N)S(\lambda E,\lambda V,\lambda N)\approx \lambda S(E,V,N)
  • SS is concave in its natural variables, ensuring stability and positive heat capacities where appropriate
  • Fluctuations of intensive variables shrink like 1/N1/\sqrt{N}, making thermodynamics sharp

Short-range interactions guarantee ensemble equivalence; long-range cases (self-gravity, unscreened Coulomb) can violate additivity and need care.


9.1.4 Microcanonical ensemble

The microcanonical density is uniform on the constant-energy hypersurface:

ρmc(Γ)=1Ω(E,V,N)δ ⁣(H(Γ)E)\rho_{\text{mc}}(\Gamma) = \frac{1}{\Omega(E,V,N)}\,\delta\!\left(H(\Gamma)-E\right)

Expectations are

Amc=1ΩdΓ  A(Γ)δ ⁣(H(Γ)E)\langle A\rangle_{\text{mc}} = \frac{1}{\Omega}\int d\Gamma\;A(\Gamma)\,\delta\!\left(H(\Gamma)-E\right)

Maximizing S=kBlnΩS=k_{B}\ln\Omega subject to constraints reproduces thermodynamic relations. For many systems this is the conceptual anchor; in practice the canonical ensemble is often easier to compute with and becomes equivalent at large NN.


9.1.5 Canonical ensemble via a heat bath

Place system SS weakly coupled to a huge reservoir RR with fixed EtotE_{\text{tot}}. The probability that SS is found in microstate ii with energy EiE_{i} is proportional to the number of reservoir states at energy EtotEiE_{\text{tot}}-E_{i}:

P(i)ΩR(EtotEi)exp ⁣[1kBSR(Etot)EikBT]eβEiP(i) \propto \Omega_{R}(E_{\text{tot}}-E_{i}) \approx \exp\!\left[\frac{1}{k_{B}}\,S_{R}(E_{\text{tot}})-\frac{E_{i}}{k_{B}T}\right] \propto e^{-\beta E_{i}}

with inverse temperature β1/(kBT)\beta \equiv 1/(k_{B}T). Normalize to get

P(i)=eβEiZ,Z(β,V,N)ieβEiP(i) = \frac{e^{-\beta E_{i}}}{Z},\qquad Z(\beta,V,N) \equiv \sum_{i} e^{-\beta E_{i}}

ZZ is the partition function. Thermodynamic potentials follow:

F(T,V,N)kBTlnZF(T,V,N) \equiv -\,k_{B}T\,\ln Z U=E=βlnZU = \langle E\rangle = -\,\frac{\partial}{\partial \beta}\ln Z S=(FT)V,N=kB(lnZ+βU)S = -\left(\frac{\partial F}{\partial T}\right)_{V,N} = k_{B}\left(\ln Z + \beta U\right)

Energy fluctuations satisfy

(ΔE)2=kBT2CV\langle (\Delta E)^{2}\rangle = k_{B}T^{2}C_{V}

so relative fluctuations scale as 1/N1/\sqrt{N} for normal systems.


9.1.6 Grand canonical ensemble: exchanging particles

If both energy and particles exchange with a reservoir, the probability of microstate ii with (Ei,Ni)(E_{i},N_{i}) is

P(i)=eβ(EiμNi)ΞP(i) = \frac{e^{-\beta(E_{i}-\mu N_{i})}}{\Xi}

with grand partition function

Ξ(β,V,μ)N=0iNeβ(EiμN)\Xi(\beta,V,\mu) \equiv \sum_{N=0}^{\infty}\sum_{i\in N} e^{-\beta(E_{i}-\mu N)}

Define the grand potential

ΦG(T,V,μ)kBTlnΞ\Phi_{G}(T,V,\mu) \equiv -\,k_{B}T\,\ln \Xi

Then

p=(ΦGV)T,μ,N=(ΦGμ)T,V,S=(ΦGT)V,μp = -\left(\frac{\partial \Phi_{G}}{\partial V}\right)_{T,\mu},\quad N = -\left(\frac{\partial \Phi_{G}}{\partial \mu}\right)_{T,V},\quad S = -\left(\frac{\partial \Phi_{G}}{\partial T}\right)_{V,\mu}

Fluctuations:

(ΔN)2=kBT(Nμ)T,V\langle (\Delta N)^{2}\rangle = k_{B}T\left(\frac{\partial N}{\partial \mu}\right)_{T,V}

and energy–number covariances come from mixed derivatives of lnΞ\ln \Xi.


9.1.7 Legendre transforms and thermodynamic potentials

Thermodynamics is organized by Legendre transforms that swap natural variables:

  • S=S(E,V,N)S=S(E,V,N)
  • F=UTSF=U-TS with natural variables (T,V,N)(T,V,N)
  • G=F+pVG=F+pV with (T,p,N)(T,p,N)
  • H=U+pVH=U+pV with (S,p,N)(S,p,N)
  • ΦG=FμN\Phi_{G}=F-\mu N with (T,V,μ)(T,V,\mu)

Maxwell relations follow from equality of mixed partial derivatives, e.g.,

(SV)T=(pT)V\left(\frac{\partial S}{\partial V}\right)_{T} = \left(\frac{\partial p}{\partial T}\right)_{V}

9.1.8 Information-theoretic foundation (Jaynes view)

Let {pi}\{p_{i}\} be probabilities over microstates. Maximize the Shannon entropy

S[p]kBipilnpi\mathcal S[p] \equiv -k_{B}\sum_{i} p_{i}\ln p_{i}

subject to known constraints. With only normalization and mean energy fixed, Lagrange multipliers give

pi=eβEiZp_{i} = \frac{e^{-\beta E_{i}}}{Z}

With energy and particle number constraints, one obtains the grand canonical form pieβ(EiμNi)p_{i}\propto e^{-\beta(E_{i}-\mu N_{i})}. Thus ensembles are the least biased distributions compatible with macroscopic data.

Relative entropy D(pq)=piln(pi/qi)0D(p\Vert q)=\sum p_{i}\ln(p_{i}/q_{i})\ge 0 measures distance to a prior qq. Many fluctuation theorems and linear-response results are compact in this language.


9.1.9 Large deviations, concentration, and ensemble equivalence

For N1N\gg 1, probabilities concentrate exponentially near thermodynamic values. If AA is an additive observable,

P ⁣(ANa)eNI(a)\mathbb P\!\left(\frac{A}{N}\approx a\right)\sim e^{-N\,I(a)}

with a rate function I(a)0I(a)\ge 0 that vanishes at the equilibrium value. Under standard convexity conditions:

  • Microcanonical, canonical, and grand canonical ensembles are equivalent in the thermodynamic limit
  • Nonconcavities signal phase coexistence; Maxwell constructions repair them at the level of thermodynamic potentials

9.1.10 Worked mini-examples

(a) Two-level system
Energy levels 00 and ϵ\epsilon for NN noninteracting sites. Canonical partition function per site

z=1+eβϵz = 1 + e^{-\beta\epsilon}

Total Z=zNZ=z^{N}. Average energy U=Nϵ/(eβϵ+1)U=N\,\epsilon/(e^{\beta\epsilon}+1) and heat capacity shows a Schottky peak near kBTϵk_{B}T\sim \epsilon.

(b) Classical harmonic oscillator
Single oscillator with En=(n+12)ωE_{n}=(n+\tfrac12)\hbar\omega gives

Z=eβω/21eβωZ = \frac{e^{-\beta\hbar\omega/2}}{1-e^{-\beta\hbar\omega}}

Average energy U=ω2+ωeβω1U=\frac{\hbar\omega}{2}+\frac{\hbar\omega}{e^{\beta\hbar\omega}-1} tends to kBTk_{B}T at high TT (equipartition) and to ω/2\hbar\omega/2 at T0T\to 0.

(c) Coin toss and Stirling
For NN independent fair coins, the multiplicity of kk heads is (Nk)exp ⁣[NH(k/N)]/2πNx(1x)\binom{N}{k}\approx \exp\!\left[N\,H(k/N)\right]/\sqrt{2\pi N x(1-x)} with x=k/Nx=k/N and binary entropy H(x)=xlnx(1x)ln(1x)H(x)=-x\ln x-(1-x)\ln(1-x). This is the large-deviation archetype.

(d) Ideal gas microcanonical temperature
For a classical ideal gas, counting states yields

Ω(E,V,N)VNE3N21N!Γ ⁣(3N2)\Omega(E,V,N) \propto \frac{V^{N}E^{\tfrac{3N}{2}-1}}{N!\,\Gamma\!\left(\tfrac{3N}{2}\right)}

so S=kBlnΩS=k_{B}\ln\Omega gives U=32NkBTU=\tfrac{3}{2}Nk_{B}T and pV=NkBTpV=Nk_{B}T after Stirling.

(e) Canonical fluctuations
Show (ΔE)2=kBT2CV\langle (\Delta E)^{2}\rangle = k_{B}T^{2}C_{V} by taking β2lnZ\partial^{2}_{\beta}\ln Z and relate the heat capacity to curvature of lnZ\ln Z.


9.1.11 Minimal problem kit

  • Starting from S=kBlnΩS=k_{B}\ln\Omega, derive 1/T=(S/E)V,N1/T=(\partial S/\partial E)_{V,N} and p/T=(S/V)E,Np/T=(\partial S/\partial V)_{E,N}
  • Derive the canonical ensemble by maximizing S[p]\mathcal S[p] with constraints pi=1\sum p_{i}=1 and piEi=U\sum p_{i}E_{i}=U
  • Show that F=kBTlnZF=-k_{B}T\ln Z implies S=F/TS=-\partial F/\partial T and p=F/Vp=-\partial F/\partial V
  • For a classical ideal gas, compute Ω(E,V,N)\Omega(E,V,N) up to a constant using phase-space volume and obtain the Sackur–Tetrode entropy
  • Prove (ΔN)2=kBT(N/μ)T,V\langle (\Delta N)^{2}\rangle = k_{B}T\,(\partial N/\partial\mu)_{T,V} in the grand canonical ensemble
  • Discuss when ensemble equivalence fails and give a physical example with long-range forces

9.2 Partition Functions & Thermodynamic Potentials

Partition functions are the generators of everything thermodynamic. Once you have ZZ or Ξ\Xi or the N ⁣pTN\!pT analogue, derivatives give you U,p,S,μU,p,S,\mu and all the response coefficients. This section builds the canonical and grand-canonical machinery, adds the isothermal–isobaric ensemble, and shows how Legendre transforms and saddle points tie them back to microcanonical entropy. Worked examples anchor the formulas.


9.2.1 Canonical partition function and cumulants

For fixed (T,V,N)(T,V,N) the canonical weights are P(i)eβEiP(i)\propto e^{-\beta E_{i}} with

Z(β,V,N)ieβEiZ(\beta,V,N) \equiv \sum_{i} e^{-\beta E_{i}}

Helmholtz free energy

F(T,V,N)kBTlnZF(T,V,N) \equiv -\,k_{B}T\,\ln Z

Standard derivatives

UE=βlnZU \equiv \langle E\rangle = -\,\frac{\partial}{\partial \beta}\ln Z S=(FT)V,N=kB(lnZ+βU)S = -\left(\frac{\partial F}{\partial T}\right)_{V,N} = k_{B}\left(\ln Z + \beta U\right) p=(FV)T,Np = -\left(\frac{\partial F}{\partial V}\right)_{T,N}

Energy cumulants follow from derivatives of lnZ\ln Z with respect to β\beta. For instance

(ΔE)2=2β2lnZ=kBT2CV\langle(\Delta E)^{2}\rangle = \frac{\partial^{2}}{\partial \beta^{2}}\ln Z = k_{B}T^{2} C_{V}

Generalized forces for a control parameter λ\lambda that enters H(λ)H(\lambda) obey

(Fλ)T,V,N=Hλ\left(\frac{\partial F}{\partial \lambda}\right)_{T,V,N} = \left\langle \frac{\partial H}{\partial \lambda}\right\rangle

Examples: M=(F/B)T,V,NM = -\left(\partial F/\partial B\right)_{T,V,N} for a magnetic field BB and p=(F/V)T,Np = -\left(\partial F/\partial V\right)_{T,N} for volume.


9.2.2 Grand partition function, fugacity, and pressure

When particle number fluctuates at fixed (T,V,μ)(T,V,\mu)

Ξ(β,V,μ)N=0iNeβ(EiμN)\Xi(\beta,V,\mu) \equiv \sum_{N=0}^{\infty}\sum_{i\in N} e^{-\beta(E_{i}-\mu N)}

Define the grand potential

ΦG(T,V,μ)kBTlnΞ\Phi_{G}(T,V,\mu) \equiv -\,k_{B}T\,\ln \Xi

Thermodynamics in this ensemble is compact

p=(ΦGV)T,μp = -\left(\frac{\partial \Phi_{G}}{\partial V}\right)_{T,\mu} N=(ΦGμ)T,VN = -\left(\frac{\partial \Phi_{G}}{\partial \mu}\right)_{T,V} S=(ΦGT)V,μS = -\left(\frac{\partial \Phi_{G}}{\partial T}\right)_{V,\mu}

It is common to use the fugacity zeβμz\equiv e^{\beta \mu}. Number fluctuations and compressibility are linked by

(ΔN)2=kBT(Nμ)T,V\langle(\Delta N)^{2}\rangle = k_{B}T\left(\frac{\partial N}{\partial \mu}\right)_{T,V}

For a uniform ideal gas one finds the handy identity pV=kBTlnΞpV = k_{B}T \ln \Xi.


9.2.3 Isothermal–isobaric ensemble and Gibbs free energy

Many experiments hold (T,p,N)(T,p,N) fixed. The isothermal–isobaric partition function is the Laplace transform of ZZ over VV

Δ(β,p,N)0dV  eβpVZ(β,V,N)\Delta(\beta,p,N) \equiv \int_{0}^{\infty} dV\; e^{-\beta p V}\, Z(\beta,V,N)

Its thermodynamic potential is the Gibbs free energy

G(T,p,N)kBTlnΔG(T,p,N) \equiv -\,k_{B}T\,\ln \Delta

Then

(Gp)T,N=V,(GT)p,N=S,(GN)T,p=μ\left(\frac{\partial G}{\partial p}\right)_{T,N} = V,\qquad \left(\frac{\partial G}{\partial T}\right)_{p,N} = -S,\qquad \left(\frac{\partial G}{\partial N}\right)_{T,p} = \mu

Fluctuations of VV at fixed (T,p,N)(T,p,N) are generated by derivatives of lnΔ\ln \Delta.


Canonical and grand-canonical ensembles are Laplace transforms of the microcanonical state count

Z(β,V,N)=dE  Ω(E,V,N)eβEZ(\beta,V,N) = \int dE\; \Omega(E,V,N)\,e^{-\beta E}

For large systems the integral is dominated by a saddle point EE^{\ast} with ElnΩE=β\partial_{E}\ln\Omega|_{E^{\ast}}=\beta, i.e., 1/T=S/E1/T = \partial S/\partial E. Evaluating at the saddle yields the Legendre relation

F(T,V,N)=UTSF(T,V,N) = U - T S

Similarly, the grand-canonical transform over both EE and NN yields

ΦG(T,V,μ)=UTSμN\Phi_{G}(T,V,\mu) = U - T S - \mu N

These are the reasons thermodynamic potentials are Legendre transforms of SS.


9.2.5 Ideal classical gas in the canonical ensemble

Take NN indistinguishable particles of mass mm with Hamiltonian H=pi2/2mH=\sum p_{i}^{2}/2m and ignore interactions. The canonical partition function factorizes into a translational part

Z(β,V,N)=1N![VλT3]NZ(\beta,V,N) = \frac{1}{N!}\left[\frac{V}{\lambda_{T}^{3}}\right]^{N}

with thermal wavelength

λT2π2mkBT\lambda_{T} \equiv \sqrt{\frac{2\pi \hbar^{2}}{m k_{B} T}}

Then

F=kBTlnZ=NkBT[ln ⁣(VNλT3)+1]F = -\,k_{B}T \ln Z = -\,Nk_{B}T\left[\ln\!\left(\frac{V}{N\,\lambda_{T}^{3}}\right) + 1\right]

Equation of state follows immediately

pV=NkBTpV = Nk_{B}T

Entropy is the Sackur–Tetrode form

S=NkB[ln ⁣(VNλT3)+52]S = Nk_{B}\left[\ln\!\left(\frac{V}{N\,\lambda_{T}^{3}}\right) + \frac{5}{2}\right]

and the internal energy is U=32NkBTU=\tfrac{3}{2}Nk_{B}T with CV=32NkBC_{V}=\tfrac{3}{2}Nk_{B}. The 1/N!1/N! cures the Gibbs paradox by enforcing indistinguishability.


9.2.6 Quantum ideal gases via the cluster expansion

For noninteracting bosons or fermions with one-particle energies {ϵα}\{\epsilon_{\alpha}\}

lnΞ=±αln ⁣(1zeβϵα)\ln \Xi = \pm \sum_{\alpha} \ln\!\left(1 \mp z\,e^{-\beta \epsilon_{\alpha}}\right)

upper sign for bosons, lower for fermions. Occupations are

nα=1z1eβϵα1\langle n_{\alpha}\rangle = \frac{1}{z^{-1}e^{\beta \epsilon_{\alpha}} \mp 1}

In the continuum limit one replaces αdϵg(ϵ)\sum_{\alpha}\to \int d\epsilon\,g(\epsilon) to obtain thermodynamics of photon, phonon, or electron gases. Interactions can be included perturbatively via virial or cluster expansions that correct lnΞ\ln \Xi by coefficients B2,B3,B_{2},B_{3},\dots.


9.2.7 Generalized forces and response matrices

If a Hamiltonian depends on external controls λ=(λ1,)\boldsymbol{\lambda}=(\lambda_{1},\dots), the matrix of second derivatives of FF gives linear response

χab(2Fλaλb)T,V,N=βΔXaΔXb\chi_{ab} \equiv -\left(\frac{\partial^{2} F}{\partial \lambda_{a}\,\partial \lambda_{b}}\right)_{T,V,N} = \beta\,\langle \Delta X_{a}\,\Delta X_{b}\rangle

where XaH/λaX_{a}\equiv -\partial H/\partial \lambda_{a} is the conjugate observable. Examples include magnetic susceptibility, compressibility, and heat capacities. Mixed derivatives encode Maxwell relations, e.g., (S/V)T,N=(p/T)V,N\left(\partial S/\partial V\right)_{T,N}=\left(\partial p/\partial T\right)_{V,N}.


9.2.8 Ising-like lattice systems and counting practice

For a finite lattice with spins si=±1s_{i}=\pm 1 and Hamiltonian H=JijsisjhisiH=-J\sum_{\langle ij\rangle}s_{i}s_{j}-h\sum_{i}s_{i} the canonical partition function is

Z(β,h)={s}exp ⁣[βJijsisj+βhisi]Z(\beta,h) = \sum_{\lbrace s\rbrace} \exp\!\left[\beta J \sum_{\langle ij\rangle}s_{i}s_{j} + \beta h \sum_{i}s_{i}\right]

While exact evaluation is hard beyond 1D, derivatives of lnZ\ln Z still define magnetization M=(lnZ/(βh))M=\left(\partial \ln Z/\partial (\beta h)\right) and susceptibility χ=M/h\chi=\partial M/\partial h. Mean-field or transfer-matrix approaches approximate ZZ systematically.


9.2.9 Gibbs–Duhem and extensivity checks

For an extensive system with U(S,V,N)U(S,V,N) homogeneous of degree one, Euler’s relation holds

U=TSpV+μNU = TS - pV + \mu N

Differentiating yields the Gibbs–Duhem equation

SdTVdp+Ndμ=0S\,dT - V\,dp + N\,d\mu = 0

This is a consistency check for any equation of state you derive from a partition function.


9.2.10 Laplace tricks, saddle points, and when ensembles match

Because ZZ and Δ\Delta are Laplace transforms, large-NN thermodynamics follows from their leading saddle. Concavity of S(E,V,N)S(E,V,N) guarantees equivalence of ensembles. At first-order transitions ZZ develops multiple competing saddles; Legendre transforms pick the convex envelope, and fluctuations become non-Gaussian in finite systems.


9.2.11 Worked mini-examples

(a) N ⁣pTN\!pT ideal gas

Compute Δ\Delta for the ideal gas using Z=(V/λT3)N/N!Z=(V/\lambda_{T}^{3})^{N}/N! to show

Δ=1N!(kBTpλT3)N\Delta = \frac{1}{N!}\left(\frac{k_{B}T}{p\,\lambda_{T}^{3}}\right)^{N}

and verify G=NkBTln ⁣(pλT3/kBT)+NkBTG=Nk_{B}T\ln\!\left(p\,\lambda_{T}^{3}/k_{B}T\right) + Nk_{B}T and V=(G/p)T,N=NkBT/pV = \left(\partial G/\partial p\right)_{T,N} = Nk_{B}T/p.

(b) Pressure from ZZ via the virial

For a classical interacting gas with pair potential u(r)u(r), show that p=NkBT/V2πn230drr3u(r)g(r)p = Nk_{B}T/V - \tfrac{2\pi n^{2}}{3}\int_{0}^{\infty} dr\, r^{3} u'(r) g(r) using p=(F/V)T,Np=-(\partial F/\partial V)_{T,N} and a uniform scaling argument, where n=N/Vn=N/V and g(r)g(r) is the pair distribution.

(c) Number fluctuations and compressibility

In the grand ensemble, verify (ΔN)2=kBT(N/μ)T,V\langle(\Delta N)^{2}\rangle = k_{B}T\,\left(\partial N/\partial \mu\right)_{T,V} and relate to the isothermal compressibility κT=1n2(n/μ)T\kappa_{T}=\frac{1}{n^{2}}\left(\partial n/\partial \mu\right)_{T}.

(d) Microcanonical ↔ canonical

Starting from Z=dEexp[S(E)/kBβE]Z=\int dE\,\exp[S(E)/k_{B}-\beta E], perform a quadratic expansion of S(E)S(E) about EE^{\ast} to obtain Gaussian energy fluctuations with variance kBT2CVk_{B}T^{2}C_{V}.

(e) Two-level system free energy

For NN independent two-level sites with energies 00 and ϵ\epsilon, compute Z=(1+eβϵ)NZ=(1+e^{-\beta\epsilon})^{N}, then FF, SS, and CVC_{V} and locate the Schottky peak.


9.2.12 Minimal problem kit

  • Derive all canonical relations for U,S,pU,S,p starting from F=kBTlnZF=-k_{B}T\ln Z and verify the Maxwell relation (S/V)T,N=(p/T)V,N\left(\partial S/\partial V\right)_{T,N}=\left(\partial p/\partial T\right)_{V,N}
  • Show that in the grand ensemble, pV=kBTlnΞpV=k_{B}T\ln\Xi for an ideal gas and obtain NN and SS from derivatives of lnΞ\ln\Xi
  • Compute the isothermal–isobaric partition function for a classical harmonic solid treated as 3N3N independent oscillators and extract G(T,p,N)G(T,p,N) in the limit of weak compressibility
  • Using lnΞ=±αln(1zeβϵα)\ln \Xi = \pm\sum_{\alpha}\ln(1\mp z e^{-\beta\epsilon_{\alpha}}), recover the Bose–Einstein and Fermi–Dirac distributions and the classical limit when z1z\ll 1
  • From extensivity, prove Euler’s relation U=TSpV+μNU=TS-pV+\mu N and then the Gibbs–Duhem equation, and test them on the ideal gas results above

9.3 Classical Ideal Gas & Equipartition

The classical ideal gas is the training montage of statistical physics: you learn to count states, take derivatives of ln\ln partition functions, and watch macroscopic laws fall out like it’s nothing. We derive the Maxwell–Boltzmann distribution, prove equipartition, recover pV=NkBTpV=Nk_{B}T, and build intuition for adiabats, sound speed, and when the classical picture breaks (hello, λT\lambda_{T}).


9.3.1 Canonical counting recap and the ideal-gas Z

For NN indistinguishable point particles of mass mm in volume VV with Hamiltonian H=i=1Npi2/2mH=\sum_{i=1}^{N}\boldsymbol p_{i}^{2}/2m, the canonical partition function factorizes:

Z(β,V,N)=1N![VλT3]NZ(\beta,V,N) = \frac{1}{N!}\left[\frac{V}{\lambda_{T}^{3}}\right]^{N}

with thermal wavelength

λT2π2mkBT\lambda_{T} \equiv \sqrt{\frac{2\pi\hbar^{2}}{m k_{B} T}}

Thermodynamics then follows immediately:

F=NkBT[ln ⁣(VNλT3)+1],U=32NkBT,pV=NkBTF = -\,Nk_{B}T\left[\ln\!\left(\frac{V}{N\lambda_{T}^{3}}\right) + 1\right],\qquad U = \frac{3}{2}Nk_{B}T,\qquad pV = Nk_{B}T

Heat capacities:

CV=(UT)V,N=32NkB,Cp=CV+NkB=52NkB,γCpCV=53C_{V} = \left(\frac{\partial U}{\partial T}\right)_{V,N} = \frac{3}{2}Nk_{B},\qquad C_{p} = C_{V} + Nk_{B} = \frac{5}{2}Nk_{B},\qquad \gamma \equiv \frac{C_{p}}{C_{V}} = \frac{5}{3}

Entropy is the Sackur–Tetrode form:

S=NkB[ln ⁣(VNλT3)+52]S = Nk_{B}\left[\ln\!\left(\frac{V}{N\lambda_{T}^{3}}\right) + \frac{5}{2}\right]

The 1/N!1/N! kills the Gibbs paradox by encoding indistinguishability.


9.3.2 Maxwell–Boltzmann distribution (velocities, speeds, energies)

Each momentum component is Gaussian:

f(px)=β2πm  exp ⁣(βpx22m)f(p_{x}) = \sqrt{\frac{\beta}{2\pi m}}\;\exp\!\left(-\frac{\beta p_{x}^{2}}{2m}\right)

Equivalently, each velocity component is Gaussian with variance vx2=kBT/m\langle v_{x}^{2}\rangle = k_{B}T/m. The speed distribution (radial in velocity space) is

f(v)=4π(m2πkBT)3/2v2exp ⁣(mv22kBT)f(v) = 4\pi \left(\frac{m}{2\pi k_{B}T}\right)^{3/2} v^{2} \exp\!\left(-\frac{m v^{2}}{2k_{B}T}\right)

Useful moments:

vmp=2kBTm,v=8kBTπm,vrms=3kBTmv_{\text{mp}} = \sqrt{\frac{2k_{B}T}{m}},\qquad \langle v\rangle = \sqrt{\frac{8k_{B}T}{\pi m}},\qquad v_{\text{rms}} = \sqrt{\frac{3k_{B}T}{m}}

The kinetic energy distribution is χ2\chi^{2}-like with 33 degrees of freedom:

f(E)=2π1(kBT)3/2EeE/kBT,E=32kBTf(E) = \frac{2}{\sqrt{\pi}}\frac{1}{(k_{B}T)^{3/2}}\sqrt{E}\,e^{-E/k_{B}T},\qquad \langle E\rangle = \frac{3}{2}k_{B}T

9.3.3 Equipartition theorem

Statement. Every quadratic degree of freedom in the Hamiltonian contributes 12kBT\tfrac{1}{2}k_{B}T to the mean energy.

Sketch of proof (canonical, integration by parts). If HH contains a term αx2\alpha x^{2}, then

xH/x=kBT\langle x\,\partial H/\partial x \rangle = k_{B}T

For H=pi2/2m+H=\sum p_{i}^{2}/2m + \cdots, each quadratic pi2/2mp_{i}^{2}/2m gives pi2/2m=12kBT\langle p_{i}^{2}/2m\rangle = \tfrac{1}{2}k_{B}T. A monatomic gas has 3N3N translational momenta → U=32NkBTU=\tfrac{3}{2}Nk_{B}T.

When it fails (and why that’s good). If a degree of freedom isn’t effectively quadratic (e.g., quantized rotations/vibrations at kBTk_{B}T below level spacings), equipartition “freezes out.” This explains the drop of specific heats at low TT and the success/failure of Dulong–Petit.


9.3.4 Kinetic-theory

From elastic impacts on a wall, the pressure can be written

p=23UVp = \frac{2}{3}\frac{U}{V}

Combine with U=32NkBTU=\tfrac{3}{2}Nk_{B}T to get pV=NkBTpV=Nk_{B}T. More generally, for pairwise interactions u(r)u(r) the virial theorem yields

pV=NkBT+13i<jrijFijpV = Nk_{B}T + \frac{1}{3}\left\langle \sum_{i<j} \boldsymbol r_{ij}\cdot \boldsymbol F_{ij} \right\rangle

The ideal gas has Fij=0\boldsymbol F_{ij}=0, so the second term vanishes.


9.3.5 Adiabats, sound speed, and quick thermodynamic moves

For a reversible adiabatic change (dQ=0dQ=0) in a monatomic ideal gas,

pVγ=const,TVγ1=const,Tγp1γ=constp V^{\gamma} = \text{const},\qquad T V^{\gamma-1} = \text{const},\qquad T^{\gamma} p^{1-\gamma} = \text{const}

with γ=5/3\gamma=5/3. The speed of sound (small adiabatic density waves) is

cs=γkBTmc_{s} = \sqrt{\gamma \frac{k_{B}T}{m}}

For mixtures, replace mm by the mean molecular mass and γ\gamma by the appropriate ratio of heat capacities.


9.3.6 External fields: barometric formula and separability

In a uniform gravitational field U(z)=mgzU(z)=mgz, the one-body Boltzmann weight factorizes:

n(z)eβmgz,n(z)=n(0)ez/H,HkBTmgn(z) \propto e^{-\beta m g z},\qquad n(z) = n(0)\,e^{-z/H},\qquad H \equiv \frac{k_{B}T}{m g}

The momentum and position parts separate, so kinetic-energy results (equipartition, MB speeds) survive untouched.


9.3.7 Transport in one screen: mean free path and friends

For hard spheres of diameter dd at number density nn, the mean free path is

12πd2n\ell \simeq \frac{1}{\sqrt{2}\,\pi d^{2}\,n}

Using typical speed vˉv\bar v \sim \langle v\rangle,

D13vˉ,η13nmvˉ,κ13nCVper particlevˉD \sim \frac{1}{3}\bar v\,\ell,\qquad \eta \sim \frac{1}{3} n m \bar v\,\ell,\qquad \kappa \sim \frac{1}{3} n C_{V}^{\text{per particle}} \bar v\,\ell

These kinetic-theory scalings are remarkably good for dilute gases; real cross sections and mixtures refine the prefactors.


9.3.8 Validity of the classical picture

Classical, nondegenerate behavior requires the diluteness parameter to be small:

nλT31n\,\lambda_{T}^{3} \ll 1

If nλT31n\lambda_{T}^{3}\gtrsim 1 (low TT or high nn), quantum statistics kicks in: bosons want Bose–Einstein, fermions go Pauli, and the MB distribution is no longer the move.


9.3.9 Beyond monatomic: what equipartition counts

For a rigid linear diatomic at intermediate TT:

  • 33 translational + 22 rotational quadratic modes → f=5f=5 effective d.o.f., so CV52NkBC_{V} \approx \tfrac{5}{2}Nk_{B}, γ7/5\gamma \approx 7/5
  • Vibrations add 22 more quadratic modes but only when kBTk_{B}T exceeds the vibrational quantum spacing; otherwise they’re frozen

Moral: count only the modes that are active at your TT.


9.3.10 Worked mini-examples

(a) RMS speed for air at room TT

For N2\mathrm{N}_{2} with m28um\approx 28\,\mathrm{u} at T=300 KT=300\ \mathrm{K},

vrms=3kBTmv_{\text{rms}}=\sqrt{\frac{3k_{B}T}{m}}

Plug numbers to get a few hundred m/s\mathrm{m/s}; compare to vmpv_{\text{mp}} and v\langle v\rangle.

(b) Adiabatic compression

A monatomic gas is compressed from (p1,V1)(p_{1},V_{1}) to (p2,V2)(p_{2},V_{2}) adiabatically. Use pVγ=constpV^{\gamma}=\text{const} to find T2/T1=(V1/V2)γ1T_{2}/T_{1}=(V_{1}/V_{2})^{\gamma-1}.

(c) Barometric scale height of helium

For He at T=300 KT=300\ \mathrm{K}, compute H=kBT/(mg)H=k_{B}T/(m g) and compare to N2\mathrm{N}_{2} to see the mass dependence pop.

(d) Mean free path at STP

Take d0.3 nmd\approx 0.3\ \mathrm{nm}, n2.5×1025 m3n\approx 2.5\times 10^{25}\ \mathrm{m}^{-3}; estimate \ell and comment on why smells diffuse fast.

(e) Equipartition sanity check

Show directly from dp  p2eβp2/2m\int dp\;p^{2} e^{-\beta p^{2}/2m} that p2/2m=12kBT\langle p^{2}/2m\rangle = \tfrac{1}{2}k_{B}T and extend to three components.


9.3.11 Minimal problem kit

  • Starting from f(v)f(v), derive vmpv_{\text{mp}}, v\langle v\rangle, and vrmsv_{\text{rms}}
  • Prove the equipartition identity xH/x=kBT\langle x\,\partial H/\partial x\rangle=k_{B}T for a quadratic coordinate and apply it to p2/2mp^{2}/2m
  • Derive p=23U/Vp=\tfrac{2}{3}U/V via wall-collision counting and recover pV=NkBTpV=Nk_{B}T
  • For a linear diatomic with rotational constant BB, estimate the temperature where rotations “turn on” by comparing kBTk_{B}T to 2/(2I)\hbar^{2}/(2I)
  • Check the classicality criterion for a cold atomic gas of density nn and temperature TT: compute nλT3n\lambda_{T}^{3} and discuss when quantum statistics is required

9.4 Fluctuations & the Fluctuation–Response Principle

Macroscopic laws look crisp because relative fluctuations scale like 1/N1/\sqrt{N}—until they don’t. This section organizes what fluctuates, how much, and why responses equal fluctuations. We connect the canonical, grand-canonical, and isothermal–isobaric ensembles; introduce static fluctuation–dissipation relations; link pair correlations to compressibility; and preview what breaks near critical points.


9.4.1 Energy and number: the two flagship variances

Canonical (fixed T,V,NT,V,N):

(ΔE)2=2β2lnZ=kBT2CV\langle(\Delta E)^{2}\rangle = \frac{\partial^{2}}{\partial \beta^{2}}\ln Z = k_{B}T^{2}\,C_{V}

Relative size shrinks as (ΔE)21/2/U1/N\langle(\Delta E)^{2}\rangle^{1/2}/U \sim 1/\sqrt{N} for normal systems.

Grand canonical (fixed T,V,μT,V,\mu):

(ΔN)2=kBT(Nμ)T,V\langle(\Delta N)^{2}\rangle = k_{B}T\left(\frac{\partial N}{\partial \mu}\right)_{T,V}

Equivalently, in terms of density n=N/Vn=N/V and isothermal compressibility κT1V(Vp)T,N\kappa_{T} \equiv -\frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{T,N},

(ΔN)2=NkBTnκT\langle(\Delta N)^{2}\rangle = N\,k_{B}T\,n\,\kappa_{T}

for a single-component fluid where dn=χμdμdn = \chi_{\mu}\,d\mu and standard thermodynamic identities have been used.


9.4.2 Volume fluctuations at fixed pressure

Holding (T,p,N)(T,p,N) fixed, the isothermal–isobaric partition function Δ\Delta generates

(ΔV)2=kBT(Vp)T,N=kBTVκT\langle(\Delta V)^{2}\rangle = k_{B}T \left(\frac{\partial V}{\partial p}\right)_{T,N} = k_{B}T\,V\,\kappa_{T}

Again, fluctuation = response: compressibility measures how much VV wiggles when pp nudges; that same coefficient sets the equilibrium variance.


9.4.3 Static fluctuation–response in one line

Let a control λ\lambda couple to an observable XX via HHλXH \to H - \lambda X (so X=H/λX=-\partial H/\partial \lambda). In the canonical ensemble,

χXX(Xλ)T,V,N=β(ΔX)2\chi_{XX} \equiv \left(\frac{\partial \langle X\rangle}{\partial \lambda}\right)_{T,V,N} = \beta\,\langle(\Delta X)^{2}\rangle

Special cases:

  • X=EX=E, λ\lambda shifts β\beta: χEE\chi_{EE} reduces to CV/TC_{V}/T and reproduces (ΔE)2=kBT2CV\langle(\Delta E)^{2}\rangle=k_{B}T^{2}C_{V}
  • X=NX=N in the grand ensemble gives (N/μ)T,V=β(ΔN)2\left(\partial N/\partial \mu\right)_{T,V}=\beta\,\langle(\Delta N)^{2}\rangle

Positivity of variances \Rightarrow stability: CV0C_{V}\ge 0, κT0\kappa_{T}\ge 0 in ensembles with fixed TT.


9.4.4 Correlations, structure factor, and the compressibility sum rule

Define the microscopic density n^(r)=iδ(rri)\hat n(\boldsymbol r)=\sum_{i}\delta(\boldsymbol r-\boldsymbol r_{i}), the pair correlation g(r)g(\boldsymbol r), and the static structure factor

S(k)1Nδn^kδn^k,δn^kd3reikr[n^(r)n]S(\boldsymbol k) \equiv \frac{1}{N}\left\langle \delta \hat n_{\boldsymbol k}\,\delta \hat n_{-\boldsymbol k}\right\rangle,\qquad \delta \hat n_{\boldsymbol k} \equiv \int d^{3}r\, e^{-i\mathbf k\cdot \mathbf r}\,[\hat n(\mathbf r)-n]

In a uniform fluid,

S(k)=1+nd3reikr[g(r)1]S(\boldsymbol k) = 1 + n \int d^{3}r\, e^{-i\boldsymbol k\cdot \boldsymbol r}\,[g(\boldsymbol r)-1]

Taking k0\boldsymbol k\to 0 connects microscopic correlations to macroscopic response:

S(0)=nkBTκTS(0) = n\,k_{B}T\,\kappa_{T}

This is the compressibility sum rule. Near a critical point, κT\kappa_{T}\to \infty and S(0)S(0) spikes—aka critical opalescence.


9.4.5 Saddle points, curvature, and large deviations

Microcanonical entropy S(E)S(E) controls fluctuations through its curvature. Expand around the most probable energy EE^{\ast}:

S(E)S(E)+12(EE)2(2SE2)ES(E) \approx S(E^{\ast}) + \frac{1}{2}(E-E^{\ast})^{2}\left(\frac{\partial^{2} S}{\partial E^{2}}\right)_{E^{\ast}}

Performing the Laplace transform to the canonical ensemble gives a Gaussian for EE with variance (kB/S)( -k_{B}/S'' ), equivalent to kBT2CVk_{B}T^{2}C_{V}. More generally, additive observables obey a large-deviation principle:

P ⁣(ANa)eNI(a)\mathbb P\!\left(\frac{A}{N}\approx a\right)\sim e^{-N I(a)}

where I(a)0I(a)\ge 0 is the rate function (Legendre transform of the cumulant generating function). At first-order transitions the relevant thermodynamic potential develops multiple competing saddles → non-Gaussian fluctuations and phase coexistence.


9.4.6 Critical fluctuations and correlation length

Let ϕ\phi be an order parameter (e.g., magnetization density). Define the connected correlator

G(r)ϕ(r)ϕ(0)cG(\boldsymbol r) \equiv \langle \phi(\boldsymbol r)\phi(\boldsymbol 0)\rangle_{c}

Away from criticality,

G(r)er/ξrd2+ηG(\boldsymbol r) \sim \frac{e^{-r/\xi}}{r^{d-2+\eta}}

with finite correlation length ξ\xi. The static susceptibility χddrG(r)\chi \equiv \int d^{d}r\, G(\boldsymbol r) scales as

χξ2η\chi \sim \xi^{2-\eta}

Therefore, in a volume VV, the variance of the spatially averaged order parameter scales like Var(ϕˉ)χ/V\mathrm{Var}(\bar\phi) \sim \chi/V. When ξ\xi is finite, relative fluctuations 1/V\sim 1/\sqrt{V}. As TTcT\to T_{c}, ξ\xi\to\infty and the 1/N1/\sqrt{N} law breaks down.


9.4.7 Finite-size scaling: when N is large but not infinite

  • For noncritical systems: Var(A)N\mathrm{Var}(A) \propto N for additive AA → relative size N1/2\propto N^{-1/2}.
  • Near criticality: with ξ\xi comparable to system size LL, replace ξ\xi by LL in scaling formulas; peaks in specific heat or susceptibility round and shift with LL in predictable ways (used to extract critical exponents numerically).

9.4.8 Example: noninteracting paramagnet and Curie law

NN spins s=12s=\tfrac12 in field BB with Hamiltonian H=μBiσiH=-\mu B\sum_{i}\sigma_{i}, σi=±1\sigma_{i}=\pm 1. The single-spin partition function is z=2cosh(βμB)z=2\cosh(\beta\mu B), magnetization per spin

mMN=μtanh(βμB)m \equiv \frac{\langle M\rangle}{N} = \mu\,\tanh(\beta\mu B)

Variance of MM is binomial:

(ΔM)2=Nμ2sech2(βμB)\langle(\Delta M)^{2}\rangle = N\,\mu^{2}\,\mathrm{sech}^{2}(\beta\mu B)

Static susceptibility χ(M/B)T\chi \equiv \left(\partial \langle M\rangle/\partial B\right)_{T} matches fluctuation–response:

χ=β(ΔM)2=Nμ2kBTsech2(βμB)\chi = \beta\,\langle(\Delta M)^{2}\rangle = \frac{N\mu^{2}}{k_{B}T}\,\mathrm{sech}^{2}(\beta\mu B)

In the weak-field limit, Curie law:

χNμ2kBT\chi \approx \frac{N\mu^{2}}{k_{B}T}

9.4.9 Example: ideal gas density fluctuations and S(0)

In a classical ideal gas, particles are uncorrelated: g(r)=1g(\boldsymbol r)=1. Then S(k)=1S(\boldsymbol k)=1 for all k\boldsymbol k and

S(0)=1=nkBTκTS(0) = 1 = n\,k_{B}T\,\kappa_{T}

so κT=1/(nkBT)\kappa_{T}=1/(n k_{B}T), the ideal-gas result. Number variance in a subvolume VV' is Poissonian: (ΔN)2=N\langle(\Delta N')^{2}\rangle=\langle N'\rangle.


9.4.10 Common pitfalls

  • Mixing ensembles. Don’t use canonical CVC_{V} with grand-canonical number fluctuations unless you’ve matched the ensemble to the measurement.
  • Forcing Gaussian fits near transitions. At first order, distributions are bimodal (two phases); at criticality, power-law tails win.
  • Ignoring constraints. Conserved totals suppress variances (e.g., fixed total NN reduces subvolume fluctuations).
  • Calling negative CVC_{V} “instability.” In microcanonical small systems with long-range forces, CV<0C_{V}<0 can appear; it signals ensemble inequivalence, not algebra failure.

9.4.11 Worked mini-examples

(a) N ⁣pTN\!pT volume variance
From Δ=dVeβpVZ\Delta=\int dV\,e^{-\beta pV}Z, show (ΔV)2=kBTVκT\langle(\Delta V)^{2}\rangle = k_{B}T\,V\,\kappa_{T} and verify it for an ideal gas.

(b) Compressibility sum rule
Starting with S(k)=1+nd3reikr[g(r)1]S(\boldsymbol k)=1+n\int d^{3}r\,e^{-i\boldsymbol k\cdot\boldsymbol r}[g(r)-1], take k0\boldsymbol k\to 0 and use thermodynamics to derive S(0)=nkBTκTS(0)=n k_{B}T\kappa_{T}.

(c) Energy large deviations
From Z=dEeS(E)/kBβEZ=\int dE\,e^{S(E)/k_{B}-\beta E}, do the quadratic expansion and read off (ΔE)2\langle(\Delta E)^{2}\rangle; identify when the Gaussian approximation fails.

(d) Paramagnet variance → Curie
Compute M\langle M\rangle and M2\langle M^{2}\rangle from zz and show χ=β(ΔM)2\chi=\beta\langle(\Delta M)^{2}\rangle; take B0B\to 0 to get χ1/T\chi\propto 1/T.

(e) Subvolume fluctuations
For an ideal gas, partition a box into two volumes VV' and VVV-V'. With fixed total NN, show Var(N)=N(V/V)(1V/V)\mathrm{Var}(N')=N(V'/V)(1-V'/V) (binomial), contrasting with grand-canonical Poisson variance.


9.4.12 Minimal problem kit

  • Prove the general static relation χXX=β(ΔX)2\chi_{XX}=\beta\langle(\Delta X)^{2}\rangle from derivatives of lnZ\ln Z with a source term
  • Derive (ΔN)2=kBT(N/μ)T,V\langle(\Delta N)^{2}\rangle = k_{B}T\left(\partial N/\partial\mu\right)_{T,V} and connect it to κT\kappa_{T} for a single-component fluid
  • Show that S(0)=nkBTκTS(0)=n k_{B}T\kappa_{T} implies S(0)S(0)\to\infty as a liquid approaches its gas–liquid critical point and explain “critical opalescence” qualitatively
  • Using a Landau ϕ4\phi^{4} free energy f=a(TTc)ϕ2+bϕ4f=a(T-T_{c})\phi^{2}+b\phi^{4}, compute Gaussian fluctuations above TcT_{c} and extract χ1/(TTc)\chi\propto 1/(T-T_{c}) at mean-field level
  • For a lattice model with conserved order parameter, discuss how conservation modifies long-time relaxation (but not equal-time static variances) at fixed TT

9.5 Quantum Gases: Bose & Fermi Statistics, Polylogs, and the Road to Degeneracy

Classical MB stats are great… until nλT3n\lambda_{T}^{3} isn’t tiny. Then indistinguishability and exchange symmetry flip the table. This section derives Bose–Einstein and Fermi–Dirac from the grand ensemble, computes thermodynamics with polylog functions, connects to the classical limit via virial expansions, and builds the degenerate Fermi-gas toolkit including the Sommerfeld expansion. We also map how the chemical potential moves with TT and nn.


9.5.1 Grand ensemble → BE/FD distributions

Put noninteracting identical particles in the grand canonical ensemble at (T,V,μ)(T,V,\mu). With fugacity zeβμz\equiv e^{\beta \mu} and one–particle levels {ϵα}\{\epsilon_{\alpha}\},

lnΞ=±αln ⁣(1zeβϵα)\ln \Xi = \pm \sum_{\alpha}\ln\!\left(1 \mp z\,e^{-\beta\epsilon_{\alpha}}\right)

upper sign for bosons, lower for fermions. Occupations are

nα=1z1eβϵα1\langle n_{\alpha}\rangle = \frac{1}{z^{-1}e^{\beta\epsilon_{\alpha}} \mp 1}

The minus in the denominator gives Bose–Einstein bunching, the plus gives Fermi–Dirac blocking.


9.5.2 Density of states in 3D and continuum replacement

For spin degeneracy gg and nonrelativistic dispersion ϵ=p2/2m\epsilon=\boldsymbol p^{2}/2m in volume VV,

g(ϵ)dϵ=gV4π2(2m2)3/2ϵdϵg(\epsilon)\,d\epsilon = g\,\frac{V}{4\pi^{2}}\left(\frac{2m}{\hbar^{2}}\right)^{3/2}\sqrt{\epsilon}\,d\epsilon

Sums become integrals αdϵg(ϵ)\sum_{\alpha}\to \int d\epsilon\,g(\epsilon).

Define the thermal wavelength

λT2π2mkBT\lambda_{T} \equiv \sqrt{\frac{2\pi \hbar^{2}}{m k_{B} T}}

and the polylogarithm Lis(x)k1xk/ks\operatorname{Li}_{s}(x) \equiv \sum_{k\ge1} x^{k}/k^{s}.


9.5.3 Polylog thermodynamics: N, U, and p

For both bosons and fermions, results compactify via polylogs. Let

gs(z)Lis(z),fs(z)Lis(z)g_{s}(z) \equiv \operatorname{Li}_{s}(z),\qquad f_{s}(z) \equiv -\,\operatorname{Li}_{s}(-z)

Then in d=3d=3 for a uniform nonrelativistic gas,

  • Bosons (BE)(\text{BE})
N=gVλT3g3/2(z),p=gkBTλT3g5/2(z),U=32pVN = \frac{gV}{\lambda_{T}^{3}}\,g_{3/2}(z),\qquad p = \frac{g k_{B}T}{\lambda_{T}^{3}}\,g_{5/2}(z),\qquad U = \frac{3}{2}\,pV
  • Fermions (FD)(\text{FD})
N=gVλT3f3/2(z),p=gkBTλT3f5/2(z),U=32pVN = \frac{gV}{\lambda_{T}^{3}}\,f_{3/2}(z),\qquad p = \frac{g k_{B}T}{\lambda_{T}^{3}}\,f_{5/2}(z),\qquad U = \frac{3}{2}\,pV

The virial identity U=32pVU=\tfrac{3}{2}pV holds for any ideal nonrelativistic gas with quadratic dispersion.


9.5.4 Classical limit and the first quantum correction

When z1z\ll 1 (high TT or low nn), both stats reduce to MB since gs(z)zg_{s}(z)\approx z and fs(z)zf_{s}(z)\approx z. Expanding further gives the leading quantum correction to the ideal–gas law

pnkBT=1±125/2nλT3+O ⁣((nλT3)2)\frac{p}{n k_{B}T} = 1 \pm \frac{1}{2^{5/2}}\,n\,\lambda_{T}^{3} + \mathcal O\!\left((n\lambda_{T}^{3})^{2}\right)

upper sign for fermions (Pauli “repulsion” raises pressure), lower sign for bosons (Bose “attraction” lowers it)


9.5.5 Chemical potential behavior

  • Bosons: μ0\mu \le 0. At fixed nn as TT decreases, μ\mu rises toward 00^{-}. For an ideal uniform gas, the excited-state capacity is capped by g3/2(1)=ζ(3/2)g_{3/2}(1)=\zeta(3/2), leading to Bose–Einstein condensation when nλT3n\lambda_{T}^{3} exceeds a constant. We detail BEC in §9.7.
  • Fermions: μ\mu is positive at low TT and approaches the Fermi energy ϵF\epsilon_{F} as T0T\to 0. With nN/Vn\equiv N/V and spin degeneracy gg,
ϵF=22m(6π2ng)2/3,kBTFϵF\epsilon_{F} = \frac{\hbar^{2}}{2m}\left(\frac{6\pi^{2} n}{g}\right)^{2/3},\qquad k_{B}T_{F} \equiv \epsilon_{F}

At finite but small T/TFT/T_{F}, μ(T)\mu(T) decreases from ϵF\epsilon_{F} by (T/TF)2\sim (T/T_{F})^{2}.


9.5.6 Degenerate Fermi gas and the Sommerfeld expansion

For TTFT\ll T_{F}, evaluate Fermi integrals asymptotically. Let xT/TF1x\equiv T/T_{F}\ll 1.

  • Chemical potential
μ(T)=ϵF[1π212x2π480x4+]\mu(T) = \epsilon_{F}\left[1 - \frac{\pi^{2}}{12}x^{2} - \frac{\pi^{4}}{80}x^{4} + \cdots\right]
  • Internal energy and pressure at low TT
U=35NϵF[1+5π212x2+],p=25nϵF[1+5π212x2+]U = \frac{3}{5}N\epsilon_{F}\left[1 + \frac{5\pi^{2}}{12}x^{2} + \cdots\right],\qquad p = \frac{2}{5} n \epsilon_{F}\left[1 + \frac{5\pi^{2}}{12}x^{2} + \cdots\right]
  • Heat capacity per particle
CVNkB=π22x+O(x3)\frac{C_{V}}{N k_{B}} = \frac{\pi^{2}}{2}\,x + \mathcal O(x^{3})

Linear in TT is the signature of a Fermi surface with a thin shell of thermally active states of thickness kBT\sim k_{B}T.


9.5.7 Nondegenerate Bose gas and the edge of BEC

Above condensation (no macroscopic ground-state occupation), μ<0\mu<0 solves nλT3=gg3/2(z)n\lambda_{T}^{3} = g\,g_{3/2}(z). Thermodynamics follows from §9.5.3 with z<1z<1. As TT is lowered at fixed nn, z1z\to 1^{-}, gs(z)ζ(s)g_{s}(z)\to \zeta(s), and the excited-state density saturates at

nex,max=gλT3ζ(3/2)n_{\text{ex,max}} = \frac{g}{\lambda_{T}^{3}}\,\zeta(3/2)

The “excess” particles must go into the ground state at and below the critical temperature found in §9.7.


9.5.8 Relativistic and massless limits (teaser)

For ϵ=pc\epsilon=pc or ϵ=p2c2+m2c4\epsilon=\sqrt{p^{2}c^{2}+m^{2}c^{4}}, the density of states changes to g(ϵ)ϵ2g(\epsilon)\propto \epsilon^{2} and UVT4U\propto VT^{4} for massless bosons (photons), with p=U/3Vp=U/3V. We develop photons and phonons in §9.6.


9.5.9 Worked mini-examples

(a) Classical limit check

Show that taking z1z\ll 1 in N=gVλT3g3/2(z)N = gV\lambda_{T}^{-3} g_{3/2}(z) and p=gkBTλT3g5/2(z)p = g k_{B}T \lambda_{T}^{-3} g_{5/2}(z) reproduces pV=NkBTpV=Nk_{B}T to leading order and yields the ±25/2nλT3\pm 2^{-5/2} n\lambda_{T}^{3} correction next.

(b) 3D density of states

Derive g(ϵ)g(\epsilon) by counting momentum states in a box and using ϵ=p2/2m\epsilon=\boldsymbol p^{2}/2m.

(c) Fermi energy numerics

For electrons in a metal with n8.5×1028 m3n\approx 8.5\times 10^{28}\ \mathrm{m}^{-3} and g=2g=2, compute ϵF\epsilon_{F} and TFT_{F}, then estimate CV/(NkB)C_{V}/(Nk_{B}) at T=300 KT=300\ \mathrm{K}.

(d) Low-TT μ(T)\mu(T) for fermions

Use the Sommerfeld expansion of dϵϵ1/2/(eβ(ϵμ)+1)\int d\epsilon\,\epsilon^{1/2}/(e^{\beta(\epsilon-\mu)}+1) to obtain the μ(T)\mu(T) series up to O(x4)\mathcal O(x^{4}).

(e) Bose gas near z1z\to 1^{-}

Evaluate g3/2(z)g_{3/2}(z) numerically close to z=1z=1 using the polylog expansion and show how nλT3n\lambda_{T}^{3} approaches gζ(3/2)g\,\zeta(3/2).


9.5.10 Minimal problem kit

  • Starting from lnΞ\ln\Xi, derive nα\langle n_{\alpha}\rangle and the BE/FD forms
  • Replace sums by dϵg(ϵ)\int d\epsilon\,g(\epsilon) to obtain the polylog formulas for NN, pp, and UU in d=3d=3
  • Expand p/(nkBT)p/(n k_{B}T) in powers of nλT3n\lambda_{T}^{3} and identify the second virial coefficient sign for bosons vs fermions
  • Derive the Fermi energy ϵF\epsilon_{F} and p(T=0)=25nϵFp(T=0)=\tfrac{2}{5}n\epsilon_{F}, then use Sommerfeld to obtain CVTC_{V}\propto T
  • At fixed nn for an ideal Bose gas, show μ0\mu\to 0^{-} as TTcT\downarrow T_{c} and compute the slope dμ/dTd\mu/dT at z1z\lesssim 1 from the number equation

9.6 Photons & Phonons: Blackbody Radiation and Heat Capacity of Solids

Two archetypal Bose gases, two very different worlds. Photons: massless, spin-1, number not conserved, μ=0\mu=0, pressure p=U/3Vp=U/3V, the T4T^{4} law. Phonons: lattice vibrations with linear dispersion at long wavelengths, three polarizations, and a Debye cutoff that enforces 3N3N modes. Together they explain the spectrum of light and why solids’ specific heats crash to zero at low TT.


9.6.1 Photon gas basics

Photons are created/annihilated freely in thermal equilibrium, so the chemical potential must be zero. Put EM waves in a large cubic cavity of volume VV with periodic boundary conditions. Counting transverse modes per angular frequency ω\omega gives the density of states

g(ω)dω=Vπ2c3ω2dωg(\omega)\,d\omega = \frac{V}{\pi^{2}c^{3}}\,\omega^{2}\,d\omega

Bose occupation with μ=0\mu=0 is nB(ω)=(eβω1)1n_{B}(\omega)=\left(e^{\beta\hbar\omega}-1\right)^{-1}, yielding the Planck energy density

u(ω)dωUV=ω3π2c3  dωeβω1u(\omega)\,d\omega \equiv \frac{U}{V} = \frac{\hbar\omega^{3}}{\pi^{2}c^{3}}\;\frac{d\omega}{e^{\beta\hbar\omega}-1}

The number spectrum follows by dividing by ω\hbar\omega

n(ω)dω=ω2π2c3  dωeβω1n(\omega)\,d\omega = \frac{\omega^{2}}{\pi^{2}c^{3}}\;\frac{d\omega}{e^{\beta\hbar\omega}-1}

9.6.2 Stefan–Boltzmann and Wien: two famous corollaries

Integrating u(ω)u(\omega) over all ω\omega gives

UVaT4,a=π2kB4153c3\frac{U}{V} \equiv a\,T^{4},\qquad a = \frac{\pi^{2}k_{B}^{4}}{15\,\hbar^{3}c^{3}}

Radiation pressure for an isotropic photon gas is

p=13UVp = \frac{1}{3}\,\frac{U}{V}

The radiative flux from a black surface is Φ=σT4\Phi=\sigma T^{4} with

σ=ac4=π2kB4603c2\sigma = \frac{a c}{4} = \frac{\pi^{2}k_{B}^{4}}{60\,\hbar^{3}c^{2}}

The wavelength of peak spectral exitance satisfies Wien’s displacement

λmaxT=b,b2.898×103 mK\lambda_{\text{max}}\,T = b,\qquad b \approx 2.898\times 10^{-3}\ \text{m}\cdot\text{K}

obtained by maximizing the λ\lambda-form of Planck’s law.

Entropy density and enthalpy follow from p=U/3Vp=U/3V:

sSV=43aT3,hU+pVV=43aT4s \equiv \frac{S}{V} = \frac{4}{3}aT^{3},\qquad h \equiv \frac{U+pV}{V} = \frac{4}{3}aT^{4}

9.6.3 Phonons: quantized sound with linear dispersion

In a crystal, small lattice displacements form normal modes. At long wavelengths the acoustic branches have

ω=vsk\omega = v_{s}\,k

with sound speed vsv_{s} and three polarizations (one longitudinal, two transverse). Treating them as a gas of noninteracting bosons, the Bose occupation is the same nB(ω)n_{B}(\omega), but the density of states uses vsv_{s} instead of cc and must be cut off to enforce a finite number of modes.

Debye’s key move: approximate the acoustic branches as linear up to a cutoff kDk_{D} chosen so that the total number of modes equals 3N3N for NN atoms

V2π20kDk2dk×3=3N    kD=(6π2NV)1/3\frac{V}{2\pi^{2}}\int_{0}^{k_{D}} k^{2}\,dk\times 3 = 3N \;\Rightarrow\; k_{D} = \left(6\pi^{2}\frac{N}{V}\right)^{1/3}

Define the Debye frequency and Debye temperature

ωDvskD,ΘDωDkB\omega_{D} \equiv v_{s}\,k_{D},\qquad \Theta_{D} \equiv \frac{\hbar\omega_{D}}{k_{B}}

A single “average” vsv_{s} captures the three polarizations to good accuracy for bulk thermodynamics.


9.6.4 Debye model

The phonon internal energy omitting zero-point pieces is

U=0ωDdω  gD(ω)ωnB(ω)U = \int_{0}^{\omega_{D}} d\omega\; g_{D}(\omega)\,\hbar\omega\,n_{B}(\omega)

with Debye density of states

gD(ω)=9NωD3ω2g_{D}(\omega) = \frac{9N}{\omega_{D}^{3}}\,\omega^{2}

Change variables xω/kBTx \equiv \hbar\omega/k_{B}T to obtain the standard Debye heat capacity

CV=9NkB(TΘD)30ΘD/Tdx  x4ex(ex1)2C_{V} = 9N k_{B}\left(\frac{T}{\Theta_{D}}\right)^{3} \int_{0}^{\Theta_{D}/T} dx\; \frac{x^{4} e^{x}}{(e^{x}-1)^{2}}

Asymptotics:

  • Low TT (TΘD)\left(T\ll \Theta_{D}\right)
CV12π45NkB(TΘD)3C_{V} \simeq \frac{12\pi^{4}}{5}\,N k_{B}\left(\frac{T}{\Theta_{D}}\right)^{3}
  • High TT (TΘD)\left(T\gg \Theta_{D}\right)
CV3NkBC_{V} \to 3N k_{B}

recovering Dulong–Petit.


9.6.5 Einstein model: optical modes and the crossover

Einstein imagined each atom as an independent oscillator of a single frequency ωE\omega_{E}. Then

CVEin=3NkB  x2ex(ex1)2,xωEkBTC_{V}^{\text{Ein}} = 3N k_{B}\;\frac{x^{2} e^{x}}{(e^{x}-1)^{2}},\qquad x \equiv \frac{\hbar\omega_{E}}{k_{B}T}

It captures the high-TT limit and the exponential freeze-out at low TT, but misses the T3T^{3} law because it lacks low-frequency acoustic modes. Real crystals often interpolate: optical modes look Einstein-like, acoustic modes look Debye-like.


9.6.6 Phonon gas thermodynamics

With ω=vsk\omega=v_{s}k and three polarizations, the energy density of a free phonon gas at low TΘDT\ll \Theta_{D} mirrors photons with cvsc\to v_{s} and an extra factor 33 from polarizations

uph(T)(kBT)43vs3u_{\text{ph}}(T) \propto \frac{(k_{B}T)^{4}}{\hbar^{3} v_{s}^{3}}

Differentiating gives CVT3C_{V}\propto T^{3} as in Debye. In a solid, the “pressure” of phonons modifies the elastic response rather than pushing on a volume like a gas, but the scaling intuition is identical.


9.6.7 Blackbodies in nature: quick tour

  • Cosmic microwave background is a near-perfect blackbody at T2.725 KT\approx 2.725\ \text{K}, a fossil photon gas filling the universe
  • Stars approximate blackbodies over parts of their spectrum; deviations encode atmospheres and lines
  • Laboratory cavities with small apertures mimic blackbody emitters and set radiometry standards

9.6.8 Common pitfalls

  • Forgetting μ=0\mu=0 for photons. Any attempt to fix photon number in equilibrium is unphysical unless you also constrain the spectrum with pumping
  • Mixing ω\omega and λ\lambda peaks. The maximum of uωu_{\omega} and of uλu_{\lambda} occur at different places; Wien’s bb refers to the wavelength peak
  • Ignoring polarizations or cutoffs. Debye needs 3N3N modes total; missing the factor 3 or the cutoff breaks CVC_{V}
  • High-TT limit with quantum units. Always check ΘD/T\Theta_{D}/T or x=ω/kBTx=\hbar\omega/k_{B}T before claiming “classical” behavior

9.6.9 Worked mini-examples

(a) Recover Stefan–Boltzmann.
Integrate u(ω)u(\omega) using 0dxx3/(ex1)=π4/15\int_{0}^{\infty} dx\,x^{3}/(e^{x}-1)=\pi^{4}/15 after x=βωx=\beta\hbar\omega to obtain U/V=aT4U/V=aT^{4} and p=U/3Vp=U/3V.

(b) Wien’s displacement.
Maximize the wavelength spectral density to show λmaxT=b\lambda_{\text{max}}T=b and extract bb numerically from the transcendental equation.

(c) Debye low-TT asymptotic.
Expand the Debye integral for ΘD/T\Theta_{D}/T\to\infty to derive CV(12π4/5)NkB(T/ΘD)3C_{V}\simeq(12\pi^{4}/5)N k_{B}(T/\Theta_{D})^{3}.

(d) Einstein vs Debye at intermediate TT.
Plot CV/(3NkB)C_{V}/(3Nk_{B}) vs T/ΘT/\Theta for both models and explain why real data usually bends between them.

(e) Photon gas entropy density.
Using p=U/3Vp=U/3V and dU=TdSpdVdU=TdS-p\,dV, derive s=(4/3)aT3s=(4/3)aT^{3}.


9.6.10 Minimal problem kit

  • Starting from the cavity mode count, derive g(ω)g(\omega) and Planck’s u(ω)u(\omega)
  • Prove p=U/3Vp=U/3V by averaging the Maxwell stress tensor for isotropic radiation
  • Show that the phonon mode count with Debye cutoff equals 3N3N and obtain kDk_{D}
  • Evaluate the Debye heat capacity integral numerically for T/ΘD={0.05,0.2,1,3}T/\Theta_{D}=\{0.05,0.2,1,3\} and compare to Einstein’s prediction at the same TT using ωE0.806ωD\omega_{E}\approx 0.806\,\omega_{D}
  • Derive the T4T^{4} scaling of the low-TT phonon energy density by mapping the photon calculation with cvsc\to v_{s}

9.7 Bose–Einstein Condensation & Degenerate Fermi Gas

Turn the dial to low temperature and high phase-space density and quantum statistics starts making executive decisions. Bosons pile into the ground state → Bose–Einstein condensation (BEC). Fermions lock into a Fermi sea with Pauli pressure. We build the ideal-gas baseline for BEC in 3D, add traps and interaction corrections, sketch superfluidity via Bogoliubov, and recap the degenerate Fermi gas with a couple of “why this matters” stops.


9.7.1 BEC in a uniform ideal Bose gas (3D)

From §9.5, for noninteracting bosons

N=gVλT3g3/2(z)N = \frac{gV}{\lambda_{T}^{3}}\,g_{3/2}(z)

with z=eβμz=e^{\beta\mu} and gsg_{s} the polylog. Because g3/2(z)g3/2(1)=ζ(3/2)g_{3/2}(z)\le g_{3/2}(1)=\zeta(3/2), the excited states can hold at most

Nex,max=gVλT3ζ(3/2)N_{\mathrm{ex,max}} = \frac{gV}{\lambda_{T}^{3}}\,\zeta(3/2)

At fixed density n=N/Vn=N/V, lowering TT increases 1/λT3T3/21/\lambda_{T}^{3}\propto T^{3/2} until the excited-state capacity saturates. The temperature where this happens is the critical temperature

Tc=2π2mkB(ngζ(3/2))2/3T_{c} = \frac{2\pi\hbar^{2}}{m k_{B}} \left(\frac{n}{g\,\zeta(3/2)}\right)^{2/3}

Below TcT_{c}, the chemical potential pins to μ=0\mu=0 and the condensed fraction is

N0N=1(TTc)3/2\frac{N_{0}}{N} = 1 - \left(\frac{T}{T_{c}}\right)^{3/2}

Thermodynamics below TcT_{c} comes entirely from excitations (N0N_{0} contributes no energy in the ideal gas):

p=gkBTλT3ζ(5/2),U=32pVp = \frac{g k_{B}T}{\lambda_{T}^{3}}\,\zeta(5/2),\qquad U = \frac{3}{2}pV

Specific heat per particle for T<TcT<T_{c}

CVNkB=154ζ(5/2)ζ(3/2)(TTc)3/2\frac{C_{V}}{N k_{B}} = \frac{15}{4}\,\frac{\zeta(5/2)}{\zeta(3/2)}\left(\frac{T}{T_{c}}\right)^{3/2}

Above TcT_{c}, μ<0\mu<0 solves nλT3=gg3/2(z)n\lambda_{T}^{3}=g\,g_{3/2}(z) and CVC_{V} approaches the classical 3NkB/23N k_{B}/2 at high TT. At TcT_{c} the ideal-gas CVC_{V} shows a cusp.


9.7.2 No BEC in uniform 1D/2D at finite TT

For a homogeneous ideal gas, Nexgd/2(z)N_{\mathrm{ex}}\propto g_{d/2}(z) in dd dimensions. The series diverges at z1z\to 1^{-} for d2d\le 2 because g1(1)g_{1}(1) and g1(z)g_{1}(z) blow up logarithmically. So no true BEC in uniform 1D/2D at finite TT; interactions in 2D can still support superfluidity via a Kosterlitz–Thouless transition, but that’s not a macroscopic population of a single k=0k=0 mode.


9.7.3 Trapped Bose gases: harmonic potential

Experiments confine atoms in approximately harmonic traps V(r)=12m(ωx2x2+ωy2y2+ωz2z2)V(\mathbf r)=\tfrac{1}{2}m(\omega_{x}^{2}x^{2}+\omega_{y}^{2}y^{2}+\omega_{z}^{2}z^{2}). The single-particle density of states becomes g(ϵ)ϵ2/(ωˉ)3g(\epsilon)\propto \epsilon^{2}/(\hbar\bar\omega)^{3} with geometric mean ωˉ(ωxωyωz)1/3\bar\omega\equiv(\omega_{x}\omega_{y}\omega_{z})^{1/3}. The critical temperature for an ideal trapped gas is

kBTctrap=ωˉ(Nζ(3))1/3k_{B}T_{c}^{\mathrm{trap}} = \hbar \bar\omega \left(\frac{N}{\zeta(3)}\right)^{1/3}

and, for T<TctrapT<T_{c}^{\mathrm{trap}},

N0N=1(TTctrap)3\frac{N_{0}}{N} = 1 - \left(\frac{T}{T_{c}^{\mathrm{trap}}}\right)^{3}

The steeper exponent reflects the different density of states.


9.7.4 Canonical vs grand-canonical fluctuations: the “catastrophe”

In the grand canonical ensemble for an ideal BEC, condensate-number fluctuations scale anomalously large near and below TcT_{c}, Var(N0)O(N2)\mathrm{Var}(N_{0})\sim \mathcal O(N^{2}), which is unphysical for finite isolated clouds. The canonical (or microcanonical) ensemble tames this to Var(N0)O(N)\mathrm{Var}(N_{0})\sim \mathcal O(N) once the fixed total NN constraint is enforced. Moral: choose the ensemble your experiment actually realizes.


9.7.5 Interactions and the Gross–Pitaevskii (GP) mean field

At ultracold TT, ss-wave scattering dominates with amplitude set by the scattering length aa. The effective coupling is

gint=4π2amg_{\mathrm{int}} = \frac{4\pi\hbar^{2} a}{m}

For a weakly interacting condensate with macroscopic wavefunction Ψ(r,t)\Psi(\mathbf r,t),

itΨ=[222m+V(r)+gintΨ2]Ψi\hbar\,\partial_{t}\Psi = \left[-\frac{\hbar^{2}\nabla^{2}}{2m}+V(\mathbf r)+g_{\mathrm{int}}\,|\Psi|^{2}\right]\Psi

This is the Gross–Pitaevskii equation. In uniform matter with density n=Ψ2n=|\Psi|^{2},

μ=gintn,c=gintnm,ξ2mc\mu = g_{\mathrm{int}}\,n,\qquad c = \sqrt{\frac{g_{\mathrm{int}}\,n}{m}},\qquad \xi \equiv \frac{\hbar}{\sqrt{2} m c}

where cc is the sound speed and ξ\xi the healing length, the scale over which the order parameter recovers near defects or boundaries.


9.7.6 Bogoliubov spectrum and superfluidity (sketch)

Linearizing fluctuations around the condensate gives the Bogoliubov dispersion

ϵk=(2k22m)2+c22k2\epsilon_{k} = \sqrt{\left(\frac{\hbar^{2}k^{2}}{2m}\right)^{2} + c^{2}\hbar^{2}k^{2}}

Low kk is phononic ϵkck\epsilon_{k}\approx \hbar c k and high kk crosses over to free-particle ϵk2k2/2m\epsilon_{k}\approx \hbar^{2}k^{2}/2m. The Landau criterion sets the critical flow speed

vc=minkϵkk=cv_{c} = \min_{k}\frac{\epsilon_{k}}{\hbar k} = c

which is why weakly interacting BECs are superfluids with frictionless flow below cc.


9.7.7 Degenerate Fermi gas recap: T=0 and small T

From §9.5, the noninteracting Fermi gas has Fermi energy

ϵF=22m(6π2ng)2/3,p(0)=25nϵF,U(0)=35NϵF\epsilon_{F} = \frac{\hbar^{2}}{2m}\left(\frac{6\pi^{2}n}{g}\right)^{2/3},\qquad p(0) = \frac{2}{5}n\,\epsilon_{F},\qquad U(0) = \frac{3}{5}N\,\epsilon_{F}

Finite-TT corrections follow from the Sommerfeld expansion: CVTC_{V}\propto T, μ(T)ϵF[1π212(T/TF)2+]\mu(T)\approx \epsilon_{F}\left[1-\frac{\pi^{2}}{12}(T/T_{F})^{2}+\cdots\right]. The pressure at T=0T=0 embodies Pauli pressure, stabilizing systems from collapse and setting scales in metals and compact objects.

Harmonic trap. For NN fermions in a 3D harmonic trap with ωˉ\bar\omega, the T=0T=0 Fermi energy is

ϵFtrap=ωˉ(6N/g)1/3\epsilon_{F}^{\mathrm{trap}} = \hbar \bar\omega\,(6N/g)^{1/3}

and the cloud radius follows from a Thomas–Fermi profile.


9.7.8 Two quick applications

  • White dwarfs (cartoon). Electrons form a highly degenerate gas. In the relativistic limit, pn4/3p\sim n^{4/3} instead of n5/3n^{5/3}, which leads to a maximum supported mass (Chandrasekhar limit). The point is qualitative here: degeneracy pressure is real, thermal pressure is optional.
  • Cold-atom labs. Feshbach resonances tune aa across zero → weakly interacting BECs, strongly interacting unitary Fermi gases, and the BEC–BCS crossover. Bogoliubov sound and collective modes are textbook confirmations of the theory above.

9.7.9 Common pitfalls

  • Forgetting that μ=0\mu=0 below TcT_{c} for the ideal Bose gas. Don’t force a negative μ\mu once the condensate forms
  • Using grand-canonical fluctuations for a fixed-NN trap. Canonical variance is the right one for cold-atom clouds
  • Claiming uniform 2D BEC. Not at finite TT; look for KT superfluidity instead
  • Treating interactions as “small” without checking na3n a^{3}. Weak-coupling GP/Bogoliubov needs na31n a^{3}\ll 1

9.7.10 Worked mini-examples

(a) TcT_{c} for a typical BEC.
Take 87Rb\mathrm{^{87}Rb} with m1.44×1025 kgm\approx 1.44\times 10^{-25}\ \mathrm{kg}, g=1g=1, and density n=1020 m3n=10^{20}\ \mathrm{m^{-3}}. Plug into the TcT_{c} formula and estimate the condensate fraction at T=0.5TcT=0.5T_{c}.

(b) No BEC in 2D.
Show Nexg1(z)N_{\mathrm{ex}} \propto g_{1}(z) diverges as z1z\to 1^{-} for a uniform 2D gas by converting the sum to an integral and identifying the logarithm.

(c) Trap TcT_{c}.
Given N=2×105N=2\times 10^{5} atoms and ωˉ=2π×100 Hz\bar\omega=2\pi\times 100\ \mathrm{Hz}, compute TctrapT_{c}^{\mathrm{trap}} and the condensed fraction at 0.7Tctrap0.7T_{c}^{\mathrm{trap}}.

(d) Healing length.
For a=5 nma=5\ \mathrm{nm} and n=1020 m3n=10^{20}\ \mathrm{m^{-3}}, estimate gintg_{\mathrm{int}}, cc, and ξ\xi.

(e) Bogoliubov slope.
Verify that ϵkck\epsilon_{k}\approx \hbar c k at k1/ξk\ll 1/\xi and find the crossover kk where the dispersion deviates by, say, 10%10\%.

(f) Fermi pressure number check.
For electrons with n=8×1028 m3n=8\times 10^{28}\ \mathrm{m^{-3}}, compute ϵF\epsilon_{F} and p(0)=25nϵFp(0)=\tfrac{2}{5}n\epsilon_{F}; compare to atmospheric pressure.


9.7.11 Minimal problem kit

  • Derive TcT_{c} and the condensed fraction for a uniform ideal Bose gas, then repeat for a harmonic trap using the appropriate density of states
  • Starting from p=(gkBT/λT3)ζ(5/2)p=(g k_{B}T/\lambda_{T}^{3})\zeta(5/2), compute CVC_{V} below TcT_{c} and show the cusp at TcT_{c}
  • Show explicitly why g3/2(1)g_{3/2}(1) converges in 3D but g1(1)g_{1}(1) diverges in 2D, and relate this to the absence of uniform 2D BEC
  • Linearize the GP equation around a uniform solution and derive the Bogoliubov dispersion and Landau critical velocity
  • Use the Sommerfeld expansion to obtain U(T)U(T) and CV(T)C_{V}(T) of a 3D Fermi gas through order (T/TF)2(T/T_{F})^{2}; derive ϵFtrap\epsilon_{F}^{\mathrm{trap}} for a harmonic confinement

9.8 Mean-Field & Landau Theory: Order Parameters, van der Waals, Curie–Weiss

Mean-field is the “group project” approximation: every degree of freedom feels the average of everyone else. It often nails the shape of phase diagrams and gives analytic critical exponents. Landau theory turns that idea into a controlled expansion in the order parameter and symmetries. This section builds Curie–Weiss magnetism, the van der Waals fluid with Maxwell construction, Landau free energies and exponents, correlation length from Landau–Ginzburg, the Ginzburg criterion, and where first-order and tricritical points sneak in.


9.8.1 What mean-field assumes

Pick an order parameter ϕ\phi that flips sign under the broken symmetry:

  • Ferromagnet: ϕm\phi \equiv m (magnetization density), symmetry mmm\to -m
  • Liquid–gas: ϕnnc\phi \equiv n-n_{c} (density deviation), no Ising symmetry but same math near TcT_{c}

Mean-field idea: Replace neighbors by their average. Fluctuations are ignored when coordination is high or the interaction is long-ranged, which gets increasingly true in higher spatial dimension dd.


9.8.2 Curie–Weiss ferromagnet (Ising-like, z neighbors)

Start with Ising spins σi=±1\sigma_{i}=\pm 1, coupling J>0J>0, field hh. In mean-field each spin sees an effective field heff=h+zJmh_{\text{eff}}=h + z J m with mσm\equiv \langle \sigma\rangle. The single-site partition function is z1=2cosh(βheff)z_{1}=2\cosh(\beta h_{\text{eff}}), giving the self-consistency equation

m=tanh ⁣[β(h+zJm)]m = \tanh\!\left[\beta\,(h + z J m)\right]

At h=0h=0, expand for small mm to get the Curie–Weiss critical temperature

kBTc=zJk_{B}T_{c} = z J

Critical behavior:

  • T>TcT>T_{c} small hh: mχhm \approx \chi h with
χ=1kB(TTc)\chi = \frac{1}{k_{B}(T-T_{c})}

so γMF=1\gamma_{\text{MF}}=1

  • T<TcT<T_{c} at h=0h=0: nonzero solutions appear
m(1TTc)1/2m \sim \left(1-\frac{T}{T_{c}}\right)^{1/2}

so βMF=1/2\beta_{\text{MF}}=1/2

  • At T=TcT=T_{c}, mh1/3m \propto h^{1/3}, so δMF=3\delta_{\text{MF}}=3

Specific heat has a finite jump (no divergence): αMF=0\alpha_{\text{MF}}=0


9.8.3 Landau free energy: symmetry dictates the form

Write an analytic expansion in the order parameter consistent with symmetries. For an Ising-like scalar mm in a uniform system

F(m;T,h)=F0(T)+a2tm2+a4m4+a6m6hmF(m;T,h) = F_{0}(T) + a_{2}\,t\,m^{2} + a_{4}\,m^{4} + a_{6}\,m^{6} - h\,m

with reduced temperature tTTct\equiv T-T_{c} and a4>0a_{4}>0 for a standard second-order transition. Minimize:

Fm=2a2tm+4a4m3h=0\frac{\partial F}{\partial m} = 2 a_{2}\,t\,m + 4 a_{4}\,m^{3} - h = 0

Consequences at h=0h=0:

  • For t>0t>0: minimum at m=0m=0
  • For t<0t<0: minima at
m(t)=±a2t2a4m_{\star}(t) = \pm \sqrt{\frac{-a_{2}\,t}{2 a_{4}}}

so βMF=1/2\beta_{\text{MF}}=1/2

Susceptibility above TcT_{c} from linear response:

χ1=2Fm2m=0=2a2t\chi^{-1} = \left.\frac{\partial^{2}F}{\partial m^{2}}\right|_{m=0} = 2 a_{2}\,t

so χt1\chi \propto t^{-1} and γMF=1\gamma_{\text{MF}}=1

At t=0t=0 with h0h\neq 0:

4a4m3=hmh1/34 a_{4}\,m^{3} = h \quad\Rightarrow\quad m \propto h^{1/3}

so δMF=3\delta_{\text{MF}}=3

Specific heat jump. Plug mm_{\star} into FF:

Fmin(t<0)=F0(T)a224a4t2F_{\min}(t<0) = F_{0}(T) - \frac{a_{2}^{2}}{4 a_{4}}\,t^{2}

Therefore C=T2F/T2C = -T\,\partial^{2}F/\partial T^{2} gets a finite step at TcT_{c}, i.e., αMF=0\alpha_{\text{MF}}=0 with a discontinuity.


9.8.4 First order and tricritical

If microscopic physics drives a4<0a_{4}<0, the m4m^{4} term destabilizes. Stabilize with a6>0a_{6}>0. Then as tt varies, two minima at ±m\pm m appear before t=0t=0 and the system jumps at coexistence t=tcoext=t_{\text{coex}} where the wells are equal depth. That jump is a first-order transition with latent heat.

The point where a4a_{4} changes sign at t=0t=0 is a tricritical point. Landau analysis with a4=0a_{4}=0 predicts different exponents, e.g., m(t)1/4m\sim (-t)^{1/4}.


9.8.5 Landau–Ginzburg functional and correlation length

Allow slow spatial variations of the order parameter. The Landau–Ginzburg free-energy functional is

F[ϕ]=ddr  [r02ϕ2+c2(ϕ)2+u4ϕ4hϕ]\mathcal F[\phi] = \int d^{d}r\;\left[\frac{r_{0}}{2}\,\phi^{2} + \frac{c}{2}\,(\nabla \phi)^{2} + \frac{u}{4}\,\phi^{4} - h\,\phi\right]

with r0tr_{0}\propto t. In the Gaussian regime (uu small, t>0t>0), the two-point correlator in Fourier space has Ornstein–Zernike form

S(k)ϕk21r0+ck2S(\boldsymbol k) \equiv \langle|\phi_{\boldsymbol k}|^{2}\rangle \propto \frac{1}{r_{0} + c k^{2}}

Define the correlation length

ξ2=cr0t1\xi^{2} = \frac{c}{r_{0}} \propto t^{-1}

hence νMF=1/2\nu_{\text{MF}}=1/2 and the real-space correlator decays as er/ξ/rd2e^{-r/\xi}/r^{d-2}, implying ηMF=0\eta_{\text{MF}}=0.


9.8.6 Ginzburg criterion: when mean-field breaks

Mean-field ignores order-parameter fluctuations. Compare fluctuation size in a correlation volume ξd\xi^{d} to m2m_{\star}^{2}. Roughly,

δϕ2ξm2kBTuξdm2\frac{\langle \delta \phi^{2}\rangle_{\xi}}{m_{\star}^{2}} \sim \frac{k_{B}T}{u\,\xi^{d}\,m_{\star}^{2}}

Using m2t/um_{\star}^{2}\sim |t|/u and ξt1/2\xi\sim t^{-1/2} gives a Ginzburg number Gi\mathrm{Gi} such that mean-field is valid only for

tGi|t| \gg \mathrm{Gi}

In short: for d>4d>4 the fluctuation integral is benign and mean-field is asymptotically exact. For d4d\le 4, a window close to TcT_{c} is fluctuation dominated and requires RG (§9.10).


9.8.7 van der Waals fluid: mean-field liquid–gas

Write the equation of state in number-density form n=N/Vn=N/V:

p(n,T)=nkBT1bnan2p(n,T) = \frac{n\,k_{B}T}{1 - b n} - a\,n^{2}

with excluded-volume bb and attraction aa. The critical point satisfies (p/n)Tc=(2p/n2)Tc=0(\partial p/\partial n)_{T_{c}}=(\partial^{2}p/\partial n^{2})_{T_{c}}=0, giving

nc=13b,kBTc=8a27b,pc=a27b2n_{c} = \frac{1}{3b},\qquad k_{B}T_{c} = \frac{8 a}{27 b},\qquad p_{c} = \frac{a}{27 b^{2}}

In molar form p=RT/(Vmb)a/Vm2p = RT/(V_{m}-b) - a/V_{m}^{2} one gets Vc=3bV_{c}=3b, Tc=8a/(27bR)T_{c}=8a/(27 b R), pc=a/(27b2)p_{c}=a/(27 b^{2}).

Define reduced variables prp/pcp_{r}\equiv p/p_{c}, vrVm/Vcv_{r}\equiv V_{m}/V_{c}, TrT/TcT_{r}\equiv T/T_{c} to get the law of corresponding states

pr=8Tr3vr13vr2p_{r} = \frac{8 T_{r}}{3 v_{r}-1} - \frac{3}{v_{r}^{2}}

Below TcT_{c}, p(V)p(V) has an S-shape. The mechanical instability region satisfies (p/V)T>0(\partial p/\partial V)_{T}>0; the spinodals solve (p/V)T=0(\partial p/\partial V)_{T}=0. The coexistence pressure at a given T<TcT<T_{c} is set by the Maxwell equal-area rule (since g=μg=\mu must be equal in both phases).

Near TcT_{c} the density difference obeys the mean-field law

Δn(TcT)1/2\Delta n \propto (T_{c}-T)^{1/2}

matching βMF=1/2\beta_{\text{MF}}=1/2.


9.8.8 Metastability, nucleation, and spinodals

Landau’s double-well for t<0t<0 with a small field hh shows two minima of unequal depth: the higher is metastable. Decay proceeds by nucleation of critical droplets, not captured by static mean-field but hinted by the barrier between minima. Approaching a spinodal the barrier vanishes; susceptibility and correlation length blow up along that limit in mean-field.


9.8.9 Cheat sheet: mean-field exponents and scaling relations

Mean-field exponents (Ising-like):

ExponentMeaningValue
α\alphaCtαC \sim \lvert t\rvert^{-\alpha}00 with jump
β\betam(t)βm \sim (-t)^{\beta}1/21/2
γ\gammaχtγ\chi \sim \lvert t\rvert^{-\gamma}11
δ\deltamh1/δm \sim h^{1/\delta} at t=0t=033
ν\nuξtν\xi \sim \lvert t\rvert^{-\nu}1/21/2
η\etaG(r)r(d2+η)G(r)\sim r^{-(d-2+\eta)} at t=0t=000

Scaling relations (satisfied by MF):

  • Widom: γ=β(δ1)\gamma = \beta(\delta-1)
  • Rushbrooke: α+2β+γ=2\alpha + 2\beta + \gamma = 2
  • Josephson (hyperscaling): 2α=dν2 - \alpha = d\,\nu
    • Holds in MF only for ddc=4d \ge d_{c}=4; below dcd_{c} hyperscaling fails without RG corrections

9.8.10 Common pitfalls

  • Forcing m0m\neq 0 above TcT_{c} at h=0h=0. In Landau with a4>0a_{4}>0, the only minimum for t>0t>0 is m=0m=0
  • Confusing coexistence with spinodal. Maxwell construction sets coexistence; spinodal is the limit of metastability where (p/V)T=0(\partial p/\partial V)_{T}=0
  • Ignoring symmetry. If the symmetry forbids odd terms, don’t insert them by hand. If a field hh breaks it, include hm-h m
  • Declaring MF “exact”. Good far from TcT_{c}, in high dd, or with long-range interactions; close to TcT_{c} in d4d\le 4 see §9.10

9.8.11 Worked mini-examples

(a) Curie–Weiss χ\chi
Linearize m=tanh[β(zJm+h)]m=\tanh[\beta(zJ m + h)] for small mm at h=0h=0 to derive χ=1/[kB(TTc)]\chi=1/[k_{B}(T-T_{c})] and identify TcT_{c}.

(b) Landau m(T)m(T) and CC-jump
Minimize F=a2tm2+a4m4F= a_{2} t m^{2} + a_{4} m^{4} at h=0h=0 for t<0t<0 to get m2=a2t/(2a4)m_{\star}^{2}=-a_{2}t/(2a_{4}). Insert into FF and compute the specific-heat discontinuity at TcT_{c}.

(c) Ornstein–Zernike
From the quadratic LG functional, derive S(k)1/(r0+ck2)S(k)\propto 1/(r_{0}+c k^{2}) and identify ξ=c/r0\xi=\sqrt{c/r_{0}}.

(d) van der Waals critical point
Using p=nkBT/(1bn)an2p = n k_{B}T/(1-b n) - a n^{2}, set np=n2p=0\partial_{n}p=\partial_{n}^{2}p=0 to find (nc,Tc,pc)(n_{c},T_{c},p_{c}); then write the reduced EOS.

(e) Tricritical scaling
Set a4=0a_{4}=0, keep a6>0a_{6}>0, minimize F=a2tm2+a6m6F=a_{2} t m^{2}+a_{6} m^{6}. Show m(t)1/4m\sim (-t)^{1/4} and extract βtri\beta_{\text{tri}}.


9.8.12 Minimal problem kit

  • Derive the MF exponents α,β,γ,δ\alpha,\beta,\gamma,\delta from the Landau polynomial and verify Widom and Rushbrooke relations
  • Starting from the LG functional, compute the real-space correlator G(r)G(r) in dd dimensions and confirm ηMF=0\eta_{\text{MF}}=0
  • Apply the Ginzburg criterion to estimate the width tGi|t|\lesssim \mathrm{Gi} of the non-MF critical region for a 3D Ising-like magnet with given uu and cc
  • For van der Waals, implement the Maxwell equal-area construction analytically near TcT_{c} and show Δn(TcT)1/2\Delta n \propto (T_{c}-T)^{1/2}
  • Explore the first-order case with a4<0a_{4}<0: find coexistence tcoext_{\text{coex}} by equating well depths and locate the spinodals where the second derivative at a minimum vanishes

9.9 The Ising Model: Transfer Matrices, Duality, and Exact 2D Criticality

If statistical physics had a mascot, it’d be the Ising model: spins σi=±1\sigma_{i}=\pm 1 talking only to neighbors, yet somehow recreating the drama of phase transitions. In 1D it’s chill (no finite-TT transition). In 2D it’s iconic: an exact critical point with non-mean-field exponents and logarithmic specific-heat blow-up. This section does the 1D exact solution, domain-wall intuition, Kramers–Wannier duality, Onsager’s highlights, and the scaling data you actually use.


9.9.1 Definition and notation

On a lattice with nearest neighbors ij\langle ij\rangle and external field HH,

H({σ})=JijσiσjHiσiH(\{\sigma\}) = -J \sum_{\langle ij\rangle} \sigma_{i}\sigma_{j} - H \sum_{i} \sigma_{i}

We will also use the dimensionless couplings

KβJ,h~βHK \equiv \beta J,\qquad \tilde h \equiv \beta H

Positive JJ favors alignment (ferromagnet). Partition function Z={σ}eβH({σ})Z=\sum_{\{\sigma\}} e^{-\beta H(\{\sigma\})}.


9.9.2 1D chain: transfer matrix exact solution

For a ring of NN spins (periodic boundary conditions), build the 2×22\times 2 transfer matrix

Tσ,σ=exp ⁣[Kσσ+h~2(σ+σ)]T_{\sigma,\sigma'} = \exp\!\left[K\,\sigma\sigma' + \frac{\tilde h}{2}(\sigma+\sigma')\right]

Its eigenvalues are

λ±=eKcoshh~±e2Ksinh2h~+e2K\lambda_{\pm} = e^{K}\cosh \tilde h \pm \sqrt{e^{2K}\sinh^{2}\tilde h + e^{-2K}}

Thermodynamics in the NN\to\infty limit is governed by λ+\lambda_{+}:

f1βNlnZ=1βlnλ+f \equiv -\frac{1}{\beta N}\ln Z = -\frac{1}{\beta}\ln \lambda_{+}

In zero field h~=0\tilde h=0,

λ+=2coshK,λ=2sinhK\lambda_{+}=2\cosh K,\qquad \lambda_{-}=2\sinh K

There is no spontaneous magnetization at any T>0T>0, i.e., m=limh~01Nh~lnZ=0m=\lim_{\tilde h\to 0}\frac{1}{N}\partial_{\tilde h}\ln Z=0. The connected two-point function decays exponentially,

σ0σrc=(λλ+)r\langle \sigma_{0}\sigma_{r}\rangle_{c} = \left(\frac{\lambda_{-}}{\lambda_{+}}\right)^{r}

so the correlation length is

ξ1=lntanhK\xi^{-1} = -\ln \tanh K

Finite at all T>0T>0, blowing up only at T=0T=0. Verdict: 1D Ising has no finite-TT phase transition.


9.9.3 Domain-wall (kink) intuition: why 1D stays disordered

Flip a contiguous block and you create two domain walls, each costing energy 2J2J. But their entropy grows like lnN\ln N (many places to put them). The free energy of a wall pair is ΔF22JkBTlnN\Delta F \sim 2\cdot 2J - k_{B}T\ln N, so for any T>0T>0 and large NN the entropic term wins → domain walls proliferate → long-range order is washed out. In 2D the energy grows with perimeter, and entropy competes differently, allowing a finite TcT_{c}.


9.9.4 2D square lattice

Kramers–Wannier duality maps the partition function at coupling KK to that at a dual coupling KK^{\star} via

sinh(2K)sinh(2K)=1\sinh(2K)\,\sinh(2K^{\star}) = 1

The critical point sits at self-duality, K=KK=K^{\star}, giving

sinh(2Kc)=1\sinh(2K_{c}) = 1

Equivalently,

kBTcJ=2ln(1+2)\frac{k_{B}T_{c}}{J} = \frac{2}{\ln(1+\sqrt{2})}

For anisotropic couplings Jx,JyJ_{x},J_{y}, the critical line is sinh(2Kx)sinh(2Ky)=1\sinh(2K_{x})\sinh(2K_{y})=1.


9.9.5 Onsager’s solution: exact critical behavior (at H=0)

The full free energy at H=0H=0 is exactly known (we won’t rewrite the integral here), but the key outputs are:

  • Specific heat diverges logarithmically
Clnt,tTTcTcC \sim -\ln|t|,\qquad t \equiv \frac{T-T_{c}}{T_{c}}

so α=0\alpha=0 with a log

  • Spontaneous magnetization for T<TcT<T_{c} (Yang)
m(T)=[1sinh4(2K)]1/8m(T) = \left[\,1 - \sinh^{-4}(2K)\,\right]^{1/8}
  • Critical exponents (2D Ising universality)
α=0,β=18,γ=74,δ=15,ν=1,η=14\alpha=0,\quad \beta=\frac{1}{8},\quad \gamma=\frac{7}{4},\quad \delta=15,\quad \nu=1,\quad \eta=\frac{1}{4}
  • Correlation length diverges as ξtν\xi \sim |t|^{-\nu} with ν=1\nu=1; the critical correlator decays as G(r)rd+2η=r1/4G(r)\sim r^{-d+2-\eta}=r^{-1/4} in 2D

The 2D Ising critical point is also a conformal field theory with central charge c=12c=\tfrac{1}{2} (useful later for finite-size scaling and exact amplitude ratios).


9.9.6 Transfer matrices, scaling, and finite-size tricks

  • Transfer-matrix ↔ quantum mapping. The 2D classical Ising model maps to the 1D quantum transverse-field Ising chain via Trotter decomposition. Classical T=TcT=T_{c} \leftrightarrow quantum zero-temperature critical point. Many finite-size formulas port directly.
  • Finite-size scaling. In a box of linear size LL, susceptibilities and order parameter moments scale at TcT_{c} as
χmax(L)Lγ/ν=L7/4,m(Tc,L)Lβ/ν=L1/8\chi_{\max}(L) \propto L^{\gamma/\nu} = L^{7/4},\qquad m(T_{c},L) \propto L^{-\beta/\nu} = L^{-1/8}

Binder cumulants and crossing methods are practical ways to locate TcT_{c} in simulations.


9.9.7 What universality buys you

You never need the lattice-level math again. Any 2D short-range system with a Z2\mathbb Z_{2} symmetry breaking (liquid–gas along the critical isotherm, uniaxial magnets, binary alloys) shares the same exponents and scaling functions near criticality. Landau mean-field (§9.8) is qualitatively right but its exponents differ; renormalization (§9.10) explains why.


9.9.8 Common pitfalls

  • Forgetting 1D has no finite-TT order. If your code finds m0m\neq 0 in 1D at T>0T>0 with H=0H=0, you’re finite-size–fooled or symmetry-broken by the algorithm
  • Mixing fields. The closed-form m(T)m(T) above is only for H=0H=0; turning on HH destroys the critical singularities in that exact way
  • Wrong TcT_{c}. It’s sinh(2Kc)=1\sinh(2K_{c})=1, not tanh\tanh or cosh\cosh; equivalently kBTc/J=2/ln(1+2)k_{B}T_{c}/J = 2/\ln(1+\sqrt{2})
  • Mean-field exponents in 2D. MF gives β=1/2\beta=1/2, γ=1\gamma=1, etc.—not correct for 2D Ising; use the exact set above
  • Correlation length sign errors. In 1D, ξ1=ln(tanhK)\xi^{-1}=-\ln(\tanh K); note the minus sign and that tanhK<1\tanh K<1 for T>0T>0

9.9.9 Worked mini-examples

(a) 1D transfer matrix at H=0H=0.
Derive λ±=2(coshK, sinhK)\lambda_{\pm}=2(\cosh K,\ \sinh K), get f=kBTln(2coshK)f=-k_{B}T\ln(2\cosh K), and show ξ1=lntanhK\xi^{-1}=-\ln\tanh K.

(b) Kramers–Wannier duality.
Starting from high-temperature and low-temperature series for lnZ\ln Z, show that up to a prefactor they coincide under sinh(2K)sinh(2K)=1\sinh(2K)\sinh(2K^{\star})=1, then deduce KcK_{c} from self-duality.

(c) Magnetization check.
Verify m(T ⁣ ⁣0)1m(T\!\to\!0)\to 1 from m(T)=[1sinh4(2K)]1/8m(T)=\left[1-\sinh^{-4}(2K)\right]^{1/8} and that m(T ⁣ ⁣Tc)t1/8m(T\!\to\!T_{c}^{-})\sim |t|^{1/8}.

(d) Finite-size scaling of χ\chi.
Assuming χ(t,L)=Lγ/νΦ(tL1/ν)\chi(t,L)=L^{\gamma/\nu}\,\Phi(t L^{1/\nu}), show that the peak value scales as L7/4L^{7/4} in 2D and how you’d extract ν\nu from data collapse.

(e) 1D domain-wall gas.
Treat kinks as noninteracting with fugacity ye2Ky\equiv e^{-2K}. Show the two-point function (12y)r\propto (1-2y)^{r} at leading order, reproducing ξ1ln(12y)\xi^{-1}\approx -\ln(1-2y) for y1y\ll 1.


9.9.10 Minimal problem kit

  • Build the 1D transfer matrix with field h~\tilde h and derive λ±\lambda_{\pm} and f(T,h~)f(T,\tilde h); compute m=h~lnλ+m=\partial_{\tilde h}\ln \lambda_{+} and check m(h~0)=0m(\tilde h\to 0)=0 for T>0T>0
  • Use duality to get the anisotropic 2D critical line sinh(2Kx)sinh(2Ky)=1\sinh(2K_{x})\sinh(2K_{y})=1; reduce to the isotropic sinh(2Kc)=1\sinh(2K_{c})=1
  • From the exact m(T)m(T), extract β=1/8\beta=1/8 and confirm the Widom relation γ=β(δ1)\gamma=\beta(\delta-1) with δ=15\delta=15 and γ=7/4\gamma=7/4
  • Show that the 2D specific heat behaves as ClntC \sim -\ln|t| by expanding Onsager’s free-energy integral near TcT_{c}
  • Map the 2D classical model to the 1D quantum transverse-field Ising chain and identify the quantum critical point in terms of transverse field and exchange

9.10 Critical Phenomena & Renormalization Group Basics

Close to a continuous phase transition the system forgets microscopic details and becomes scale invariant. Correlation length ξ\xi blows up, fluctuations span all sizes, and thermodynamics develops power laws with universal exponents. The renormalization group (RG) turns this into math: coarse-grain, rescale, track how couplings flow. Fixed points rule the neighborhood, and their eigenvalues set the exponents.


9.10.1 Why mean-field breaks

Mean-field pretends fluctuations are small. But at a critical point the correlation length

ξtν,tTTcTc\xi \sim |t|^{-\nu},\qquad t \equiv \frac{T-T_{c}}{T_{c}}

diverges, the correlator at t=0t=0 becomes a power law

G(r)ϕ(r)ϕ(0)c1rd2+ηG(r) \equiv \langle \phi(\mathbf r)\phi(\mathbf 0)\rangle_{c} \sim \frac{1}{r^{d-2+\eta}}

and the free-energy density’s singular part scales like fsξdf_{s}\sim \xi^{-d}. You cannot expand around a single length scale when there isn’t one.


9.10.2 Scaling hypothesis and homogeneity laws

Assume the singular free-energy density satisfies a homogeneity relation under rescaling by factor b>0b>0

fs(t,h)=bdfs ⁣(tbyt,hbyh)f_{s}(t,h) = b^{-d}\, f_{s}\!\left(t\,b^{y_{t}},\, h\,b^{y_{h}}\right)

Here hh is the field conjugate to the order parameter. Choose b=t1/ytb=|t|^{-1/y_{t}} and identify exponents:

  • Specific heat C2fs/T2tαC \sim -\partial^{2}f_{s}/\partial T^{2} \sim |t|^{-\alpha} with
2α=dytdν2 - \alpha = \frac{d}{y_{t}} \equiv d\,\nu
  • Order parameter mfs/hh0tβm \equiv -\partial f_{s}/\partial h|_{h\to 0} \sim |t|^{\beta} with
β=dyhyt\beta = \frac{d - y_{h}}{y_{t}}
  • Susceptibility χm/hh0tγ\chi \equiv \partial m/\partial h|_{h\to 0} \sim |t|^{-\gamma} with
γ=2yhdyt\gamma = \frac{2 y_{h} - d}{y_{t}}

At t=0t=0, pick b=h1/yhb=h^{-1/y_{h}} to get mh1/δm \sim h^{1/\delta} with

δ=yhdyh\delta = \frac{y_{h}}{d - y_{h}}

Fisher’s relation links η\eta to yhy_{h} via the critical correlator:

2η=2yhytν2 - \eta = \frac{2 y_{h}}{y_{t}}\,\nu

Collecting the classic scaling relations

  • Widom: γ=β(δ1)\gamma = \beta(\delta - 1)
  • Rushbrooke: α+2β+γ=2\alpha + 2\beta + \gamma = 2
  • Josephson (hyperscaling): 2α=dν2 - \alpha = d\,\nu
  • Fisher: γ=(2η)ν\gamma = (2 - \eta)\,\nu

Hyperscaling typically holds below the upper critical dimension dcd_{c}.


9.10.3 Kadanoff’s block spin: the cartoon that works

Coarse-grain spins into blocks of size bb, average to get a new field ϕ\phi', and rescale lengths back to the original lattice. Couplings {K}\{K\} flow to new couplings {K}\{K'\}. Iterate:

  • Relevant directions grow under RG (K ⁣ ⁣Kby(KK)K'\!-\!K^{\star}\sim b^{y} (K-K^{\star}) with y>0y>0)
  • Irrelevant directions shrink (y<0y<0)
  • Marginal need higher-order analysis

Criticality is a flow to a nontrivial fixed point KK^{\star} with at least one relevant direction (temperature). The correlation length exponent is the inverse of the thermal eigenvalue:

ν=1yt\nu = \frac{1}{y_{t}}

and the field eigenvalue controls yhy_{h} and thus η,β,γ,δ\eta,\beta,\gamma,\delta via §9.10.2.


9.10.4 Wilson’s momentum-shell RG and beta functions

For a scalar order parameter, the Landau–Ginzburg–Wilson (LGW) functional

F[ϕ]=ddx[12(ϕ)2+r2ϕ2+u4!ϕ4hϕ]\mathcal F[\phi] = \int d^{d}x\left[\frac{1}{2}(\nabla\phi)^{2} + \frac{r}{2}\phi^{2} + \frac{u}{4!}\phi^{4} - h\,\phi\right]

Coarse-grain by integrating out modes in a momentum shell Λ/b<k<Λ\Lambda/b < |\boldsymbol k| < \Lambda, then rescale xx/b\boldsymbol x\to \boldsymbol x/b and ϕb(d2+η)/2ϕ\phi\to b^{(d-2+\eta)/2}\phi. Couplings flow with “RG time” lnb\ell\equiv \ln b:

drd=2rc1u+,dud=β(u)=ϵu+c2u2+\frac{dr}{d\ell} = 2 r - c_{1} u + \cdots,\qquad \frac{du}{d\ell} = \beta(u) = -\epsilon\,u + c_{2}\,u^{2} + \cdots

with ϵ4d\epsilon \equiv 4 - d. Constants c1,c2c_{1},c_{2} depend on conventions; what matters:

  • For d<4d<4 (ϵ>0\epsilon>0) there is a Wilson–Fisher fixed point uϵ/c2u^{\star}\sim \epsilon/c_{2} giving non-mean-field exponents
  • For d>4d>4 (ϵ<0\epsilon<0) the Gaussian fixed point u=0u^{\star}=0 is stable and mean-field exponents hold (with possible logs at d=4d=4)

At one loop for the O(N)O(N) model, the ϵ\epsilon-expansion yields

ν=12+N+24(N+8)ϵ+O(ϵ2)\nu = \frac{1}{2} + \frac{N+2}{4(N+8)}\,\epsilon + \mathcal O(\epsilon^{2}) η=(N+2)2(N+8)2ϵ2+O(ϵ3)\eta = \frac{(N+2)}{2(N+8)^{2}}\,\epsilon^{2} + \mathcal O(\epsilon^{3})

For Ising universality (N=1N=1):

ν=12+112ϵ+,η=154ϵ2+\nu = \frac{1}{2} + \frac{1}{12}\,\epsilon + \cdots,\qquad \eta = \frac{1}{54}\,\epsilon^{2} + \cdots

Plugging ϵ=1\epsilon=1 gives rough 3D estimates; higher loops plus resummation refine them.


9.10.5 Upper critical dimension and dangerous irrelevant variables

The ϕ4\phi^{4} coupling uu is marginal at d=4d=4 and irrelevant for d>4d>4. Yet for d>4d>4 it can be dangerously irrelevant: it vanishes under RG but still controls the amplitude of mm below TcT_{c}, modifying naive hyperscaling. Consequences:

  • Mean-field exponents α=0,β=1/2,γ=1,δ=3,ν=1/2,η=0\alpha=0,\beta=1/2,\gamma=1,\delta=3,\nu=1/2,\eta=0 hold for ddc=4d\ge d_{c}=4
  • Hyperscaling 2α=dν2-\alpha=d\nu fails for d>4d>4 (free-energy density does not scale as ξd\xi^{-d} without uu-dependent rescaling)

Logs at d=4d=4 show up, e.g., ClntpC \sim |\ln t|^{p}.


9.10.6 Finite-size scaling and data collapse

With periodic box size LL, RG/homogeneity imply the finite-size scaling (FSS) ansatz for the singular part of the free energy

fs(t,h;L)=LdF ⁣(tL1/ν, hLyh)f_{s}(t,h;L) = L^{-d}\,\mathcal F\!\left(t\,L^{1/\nu},\ h\,L^{y_{h}}\right)

Observable templates:

  • Order parameter: m(t,h=0;L)=Lβ/νM(tL1/ν)m(t,h=0;L) = L^{-\beta/\nu}\,\mathcal M(t\,L^{1/\nu})
  • Susceptibility: χ(t;L)=Lγ/νX(tL1/ν)\chi(t;L) = L^{\gamma/\nu}\,\mathcal X(t\,L^{1/\nu})
  • Specific heat: C(t;L)=Lα/νC(tL1/ν)C(t;L) = L^{\alpha/\nu}\,\mathcal C(t\,L^{1/\nu}) up to additive analytic background

Practical wins: Curves for different LL collapse when plotted vs tL1/νt L^{1/\nu} with vertical rescaling by Lβ/νL^{-\beta/\nu} or Lγ/νL^{-\gamma/\nu}, and pseudo-critical temperatures shift as

Tc(L)Tc()L1/νT_{c}(L) - T_{c}(\infty) \propto L^{-1/\nu}

9.10.7 Universality classes (quick map)

  • Ising (O(1)O(1), Z2\mathbb Z_{2} symmetry): uniaxial magnets, liquid–gas along the critical isotherm, binary alloys
  • XY (O(2)O(2)): superfluid helium, planar spins, thin-film superconductors; in 2D has a Kosterlitz–Thouless transition (essential, not power-law)
  • Heisenberg (O(3)O(3)): isotropic magnets
  • Percolation: geometric connectivity transition with its own exponents
  • Directed percolation: absorbing-state phase transitions (nonequilibrium)

Same dimension + same symmetries + similar conservation rules → same exponents.


9.10.8 KT interlude (2D XY)

In 2D, vortices unbind at TKTT_{\mathrm{KT}} with correlation length

ξexp ⁣(bTTKT)\xi \sim \exp\!\left(\frac{b}{\sqrt{|T-T_{\mathrm{KT}}|}}\right)

No standard power-law exponents; instead there is a universal jump of the superfluid stiffness. This is still RG—just on a different flow diagram with vortex fugacity and stiffness as couplings.


9.10.9 Worked mini-examples

(a) Scaling relations in two lines
Start from fs(t,h)=bdfs(tbyt,hbyh)f_{s}(t,h)=b^{-d} f_{s}(t b^{y_{t}}, h b^{y_{h}}). Derive α,β,γ,δ\alpha,\beta,\gamma,\delta and verify Widom and Rushbrooke.

(b) One-loop ϕ4\phi^{4} shell RG
Integrate modes in Λ/b<k<Λ\Lambda/b<k<\Lambda for LGW, rescale, and show schematically du/d=ϵu+c2u2du/d\ell=-\epsilon u + c_{2}u^{2} and dr/d=2rc1udr/d\ell=2r - c_{1}u. Identify uu^{\star} and extract ν=1/yt\nu=1/y_{t} to O(ϵ)\mathcal O(\epsilon).

(c) Finite-size shift
Assuming χ(t,L)=Lγ/νX(tL1/ν)\chi(t,L)=L^{\gamma/\nu}\,\mathcal X(t L^{1/\nu}), show the susceptibility peak location t(L)L1/νt^{\ast}(L)\propto L^{-1/\nu}.

(d) Data collapse recipe
Given Monte Carlo data for m(T;L)m(T;L) in 2D Ising at H=0H=0, rescale mLβ/νmm\to L^{\beta/\nu} m and T(TTc)L1/νT\to (T-T_{c}) L^{1/\nu} with β=1/8\beta=1/8, ν=1\nu=1. Explain the adjustable parameters and quality checks.

(e) Dangerous irrelevant uu above dcd_{c}
In d>4d>4, show that m(t)1/2u1/2m\propto (-t)^{1/2} u^{-1/2} at mean-field level, illustrating how an “irrelevant” coupling sets amplitudes and breaks naive hyperscaling.


9.10.10 Common pitfalls

  • Using hyperscaling where it fails. Above dc=4d_{c}=4 for Ising-like systems, 2α=dν2-\alpha=d\nu is not valid without caveats
  • Confusing crossover with criticality. A flow near but not at a fixed point shows effective exponents that drift with scale
  • Ignoring analytic backgrounds. Specific heat often has a non-singular part that can mask small α\alpha
  • Forgetting boundary conditions in FSS. Exponents are universal; scaling functions and subleading corrections are not
  • Mixing KT with power-law scaling. In 2D XY, expect essential singularities and a stiffness jump, not β,γ\beta,\gamma etc.

9.10.11 Minimal problem kit

  • From the homogeneity hypothesis, derive the four standard scaling relations and identify the independent exponent set
  • Compute ν\nu and η\eta to O(ϵ)\mathcal O(\epsilon) for the O(N)O(N) model and specialize to N=1,2,3N=1,2,3
  • Show that at d=4d=4 the Gaussian fixed point is marginal and derive a logarithmic correction to CC in ϕ4\phi^{4} theory
  • Use FSS to estimate ν\nu from synthetic or simulated Ising susceptibility peaks at sizes L={16,32,64,128}L=\{16,32,64,128\}
  • For the KT RG flow, write the lowest-order equations for stiffness KK and vortex fugacity yy, and show the separatrix corresponding to TKTT_{\mathrm{KT}}

9.11 Linear Response & Fluctuation–Dissipation

Kick a system gently, watch it wiggle. Linear response theory turns small perturbations into universal formulas linking dissipation to equilibrium fluctuations. The retarded correlator sets the response, causality gives Kramers–Kronig, and the fluctuation–dissipation theorem (FDT) glues noise spectra to the imaginary part of susceptibilities. Transport coefficients then drop out from time integrals of current autocorrelations (Green–Kubo).


9.11.1 Setup: source, observable, and response kernel

Perturb an equilibrium Hamiltonian by a weak time-dependent field h(t)h(t) coupled to an operator AA:

H(t)=h(t)AH'(t) = -\,h(t)\,A

Measure another observable BB. To linear order,

δB(t)=dt  χBA(tt)h(t)\delta\langle B(t)\rangle = \int_{-\infty}^{\infty} dt'\; \chi_{BA}(t-t')\,h(t')

with causal susceptibility χBA(t<0)=0\chi_{BA}(t<0)=0. In frequency space,

δB(ω)=χBA(ω)h(ω)\delta\langle B(\omega)\rangle = \chi_{BA}(\omega)\,h(\omega)

The power absorbed by a sinusoidal drive h(t)=h0cosωth(t)=h_{0}\cos\omega t is set by the dissipative part χBA(ω)\chi_{BA}''(\omega).


9.11.2 Kubo formula: quantum and classical limits

In the Heisenberg picture with equilibrium average \langle\cdots\rangle, the retarded susceptibility is

χBA(t)=iΘ(t)[B(t),A(0)]\chi_{BA}(t) = \frac{i}{\hbar}\,\Theta(t)\,\langle\,[B(t),A(0)]\,\rangle

Equivalently,

χBA(ω)=0dt  eiωti[B(t),A(0)]\chi_{BA}(\omega) = \int_{0}^{\infty} dt\; e^{i\omega t}\,\frac{i}{\hbar}\,\langle\,[B(t),A(0)]\,\rangle

Classical limit. Replace commutators by ii\hbar times Poisson brackets and obtain

χBA(t)=Θ(t)βddtCBA(t),CBA(t)δB(t)δA(0)\chi_{BA}(t) = \Theta(t)\,\beta\,\frac{d}{dt}\,C_{BA}(t),\qquad C_{BA}(t)\equiv \langle \delta B(t)\,\delta A(0)\rangle

This is the time-domain FDT form for conjugate variables.


9.11.3 Causality ⇒ Kramers–Kronig

Causality makes χBA(ω)\chi_{BA}(\omega) analytic in the upper half-plane. Real and imaginary parts are Hilbert transforms:

Reχ(ω)=1πP ⁣dωχ(ω)ωω\mathrm{Re}\,\chi(\omega) = \frac{1}{\pi}\,\mathcal P\!\int_{-\infty}^{\infty} d\omega'\,\frac{\chi''(\omega')}{\omega'-\omega} χ(ω)=1πP ⁣dωReχ(ω)ωω\chi''(\omega) = -\frac{1}{\pi}\,\mathcal P\!\int_{-\infty}^{\infty} d\omega'\,\frac{\mathrm{Re}\,\chi(\omega')}{\omega'-\omega}

Dissipation \Leftrightarrow dispersion: you cannot have one without the other.


9.11.4 Fluctuation–dissipation theorem in frequency space

Define the symmetrized equilibrium spectral density of AA,

SAA(ω)dt  eiωt12{δA(t),δA(0)}S_{AA}(\omega) \equiv \int_{-\infty}^{\infty} dt\; e^{i\omega t}\,\frac{1}{2}\,\langle \{ \delta A(t),\delta A(0) \} \rangle

Then the quantum FDT states

SAA(ω)=coth ⁣(βω2)χAA(ω)S_{AA}(\omega) = \hbar\,\coth\!\left(\frac{\beta\hbar\omega}{2}\right)\,\chi_{AA}''(\omega)

Classical limit βω1\beta\hbar\omega\ll 1 gives

SAA(ω)=2kBTωχAA(ω)S_{AA}(\omega) = \frac{2 k_{B} T}{\omega}\,\chi_{AA}''(\omega)

Interpretation: the same modes that dissipate energy when driven also fluctuate at equilibrium, and the ratio is fixed by TT (or by ω\hbar\omega at high frequency).


9.11.5 Green–Kubo relations: transport from autocorrelations

Transport coefficients are time integrals of current autocorrelations.

  • Self-diffusion in dd dimensions with velocity v\boldsymbol v:
D=1d0dt  v(t)v(0)D = \frac{1}{d}\int_{0}^{\infty} dt\;\langle \boldsymbol v(t)\cdot \boldsymbol v(0)\rangle
  • Electrical conductivity for total current JxJ_{x} in volume VV:
σ=βV0dt  Jx(t)Jx(0)\sigma = \frac{\beta}{V}\int_{0}^{\infty} dt\;\langle J_{x}(t)\,J_{x}(0)\rangle
  • Shear viscosity from the stress tensor component σxy\sigma_{xy}:
η=βV0dt  σxy(t)σxy(0)\eta = \frac{\beta}{V}\int_{0}^{\infty} dt\;\langle \sigma_{xy}(t)\,\sigma_{xy}(0)\rangle
  • Thermal conductivity with heat current JQ\boldsymbol J_{Q}:
κ=β2V0dt  JQ(t)JQ(0)\kappa = \frac{\beta^{2}}{V}\int_{0}^{\infty} dt\;\langle \boldsymbol J_{Q}(t)\cdot \boldsymbol J_{Q}(0)\rangle

All are equilibrium averages with the unperturbed dynamics.


9.11.6 Langevin, Fokker–Planck, and Einstein relation

For a Brownian particle of position xx in a viscous fluid,

mv˙=γv+ξ(t)m\dot v = -\gamma v + \xi(t)

with Gaussian white noise ξ(t)ξ(0)=2γkBTδ(t)\langle \xi(t)\xi(0)\rangle = 2\gamma k_{B}T\,\delta(t). Solving gives the Einstein relation

D=kBTγD = \frac{k_{B} T}{\gamma}

and velocity autocorrelation v(t)v(0)=(kBT/m)et/τ\langle v(t)v(0)\rangle = (k_{B}T/m)\,e^{-t/\tau} with τ=m/γ\tau=m/\gamma. The associated Fokker–Planck equation evolves the probability density and relaxes to the Maxwell–Boltzmann steady state.


9.11.7 Noise examples: Johnson–Nyquist and friends

For a resistor RR in classical regime ωkBT\hbar\omega\ll k_{B}T, the voltage noise spectral density is

SV(ω)=4kBTRS_{V}(\omega) = 4 k_{B} T\,R

The quantum correction multiplies by ω2kBTcoth(βω/2)\tfrac{\hbar\omega}{2k_{B}T}\coth(\beta\hbar\omega/2). Current noise follows as SI(ω)=4kBT/RS_{I}(\omega)=4 k_{B} T/R for a resistor alone. Same logic underlies magnetic noise in NMR, force noise in AFM, and shot-noise limits in mesoscopic conductors.


9.11.8 Onsager reciprocity and microscopic reversibility

When generalized fluxes JiJ_{i} respond to thermodynamic forces XjX_{j} linearly,

Ji=jLijXjJ_{i} = \sum_{j} L_{ij}\,X_{j}

time-reversal symmetry implies Onsager–Casimir relations

Lij(B)=ϵiϵjLji(B)L_{ij}(B) = \epsilon_{i}\epsilon_{j}\,L_{ji}(-B)

where ϵi=±1\epsilon_{i}=\pm 1 is the time-reversal parity of JiJ_{i}. At zero magnetic field and for forces/fluxes with the same parity, Lij=LjiL_{ij}=L_{ji}. Reciprocity is a macroscopic echo of equilibrium correlation symmetry.


9.11.9 Causality, passivity, and sum rules

  • Passivity: χ(ω)0\chi''(\omega)\ge 0 for ω>0\omega>0 for stable, passive systems; it encodes positive dissipation.
  • Sum rules: Moments of χ(ω)\chi''(\omega) tie to equal-time commutators or static susceptibilities, e.g., an ff-sum rule for charge response.
  • High-frequency tails: Often fixed by short-time operator algebra; useful for checking numerics or truncated models.

9.11.10 Common pitfalls

  • Using symmetric instead of retarded correlators in χ\chi. The Kubo χ\chi needs the commutator and Θ(t)\Theta(t).
  • Forgetting ensemble choice. Green–Kubo averages are taken in the equilibrium ensemble matching conserved quantities.
  • Abusing classical FDT at high frequency. Use the quantum coth(βω/2)\coth(\beta\hbar\omega/2) factor when ωkBT\hbar\omega\gtrsim k_{B}T.
  • Mixing response channels. Coupling H=hAH'=-hA means the correct FDT relates SAAS_{AA} to χAA\chi_{AA}'', not to some other operator.
  • Ignoring conservation laws. Currents tied to conserved densities have hydrodynamic long-time tails that can alter naive integral convergence.

9.11.11 Worked mini-examples

(a) Kramers–Kronig from causality
Assume χ(t)=0\chi(t)=0 for t<0t<0, Fourier transform, and show the Hilbert-transform relations for Reχ\mathrm{Re}\,\chi and χ\chi''.

(b) Classical FDT in time domain
With H=h(t)AH'=-h(t)A and B=AB=A, prove χAA(t)=Θ(t)βddtCAA(t)\chi_{AA}(t)=\Theta(t)\beta\,\frac{d}{dt}C_{AA}(t) and recover SAA(ω)=2kBTωχAA(ω)S_{AA}(\omega)=\tfrac{2k_{B}T}{\omega}\chi_{AA}''(\omega).

(c) Johnson noise
Take a circuit with a resistor only. Using I(ω)=V(ω)/RI(\omega)=V(\omega)/R and FDT, derive SV=4kBTRS_{V}=4k_{B}TR in the classical band.

(d) Diffusion from velocity correlations
For an exponential velocity correlator v(t)v(0)=v02et/τ\langle v(t)v(0)\rangle=v_{0}^{2}e^{-t/\tau}, integrate to get D=v02τ/dD=v_{0}^{2}\tau/d and compare to D=kBT/γD=k_{B}T/\gamma with v02=kBT/mv_{0}^{2}=k_{B}T/m and τ=m/γ\tau=m/\gamma.

(e) Conductivity Green–Kubo
Show that a uniform electric field couples as H=ExJxtH'=-E_{x}J_{x}t in the appropriate gauge and derive σ=(β/V)0dtJx(t)Jx(0)\sigma=(\beta/V)\int_{0}^{\infty}dt\,\langle J_{x}(t)J_{x}(0)\rangle.

(f) Onsager reciprocity check
For thermoelectric transport with electric and heat currents, write the 2×22\times 2 LL-matrix and argue L12(B)=L21(B)L_{12}(B)=L_{21}(-B).


9.11.12 Minimal problem kit

  • Starting from the Kubo formula, prove the quantum FDT SAA(ω)=coth(βω/2)χAA(ω)S_{AA}(\omega)=\hbar\coth(\beta\hbar\omega/2)\chi_{AA}''(\omega)
  • Derive the Einstein relation D=μkBTD=\mu k_{B}T by relating mobility μ\mu to the low-frequency limit of the velocity response
  • Show that χ(ω)0\chi''(\omega)\ge 0 for a passive system by evaluating average absorbed power for a sinusoidal drive
  • From the continuity equation, connect long-time tails of current correlations to hydrodynamic diffusion and discuss convergence of Green–Kubo integrals
  • Compute η\eta from a microscopic hard-sphere model via the stress autocorrelation and compare to kinetic-theory scaling

9.12 Kinetic Theory & the Boltzmann Equation

Stat mech taught us what equilibrium looks like. Kinetic theory explains how systems move toward it, and what happens while they’re not there. The star is the Boltzmann equation, a mesoscopic evolution law for the one-particle distribution f(r,v,t)f(\boldsymbol r,\boldsymbol v,t). From it, we’ll derive transport coefficients, connect to hydrodynamics, and see where and why the theory breaks.


9.12.1 From micro to meso: the distribution function

Define f(r,v,t)f(\boldsymbol r,\boldsymbol v,t) so that fd3rd3vf\,d^{3}r\,d^{3}v is the expected number of particles in the phase-space cell. Normalize to N=d3rd3vfN=\int d^{3}r\,d^{3}v\,f. Conserved densities are moments of ff:

n(r,t)=d3vf,u(r,t)=1nd3vvf,ε(r,t)=d3v12mv2fn(\boldsymbol r,t) = \int d^{3}v\, f,\quad \boldsymbol u(\boldsymbol r,t) = \frac{1}{n}\int d^{3}v\, \boldsymbol v\, f,\quad \varepsilon(\boldsymbol r,t) = \int d^{3}v\, \tfrac{1}{2} m v^{2}\, f

Pressure tensor and heat flux are higher moments of the peculiar velocity cvu\boldsymbol c\equiv\boldsymbol v-\boldsymbol u.


9.12.2 The Boltzmann equation and molecular chaos

In external force F\boldsymbol F,

ft+vrf+Fmvf=C[f]\frac{\partial f}{\partial t} + \boldsymbol v\cdot\nabla_{\boldsymbol r} f + \frac{\boldsymbol F}{m}\cdot\nabla_{\boldsymbol v} f = C[f]

The collision integral C[f]C[f] encodes binary collisions. For short-range interactions and molecular chaos (Stosszahlansatz: pre-collision velocities uncorrelated),

C[f](r,v)=d3v2dΩ  gσ(g,Ω)[ff2ff2]C[f](\boldsymbol r,\boldsymbol v) = \int d^{3}v_{2}\int d\Omega\; g\,\sigma(g,\Omega)\,\big[f'f_{2}' - f f_{2}\big]

with g=vv2g=|\boldsymbol v-\boldsymbol v_{2}|, differential cross section σ\sigma, and f=f(r,v,t)f'=f(\boldsymbol r,\boldsymbol v',t) the post-collision values determined by energy–momentum conservation.

Collision invariants. For any ψ(v)\psi(\boldsymbol v) equal to 11, v\boldsymbol v, or v2v^{2},

d3v  ψ(v)C[f]=0\int d^{3}v\; \psi(\boldsymbol v)\,C[f] = 0

leading to local conservation of mass, momentum, and energy.


9.12.3 Local equilibrium = Maxwell–Boltzmann

C[f]=0C[f]=0 (for all cross sections) iff ff is a local Maxwellian,

fLE(r,v,t)=n(r,t)(m2πkBT(r,t))3/2exp ⁣[mvu(r,t)22kBT(r,t)]f_{\mathrm{LE}}(\boldsymbol r,\boldsymbol v,t) = n(\boldsymbol r,t) \left(\frac{m}{2\pi k_{B} T(\boldsymbol r,t)}\right)^{3/2} \exp\!\left[-\frac{m|\boldsymbol v-\boldsymbol u(\boldsymbol r,t)|^{2}}{2k_{B}T(\boldsymbol r,t)}\right]

This fLEf_{\mathrm{LE}} maximizes local entropy density given n,u,Tn,\boldsymbol u,T and is the fixed point of collisions.


9.12.4 The H-theorem: entropy grows

Define Boltzmann’s Hd3rd3vflnfH\equiv \int d^{3}r\,d^{3}v\, f\ln f. Using the collision integral,

dHdt=d3rd3vC[f](1+lnf)0\frac{dH}{dt} = \int d^{3}r\,d^{3}v\, C[f]\,(1+\ln f) \le 0

with equality only for f=fLEf=f_{\mathrm{LE}}. Identifying S=kBH+constS=-k_{B} H + \text{const} gives monotone entropy increase toward local equilibrium. Caveats: relies on molecular chaos and binary, dilute collisions; long-range forces and correlated pre-collisions can spoil the proof.


9.12.5 Hydrodynamics as moment equations

Take moments of Boltzmann and use collision invariants to get the Euler (ideal) or Navier–Stokes–Fourier (dissipative) equations. Let ρ=mn\rho=mn, p=nkBTp=nk_{B}T.

  • Continuity
tρ+(ρu)=0\partial_{t}\rho + \nabla\cdot(\rho\,\boldsymbol u) = 0
  • Momentum
t(ρu)+(ρuu+pI+Π)=nF\partial_{t}(\rho \boldsymbol u) + \nabla\cdot(\rho \boldsymbol u\boldsymbol u + p\,\mathbb I + \boldsymbol{\Pi}) = n\boldsymbol F
  • Energy
tε+[(ε+p)u+Πu+q]=nFu\partial_{t}\varepsilon + \nabla\cdot\big[(\varepsilon+p)\boldsymbol u + \boldsymbol{\Pi}\cdot\boldsymbol u + \boldsymbol q\big] = n\boldsymbol F\cdot\boldsymbol u

Π\boldsymbol{\Pi} is the viscous stress, q\boldsymbol q the heat flux. To close these, expand around fLEf_{\mathrm{LE}}.


9.12.6 Chapman–Enskog and transport coefficients

Assume weak gradients (small Knudsen number K ⁣n/LK\!n\equiv \ell/L) and write f=fLE+f(1)+f=f_{\mathrm{LE}} + f^{(1)} + \cdots where f(1)f^{(1)} is linear in gradients. Solving Boltzmann at first order gives Newtonian viscosity and Fourier heat law:

Π=η[u+(u)T23(u)I]ζ(u)I\boldsymbol{\Pi} = -\eta\left[\nabla\boldsymbol u + (\nabla\boldsymbol u)^{T} - \tfrac{2}{3}(\nabla\cdot\boldsymbol u)\mathbb I\right] - \zeta\,(\nabla\cdot\boldsymbol u)\,\mathbb I q=κT\boldsymbol q = -\kappa\,\nabla T

For hard spheres (mass mm, diameter dd) at number density nn and temperature TT,

η516d2mkBTπ,κ75kB64d2kBTπm\eta \sim \frac{5}{16 d^{2}} \sqrt{\frac{m k_{B} T}{\pi}},\qquad \kappa \sim \frac{75 k_{B}}{64 d^{2}} \sqrt{\frac{k_{B} T}{\pi m}}

up to modest Sonine-polynomial correction factors; the bulk viscosity ζ\zeta vanishes for monatomic ideal gases.


9.12.7 Relaxation-time and BGK models

Exact collision integrals are painful. The Bhatnagar–Gross–Krook (BGK) model replaces C[f]C[f] by a single relaxation rate toward local equilibrium:

CBGK[f]=ffLEτC_{\mathrm{BGK}}[f] = -\frac{f - f_{\mathrm{LE}}}{\tau}

Choosing τ\tau to match η\eta and κ\kappa reproduces first-order hydrodynamics and is widely used in lattice Boltzmann and semiconductor transport. The simpler relaxation time approximation in solids does the same for electron scattering.


9.12.8 Knudsen number, boundary layers, and breakdowns

  • Hydro regime K ⁣n1K\!n\ll 1: Navier–Stokes–Fourier valid.
  • Transition K ⁣n0.1K\!n \sim 0.1: slip flows, Knudsen layers near walls; need kinetic boundary conditions.
  • Ballistic K ⁣n1K\!n\gtrsim 1: free-molecular flow; Boltzmann without gradient expansions or even collisionless Vlasov dynamics (for long-range forces).
  • Dense gases: Enskog corrections modify C[f]C[f] and transport.
  • Long-range Coulomb: many small-angle deflections → Landau (Fokker–Planck) collision operator.

9.12.9 Quantum Boltzmann and Pauli/Bose factors

For dilute quantum gases at high enough TT that coherence lengths are short, the collision term acquires stimulated/blocked final states:

CQ[f] ⁣dΓ  [ff2(1±f)(1±f2)ff2(1±f)(1±f2)]C_{\mathrm{Q}}[f] \propto \int\! d\Gamma\; \big[f' f_{2}' (1 \pm f)(1 \pm f_{2}) - f f_{2} (1 \pm f')(1 \pm f_{2}')\big]

upper sign for bosons, lower for fermions. This is the Uehling–Uhlenbeck modification. At very low TT in superfluids, kinetic theory must include collective modes (phonons, Bogoliubov quasiparticles).


9.12.10 Boltzmann transport in solids (electrons and phonons)

In crystals, f(k,r,t)f(\boldsymbol k,\boldsymbol r,t) evolves in k\boldsymbol k-space with band velocity vk=kϵk/\boldsymbol v_{\boldsymbol k}=\nabla_{\boldsymbol k}\epsilon_{\boldsymbol k}/\hbar and forces from electric fields and Berry curvature (advanced topic). With a relaxation time τ(k)\tau(\boldsymbol k),

σxx=e2 ⁣ ⁣d3k(2π)3  τ(k)vx2(f0ϵ)\sigma_{xx} = e^{2}\!\int\!\frac{d^{3}k}{(2\pi)^{3}}\; \tau(\boldsymbol k)\, v_{x}^{2}\, \left(-\frac{\partial f_{0}}{\partial \epsilon}\right)

reproduces Drude in free-electron bands, while phonon Boltzmann explains lattice thermal conductivity via three-phonon scattering and boundary limits.


9.12.11 Linear response vs kinetic theory: same game, different screens

Green–Kubo (9.11) says transport = time integrals of autocorrelators. Chapman–Enskog says transport = solutions of Boltzmann to first order in gradients. They agree when both apply. Kinetic theory wins when mean free paths and cross sections are explicit; Green–Kubo wins for interacting liquids where quasiparticles are murky but simulations can measure correlations.


9.12.12 Worked mini-examples

(a) Mean free path and viscosity
For hard spheres at number density nn with cross section σ=πd2\sigma=\pi d^{2}, show 1/(2nσ)\ell\simeq 1/(\sqrt{2}\,n\sigma) and estimate η13nmvˉ\eta\sim \tfrac{1}{3} n m \bar v\,\ell; compare to the Chapman–Enskog value.

(b) BGK to Navier–Stokes
Insert CBGKC_{\mathrm{BGK}} into Boltzmann, solve for f(1)f^{(1)}, and derive η=pτ\eta = p \tau and κ=52kBnτ/m\kappa = \tfrac{5}{2} k_{B} n \tau/m for a monatomic gas.

(c) Sound attenuation
Linearize Boltzmann around fLEf_{\mathrm{LE}} for a plane wave ei(krωt)e^{i(\mathbf k\cdot\mathbf r - \omega t)} and compute the attenuation coefficient in terms of η\eta and κ\kappa.

(d) Lorentz gas
For light particles scattering elastically off fixed hard centers of density nin_{i}, compute DD and show how anisotropic cross sections reshape transport.

(e) Quantum modification
Derive detailed balance with Pauli blocking for a two-state fermion collision and show how it suppresses σ\sigma-limited conductivity at low TT.


9.12.13 Common pitfalls

  • Forgetting conservation constraints in C[f]C[f]. Any model collision operator must conserve mass, momentum, and energy (BGK does by construction if fLEf_{\mathrm{LE}} matches local moments).
  • Using Navier–Stokes at large K ⁣nK\!n. Expect slip, Knudsen layers, or ballistic behavior; go kinetic.
  • Assuming ζ=0\zeta=0 generically. True only for monatomic ideal gases; internal modes create bulk viscosity.
  • Ignoring boundary conditions. Specular vs diffuse reflection at walls changes near-wall transport by order one.
  • Treating Coulomb like hard spheres. Many-body small-angle dominance calls for Landau/Fokker–Planck, not Boltzmann with rare hard kicks.

9.12.14 Minimal problem kit

  • Starting from the Boltzmann equation, derive the continuity, momentum, and energy equations and identify Π\boldsymbol{\Pi} and q\boldsymbol q
  • Carry out the first Sonine-polynomial step for hard spheres to compute η\eta to leading order and compare to the heuristic 13nmvˉ\tfrac{1}{3} n m \bar v \ell
  • Implement BGK and show η=pτ\eta=p\tau; fit τ\tau to a measured viscosity to predict thermal conductivity and Prandtl number
  • For a dilute electron gas with elastic impurity scattering, compute σ\sigma in the relaxation-time approximation and discuss the Matthiessen rule qualitatively
  • Derive the Uehling–Uhlenbeck collision term for indistinguishable particles and show how it reduces to classical Boltzmann when f1f\ll 1

In summary: Boltzmann’s equation is the bridge from micro collisions to macro flows. With molecular chaos and short-range scatter, it drives ff to a local Maxwellian, whose slow gradients birth hydrodynamics with η,κ,ζ\eta,\kappa,\zeta computable from cross sections. Push the mean free path too large, the density too high, or the interactions too long-ranged, and the equation changes costume—but the core idea survives: dynamics of distributions, not just particles, rule nonequilibrium physics