10 Condensed matter physics

10.1 Crystals & Reciprocal Space

Crystalline solids are periodic in space. Periodicity gives you enormous leverage: lattices turn sums into selection rules, Fourier transforms quantize kk, and diffraction fingerprints structure. This section lays down the lattice–basis language, reciprocal vectors, Brillouin zones, Miller indices, and the diffraction toolkit (Bragg, Laue, Ewald, structure factors). We’ll tie together fcc/bcc duality, selection rules for common crystals, and quick ways to not get lost in indices.


10.1.1 Lattice, basis, crystal structure

A Bravais lattice is the set of points

R=n1a1+n2a2+n3a3,niZ\boldsymbol R = n_{1}\,\boldsymbol a_{1} + n_{2}\,\boldsymbol a_{2} + n_{3}\,\boldsymbol a_{3},\qquad n_{i}\in\mathbb Z

with three linearly independent primitive vectors ai\boldsymbol a_{i}. A crystal structure = lattice plus a basis of atoms at positions {τj}\{\boldsymbol\tau_{j}\} inside a primitive cell; the atomic positions are r=R+τj\boldsymbol r = \boldsymbol R + \boldsymbol\tau_{j}.

  • Wigner–Seitz cell: the set of points closer to a given lattice point than any other.
  • Conventional vs primitive: conventional cubic cells are nicer to look at; primitive cells are minimal-volume and great for theory.

Pro tip: never say “the diamond lattice.” Diamond has an fcc lattice with a two-point basis {(0,0,0),(12,12,12)a}\{(0,0,0),\,(\tfrac12,\tfrac12,\tfrac12)a\}.


10.1.2 Symmetry: point groups and space groups (one-screenshot tour)

  • Point group: rotations, reflections, inversion that leave at least one point fixed.
  • Space group: point-group ops + translations + screw rotations and glide reflections.
  • In 3D there are 14 Bravais lattices, 32 crystallographic point groups, and 230 space groups. For band theory you mostly need the lattice and the little groups of high-symmetry kk points.

10.1.3 Reciprocal lattice and Brillouin zone

Define the reciprocal primitive vectors bi\boldsymbol b_{i} by aibj=2πδij\boldsymbol a_{i}\cdot \boldsymbol b_{j}=2\pi\delta_{ij}. In vector form

b1=2πa2×a3a1(a2×a3),b2=2πa3×a1a1(a2×a3),b3=2πa1×a2a1(a2×a3)\boldsymbol b_{1} = 2\pi \frac{\boldsymbol a_{2}\times \boldsymbol a_{3}}{\boldsymbol a_{1}\cdot (\boldsymbol a_{2}\times \boldsymbol a_{3})},\quad \boldsymbol b_{2} = 2\pi \frac{\boldsymbol a_{3}\times \boldsymbol a_{1}}{\boldsymbol a_{1}\cdot (\boldsymbol a_{2}\times \boldsymbol a_{3})},\quad \boldsymbol b_{3} = 2\pi \frac{\boldsymbol a_{1}\times \boldsymbol a_{2}}{\boldsymbol a_{1}\cdot (\boldsymbol a_{2}\times \boldsymbol a_{3})}

The reciprocal lattice consists of vectors G=hb1+kb2+b3\boldsymbol G = h\,\boldsymbol b_{1} + k\,\boldsymbol b_{2} + \ell\,\boldsymbol b_{3} with integers (h,k,)(h,k,\ell).

The first Brillouin zone (BZ) is the Wigner–Seitz cell of the reciprocal lattice. It is the primitive cell in kk-space and the natural domain for Bloch wavevectors.

  • Duality: fcc \leftrightarrow bcc. The reciprocal of fcc is bcc and vice versa. Simple cubic is self-dual.

10.1.4 Bloch theorem in one line

In a periodic potential V(r+R)=V(r)V(\boldsymbol r+\boldsymbol R)=V(\boldsymbol r), single-particle eigenstates are Bloch waves:

ψnk(r)=eikrunk(r),unk(r+R)=unk(r)\psi_{n\boldsymbol k}(\boldsymbol r) = e^{i\boldsymbol k\cdot \boldsymbol r}\,u_{n\boldsymbol k}(\boldsymbol r),\qquad u_{n\boldsymbol k}(\boldsymbol r+\boldsymbol R) = u_{n\boldsymbol k}(\boldsymbol r)

Wavevectors k\boldsymbol k can be chosen in the first BZ. Translating kk+G\boldsymbol k\to \boldsymbol k+\boldsymbol G leaves physics invariant.


10.1.5 Planes, directions, and indices (Miller crash course)

  • Directions use square brackets [uvw][u\,v\,w] for a lattice vector proportional to ua1+va2+wa3u\,\boldsymbol a_{1}+v\,\boldsymbol a_{2}+w\,\boldsymbol a_{3}. Families of equivalent directions: uvw\langle u\,v\,w\rangle.
  • Planes use parentheses (hk)(h\,k\,\ell), the Miller indices, defined as reciprocals of intercepts along ai\boldsymbol a_{i}. Families: {hk}\{h\,k\,\ell\}.
  • Spacing to plane family (hk)(h\,k\,\ell) is tied to the reciprocal vector Ghkl\boldsymbol G_{hkl}:
Ghkl=hb1+kb2+b3\boldsymbol G_{hkl} = h\,\boldsymbol b_{1} + k\,\boldsymbol b_{2} + \ell\,\boldsymbol b_{3} dhkl=2πGhkld_{hkl} = \frac{2\pi}{|\boldsymbol G_{hkl}|}

For cubic a1=ax^a_{1}=a\,\hat x etc., dhkl=a/h2+k2+2d_{hkl}= a/\sqrt{h^{2}+k^{2}+\ell^{2}}.

Hexagonal often uses Miller–Bravais (hki)(h\,k\,i\,\ell) with i=(h+k)i=-(h+k) to make basal-plane symmetry manifest.


10.1.6 Diffraction: Bragg, Laue, Ewald are the same vibe

Elastic scattering conserves k=k|\boldsymbol k|=|\boldsymbol k'| and transfers crystal momentum Δk=kk\Delta\boldsymbol k=\boldsymbol k'-\boldsymbol k equal to a reciprocal vector:

Δk=G\Delta \boldsymbol k = \boldsymbol G

Bragg condition in real space:

2dhklsinθ=nλ2 d_{hkl}\,\sin\theta = n\,\lambda

Laue condition in kk-space is exactly kk=G\boldsymbol k'-\boldsymbol k=\boldsymbol G. Ewald sphere construction: draw a sphere of radius k=2π/λ|\boldsymbol k|=2\pi/\lambda anchored at the tail of k\boldsymbol k; reciprocal-lattice points on the sphere give allowed reflections.

Structure factor encodes the basis:

F(G)=j=1mfjeiGτjF(\boldsymbol G) = \sum_{j=1}^{m} f_{j}\,e^{-i\boldsymbol G\cdot \boldsymbol\tau_{j}}

with atomic form factors fjf_{j}. The diffracted intensity is

I(G)F(G)2I(\boldsymbol G) \propto |F(\boldsymbol G)|^{2}

The lattice sum enforces G\boldsymbol G belongs to the reciprocal lattice; FF decides if it actually lights up.


10.1.7 Selection rules for common structures

  • Simple cubic (sc) with one atom at (0,0,0)(0,0,0):
F(G)=f,no systematic absencesF(\mathbf G)=f,\quad \text{no systematic absences}
  • Body-centered cubic (bcc), basis {(0,0,0),(12,12,12)a}\{(0,0,0),\,(\tfrac12,\tfrac12,\tfrac12)a\}:
F=f[1+eiπ(h+k+)]={2fh+k+ even0h+k+ oddF = f\left[1 + e^{-i\pi(h+k+\ell)}\right] = \begin{cases} 2f & h+k+\ell\ \text{even} \\ 0 & h+k+\ell\ \text{odd} \end{cases}
  • Face-centered cubic (fcc), basis {(0,0,0),(12,12,0)a,(12,0,12)a,(0,12,12)a}\{(0,0,0),\,(\tfrac12,\tfrac12,0)a,\,(\tfrac12,0,\tfrac12)a,\,(0,\tfrac12,\tfrac12)a\}:
F=f[1+eiπ(h+k)+eiπ(h+)+eiπ(k+)]={4fh,k, all even or all odd0otherwiseF = f\left[1 + e^{-i\pi(h+k)} + e^{-i\pi(h+\ell)} + e^{-i\pi(k+\ell)}\right] = \begin{cases} 4f & h,k,\ell\ \text{all even or all odd} \\ 0 & \text{otherwise} \end{cases}
  • NaCl (rock salt): fcc lattice with two-species basis {(0,0,0) Na, (12,12,12)a Cl}\{(0,0,0)\ \text{Na},\ (\tfrac12,\tfrac12,\tfrac12)a\ \text{Cl}\}. Then
F=fNa+fCleiπ(h+k+)F = f_{\mathrm{Na}} + f_{\mathrm{Cl}}\,e^{-i\pi(h+k+\ell)}

Intensity alternates with h+k+h+k+\ell depending on ff-contrast.

  • Diamond: fcc lattice with basis {(0,0,0),(14,14,14)a}\{(0,0,0),\,(\tfrac14,\tfrac14,\tfrac14)a\}. Extra phase leads to additional systematic zeros beyond fcc.

10.1.8 fcc–bcc duality and high-symmetry kk points

If {ai}\{\boldsymbol a_{i}\} are fcc vectors, the reciprocal {bi}\{\boldsymbol b_{i}\} generate a bcc lattice in kk-space (and vice versa). Learn the standard labels:

  • fcc BZ: Γ\Gamma center, XX at face centers, LL at BZ corners, WW mid-edges.
  • bcc BZ: Γ\Gamma, HH, NN, PP points.

These are where band extrema and degeneracies love to appear and where experimentalists measure ARPES and quantum oscillations.


10.1.9 Real ↔ reciprocal geometry cheats

  • Volume relation:
VcellVcell=(2π)3V_{\mathrm{cell}}\,V_{\mathrm{cell}}^{\ast} = (2\pi)^{3}
  • Plane normal: the reciprocal vector Ghkl\boldsymbol G_{hkl} is normal to the (hkl)(hkl) planes.
  • Angle between planes (h1k11)(h_{1}k_{1}\ell_{1}) and (h2k22)(h_{2}k_{2}\ell_{2}):
cosθ=G1G2G1G2\cos\theta = \frac{\boldsymbol G_{1}\cdot \boldsymbol G_{2}}{|\boldsymbol G_{1}|\,|\boldsymbol G_{2}|}
  • Spacing again: dhkl=2π/Ghkld_{hkl}=2\pi/|\boldsymbol G_{hkl}|.

10.1.10 Worked mini-examples

(a) Reciprocal of bcc is fcc
Start with bcc conventional cubic cell of side aa and primitive vectors a1=a2(y^+z^x^)\boldsymbol a_{1}=\tfrac a2(\hat y+\hat z-\hat x) etc. Compute bi\boldsymbol b_{i} and show the reciprocal points sit at face centers of a cubic net with spacing 2π/a2\pi/a.

(b) dhkld_{hkl} for cubic
Using Ghkl=(2π/a)(hx^+ky^+z^)\boldsymbol G_{hkl}=(2\pi/a)(h\,\hat x + k\,\hat y + \ell\,\hat z), show dhkl=a/h2+k2+2d_{hkl}=a/\sqrt{h^{2}+k^{2}+\ell^{2}}. Bragg gives 2dsinθ=nλ2d\sin\theta=n\lambda; solve for the (111)(111) first-order reflection angle for given λ/a\lambda/a.

(c) fcc systematic absences
Evaluate FF for (100)(100), (110)(110), (111)(111) families and confirm only “all even” or “all odd” survive.

(d) Diamond zeros
Diamond has F[1+eiπ(h+k+)/2]×(fcc factor)F \propto \left[1 + e^{-i\pi(h+k+\ell)/2}\right]\times(\text{fcc factor}). Show that (222)(222) is extinct but (111)(111) is allowed, consistent with X-ray tables.

(e) Ewald construction sketch
For λ=1.54 A˚\lambda=1.54\ \text{\AA} (Cu Kα\alpha), draw a circle of radius 2π/λ2\pi/\lambda through the origin of the reciprocal fcc lattice and mark which G\boldsymbol G vectors on the circle predict reflections for a given crystal orientation.


10.1.11 Common pitfalls

  • Confusing lattice with structure: lattice is the periodic scaffold; the basis puts atoms on it.
  • Wrong units in reciprocal space: crystallographers often use 2π2\pi-free convention; here aibj=2πδija_{i}\cdot b_{j}=2\pi\delta_{ij}. Pick one and stay loyal.
  • Miller vs direction indices: (hkl)(hkl) is a plane, [uvw][uvw] is a direction. For cubic, (hkl)[hkl](hkl)\perp [hkl]; not true in general.
  • Forgetting the form factor: even if F0F\neq 0, fj(G)f_{j}(\boldsymbol G) decays with G|\boldsymbol G|, reshaping intensities.
  • BZ confusion: the “reduced zone” scheme folds kk into the first BZ; if bands look doubled, you probably mixed extended vs reduced plots.

10.1.12 Minimal problem kit

  • Derive the reciprocal vectors bi\boldsymbol b_{i} from ai\boldsymbol a_{i} and prove VcellVcell=(2π)3V_{\mathrm{cell}}\,V_{\mathrm{cell}}^{\ast}=(2\pi)^{3}
  • Show that the Laue condition kk=G\boldsymbol k'-\boldsymbol k=\boldsymbol G and Bragg’s law 2dsinθ=nλ2d\sin\theta=n\lambda are equivalent via G\boldsymbol G geometry
  • Compute the structure factor and selection rules for NaCl and CsCl; compare which (hkl)(hkl) vanish in each
  • For a hexagonal lattice with aa and cc, write bi\boldsymbol b_{i} and obtain dhkld_{hkl} in terms of (hki)(h\,k\,i\,\ell)
  • Locate the high-symmetry points Γ\Gamma, XX, LL, WW for fcc in Cartesian kk-coordinates given aa, and sketch the first BZ

10.2 Lattice Vibrations & Phonons

Atoms in a crystal aren’t statues; they wiggle. Those collective wiggles are phonons: quantized normal modes of the lattice. In this section we go from 1D chains to full 3D dynamical matrices, split acoustic vs optical branches, compute dispersions, group velocities, and densities of states, then quantize. We also hit thermal transport, anharmonicity, thermal expansion via the Grüneisen parameter, electron–phonon coupling, and quick experimental fingerprints (neutron, Raman, IR).


10.2.1 Harmonic lattice: setting the stage

Let R\boldsymbol R label Bravais sites and κ\kappa the basis atom index inside a unit cell. The displacement of atom (R,κ)(\boldsymbol R,\kappa) is uκ(R,t)\boldsymbol u_{\kappa}(\boldsymbol R,t). In the harmonic approximation the potential energy is quadratic in displacements:

U=12RκαRκβΦκα,κβ(RR)uκα(R)uκβ(R)U = \tfrac{1}{2}\sum_{\boldsymbol R\kappa\alpha}\sum_{\boldsymbol R'\kappa'\beta} \Phi_{\kappa\alpha,\kappa'\beta}(\boldsymbol R-\boldsymbol R')\,u_{\kappa\alpha}(\boldsymbol R)\,u_{\kappa'\beta}(\boldsymbol R')

where Φ\Phi are force constants and α,β{x,y,z}\alpha,\beta\in\{x,y,z\}. Newton:

Mκu¨κα(R,t)=RκβΦκα,κβ(RR)uκβ(R,t)M_{\kappa}\,\ddot u_{\kappa\alpha}(\boldsymbol R,t) = -\sum_{\boldsymbol R'\kappa'\beta} \Phi_{\kappa\alpha,\kappa'\beta}(\boldsymbol R-\boldsymbol R')\,u_{\kappa'\beta}(\boldsymbol R',t)

Try plane-wave solutions uκα(R,t)=eκα(q)ei(qRωt)u_{\kappa\alpha}(\boldsymbol R,t)=e_{\kappa\alpha}(\boldsymbol q)\,e^{i(\boldsymbol q\cdot \boldsymbol R - \omega t)}. You get an eigenproblem

κβDκα,κβ(q)eκβ(q)=ω2(q)eκα(q)\sum_{\kappa'\beta} D_{\kappa\alpha,\kappa'\beta}(\boldsymbol q)\,e_{\kappa'\beta}(\boldsymbol q) = \omega^{2}(\boldsymbol q)\,e_{\kappa\alpha}(\boldsymbol q)

with the dynamical matrix

Dκα,κβ(q)=1MκMκRΦκα,κβ(R)eiqRD_{\kappa\alpha,\kappa'\beta}(\boldsymbol q) = \frac{1}{\sqrt{M_{\kappa}M_{\kappa'}}} \sum_{\boldsymbol R} \Phi_{\kappa\alpha,\kappa'\beta}(\boldsymbol R)\,e^{i\boldsymbol q\cdot \boldsymbol R}

For rr atoms per cell you get 3r3r branches: 3 acoustic ω0\omega\to 0 as q0\boldsymbol q\to 0, and 3r33r-3 optical.


10.2.2 1D monatomic chain: the cleanest demo

Masses mm connected by springs KK with lattice spacing aa. Let unu_{n} be the displacement of site nn. Equation of motion:

mu¨n=K(un+12un+un1)m\,\ddot u_{n} = K\,(u_{n+1} - 2u_{n} + u_{n-1})

Ansatz un=u0ei(nkaωt)u_{n}=u_{0}e^{i(nka-\omega t)} gives the dispersion

ω(k)=2Km  sinka2\omega(k) = 2\sqrt{\frac{K}{m}}\;\left|\sin\frac{ka}{2}\right|

Key takeaways:

  • Acoustic branch only, linear near k=0k=0 with sound speed vs=aK/mv_{s}=a\sqrt{K/m}
  • Brillouin zone is π/a<kπ/a-\pi/a<k\le \pi/a; ω(k)\omega(k) is periodic with 2π/a2\pi/a
  • Group velocity vg(k)=ω/kv_{g}(k)=\partial\omega/\partial k controls energy flow

Density of states in 1D for a single branch:

g(ω)=Lπ1dω/dkg(\omega) = \frac{L}{\pi}\,\frac{1}{|d\omega/dk|}

So g(ω)g(\omega) diverges at BZ edges where dω/dk0d\omega/dk\to 0 (van Hove singularity).


10.2.3 1D diatomic chain: acoustic + optical

Two alternating masses m1,m2m_{1},m_{2}, same spring KK, unit cell length aa containing both atoms. With nearest-neighbor coupling, the two-branch dispersion is

ω2(k)=K(1m1+1m2)±K2(1m1+1m2)24K2m1m2sin2 ⁣ka2\omega^{2}(k) = K\left(\frac{1}{m_{1}}+\frac{1}{m_{2}}\right) \pm \sqrt{K^{2}\left(\frac{1}{m_{1}}+\frac{1}{m_{2}}\right)^{2} - \frac{4K^{2}}{m_{1}m_{2}}\sin^{2}\!\frac{ka}{2}}
  • The minus sign is the acoustic branch with ω0\omega\to 0 as k0k\to 0
  • The plus sign is the optical branch with ω(k=0)=2K(1/m1+1/m2)\omega(k=0)=\sqrt{2K(1/m_{1}+1/m_{2})}

At k=0k=0 the acoustic mode moves the two atoms in phase; the optical mode moves them out of phase. A gap opens at the zone boundary.


10.2.4 3D crystals: polarizations and sound speeds

For r=1r=1 (monatomic Bravais) you have 3 acoustic branches: one longitudinal (LA) and two transverse (TA). Near q=0\boldsymbol q=0 they are sound waves with linear dispersion

ωLA(q)vLq,ωTA(q)vTq\omega_{\text{LA}}(\boldsymbol q) \approx v_{L}|\boldsymbol q|,\qquad \omega_{\text{TA}}(\boldsymbol q) \approx v_{T}|\boldsymbol q|

In cubic crystals the speeds relate to elastic constants and mass density ρ\rho (continuum limit):

vL=C11ρ,vT=C44ρv_{L} = \sqrt{\frac{C_{11}}{\rho}},\qquad v_{T} = \sqrt{\frac{C_{44}}{\rho}}

More generally, the Christoffel equation gives direction-dependent velocities.

For r>1r>1 you add 3r33r-3 optical branches that typically have weak q\boldsymbol q-dispersion near the zone center. If the basis carries opposite charges, some optical modes are IR active (see §10.14).


10.2.5 Quantization: phonons as bosons

Promote each normal mode (q,s)(\boldsymbol q,s) to a harmonic oscillator:

H=q,sωqs(bqsbqs+12)H = \sum_{\boldsymbol q,s} \hbar\omega_{\boldsymbol q s}\left(b^{\dagger}_{\boldsymbol q s} b_{\boldsymbol q s} + \tfrac{1}{2}\right)

with [bqs,bqs]=δqqδss[b_{\boldsymbol q s},b^{\dagger}_{\boldsymbol q' s'}]=\delta_{\boldsymbol q\boldsymbol q'}\delta_{ss'}. The displacement operator is

uκ(R)=q,s2MκNωqs  eκs(q)(bqseiqR+bqseiqR)\boldsymbol u_{\kappa}(\boldsymbol R) = \sum_{\boldsymbol q,s} \sqrt{\frac{\hbar}{2M_{\kappa} N \omega_{\boldsymbol q s}}}\; \boldsymbol e_{\kappa s}(\boldsymbol q)\,\left(b_{\boldsymbol q s} e^{i\boldsymbol q\cdot \boldsymbol R} + b^{\dagger}_{\boldsymbol q s} e^{-i\boldsymbol q\cdot \boldsymbol R}\right)

Phonon occupation follows Bose statistics with chemical potential μ=0\mu=0 in equilibrium.


10.2.6 Heat capacity and Debye revisit

Total phonon energy per volume U/V=sdωωgs(ω)nB(ω)U/V=\sum_{s}\int d\omega\,\hbar\omega\,g_{s}(\omega)\,n_{B}(\omega) with nB=(eβω1)1n_{B}=(e^{\beta\hbar\omega}-1)^{-1}. Low-TT acoustic modes \Rightarrow Debye law CVT3C_{V}\propto T^{3} in 3D; high TT \Rightarrow Dulong–Petit 3NkB3Nk_{B} (see §9.6 for the full Debye integral). Optical modes add Schottky-like bumps when their gaps become thermally accessible.


10.2.7 Thermal transport: κ\kappa from kinetic theory

In a crystal, heat is mostly carried by acoustic phonons. Kinetic theory gives

κ=13sdω  Cωsvg2(ω,s)τ(ω,s)\kappa = \frac{1}{3}\sum_{s}\int d\omega\; C_{\omega s}\,v_{g}^{2}(\omega,s)\,\tau(\omega,s)

where Cωs=ωnB/TC_{\omega s}=\hbar\omega\,\partial n_{B}/\partial T is the mode heat capacity, vg=ω/qv_{g}=\partial\omega/\partial q the group velocity, and τ\tau the lifetime. Scattering channels:

  • Phonon–phonon (anharmonic). Split into Normal (N) processes conserving crystal momentum and Umklapp (U) with q1+q2=q3+G\boldsymbol q_{1}+\boldsymbol q_{2}=\boldsymbol q_{3}+\boldsymbol G that degrade heat current
  • Boundary scattering (dominant in nanoscale or at very low TT)
  • Isotope/disorder scattering
  • Phonon–electron and phonon–defect scattering in metals

Empirically, κ(T)\kappa(T) often rises as T3\sim T^{3} at very low TT (ballistic/Boundary), peaks, then falls roughly as 1/T1/T at high TT (Umklapp-limited).


10.2.8 Anharmonicity, frequency shifts, and lifetimes

Real crystals aren’t perfectly harmonic. Cubic and quartic terms cause:

  • Thermal expansion (quasi-harmonic effect)
  • Frequency renormalization ω(q,T)\omega(\boldsymbol q,T)
  • Finite lifetimes τqs\tau_{\boldsymbol q s} via three-phonon and four-phonon scattering

A handy bulk measure is the Grüneisen parameter

γqslnωqslnV,γqsγqsCqsqsCqs\gamma_{\boldsymbol q s} \equiv -\frac{\partial \ln \omega_{\boldsymbol q s}}{\partial \ln V},\qquad \gamma \equiv \frac{\sum_{\boldsymbol q s} \gamma_{\boldsymbol q s} C_{\boldsymbol q s}}{\sum_{\boldsymbol q s} C_{\boldsymbol q s}}

The volumetric thermal expansion coefficient α\alpha relates to γ\gamma and the bulk modulus BB:

α=γCVBV\alpha = \frac{\gamma\,C_{V}}{B\,V}

(Quasi-harmonic approximation; works surprisingly well for many simple solids.)


10.2.9 Electron–phonon coupling: the vibes behind resistivity and pairing

Displacements modulate the electronic Hamiltonian HeH_{e}, producing an interaction He-phH_{e\text{-}ph}. Two archetypes:

  • Deformation potential (short-range): δVΞu\delta V \sim \Xi\,\nabla\cdot \boldsymbol u couples to LA phonons
  • Fröhlich (long-range polar): in ionic crystals, LO phonons create macroscopic electric fields that couple strongly to electrons

Consequences:

  • Metal resistivity: at low TΘDT\ll \Theta_{D}, Bloch–Grüneisen law ρT5\rho \propto T^{5}; at high TT roughly ρT\rho\propto T
  • Superconductivity: virtual phonon exchange provides an effective attraction that can overcome Coulomb repulsion near the Fermi surface (BCS)

10.2.10 Instabilities and soft modes: Peierls in 1D

In 1D metals the electronic susceptibility peaks at q=2kFq=2k_{F}. Electron–phonon coupling softens the phonon at that wavevector; below a critical TT the ω(2kF)0\omega(2k_{F})\to 0 soft mode drives a lattice distortion with doubled period (dimerization), opening a gap: a Peierls transition. In higher dimensions, charge-density waves and structural transitions follow similar logic.


10.2.11 Experimental fingerprints of phonons

  • Inelastic neutron scattering (INS): measures the dynamic structure factor S(q,ω)S(\boldsymbol q,\omega), mapping full ω(q)\omega(\boldsymbol q) and eigenvectors
  • Raman spectroscopy: probes zone-center optical phonons; symmetry selection rules apply
  • Infrared (IR) absorption/reflectivity: IR-active optical phonons in polar crystals split LO vs TO at Γ\Gamma (Lyddane–Sachs–Teller relation)
  • Brillouin light scattering: accesses long-wavelength acoustic phonons (GHz)
  • Specific heat and thermal conductivity vs TT: bulk thermodynamic signatures of DOS and scattering

10.2.12 Worked mini-examples

(a) Group velocity in the monatomic chain
From ω(k)=2K/msin(ka/2)\omega(k)=2\sqrt{K/m}\,\left|\sin(ka/2)\right|, compute vg(k)v_{g}(k) and show vgvs=aK/mv_{g}\to v_{s}=a\sqrt{K/m} as k0k\to 0.

(b) DOS edge singularity
Use g(ω)=(L/π)/dω/dkg(\omega)=(L/\pi)/|d\omega/dk| to show the 1D DOS diverges at k=π/ak=\pi/a.

(c) Diatomic gap at Γ\Gamma
Evaluate ωopt(0)\omega_{\text{opt}}(0) and ωac(0)\omega_{\text{ac}}(0) for the diatomic chain and interpret the eigenvectors.

(d) Debye T3T^{3}
Assuming three acoustic branches with linear dispersion up to a Debye cutoff qDq_{D}, integrate to get CVT3C_{V}\propto T^{3} at low TT.

(e) Umklapp threshold
Show that an Umklapp process requires q1+q2\boldsymbol q_{1}+\boldsymbol q_{2} to leave the first BZ; estimate the minimum temperature where such q\boldsymbol q become thermally populated in a simple cubic with Debye wavevector qDq_{D}.

(f) Thermal expansion from γ\gamma
Starting from F(T,V)=Uel(V)+qs[12ωqs+kBTln(1eβωqs)]F(T,V)=U_{\text{el}}(V)+\sum_{\boldsymbol q s}\left[\tfrac{1}{2}\hbar\omega_{\boldsymbol q s} + k_{B}T\ln(1-e^{-\beta\hbar\omega_{\boldsymbol q s}})\right], show that α=γCV/(BV)\alpha=\gamma C_{V}/(BV) in the quasi-harmonic approximation.


10.2.13 Common pitfalls

  • “Optical requires ions.” No: you only need r>1r>1 atoms per cell. IR activity requires polarity; optical phonons exist regardless
  • Phase vs group velocity: energy transport uses vg=ω/kv_{g}=\partial\omega/\partial k, not ω/k\omega/k
  • Forgetting selection rules: Raman sees only Raman-active symmetries; IR sees IR-active modes; neutrons couple to nuclear positions and magnetic moments
  • Mixing BZ schemes: reduced vs extended zone pictures change where branches appear but not the physics
  • Overusing harmonic approximation: heat transport and expansion need anharmonicity; include N vs U processes when modeling κ(T)\kappa(T)

10.2.14 Minimal problem kit

  • Derive the monatomic and diatomic 1D dispersions and sketch both within the first BZ
  • Build a 6×66\times 6 dynamical matrix for a 2-atom cubic basis with nearest-neighbor central forces and extract ω(q)\omega(\boldsymbol q) along a high-symmetry line
  • Compute g(ω)g(\omega) for a Debye model in d=1,2,3d=1,2,3 and show the corresponding low-TT specific heat scalings
  • Use Matthiessen’s rule with boundary, isotope, and Umklapp scattering rates to fit a synthetic κ(T)\kappa(T) curve; identify the peak temperature scaling with sample size
  • Estimate the Bloch–Grüneisen temperature and the T5T^{5} law crossover for a metal with given vsv_{s} and kFk_{F}

10.3 Drude–Sommerfeld Metals

Metals look messy at the atomic scale, but their electrons act surprisingly lawful. Drude gave the first kinetic picture (1900): classical electrons as a gas of charged billiard balls with occasional collisions. Sommerfeld (1928) fixed the big misses using Fermi–Dirac statistics. Together they still power a shocking amount of metal physics: DC/AC conductivity, Hall effect, magnetotransport, specific heat, Pauli paramagnetism, and the Wiedemann–Franz law.


10.3.1 Drude model: one relaxation time, huge mileage

Assume conduction electrons of (effective) mass mm^{\ast} and charge e-e experience randomizing collisions with mean time τ\tau.

Equation of motion (steady drive) for average velocity v\boldsymbol v under fields (E,B)(\boldsymbol E,\boldsymbol B):

mdvdt=e(E+v×B)mvτm^{\ast}\,\frac{d\boldsymbol v}{dt} = -e\left(\boldsymbol E + \boldsymbol v\times \boldsymbol B\right) - \frac{m^{\ast}\boldsymbol v}{\tau}
  • DC conductivity at B=0B=0 with current density J=nev\boldsymbol J = -ne\,\boldsymbol v:
σ0=ne2τm,ρ01σ0=mne2τ\sigma_{0} = \frac{ne^{2}\tau}{m^{\ast}},\qquad \rho_{0} \equiv \frac{1}{\sigma_{0}} = \frac{m^{\ast}}{ne^{2}\tau}

Define mobility μeτ/m\mu \equiv e\tau/m^{\ast} so σ0=neμ\sigma_{0}=ne\mu.

  • AC (Drude) conductivity for field eiωt\propto e^{-i\omega t}:
σ(ω)=σ01iωτ\sigma(\omega) = \frac{\sigma_{0}}{1 - i\omega\tau}

Real part is a Lorentzian of width 1/τ1/\tau; imaginary part switches sign at ω1/τ\omega\sim 1/\tau.

  • Plasma frequency (free-electron pole in optics):
ωp2=ne2ε0m\omega_{p}^{2} = \frac{ne^{2}}{\varepsilon_{0} m^{\ast}}

For ωωp\omega\ll\omega_{p}, metals reflect; for ω>ωp\omega>\omega_{p}, they transmit like a dielectric.

  • Skin depth in the good-conductor limit (ωτ1\omega\tau\ll 1):
δ=2μ0ωσ0\delta = \sqrt{\frac{2}{\mu_{0}\,\omega\,\sigma_{0}}}

10.3.2 Hall effect and the conductivity tensor

With B=Bz^\boldsymbol B = B\hat z, the Drude solution gives the conductivity tensor

σ=(σ01+(ωcτ)2σ0ωcτ1+(ωcτ)20σ0ωcτ1+(ωcτ)2σ01+(ωcτ)2000σ0),ωceBm\sigma = \begin{pmatrix} \dfrac{\sigma_{0}}{1+(\omega_{c}\tau)^{2}} & \dfrac{\sigma_{0}\,\omega_{c}\tau}{1+(\omega_{c}\tau)^{2}} & 0 \\ -\dfrac{\sigma_{0}\,\omega_{c}\tau}{1+(\omega_{c}\tau)^{2}} & \dfrac{\sigma_{0}}{1+(\omega_{c}\tau)^{2}} & 0 \\ 0 & 0 & \sigma_{0} \end{pmatrix}, \qquad \omega_{c} \equiv \frac{eB}{m^{\ast}}

Inverting gives the resistivity tensor. Two key outputs:

  • Hall coefficient (weak field):
RHEyJxB=1neR_{H} \equiv \frac{E_{y}}{J_{x}B} = -\,\frac{1}{ne}
  • Magnetoresistance in the single-band, isotropic Drude model is zero: ρxx(B)=ρ0\rho_{xx}(B)=\rho_{0} despite σxx\sigma_{xx} shrinking, because the tensor inversion cancels it. Real materials show finite magnetoresistance due to multiple bands, anisotropy, or τ(k)\tau(\boldsymbol k) variations.

10.3.3 Drude thermal transport and Wiedemann–Franz (preview)

A simple kinetic estimate gives electronic thermal conductivity κe13CevF2τ\kappa_{e}\sim \tfrac{1}{3} C_{e} v_{F}^{2}\tau (once we introduce vFv_{F}). Drude naively predicts the Wiedemann–Franz ratio

κeσTL=?const\frac{\kappa_{e}}{\sigma T} \equiv L \stackrel{?}{=}\text{const}

But only with Sommerfeld statistics do we get the correct Lorenz number

L0=π23(kBe)2L_{0} = \frac{\pi^{2}}{3}\left(\frac{k_{B}}{e}\right)^{2}

We’ll derive this properly after quantizing the electron gas.


10.3.4 The fix: Sommerfeld free-electron gas

Electrons are fermions. Put NN of them in a box of volume VV and fill momentum states up to the Fermi sphere.

  • Fermi wavevector and energy (spin degeneracy g=2g=2):
kF=(3π2n)1/3,ϵF=2kF22m,vF=kFmk_{F} = \left(3\pi^{2} n\right)^{1/3},\qquad \epsilon_{F} = \frac{\hbar^{2} k_{F}^{2}}{2 m^{\ast}},\qquad v_{F}=\frac{\hbar k_{F}}{m^{\ast}}
  • Density of states per volume for a parabolic band:
g(ϵ)=12π2(2m2)3/2ϵg(\epsilon) = \frac{1}{2\pi^{2}}\left(\frac{2m^{\ast}}{\hbar^{2}}\right)^{3/2}\sqrt{\epsilon}

At low TTFT\ll T_{F}, only a shell of width kBT\sim k_{B}T around ϵF\epsilon_{F} is thermally active. This explains why most classical Drude thermodynamics failed.


10.3.5 Electronic specific heat: linear in TT

Using Sommerfeld expansion,

Ce=γT,γ=π23kB2g(ϵF)C_{e} = \gamma\,T,\qquad \gamma = \frac{\pi^{2}}{3}\,k_{B}^{2}\,g(\epsilon_{F})

Linear TT matches experiments and is tiny compared to 3NkB3Nk_{B} at room temperature because only (T/TF)\sim (T/T_{F}) of electrons participate.


10.3.6 Wiedemann–Franz, done right

Kinetic theory with Fermi statistics yields

κe=13vF2τCe\kappa_{e} = \frac{1}{3} v_{F}^{2}\tau\,C_{e}

Combine with σ0=ne2τ/m\sigma_{0}=ne^{2}\tau/m^{\ast} and n=ϵFg(ϵ)dϵn=\int^{\epsilon_{F}} g(\epsilon)\,d\epsilon to obtain, for elastic scattering with energy-independent τ\tau,

κeσT=π23(kBe)2L0\frac{\kappa_{e}}{\sigma T} = \frac{\pi^{2}}{3}\left(\frac{k_{B}}{e}\right)^{2} \equiv L_{0}

Deviations of LL from L0L_{0} diagnose inelastic scattering, strong energy dependence of τ\tau, or additional heat carriers (phonons, magnons).


10.3.7 Pauli paramagnetism and Landau diamagnetism (sketch)

  • Pauli paramagnetism from spin polarization in a field BB:
χPauli=μ0μB2g(ϵF)\chi_{\text{Pauli}} = \mu_{0}\,\mu_{B}^{2}\,g(\epsilon_{F})

Small, TT-independent to leading order.

  • Landau diamagnetism from orbital motion in quantizing fields gives
χLandau=13χPauli\chi_{\text{Landau}} = -\frac{1}{3}\,\chi_{\text{Pauli}}

for a free-electron gas. In real metals the net χ\chi also includes core diamagnetism and possible Van Vleck terms.


10.3.8 Thermopower and the Mott formula (one-liner)

The Seebeck coefficient SS (voltage from a temperature gradient) vanishes in Drude’s classical picture but, with Fermi statistics,

S=π23kB2Tedlnσ(ϵ)dϵϵFS = -\frac{\pi^{2}}{3}\,\frac{k_{B}^{2}T}{e}\left.\frac{d\ln \sigma(\epsilon)}{d\epsilon}\right|_{\epsilon_{F}}

This Mott formula makes SS a sensitive probe of particle–hole asymmetry and scattering near ϵF\epsilon_{F}.


10.3.9 What Drude gets wrong (and why Sommerfeld wins)

  • Heat capacity: classical Drude predicts CeNkBC_{e}\sim Nk_{B}; data show CeTNkBC_{e}\propto T\ll Nk_{B}. Sommerfeld fixes it.
  • Hall sign: single-band Drude gives RH=1/neR_{H}=-1/ne (electrons). Many materials have positive RHR_{H} due to hole-like Fermi surfaces or multiband compensation.
  • Magnetoresistance: zero in single-band Drude, but finite in reality from anisotropy, multiple pockets, or τ(k)\tau(\boldsymbol k).
  • High-frequency optics: Drude alone misses interband absorption and bound-charge contributions; add Lorentz oscillators for realistic ε(ω)\varepsilon(\omega).

10.3.10 Connecting to band structure

Replace mm^{\ast} by the band effective mass from the dispersion ϵ(k)\epsilon(\boldsymbol k) near the Fermi surface:

1mαβ=122ϵ(k)kαkβkF\frac{1}{m^{\ast}_{\alpha\beta}} = \frac{1}{\hbar^{2}}\left.\frac{\partial^{2}\epsilon(\boldsymbol k)}{\partial k_{\alpha}\,\partial k_{\beta}}\right|_{\boldsymbol k_{F}}

Anisotropic mm^{\ast} makes σ\sigma a tensor even at B=0B=0. Real transport uses Boltzmann theory linearized around the Fermi surface with τ(k)\tau(\boldsymbol k) (see §10.7).


10.3.11 Optical sum rule and spectral weight

The f-sum rule ties the integrated Drude weight to carrier density:

0dω  Reσ(ω)=πne22m\int_{0}^{\infty} d\omega\;\mathrm{Re}\,\sigma(\omega) = \frac{\pi ne^{2}}{2 m^{\ast}}

In correlated systems, Drude weight can be suppressed and shifted to mid-infrared interband-like features—often read as a fingerprint of reduced quasiparticle weight.


10.3.12 Worked mini-examples

(a) Numbers you can trust
Take n=8.5×1028 m3n=8.5\times 10^{28}\ \mathrm{m^{-3}}, τ=3×1014 s\tau=3\times 10^{-14}\ \mathrm{s}, m=mem^{\ast}=m_{e}. Compute σ0\sigma_{0}, ρ0\rho_{0}, ωp\omega_{p}, and δ\delta at ω/2π=10 GHz\omega/2\pi=10\ \mathrm{GHz}.

(b) Hall bar
For a sample with thickness tt, width ww, current II along xx, and field Bz^B\hat z, show the Hall voltage VH=IB/(net)V_{H} = I\,B/(ne\,t) in the Drude single-band limit. Estimate VHV_{H} for copper with the numbers from (a).

(c) Sommerfeld CeC_{e}
Derive Ce=γTC_{e}=\gamma T using the Sommerfeld expansion and express γ\gamma in terms of nn and mm^{\ast}.

(d) Wiedemann–Franz from Boltzmann
Linearize the Boltzmann equation with elastic impurity scattering and show L=L0L=L_{0} when τ(ϵ)\tau(\epsilon) is constant near ϵF\epsilon_{F}.

(e) Mott thermopower sign
Assuming σ(ϵ)ϵs\sigma(\epsilon)\propto \epsilon^{s} near ϵF\epsilon_{F}, show S=(π2/3)(kB2T/e)s/ϵFS = -(\pi^{2}/3)(k_{B}^{2}T/e)\,s/\epsilon_{F} and interpret S>0S>0 vs S<0S<0.


10.3.13 Common pitfalls

  • Forgetting effective mass: always use mm^{\ast} from the band structure, not bare mem_{e}, in transport formulae.
  • Reading magnetoresistance from σxx\sigma_{xx}: the measured ρxx\rho_{xx} can be BB-independent in single-band Drude even though σxx\sigma_{xx} shrinks; invert the tensor.
  • Assuming RH=1/neR_{H}=-1/ne universally: multiband compensation and warped Fermi surfaces can flip the sign or reduce magnitude.
  • Overusing constant τ\tau: energy- or angle-dependent scattering changes LL, SS, and MR; be cautious outside low-TT impurity-limited regimes.
  • Mixing optical and transport masses: cyclotron, thermodynamic, and band masses can differ in anisotropic or interacting systems.

10.4 Bloch Theorem & Band Structure

Periodic potentials take messy atomic chaos and compress it into clean momentum-space rules. The headline: Bloch’s theorem says electrons in a crystal move as plane waves modulated by a lattice-periodic envelope. From that, gaps pop open at Brillouin-zone boundaries, dispersions En(k)E_{n}(\boldsymbol k) define velocities and effective masses, and both “nearly free” and “tight-binding” lenses become the same story, zoomed in from opposite sides.


10.4.1 Bloch states and the central equation

For a single electron in a periodic potential V(r+R)=V(r)V(\boldsymbol r+\boldsymbol R)=V(\boldsymbol r),

ψnk(r)=eikrunk(r),unk(r+R)=unk(r)\psi_{n\boldsymbol k}(\boldsymbol r) = e^{i\boldsymbol k\cdot \boldsymbol r}\,u_{n\boldsymbol k}(\boldsymbol r),\qquad u_{n\boldsymbol k}(\boldsymbol r+\boldsymbol R)=u_{n\boldsymbol k}(\boldsymbol r)

Expand the lattice-periodic part in reciprocal vectors G\boldsymbol G:

unk(r)=Gcn(k+G)eiGru_{n\boldsymbol k}(\boldsymbol r) = \sum_{\boldsymbol G} c_{n}(\boldsymbol k+\boldsymbol G)\,e^{i\boldsymbol G\cdot \boldsymbol r}

and the potential as V(r)=GUGeiGrV(\boldsymbol r)=\sum_{\boldsymbol G} U_{\boldsymbol G}\,e^{i\boldsymbol G\cdot \boldsymbol r}. Insert into Schrödinger’s equation to get the central equation

[22mk+G2E]c(k+G)+GUGGc(k+G)=0\left[\frac{\hbar^{2}}{2m}\,|\boldsymbol k+\boldsymbol G|^{2} - E\right] c(\boldsymbol k+\boldsymbol G) + \sum_{\boldsymbol G'} U_{\boldsymbol G-\boldsymbol G'}\,c(\boldsymbol k+\boldsymbol G') = 0

It’s an infinite linear system in the plane-wave coefficients cc. Truncate sensibly and you get bands.


10.4.2 Zone folding and reduced vs extended zone

Free electrons have parabolas E=2k2/2mE=\hbar^{2}|\boldsymbol k|^{2}/2m that repeat every reciprocal vector: k\boldsymbol k and k+G\boldsymbol k+\boldsymbol G are equivalent. You can:

  • Extended-zone: keep E(k)E(\boldsymbol k) as endlessly repeated parabolas in kk-space
  • Reduced-zone: fold all momenta back into the first Brillouin zone; copies become separate bands indexed by nn

When V0V\neq 0, states that were degenerate under folding mix and gaps open at the crossings.


10.4.3 Gaps from Bragg reflection: a 2×2 demo

At a boundary where k\boldsymbol k and k=kG\boldsymbol k'=\boldsymbol k-\boldsymbol G are degenerate for the free electron, keep only those two plane waves in the central equation. The eigenproblem is

(ϵkUGUGϵk)(ckck)=E(ckck)\begin{pmatrix} \epsilon_{\boldsymbol k} & U_{\boldsymbol G} \\ U_{\boldsymbol G}^{\ast} & \epsilon_{\boldsymbol k'} \end{pmatrix} \begin{pmatrix} c_{\boldsymbol k} \\ c_{\boldsymbol k'} \end{pmatrix} = E \begin{pmatrix} c_{\boldsymbol k} \\ c_{\boldsymbol k'} \end{pmatrix}

with ϵq=2q2/2m\epsilon_{\boldsymbol q}=\hbar^{2}|\boldsymbol q|^{2}/2m. At the exact zone boundary where ϵk=ϵk\epsilon_{\boldsymbol k}=\epsilon_{\boldsymbol k'}, the two energies split as

E±=ϵ0±UGE_{\pm} = \epsilon_{0} \pm |U_{\boldsymbol G}|

so the band gap is

Eg=2UGE_{g} = 2\,|U_{\boldsymbol G}|

Moral: stronger Fourier components of the lattice potential at reciprocal vectors G\boldsymbol G carve bigger gaps at the corresponding Brillouin-zone planes.


10.4.4 Kronig–Penney in one line (1D intuition)

For a 1D periodic array of barriers (lattice constant aa), the dispersion satisfies a transcendental relation

cos(ka)=cos(αa)+Pαasin(αa)\cos(ka) = \cos(\alpha a) + \frac{P}{\alpha a}\,\sin(\alpha a)

where α=2mE/\alpha=\sqrt{2mE}/\hbar in the wells and PP encodes barrier strength. Regions where the right-hand side has magnitude 1\le 1 are allowed bands; where it exceeds 1 are forbidden gaps. Tuning PP moves you smoothly from nearly free (P0P\to 0) to tight-binding (P1P\gg 1).


10.4.5 Group velocity, effective mass, and DOS

Band dispersion is not just eye candy; its derivatives are physics:

  • Group velocity
vn(k)=1kEn(k)\boldsymbol v_{n}(\boldsymbol k) = \frac{1}{\hbar}\,\nabla_{\boldsymbol k} E_{n}(\boldsymbol k)
  • Effective mass tensor near an extremum
(mn)αβ1=122Enkαkβk0\left(m_{n}^{\ast}\right)^{-1}_{\alpha\beta} = \frac{1}{\hbar^{2}}\,\left.\frac{\partial^{2}E_{n}}{\partial k_{\alpha}\,\partial k_{\beta}}\right|_{\boldsymbol k_{0}}

Curvature \uparrow ⇒ light mass; curvature \downarrow ⇒ heavy mass. Negative curvature near a valence-band maximum means holes with positive charge and positive mass run the transport.

  • Density of states for a single band in 3D
g(E)=14π3En(k)=EdSkkEng(E) = \frac{1}{4\pi^{3}} \int_{E_{n}(\boldsymbol k)=E} \frac{dS_{k}}{|\nabla_{\boldsymbol k} E_{n}|}

The DOS diverges where the gradient vanishes along the constant-energy surface (van Hove singularities).


10.4.6 Nearly free vs tight-binding: two zoom levels, same film

  • Nearly free electron (NFE): start from parabolas and open gaps perturbatively where they cross. Good when UGϵF|U_{\boldsymbol G}| \ll \epsilon_{F}.
  • Tight-binding (TB): start from localized orbitals with hopping tt between sites; bands emerge from bonding/antibonding combinations. Best when orbitals are compact and tt is small compared to on-site energies.

Both are limits of the same Bloch machinery. NFE explains gap openings and light masses in simple metals; TB nails band shapes in covalent/ionic solids and will be our §10.5 focus (Wannier functions live here).


10.4.7 kpk\cdot p method: local expansion near a band edge

Near a high-symmetry k0\boldsymbol k_{0}, expand the Hamiltonian as H(k)H(k0)+mkp+H(\boldsymbol k)\approx H(\boldsymbol k_{0})+\frac{\hbar}{m}\,\boldsymbol k\cdot \boldsymbol p + \cdots. Second-order perturbation yields

1mαβ=1mδαβ+2m2mnn0pαm0m0pβn0En0Em0\frac{1}{m^{\ast}_{\alpha\beta}} = \frac{1}{m}\delta_{\alpha\beta} + \frac{2}{m^{2}} \sum_{m\neq n} \frac{\langle n0|p_{\alpha}|m0\rangle \langle m0|p_{\beta}|n0\rangle}{E_{n0}-E_{m0}}

where {m0}\{|m0\rangle\} are Bloch states at k0\boldsymbol k_{0}. This connects effective mass to interband matrix elements and symmetry selection rules (e.g., heavy vs light holes in semiconductors).


10.4.8 Symmetry, degeneracies, and spin–orbit spice

  • Crystal symmetry protects degeneracies at high-symmetry kk points and along lines. Breaking symmetry (strain, fields) splits them.
  • Inversion and time reversal together enforce Kramers doublets with spin.
  • Spin–orbit coupling (SOC) mixes spin and orbital character; with broken inversion you can get Rashba-like splittings. We stay light here and return to topology in §10.11.

10.4.9 Fermi surfaces and transport pointers

The Fermi surface is the constant-energy surface En(k)=ϵFE_{n}(\boldsymbol k)=\epsilon_{F}. Its curvature and connectivity control:

  • Cyclotron masses and quantum oscillations (de Haas–van Alphen, Shubnikov–de Haas)
  • Anisotropic conductivity and Hall effect (via velocities and lifetimes on the FS)
  • Nesting vectors that can amplify density-wave instabilities

Bloch dispersion + scattering time is the whole transport game in §10.7.


10.4.10 Wannier functions: real-space avatars of bands

A unitary transform of Bloch states over the BZ gives Wannier functions n,R|n,\boldsymbol R\rangle localized near cell R\boldsymbol R. Smooth gauge choices yield maximally localized Wanniers that act like tight-binding orbitals and are perfect for model building, interpolation, and computing real-space responses. We build on this in §10.5.


10.4.11 ARPES and band mapping (how we see E(k)E(\boldsymbol k))

Angle-resolved photoemission (ARPES) measures the occupied part of E(k)E(\boldsymbol k) directly. In 2D materials it’s almost literal; in 3D there’s a kzk_{z} wrinkle. Combine with quantum oscillations and optical conductivity to reconstruct the low-energy band structure that transport actually feels.


10.4.12 Worked mini-examples

(a) Gap size from a single Fourier component
Take a 1D cosine potential V(x)=2UGcos(Gx)V(x)=2U_{G}\cos(Gx) with G=2π/aG=2\pi/a. Do the 2×2 central-equation truncation at k=G/2k=G/2 and show Eg=2UGE_{g}=2|U_{G}|.

(b) Effective mass near a band minimum
For E(k)E0+αk2+βk4E(k)\approx E_{0}+\alpha k^{2}+\beta k^{4} in 1D, extract m=2/(2α)m^{\ast}=\hbar^{2}/(2\alpha) and discuss how the k4k^{4} term skews DOS but not mm^{\ast} at the minimum.

(c) DOS singularity in 2D
On a square lattice TB band E(k)=2t(coskxa+coskya)E(\boldsymbol k)=-2t(\cos k_{x}a+\cos k_{y}a), locate the saddle points and show the logarithmic van Hove singularity in g(E)g(E).

(d) Zone-folding check
Start from a free-electron parabola and fold by ±G\pm G into the first BZ. Draw where degeneracies occur and mark which get lifted when UG0U_{G}\neq 0.

(e) kpk\cdot p for a direct-gap semiconductor
At Γ\Gamma, use symmetry to argue that conduction and valence states have opposite parity so the linear-in-kk coupling vanishes; show why the conduction-band mass is then dominated by second-order coupling to remote bands.


10.4.13 Common pitfalls

  • Mixing reduced and extended zones: keep track of whether a “crossing” is real or just a folded copy.
  • Equating curvature with speed: velocity is first derivative, not curvature; flat band \Rightarrow heavy mass and slow carriers.
  • Forgetting the basis: selection rules and gap sizes depend on the Fourier components UGU_{\boldsymbol G}, which encode the atomic basis and bonding.
  • Assuming parabolic everywhere: effective-mass fits work only near extrema; far away, nonparabolicity and anisotropy matter.
  • Ignoring SOC where it’s large: in heavy elements or broken-inversion crystals, SOC reshapes band order and splitting—critical for topology.

10.4.14 Minimal problem kit

  • Derive the central equation from the Bloch expansion and write it as a matrix eigenproblem for a finite set of G\boldsymbol G
  • Perform the 2×2 gap-opening calculation at a generic zone plane with normal G\boldsymbol G; obtain Eg=2UGE_{g}=2|U_{\boldsymbol G}|
  • Compute mm^{\ast} tensor for an ellipsoidal constant-energy surface and relate it to cyclotron mass in a field along an arbitrary axis
  • For a 1D Kronig–Penney with delta barriers, plot allowed bands for several PP and identify the NFE and TB limits
  • Using a simple TB model on honeycomb, show how Dirac cones arise and how a staggered sublattice potential opens a mass gap

10.5 Tight-Binding & Wannier: From Local Orbitals to Bands

When electrons are more “homebody” than “frequent flyer,” tight-binding (TB) is the lens. Start with localized atomiclike orbitals, let them hop between sites, and you get bands with shapes and symmetries carved by real-space chemistry. Wannier functions are the clean, orthonormal real-space avatars of Bloch bands, perfect for building models, computing responses, and adding fields or disorder. This section builds TB from scratch, adds multi-orbital and Slater–Koster rules, handles magnetic fields via Peierls phases, and shows how Wanniers tie to polarization and edge physics (SSH preview).


10.5.1 From orbitals to Bloch sums

Pick orbitals {ϕα(rRτα)}\{\phi_{\alpha}(\boldsymbol r-\boldsymbol R-\boldsymbol\tau_{\alpha})\} centered on basis sites τα\boldsymbol\tau_{\alpha} in cell R\boldsymbol R, with orbital label α\alpha (e.g., s,px,py,pzs,p_{x},p_{y},p_{z}).

Form Bloch sums

φαk(r)=1NReik(R+τα)ϕα(rRτα)\varphi_{\alpha\boldsymbol k}(\boldsymbol r) = \frac{1}{\sqrt{N}} \sum_{\boldsymbol R} e^{i\boldsymbol k\cdot(\boldsymbol R+\boldsymbol\tau_{\alpha})}\,\phi_{\alpha}(\boldsymbol r-\boldsymbol R-\boldsymbol\tau_{\alpha})

Expand eigenstates as ψnk=αcnα(k)φαk\psi_{n\boldsymbol k}=\sum_{\alpha} c_{n\alpha}(\boldsymbol k)\,\varphi_{\alpha\boldsymbol k}. Matrix elements

Hαβ(k)=Reik(R+τβτα)tαβ(R),Sαβ(k)=Reik(R+τβτα)sαβ(R)H_{\alpha\beta}(\boldsymbol k)=\sum_{\boldsymbol R} e^{i\boldsymbol k\cdot(\boldsymbol R+\boldsymbol\tau_{\beta}-\boldsymbol\tau_{\alpha})}\,t_{\alpha\beta}(\boldsymbol R),\qquad S_{\alpha\beta}(\boldsymbol k)=\sum_{\boldsymbol R} e^{i\boldsymbol k\cdot(\boldsymbol R+\boldsymbol\tau_{\beta}-\boldsymbol\tau_{\alpha})}\,s_{\alpha\beta}(\boldsymbol R)

with hopping tαβ(R)ϕα0HϕβRt_{\alpha\beta}(\boldsymbol R)\equiv\langle \phi_{\alpha 0}|H|\phi_{\beta\boldsymbol R}\rangle, overlap sαβ(R)ϕα0ϕβRs_{\alpha\beta}(\boldsymbol R)\equiv\langle \phi_{\alpha 0}|\phi_{\beta\boldsymbol R}\rangle. Solve the generalized eigenproblem

H(k)cn(k)=En(k)S(k)cn(k)H(\boldsymbol k)\,c_{n}(\boldsymbol k)=E_{n}(\boldsymbol k)\,S(\boldsymbol k)\,c_{n}(\boldsymbol k)

In orthogonal TB we approximate S=IS=\mathbb I.


10.5.2 1D monatomic chain (orthogonal TB)

One orbital per site, lattice constant aa, on-site energy ε0\varepsilon_{0}, nearest-neighbor hopping t-t.

E(k)=ε02tcos(ka)E(k)=\varepsilon_{0}-2t\cos(ka)

Bandwidth W=4tW=4t. Effective mass near an extremum at k=0k=0 is

1m=122Ek2=2ta22\frac{1}{m^{\ast}}=\frac{1}{\hbar^{2}}\frac{\partial^{2}E}{\partial k^{2}}=\frac{2ta^{2}}{\hbar^{2}}

Sign of tt sets band curvature and which end is bonding vs antibonding.


10.5.3 Diatomic chain and the SSH teaser

Two sites per cell, AA at 00, BB at a/2a/2, hoppings t1t_{1} and t2t_{2} alternating. In the (A,B)(A,B) basis

H(k)=(εAt1t2eikat1t2eikaεB)H(k)= \begin{pmatrix} \varepsilon_{A} & -t_{1}-t_{2}\,e^{-ika} \\ -t_{1}-t_{2}\,e^{ika} & \varepsilon_{B} \end{pmatrix}

For εA=εB\varepsilon_{A}=\varepsilon_{B} the dispersion is

E±(k)=ε0±t12+t22+2t1t2cos(ka)E_{\pm}(k)=\varepsilon_{0}\pm \sqrt{t_{1}^{2}+t_{2}^{2}+2t_{1}t_{2}\cos(ka)}

Gap at the zone boundary k=π/ak=\pi/a is 2t1t22|t_{1}-t_{2}|. The SSH limit (εA=εB\varepsilon_{A}=\varepsilon_{B}) supports topological edge states when t2>t1|t_{2}|>|t_{1}| with open boundaries; we revisit the topology in §10.11.


10.5.4 Square and honeycomb lattices (classics)

Square (1 orbital, nn)

E(k)=ε02t[cos(kxa)+cos(kya)]E(\boldsymbol k)=\varepsilon_{0}-2t\big[\cos(k_{x}a)+\cos(k_{y}a)\big]

Honeycomb (graphene, 1 pzp_{z} per site, nn). Two-site basis A,BA,B and three nn vectors {δi}\{\boldsymbol\delta_{i}\}:

H(k)=(0ti=13eikδiti=13eikδi0)H(\boldsymbol k)= \begin{pmatrix} 0 & -t\,\sum_{i=1}^{3}e^{i\boldsymbol k\cdot \boldsymbol\delta_{i}} \\ -t\,\sum_{i=1}^{3}e^{-i\boldsymbol k\cdot \boldsymbol\delta_{i}} & 0 \end{pmatrix}

Eigenvalues

E±(k)=±t1+eika1+eika2E_{\pm}(\boldsymbol k)=\pm t\,\Big|1+e^{i\boldsymbol k\cdot \boldsymbol a_{1}}+e^{i\boldsymbol k\cdot \boldsymbol a_{2}}\Big|

Touch at Dirac cones K,KK,K' with linear dispersion and Fermi velocity vF=32at/v_{F}=\tfrac{\sqrt{3}}{2}at/\hbar.


10.5.5 Multi-orbital TB and Slater–Koster rules

Hoppings depend on orbital orientation relative to the bond. Slater–Koster parameterizes two-center integrals using direction cosines (l,m,n)(l,m,n) between sites.

Examples:

  • ssss: tssσt_{ss\sigma}
  • sspp: tspσ(l,m,n)t_{sp\sigma}\,(l,m,n) projected on the pp axis
  • pppp: σ\sigma and π\pi channels

For a bond along x^\hat x:

pxHpx=tppσ,pyHpy=pzHpz=tppπ\langle p_{x}|H|p_{x}\rangle = t_{pp\sigma},\qquad \langle p_{y}|H|p_{y}\rangle = \langle p_{z}|H|p_{z}\rangle = t_{pp\pi}

Cross-hoppings like sHpxtspσ\langle s|H|p_{x}\rangle\propto t_{sp\sigma} times the appropriate direction cosine. This symmetry bookkeeping lets you write H(k)H(\boldsymbol k) for real crystals quickly and keeps you from illegal hoppings.


10.5.6 Beyond nearest neighbors and van Hove moves

Adding next-nearest neighbors tt' etc. reshapes curvature and DOS singularities. On the square lattice

E(k)=ε02t[coskxa+coskya]4tcoskxacoskyaE(\boldsymbol k)=\varepsilon_{0}-2t[\cos k_{x}a+\cos k_{y}a]-4t'\cos k_{x}a\cos k_{y}a

Tuning tt' slides the van Hove saddle through ϵF\epsilon_{F}, a common route to instabilities (density waves, superconductivity) in layered materials.


10.5.7 Magnetic field: Peierls substitution

Minimal coupling in TB enters via Peierls phases on hoppings

tij  tijexp ⁣(ierirjAd)t_{ij}\ \longrightarrow\ t_{ij}\,\exp\!\left(\frac{ie}{\hbar}\int_{\boldsymbol r_{i}}^{\boldsymbol r_{j}} \boldsymbol A\cdot d\boldsymbol\ell\right)

So the product of phases around a plaquette equals the magnetic flux 2πΦ/Φ02\pi\Phi/\Phi_{0}. On a square lattice with rational flux Φ/Φ0=p/q\Phi/\Phi_{0}=p/q, translation symmetry enlarges to a magnetic unit cell and the spectrum fractures into Hofstadter’s butterfly.


10.5.8 Disorder and real-space Green’s functions

Real materials have impurities. Add site randomness εi\varepsilon_{i} or bond randomness tijtij+δtijt_{ij}\to t_{ij}+\delta t_{ij}. The TB Hamiltonian in real space is perfect for:

  • Anderson localization studies (see §10.13)
  • Surface/edge spectra via finite ribbons and recursive Green’s functions
  • Transport with Landauer–Büttiker, where G=(2e2/h)T(EF)G=(2e^{2}/h)\,T(E_{F}) and TT comes from real-space TB + leads

10.5.9 Wannier functions: localized avatars of bands

Given Bloch states {unk}\{|u_{n\boldsymbol k}\rangle\} in a band (or group) subspace, define Wannier functions

R,n=VBZ(2π)3BZd3k  eikRmUmn(k)umk|\,\boldsymbol R, n\,\rangle = \frac{V_{\text{BZ}}}{(2\pi)^{3}}\int_{\text{BZ}} d^{3}k\; e^{-i\boldsymbol k\cdot \boldsymbol R}\, \sum_{m} U_{mn}(\boldsymbol k)\,|\,u_{m\boldsymbol k}\rangle

where U(k)U(\boldsymbol k) is a gauge unitary within the subspace. Properties:

  • Orthonormal and cell-localized. The choice of U(k)U(\boldsymbol k) controls spread
  • Maximally localized Wanniers (MLWFs) choose UU to minimize total spread Ω=n[r2nrn2]\Omega=\sum_{n}\big[\langle r^{2}\rangle_{n}-|\langle \boldsymbol r\rangle_{n}|^{2}\big]
  • For isolated bands in trivial topology, exponentially localized Wanniers exist and reproduce the band exactly

Tight-binding from Wannier. Hoppings are matrix elements between Wanniers

tnm(R)=0,nHR,mt_{nm}(\boldsymbol R)=\langle \boldsymbol 0,n|H|\boldsymbol R,m\rangle

The interpolation H(k)=ReikRt(R)H(\boldsymbol k)=\sum_{\boldsymbol R} e^{i\boldsymbol k\cdot \boldsymbol R}\,t(\boldsymbol R) gives bands indistinguishable from the ab initio subspace but at TB cost.


10.5.10 Polarization and Wannier centers (Berry twist)

Electronic polarization per cell of an insulator relates to Wannier centers rn=0,nr^0,n\boldsymbol r_{n}=\langle \boldsymbol 0,n|\hat{\boldsymbol r}| \boldsymbol 0,n\rangle:

P=eVcellnoccrnmodulo a quantum\boldsymbol P = -\frac{e}{V_{\text{cell}}}\sum_{n}^{\text{occ}} \boldsymbol r_{n}\quad \text{modulo a quantum}

Equivalent Berry-phase formula uses the Bloch gauge. Practically, track Wannier center flow under adiabatic changes to compute polarization differences and piezoelectric responses (topology returns in §10.11).


10.5.11 Edge physics with TB: the SSH demo in 15 seconds

Open a finite chain with alternating t1,t2t_{1},t_{2}. In the topological phase t2>t1|t_{2}|>|t_{1}| you get midgap edge modes exponentially localized with decay (t1/t2)x\sim (t_{1}/t_{2})^{x}. Chiral symmetry pins them at E=0E=0 until perturbations that break sublattice symmetry lift them.


10.5.12 Worked mini-examples

(a) Wannier from a 1D cosine band
For E(k)=ε02tcoskaE(k)=\varepsilon_{0}-2t\cos ka, show the Wannier w0(x)w_{0}(x) is proportional to a modified Bessel K0(x/ξ)K_{0}(|x|/\xi) asymptotically, with localization length ξat/Δ\xi\sim a\sqrt{t/\Delta} set by the nearest gap.

(b) Graphene Fermi velocity
From the honeycomb nn Hamiltonian, linearize near KK to get HDirac=vFσqH_{\text{Dirac}}=\hbar v_{F}\,\boldsymbol\sigma\cdot \boldsymbol q and extract vF=32at/v_{F}=\tfrac{\sqrt{3}}{2}at/\hbar.

(c) Slater–Koster on a cubic bond
For a cubic crystal with one ss and three pp orbitals per site and nn bonds along axes, write H(k)H(\boldsymbol k) using tssσ,tspσ,tppσ,tppπt_{ss\sigma},t_{sp\sigma},t_{pp\sigma},t_{pp\pi}.

(d) Peierls phase check
On a square lattice in Landau gauge A=(0,Bx,0)\boldsymbol A=(0,Bx,0), show the hopping along y^\hat y picks a phase ei2πϕme^{i 2\pi \phi\, m} with ϕΦ/Φ0\phi\equiv \Phi/\Phi_{0} and mm the xx-index, while x^\hat x hoppings stay real.

(e) Overlap matters
Include a nearest-neighbor overlap ss in the 1D chain and solve the generalized eigenproblem to show the dispersion becomes E(k)=(ε02tcoska)/(1+2scoska)E(k)=(\varepsilon_{0}-2t\cos ka)/(1+2s\cos ka).

(f) SSH edge states
Diagonalize a 40-site open SSH chain numerically (TB) and plot the wavefunctions of the two midgap states; verify exponential localization and parity on opposite ends.


10.5.13 Common pitfalls

  • Forgetting overlap: many intro TBs set S=IS=\mathbb I; when orbitals are not orthogonal, use the generalized problem or orthogonalize first
  • Illegal hoppings: if symmetry forbids a matrix element, do not sneak it in. Use Slater–Koster or group theory to zero the right terms
  • Gauge sloppiness in Wanniers: random phases across k\boldsymbol k give delocalized Wanniers. MLWFs fix the gauge—do that before trusting real-space hoppings
  • Peierls without flux accounting: ensure the product of link phases equals 2πΦ/Φ02\pi\Phi/\Phi_{0} per plaquette, independent of gauge
  • Overfitting TB: too many parameters hide physics and blow up outside the fit window. Start minimal; add neighbors only as data demand

10.5.14 Minimal problem kit

  • Build the 1D and diatomic TB dispersions and extract effective masses around band edges
  • Write the honeycomb nn Hamiltonian, find K,KK,K', linearize to obtain the Dirac Hamiltonian and vFv_{F}
  • Use Slater–Koster to construct a sp2sp^{2} TB for graphene including ppπpp\pi and ppσpp\sigma sectors; discuss why pzp_{z} decouples in the flat sheet
  • Implement Peierls phases for square-lattice TB at rational flux p/qp/q and plot the Hofstadter spectrum for q10q\le 10
  • Generate MLWFs for a two-band toy model, compute t(R)t(\boldsymbol R), and show Wannier-interpolated E(k)E(\boldsymbol k) matches the exact band to machine precision
  • Compute the polarization change from a sliding dimerization in the SSH model via Wannier center shift and compare with the Berry-phase result

10.6 Semiconductors & Doping

Semiconductors are metals on “hard mode”: carriers are scarce until you tune the band gap, sprinkle dopants, or shine photons. This section builds the effective-mass picture, intrinsic vs doped statistics, drift–diffusion with the Einstein relation, recombination channels, optical fingerprints, and the minimalist pn-junction toolkit. We keep it band-structure real while staying device-friendly.


10.6.1 Band edges, effective masses, and densities of states

Near a conduction minimum and a valence maximum,

Ec(k)Ec0+22kTme1k,Ev(k)Ev022kTmh1kE_{c}(\boldsymbol k) \approx E_{c0} + \frac{\hbar^{2}}{2}\,\boldsymbol k^{\mathsf T} m_{e}^{\ast -1}\,\boldsymbol k,\qquad E_{v}(\boldsymbol k) \approx E_{v0} - \frac{\hbar^{2}}{2}\,\boldsymbol k^{\mathsf T} m_{h}^{\ast -1}\,\boldsymbol k

Define the band gap Eg=Ec0Ev0E_{g}=E_{c0}-E_{v0}. Effective-mass tensors me,mhm_{e}^{\ast},m_{h}^{\ast} may be anisotropic; often one uses DOS masses me,DOS,mh,DOSm_{e}^{\ast,\mathrm{DOS}}, m_{h}^{\ast,\mathrm{DOS}} to define 3D densities of states

Nc(T)=2(me,DOSkBT2π2)3/2,Nv(T)=2(mh,DOSkBT2π2)3/2N_{c}(T) = 2\left(\frac{m_{e}^{\ast,\mathrm{DOS}} k_{B} T}{2\pi\hbar^{2}}\right)^{3/2},\qquad N_{v}(T) = 2\left(\frac{m_{h}^{\ast,\mathrm{DOS}} k_{B} T}{2\pi\hbar^{2}}\right)^{3/2}

These feed the carrier statistics. e carrier statistics.


10.6.2 Intrinsic carriers and the mass-action law

With Fermi level EFE_{F}, nondegenerate (Boltzmann) populations are

n=Nce(Ec0EF)/kBT,p=Nve(EFEv0)/kBTn = N_{c}\,e^{-(E_{c0}-E_{F})/k_{B}T},\qquad p = N_{v}\,e^{-(E_{F}-E_{v0})/k_{B}T}

Multiply to get the mass-action law

np=ni2,ni(T)=NcNv  eEg/(2kBT)T3/2eEg/(2kBT)np = n_{i}^{2},\qquad n_{i}(T) = \sqrt{N_{c}N_{v}}\;e^{-E_{g}/(2k_{B}T)} \propto T^{3/2}\,e^{-E_{g}/(2k_{B}T)}

In an intrinsic crystal n=p=nin=p=n_{i} and EFE_{F} sits near midgap skewed by DOS masses.


10.6.3 Doping: donors, acceptors, and charge neutrality

Substitutional dopants add shallow hydrogenic levels. Effective-mass theory gives

EDmeme1εr213.6 eV,aDεrmemea0E_{D} \approx \frac{m_{e}^{\ast}}{m_{e}} \frac{1}{\varepsilon_{r}^{2}}\,13.6\ \mathrm{eV},\qquad a_{D}^{\ast} \approx \varepsilon_{r}\,\frac{m_{e}}{m_{e}^{\ast}}\,a_{0}

Similarly for acceptors with mhm_{h}^{\ast}. At room TT, shallow dopants are usually ionized.

For an n-type crystal with donor density NDN_{D} and acceptors NAN_{A},

np+NAND+=0n - p + N_{A}^{-} - N_{D}^{+} = 0

Assuming full ionization in the nondegenerate regime, ND+NDN_{D}^{+}\approx N_{D} and NANAN_{A}^{-}\approx N_{A}. Then

n12[(NDNA)+(NDNA)2+4ni2],p=ni2nn \approx \frac{1}{2}\left[(N_{D}-N_{A}) + \sqrt{(N_{D}-N_{A})^{2} + 4 n_{i}^{2}}\right],\qquad p = \frac{n_{i}^{2}}{n}

In the majority-carrier limit NDniN_{D}\gg n_{i}, nNDNAn\approx N_{D}-N_{A} and pni2/np\approx n_{i}^{2}/n.

At low TT, partial ionization leads to a freeze-out region where n<NDn<N_{D}; use Fermi–Dirac occupancy of dopant levels.


10.6.4 Mobilities, scattering, and conductivity

Each carrier has mobility μ\mu with drift velocity vd=±μE\boldsymbol v_{d}=\pm \mu \boldsymbol E. Conductivity is

σ=q(nμn+pμp),ρ=1σ\sigma = q\,(n\mu_{n} + p\mu_{p}),\qquad \rho = \frac{1}{\sigma}

Scattering sources:

  • Lattice (phonons): typically μT3/2\mu \propto T^{-3/2} in 3D
  • Ionized impurities: μT+3/2\mu \propto T^{+3/2} at low TT
  • Matthiessen’s rule for rates: μ1μph1+μimp1+\mu^{-1}\approx \mu_{\mathrm{ph}}^{-1}+\mu_{\mathrm{imp}}^{-1}+\cdots

Two-carrier Hall coefficient (weak field) shows compensation:

RH=pμp2nμn2q(nμn+pμp)2R_{H} = \frac{p\mu_{p}^{2} - n\mu_{n}^{2}}{q\,(n\mu_{n}+p\mu_{p})^{2}}

Sign reveals the dominant carrier type.


10.6.5 Drift–diffusion and Einstein relation

Carrier continuity with drift and diffusion reads

Jn=qnμnE+qDnn,Jp=qpμpEqDpp\boldsymbol J_{n} = q\,n\mu_{n}\boldsymbol E + q D_{n}\nabla n,\qquad \boldsymbol J_{p} = q\,p\mu_{p}\boldsymbol E - q D_{p}\nabla p

In nondegenerate statistics,

Dn=kBTqμn,Dp=kBTqμpD_{n} = \frac{k_{B}T}{q}\,\mu_{n},\qquad D_{p} = \frac{k_{B}T}{q}\,\mu_{p}

This Einstein relation is the semiconductor face of fluctuation–dissipation.


10.6.6 Recombination, lifetime, and quasi-Fermi levels

Out of equilibrium (illumination, injection), electrons and holes can be described by quasi-Fermi levels EFn,EFpE_{Fn},E_{Fp} with

n=Nce(Ec0EFn)/kBT,p=Nve(EFpEv0)/kBTn = N_{c}\,e^{-(E_{c0}-E_{Fn})/k_{B}T},\qquad p = N_{v}\,e^{-(E_{Fp}-E_{v0})/k_{B}T}

Net recombination–generation channels:

  • Radiative (band-to-band): Rrad=B(npni2)R_{\mathrm{rad}} = B\,(np - n_{i}^{2})
  • Shockley–Read–Hall (trap-mediated): roughly RSRHnpni2τp(n+n1)+τn(p+p1)R_{\mathrm{SRH}} \approx \frac{np - n_{i}^{2}}{\tau_{p}(n+n_{1}) + \tau_{n}(p+p_{1})} with trap parameters n1,p1n_{1},p_{1}
  • Auger: RAuger=Cnn(npni2)+Cpp(npni2)R_{\mathrm{Auger}} = C_{n} n (np-n_{i}^{2}) + C_{p} p (np-n_{i}^{2})

Define minority-carrier lifetime τ\tau via small-signal Rδn/τR\approx \delta n/\tau. Lifetimes set diffusion lengths L=DτL=\sqrt{D\tau} that control photodiode and solar-cell collection.


10.6.7 Direct vs indirect gaps, absorption, and excitons

Optical absorption coefficient near the band edge:

  • Direct gap: α(ω)ωEg\alpha(\hbar\omega)\propto \sqrt{\hbar\omega - E_{g}}
  • Indirect gap: phonon-assisted, α\alpha rises slowly and shows temperature dependence

Binding of electron–hole pairs gives excitons with hydrogenic series. Effective-mass theory estimates

EXμme1εr213.6 eV,aXεrmeμa0E_{X} \approx \frac{\mu^{\ast}}{m_{e}}\,\frac{1}{\varepsilon_{r}^{2}}\,13.6\ \mathrm{eV},\qquad a_{X}^{\ast} \approx \varepsilon_{r}\,\frac{m_{e}}{\mu^{\ast}}\,a_{0}

with reduced mass μ=(memh)/(me+mh)\mu^{\ast}=(m_{e}^{\ast} m_{h}^{\ast})/(m_{e}^{\ast}+m_{h}^{\ast}). Excitonic peaks decorate the absorption edge in clean crystals and 2D materials.


10.6.8 Minimal pn junction: built-in potential, depletion, and diode law

Bring p and n regions together; carriers diffuse, leaving behind ionized dopants and forming a depletion region with built-in field. At equilibrium,

Vbi=kBTqln ⁣(NANDni2)V_{\mathrm{bi}} = \frac{k_{B}T}{q}\,\ln\!\left(\frac{N_{A} N_{D}}{n_{i}^{2}}\right)

Ignoring free charge in depletion (abrupt junction), the total depletion width under applied bias VV is

W=2εsqNA+NDNAND(VbiV)W = \sqrt{\frac{2\varepsilon_{s}}{q}\,\frac{N_{A}+N_{D}}{N_{A}N_{D}}\,(V_{\mathrm{bi}} - V)}

Forward bias reduces the barrier; the ideal Shockley diode equation:

I=I0(eqV/(kBT)1)I = I_{0}\left(e^{qV/(k_{B}T)} - 1\right)

where I0ni2I_{0}\propto n_{i}^{2} scales inversely with diffusion lengths and doping. Reverse bias nearly clamps at I0-I_{0} until breakdown mechanisms kick in (Zener/tunneling or avalanche).


10.6.9 Metal–semiconductor contacts: ohmic vs Schottky in one breath

A metal on a lightly doped semiconductor forms either:

  • Ohmic contact if the barrier is thin/low enough for tunneling or if heavy doping bends bands into easy injection
  • Schottky barrier with rectification otherwise. Barrier height is set by metal work function, semiconductor electron affinity, and interface pinning

Transport across a Schottky contact at forward bias often follows thermionic emission with IT2eqΦB/kBT(eqV/kBT1)I\propto T^{2} e^{-q\Phi_{B}/k_{B}T}\,(e^{qV/k_{B}T}-1).


10.6.10 Temperature regimes (the three-act play)

For a doped semiconductor as TT rises:

  1. Freeze-out: dopants partially ionized, n<predictably<NDn<predictably< N_{D}
  2. Extrinsic: plateau where nNDn\approx N_{D} (or pNAp\approx N_{A}) and npni2np\approx n_{i}^{2}
  3. Intrinsic: npnin\approx p\approx n_{i} once nin_{i} surpasses dopant densities exponentially

Resistivity and Hall data vs TT let you read off activation energies and dominant scattering.


10.6.11 Worked mini-examples

(a) Intrinsic carrier density
Given me,DOS=0.26mem_{e}^{\ast,\mathrm{DOS}}=0.26\,m_{e}, mh,DOS=0.39mem_{h}^{\ast,\mathrm{DOS}}=0.39\,m_{e}, Eg=1.12 eVE_{g}=1.12\ \mathrm{eV} at 300 K300\ \mathrm{K}, estimate nin_{i} for Si using the T3/2eEg/2kBTT^{3/2} e^{-E_{g}/2k_{B}T} form.

(b) Fermi level in doped Si
For ND=1016 cm3N_{D}=10^{16}\ \mathrm{cm^{-3}} and ni=1010 cm3n_{i}=10^{10}\ \mathrm{cm^{-3}} at 300 K300\ \mathrm{K}, find n,pn,p and locate EFE_{F} relative to Ec0E_{c0} via n=Nce(Ec0EF)/kBTn=N_{c} e^{-(E_{c0}-E_{F})/k_{B}T}.

(c) Einstein relation check
With measured μn=1350 cm2/Vs\mu_{n}=1350\ \mathrm{cm^{2}/Vs} in Si at 300 K300\ \mathrm{K}, compute DnD_{n} and a diffusion length for τn=1 μs\tau_{n}=1\ \mu\mathrm{s}.

(d) Depletion width
For a Si pn junction with NA=1017N_{A}=10^{17}, ND=1016 cm3N_{D}=10^{16}\ \mathrm{cm^{-3}}, εs=11.7ε0\varepsilon_{s}=11.7\,\varepsilon_{0}, compute WW at V=0V=0 and at V=0.6 VV=0.6\ \mathrm{V}.

(e) Radiative vs SRH
Given B=1010 cm3/sB=10^{-10}\ \mathrm{cm^{3}/s}, n=p=1015 cm3n=p=10^{15}\ \mathrm{cm^{-3}}, estimate RradR_{\mathrm{rad}} and compare to an SRH rate with τ=1 μs\tau=1\ \mu\mathrm{s}.


10.6.12 Common pitfalls

  • Forgetting mass-action: in equilibrium np=ni2np=n_{i}^{2} even in doped material; violations signal injection or illumination
  • Using Einstein beyond its lane: standard D=μkBT/qD=\mu k_{B}T/q needs nondegenerate statistics; degenerate semiconductors require Fermi–Dirac corrections
  • Assuming full ionization: at low TT or for deeper dopants, solve dopant occupancy self-consistently
  • Mixing DOS and conductivity masses: Nc,v(T)N_{c,v}(T) want DOS masses; transport needs conductivity (tensor) masses and scattering
  • Ideal diode dogma: real diodes have series resistance, recombination in depletion (ideality factor nid>1n_{\mathrm{id}}>1), and interface states

10.7 Fermi Surfaces & Transport

The Fermi surface (FS) is the VIP section of a metal: at low temperature almost everything transporty—conductivity, Hall, magnetoresistance, quantum oscillations—comes from quasiparticles living right on it. Geometry in kk-space becomes measurable physics in the lab. This section builds the semiclassical equations of motion, the Boltzmann/Chambers machinery, magnetotransport including multiband and compensation, and the quantum‐oscillation toolkit (Onsager + Lifshitz–Kosevich). We’ll also flag how FS topology (closed vs open orbits) controls magnetoresponse.


10.7.1 Semiclassical equations of motion (no Berry yet)

For a band ϵ(k)\epsilon(\boldsymbol k) in fields (E,B)(\boldsymbol E,\boldsymbol B) the semiclassical dynamics are

k˙=e(E+r˙×B),r˙=1kϵ(k)\hbar\,\dot{\boldsymbol k} = -e\left(\boldsymbol E + \dot{\boldsymbol r}\times \boldsymbol B\right),\qquad \dot{\boldsymbol r} = \frac{1}{\hbar}\,\nabla_{\boldsymbol k}\epsilon(\boldsymbol k)

With E=0\boldsymbol E=0 you get cyclotron motion in kk-space along constant-energy contours perpendicular to B\boldsymbol B; real-space velocity is v=r˙\boldsymbol v=\dot{\boldsymbol r}. We postpone Berry curvature and anomalous velocity to §10.11.


10.7.2 Cyclotron frequency and mass

Closed orbits have frequency

ωc=eBmc\omega_{c} = \frac{eB}{m_{c}}

where the cyclotron mass is set by the extremal kk-space orbit area A(E)A(E) normal to B\boldsymbol B:

mc(E)=22πA(E)EE=ϵFm_{c}(E) = \frac{\hbar^{2}}{2\pi}\,\frac{\partial A(E)}{\partial E}\Big|_{E=\epsilon_{F}}

For an isotropic parabola E=2k2/2mE=\hbar^{2}k^{2}/2m^{\ast}, mc=mm_{c}=m^{\ast}; for anisotropic ellipsoids, mcm_{c} depends on field orientation.


10.7.3 Boltzmann transport: relaxation time and Chambers

Linearizing the Boltzmann equation in a steady field with relaxation time τ(k)\tau(\boldsymbol k) gives the current

J=ed3k4π3v(k)δf(k),δf=(f0ϵ)ΛE\boldsymbol J = -e \int \frac{d^{3}k}{4\pi^{3}}\,\boldsymbol v(\boldsymbol k)\,\delta f(\boldsymbol k),\qquad \delta f = -\left(-\frac{\partial f_{0}}{\partial \epsilon}\right)\,\boldsymbol{\Lambda}\cdot \boldsymbol E

In Chambers’ formula the vector mean free path is a history average along the orbit

Λ(k)=0dt  et/τ(kt)v(kt)\boldsymbol{\Lambda}(\boldsymbol k) = \int_{0}^{\infty} dt\; e^{-t/\tau(\boldsymbol k_{-t})}\,\boldsymbol v(\boldsymbol k_{-t})

leading to the conductivity tensor (schematically)

σαβ=e24π3FSdSvvαvˉβ\sigma_{\alpha\beta} = \frac{e^{2}}{4\pi^{3}}\int_{\mathrm{FS}} \frac{dS}{\hbar |\boldsymbol v|}\, v_{\alpha}\,\bar v_{\beta}

where vˉβ\bar v_{\beta} is the orbit-averaged, decay-weighted velocity. For an isotropic, single-τ\tau metal this collapses to Drude.


10.7.4 Hall angle, mobility tensor, mean free path

Define mobility μ=τm1e\boldsymbol{\mu}=\tau\,\boldsymbol m^{\ast -1}\,e and Hall angle tanθH=σxy/σxxωcτ\tan\theta_{H}=\sigma_{xy}/\sigma_{xx}\approx \omega_{c}\tau (single band). The mean free path

=vFτ,kF1 for good quasiparticles\ell = v_{F}\,\tau,\qquad k_{F}\ell \gg 1 \ \text{for good quasiparticles}

Close to the Ioffe–Regel limit kF1k_{F}\ell\sim 1 metals “bad” out, and semiclassical transport starts to crumble.


10.7.5 Magnetoresistance: why single-band Drude lies flat

In a single, isotropic band with constant τ\tau, the resistivity ρxx\rho_{xx} is field-independent even though σxx\sigma_{xx} shrinks; tensor inversion cancels it. Real materials show MR because of (i) multiple carrier types, (ii) FS anisotropy/warping, (iii) angle-dependent τ(k)\tau(\boldsymbol k), or (iv) open orbits. A practical summary is Kohler’s rule

Δρρ0ρ(B)ρ0ρ0=F ⁣(Bρ0)F(ωcτ)\frac{\Delta \rho}{\rho_{0}} \equiv \frac{\rho(B)-\rho_{0}}{\rho_{0}} = F\!\left(\frac{B}{\rho_{0}}\right) \approx F(\omega_{c}\tau)

If a single scattering time controls everything, curves at different TT collapse when plotted vs B/ρ0(T)B/\rho_{0}(T).


10.7.6 Two-band model (electrons + holes)

Let (ne,μe)(n_{e},\mu_{e}) and (nh,μh)(n_{h},\mu_{h}) be densities and mobilities; se=1s_{e}=-1, sh=+1s_{h}=+1 charge signs. The conductivities sum:

σxx=i=e,hnieμi1+(μiB)2,σxy=i=e,hsinieμi2B1+(μiB)2\sigma_{xx} = \sum_{i=e,h} \frac{n_{i} e \mu_{i}}{1+(\mu_{i}B)^{2}},\qquad \sigma_{xy} = \sum_{i=e,h} \frac{s_{i} n_{i} e \mu_{i}^{2} B}{1+(\mu_{i}B)^{2}}

Invert to get ρxx,ρxy\rho_{xx},\rho_{xy}. Key regimes:

  • Compensation nenhn_{e}\approx n_{h} with high mobilities ⇒ large, nearly quadratic unsaturating MR
  • Dominant carrier nenhn_{e}\gg n_{h} ⇒ Hall sign fixed, MR smaller and tends to saturate

Multiband fits to ρxx(B)\rho_{xx}(B) and ρxy(B)\rho_{xy}(B) are the day-one way to read carrier types and mobilities.


10.7.7 Open vs closed orbits; angle dependence

  • Closed orbits on closed FS pockets give conventional ωc\omega_{c} and oscillations.
  • Open orbits (e.g., warped quasi-1D sheets) allow carriers to stream along a direction; MR becomes strongly anisotropic and can grow linearly or quadratically without saturation depending on geometry.
  • AMRO (angle-dependent MR oscillations) in layered metals arise when tilting B\boldsymbol B threads Landau orbits across a warped quasi-2D FS; minima occur at Yamaji angles set by interlayer warping.

10.7.8 Quantum oscillations: Onsager + Lifshitz–Kosevich

Landau quantization turns orbit areas into discrete levels. The Onsager relation ties the oscillation frequency FF (in 1/B1/B) to the extremal FS cross-section area AFA_{F} perpendicular to B\boldsymbol B:

F=2πeAFF = \frac{\hbar}{2\pi e}\,A_{F}

The Lifshitz–Kosevich (LK) amplitude has three classic factors:

ΔM, Δρ  RTRDRScos ⁣(2πFBγ)\Delta M,\ \Delta \rho \ \propto\ R_{T}\,R_{D}\,R_{S}\,\cos\!\left(2\pi\frac{F}{B} - \gamma\right)

with

RT=XsinhX,X=2π2kBTωcR_{T} = \frac{X}{\sinh X},\quad X=\frac{2\pi^{2} k_{B} T}{\hbar \omega_{c}} RD=exp ⁣(πωcτ)=exp ⁣(2π2kBTDωc)R_{D} = \exp\!\left(-\frac{\pi}{\omega_{c}\tau}\right) = \exp\!\left(-\frac{2\pi^{2} k_{B} T_{D}}{\hbar \omega_{c}}\right) RS=cos ⁣(πgmc2me)R_{S} = \cos\!\left(\frac{\pi g m_{c}}{2 m_{e}}\right)

RTR_{T} gives the mass plot to read mcm_{c}, RDR_{D} the Dingle factor for scattering, and RSR_{S} the spin-splitting modulation. The phase γ\gamma can encode Berry phase for Dirac/Weyl systems (see §10.11).


10.7.9 Cyclotron resonance and optical probes

In clean samples, an AC field picks up resonances at ω=nωc\omega=n\,\omega_{c} giving mcm_{c} directly. Optical conductivity decomposes into a Drude piece and interband contributions; the f-sum rule sets the total spectral weight (see §10.3). In layered conductors, interlayer conductivity vs tilt reveals FS warping.


10.7.10 Practical FS metrology (ARPES vs oscillations)

  • ARPES maps occupied bands and FS contours directly in kk_{\parallel}; great for 2D and surfaces, trickier for bulk kzk_{z}.
  • Quantum oscillations see only extremal areas but with exquisite precision and bulk sensitivity.
  • Quantum oscillation angle sweeps reconstruct full FS shapes from F(θ)F(\theta); combine with ARPES to de-ambiguate pockets and 3D warping.

10.7.11 Thermoelectric and Nernst hints

Near the FS, the Mott formula relates the Seebeck coefficient to the energy derivative of conductivity at ϵF\epsilon_{F}:

S=π23kB2Tedlnσ(ϵ)dϵϵFS = -\frac{\pi^{2}}{3}\,\frac{k_{B}^{2}T}{e}\left.\frac{d\ln \sigma(\epsilon)}{d\epsilon}\right|_{\epsilon_{F}}

The Nernst effect (transverse voltage from a thermal gradient in BB) is extremely sensitive to FS curvature and quasiparticle lifetime asymmetries—often a canary for small pockets.


10.7.12 Strong-field and quantum-limit quirks

At very large BB, carriers collapse into a few lowest Landau levels. In the quantum limit (only n=0n=0 occupied), MR and thermoelectrics can behave non-Drude, sometimes linear in BB or with sign flips. Dirac/Weyl materials can show a chiral-anomaly contribution to longitudinal MR when EB\boldsymbol E\parallel \boldsymbol B (details in §10.11).


10.7.13 Worked mini-examples

(a) Onsager area from frequency
You measure F=1.0 kTF=1.0\ \mathrm{kT} for Bz^\mathbf B\parallel \hat z. Compute the extremal area

AF=2πeF0.095 A˚2A_{F} = \frac{2\pi e}{\hbar} F \approx 0.095\ \mathrm{\AA^{-2}}

and an equivalent circular Fermi wavevector kF=AF/πk_{F}=\sqrt{A_{F}/\pi}.

(b) LK mass plot
Given oscillation amplitudes vs TT at fixed BB, fit RT=X/sinhXR_{T}=X/\sinh X to extract mcm_{c} from X=(2π2kBTmc)/(eB)X=(2\pi^{2}k_{B}T m_{c})/(e\hbar B).

(c) Two-band MR
Take ne=nh=5×1025 m3n_{e}=n_{h}=5\times 10^{25}\ \mathrm{m^{-3}}, μe=1.5 m2/Vs\mu_{e}=1.5\ \mathrm{m^{2}/Vs}, μh=0.8 m2/Vs\mu_{h}=0.8\ \mathrm{m^{2}/Vs}. Evaluate ρxx(B)\rho_{xx}(B) and show quadratic growth up to several tesla without saturation.

(d) AMRO/Yamaji
For a quasi-2D FS with interlayer dispersion 2tcoskzc-2t_{\perp}\cos k_{z}c, show that MR minima occur when tanθπν/(kFc)\tan\theta \approx \pi \nu/(k_{F} c) with integer ν\nu, giving kFk_{F} from the angular period.

(e) Ioffe–Regel check
With kF=0.8 A˚1k_{F}=0.8\ \mathrm{\AA^{-1}}, vF=1.5×105 m/sv_{F}=1.5\times 10^{5}\ \mathrm{m/s}, and ρ0=150 μΩ cm\rho_{0}=150\ \mu\Omega\ \mathrm{cm} in a single band, estimate τ=m/(ne2ρ0)\tau=m^{\ast}/(ne^{2}\rho_{0}), compute =vFτ\ell=v_{F}\tau, and comment on kFk_{F}\ell.


10.7.14 Common pitfalls

  • Mixing masses: mm^{\ast} (band curvature), mcm_{c} (cyclotron), and optical/thermodynamic masses differ in anisotropic or multiband systems
  • Forgetting tensor inversion: trends in σxx\sigma_{xx} don’t translate 1:1 to ρxx\rho_{xx}; always invert the full tensor
  • Assuming single τ\tau: hot/cold spots and anisotropic scattering wreck Kohler collapse and mimic multiband MR
  • Misreading oscillation phase: Berry phase extraction needs careful background subtraction and index conventions; spin zeros can shift apparent phase
  • Ignoring geometry: sample misalignment and current jetting can fake huge linear MR in high-mobility, compensated crystals
  • Open-orbit blinders: seeing “saturating MR” and blaming impurities when the real culprit is FS topology and field orientation

10.8 Magnetism

Magnetism comes in two big flavors: localized moments that behave like tiny quantum spins coupled by exchange, and itinerant magnetism where the Fermi sea itself polarizes. From Curie–Weiss laws to spin waves and ferromagnetic resonance, this section builds the standard playbook: Heisenberg models and exchange mechanisms (direct, superexchange, double exchange, RKKY), Stoner ferromagnets, magnons in ferro/antiferromagnets, anisotropy and domains, and the Mermin–Wagner no-go in 1D/2D without symmetry breaking fields. We keep the math tight and the takeaways actionable.


10.8.1 Magnetic moments, gg factors, and units

An electron carries spin S\boldsymbol S and possibly orbital L\boldsymbol L contributions. The magnetic moment operator is

μ=gSμBS/gLμBL/\boldsymbol{\mu} = -g_{S}\mu_{B}\,\boldsymbol S/\hbar - g_{L}\mu_{B}\,\boldsymbol L/\hbar

with Bohr magneton μB=e/(2me)\mu_{B}=e\hbar/(2m_{e}) and gS2g_{S}\approx 2. In solids, quenching of orbital angular momentum by the crystal field often leaves effective spin moments, sometimes with spin–orbit dressing that yields anisotropy.

The magnetization M\boldsymbol M is moment per volume. Linear response defines susceptibility χ\chi via M=χH\boldsymbol M=\chi \boldsymbol H (isotropic case), with B=μ0(H+M)\boldsymbol B=\mu_{0}(\boldsymbol H+\boldsymbol M).


10.8.2 Local-moment models: Ising/Heisenberg/XXZ

For localized spins {Si}\{\boldsymbol S_{i}\} on a lattice the minimal Hamiltonians are

  • Ising
H=ijJijSizSjzhiSizH = -\sum_{\langle ij\rangle} J_{ij}\,S_{i}^{z} S_{j}^{z} - h\sum_{i} S_{i}^{z}
  • Heisenberg (SU(2)-symmetric)
H=ijJijSiSjgμBBiSizH = -\sum_{\langle ij\rangle} J_{ij}\,\boldsymbol S_{i}\cdot \boldsymbol S_{j} - g\mu_{B} B \sum_{i} S_{i}^{z}
  • XXZ anisotropy
H=ij[J(SixSjx+SiySjy)+JzSizSjz]H = -\sum_{\langle ij\rangle} \left[J_{\perp}\,(S_{i}^{x} S_{j}^{x}+S_{i}^{y} S_{j}^{y}) + J_{z}\,S_{i}^{z} S_{j}^{z}\right]

Here J>0J>0 favors ferromagnetism (FM) and J<0J<0 antiferromagnetism (AFM) in our sign convention.

Curie–Weiss phenomenology for paramagnets: above the ordering temperature,

χ(T)=CTΘCW,C=μ0Ng2μB2S(S+1)3kB\chi(T) = \frac{C}{T-\Theta_{\mathrm{CW}}},\qquad C = \frac{\mu_{0} N g^{2} \mu_{B}^{2} S(S+1)}{3k_{B}}

The sign of ΘCW\Theta_{\mathrm{CW}} hints at dominant FM (>0>0) or AFM (<0<0) exchange.


10.8.3 Exchange mechanisms: where JJ comes from

  • Direct exchange: overlap of neighboring local orbitals; Pauli + Coulomb yield an energy difference between parallel and antiparallel spin alignments.
  • Superexchange: virtual hopping via a nonmagnetic ligand (e.g., TM–O–TM). For a half-filled Hubbard-like situation,
J4t2UJ \sim \frac{4t^{2}}{U}

often antiferromagnetic.

  • Double exchange (mixed valence manganites): itinerant electrons align local core spins to maximize hopping, favoring FM and metallicity.
  • RKKY (in metals with localized moments): conduction electrons mediate an oscillatory coupling
J(r)cos(2kFr)r3J(r) \propto \frac{\cos(2k_{F} r)}{r^{3}}

driving complex order or spin glass when disorder scrambles phases.


10.8.4 Itinerant ferromagnetism: Stoner criterion

In a simple band, exchange lowers energy if spin polarization splits the bands by ±Δ/2\pm \Delta/2. The Stoner mean-field yields a criterion

IN(ϵF)>1I\,N(\epsilon_{F}) > 1

with II an interaction parameter and N(ϵF)N(\epsilon_{F}) the DOS per spin at the Fermi level. Consequences:

  • Finite Pauli susceptibility enhanced to
χ=χ01IN(ϵF)\chi = \frac{\chi_{0}}{1 - I N(\epsilon_{F})}
  • Spin-wave stiffness emerges in the broken-symmetry phase (see §10.8.6)

Itinerant magnets mix band structure with interactions; ARPES, quantum oscillations, and optical sum rules help disentangle.


10.8.5 Mean-field ordering temperatures

On a zz-coordination lattice with exchange JJ between near neighbors, Weiss mean-field predicts

  • FM Heisenberg
kBTCMF=23zJS(S+1)k_{B} T_{C}^{\mathrm{MF}} = \frac{2}{3}\,z J S(S+1)
  • AFM (bipartite)
kBTNMF=23zJS(S+1)k_{B} T_{N}^{\mathrm{MF}} = \frac{2}{3}\,z |J| S(S+1)

Real TC/NT_{C/N} are reduced by fluctuations, especially in low dimensions or small SS.


10.8.6 Spin waves (magnons) in ferromagnets

In a FM Heisenberg model, small transverse oscillations are magnons. Holstein–Primakoff mapping gives, to leading order,

HE0+kωkbkbkH \approx E_{0} + \sum_{\boldsymbol k} \hbar \omega_{\boldsymbol k}\, b_{\boldsymbol k}^{\dagger} b_{\boldsymbol k}

with dispersion near k=0\boldsymbol k=0

ωkDk2+Δani\hbar \omega_{\boldsymbol k} \approx D\,k^{2} + \Delta_{\mathrm{ani}}

Here DD is the spin-wave stiffness set by JJ and lattice geometry, and Δani\Delta_{\mathrm{ani}} comes from anisotropy (otherwise Goldstone gapless). Thermal excitation of magnons reduces the magnetization:

M(T)M(0)[1AT3/2]M(T) \approx M(0)\left[1 - A\,T^{3/2}\right]

This is the Bloch T3/2T^{3/2} law in 3D FMs.

Specific heat of magnons in 3D scales as CmagT3/2C_{\mathrm{mag}}\propto T^{3/2} at low TT (quadratic dispersion → DOS E\propto \sqrt{E}).


10.8.7 Antiferromagnets: two-sublattice spin waves

For a collinear AFM on a bipartite lattice, linear spin-wave theory yields two branches with linear dispersion at small kk:

ωkck\hbar \omega_{\boldsymbol k} \approx c\,|\boldsymbol k|

Velocity cc depends on J|J|, SS, and lattice spacing. Anisotropy produces a small gap. The uniform susceptibility often shows a cusp at TNT_{N}; below, transverse and longitudinal components differ.


10.8.8 Magnetic anisotropy, demagnetizing fields, and domains

  • Magnetocrystalline anisotropy from spin–orbit coupling selects easy axes/planes. Minimal model
Eani=K1sin2θ+K2sin4θ+E_{\mathrm{ani}} = K_{1} \sin^{2}\theta + K_{2} \sin^{4}\theta + \cdots

affects coercivity and domain structure.

  • Demagnetizing fields arise because M\nabla\cdot \boldsymbol M at surfaces acts like “magnetic charge.” Shape defines a demag tensor N\mathsf N with field Hd=NM\boldsymbol H_{d}=-\mathsf N \boldsymbol M.
  • Domains form to balance wall energy and magnetostatic energy. Domain walls have width δA/K\delta \sim \sqrt{A/K} and energy σAK\sigma \sim \sqrt{A K} where AA is exchange stiffness and KK anisotropy.

10.8.9 Ferromagnetic resonance (FMR) and dynamics

Magnetization dynamics obey Landau–Lifshitz–Gilbert (LLG)

dMdt=γM×Heff+αMsM×dMdt\frac{d\boldsymbol M}{dt} = -\gamma\,\boldsymbol M\times \boldsymbol H_{\mathrm{eff}} + \frac{\alpha}{M_{s}}\,\boldsymbol M\times \frac{d\boldsymbol M}{dt}

with gyromagnetic ratio γ\gamma and damping α\alpha. For a uniform mode in an in-plane thin film (Kittel formula, simple case),

ω2=γ2(H+Hk)(H+Hk+4πMs)\omega^{2} = \gamma^{2}\,(H+H_{k})(H+H_{k}+4\pi M_{s})

FMR reads out anisotropy fields HkH_{k} and damping.


10.8.10 Dimensionality: Mermin–Wagner and Kosterlitz–Thouless

Continuous-symmetry magnets in strictly 1D/2D with short-range exchange cannot order at finite TT (no gap, Goldstone modes kill long-range order): the Mermin–Wagner theorem. Workarounds:

  • Add anisotropy (Ising-like discretization), long-range dipolar terms, or interlayer coupling
  • In 2D XY, expect a Kosterlitz–Thouless transition with vortex unbinding and algebraic order rather than conventional TCT_{C}

10.8.11 Magnetotransport cliffs notes

  • Anomalous Hall effect (AHE) in FMs: ρxy=R0B+RsM\rho_{xy}=R_{0}B + R_{s} M, with intrinsic Berry-curvature and extrinsic skew/side-jump mechanisms.
  • Spin waves and resistivity: additional T2T^{2} or T3/2T^{3/2} contributions at low TT in FMs from magnon scattering.
  • Spin caloritronics: spin Seebeck and magnon-drag signals link magnon and charge/heat transport.

10.8.12 Worked mini-examples

(a) Curie–Weiss slope
Given S=5/2S=5/2, g=2g=2, number density NN, show that the high-TT slope of 1/χ(T)1/\chi(T) is kB/(C)k_{B}/(C) with C=μ0Ng2μB2S(S+1)/3kBC=\mu_{0} N g^{2}\mu_{B}^{2} S(S+1)/3k_{B}. Fit data to extract NN or SS.

(b) Spin-wave stiffness in a cubic FM
Nearest-neighbor Heisenberg on cubic lattice (spacing aa): show

ωk=2JSz(1γk)\hbar \omega_{\mathbf k} = 2 J S z \left(1-\gamma_{\mathbf k}\right)

with z=6z=6 and γk=13[cos(kxa)+cos(kya)+cos(kza)]\gamma_{\mathbf k}=\tfrac{1}{3}[\cos(k_{x}a)+\cos(k_{y}a)+\cos(k_{z}a)]. Expand for small kk to extract DD.

(c) AFM two-sublattice dispersion
For a bipartite AFM with nn exchange J<0J<0, derive the linear small-kk magnon velocity c=a2JSz/2c=a\sqrt{2|J|S z/\hbar^{2}} up to geometry factors.

(d) Stoner enhancement
Given χ/χ0=5\chi/\chi_{0}=5, estimate IN(ϵF)I N(\epsilon_{F}) and comment on proximity to instability =1=1.

(e) Domain wall width
With A=1×1011 J/mA=1\times 10^{-11}\ \mathrm{J/m} and K=5×104 J/m3K=5\times 10^{4}\ \mathrm{J/m^{3}}, estimate δ=A/K\delta=\sqrt{A/K} and σAK\sigma\sim\sqrt{A K}.

(f) RKKY sign changes
For kF=1.2 A˚1k_{F}=1.2\ \mathrm{\AA^{-1}}, compute the first few zeros/sign flips of cos(2kFr)\cos(2k_{F} r) to see at what separations the coupling toggles FM/AFM.


10.8.13 Common pitfalls

  • Sign conventions for JJ: check whether your Hamiltonian uses JSiSj-J \boldsymbol S_{i}\cdot \boldsymbol S_{j} or +J+J; FM/AFM flip with the choice.
  • Curie–Weiss overreach: it is a high-TT law; near and below TC/NT_{C/N} critical fluctuations break mean-field straightness.
  • Ignoring anisotropy: in 2D, no finite-TT order without anisotropy or long-range terms; tiny KK matters a lot.
  • Confusing Pauli and Curie: itinerant paramagnets (Pauli) have TT-independent χ\chi; local moments (Curie) go like 1/T1/T. Real compounds can have both.
  • Mixing demag and intrinsic: raw magnetization loops are geometry-dependent; correct for demagnetization to get intrinsic M(H)M(H).
  • Magnon gap blind spot: any observed finite-frequency intercept in FM/AFM spin-wave spectra signals anisotropy or Zeeman gaps—don’t fit as purely gapless.

10.8.14 Minimal problem kit

  • Derive Curie–Weiss susceptibility from a Weiss field Heff=λMH_{\mathrm{eff}}=\lambda M for spins SS and connect ΘCW\Theta_{\mathrm{CW}} to λ\lambda and microscopic JJ
  • Compute TCMFT_{C}^{\mathrm{MF}} and M(T)M(T) within mean-field for a cubic Heisenberg FM; compare to Bloch T3/2T^{3/2} at low TT and discuss regimes of validity
  • Do linear spin-wave theory (Holstein–Primakoff) for a 1D FM chain, obtain ωk\omega_{k} and show why 1D FM lacks finite-TT long-range order (infrared divergence)
  • Starting from LLG, derive the Kittel formula for a thin film with uniaxial anisotropy and identify Hk=2K/μ0MsH_{k}=2K/\mu_{0} M_{s}
  • For a simple parabolic band with Stoner parameter II, compute the magnetization mm vs TT at mean-field and sketch the phase diagram in (IN(ϵF),T)(I N(\epsilon_{F}),T) space

10.9 Superconductivity

Superconductors are the drama queens of condensed matter: below a critical temperature they drop their DC resistance to (effectively) zero and expel magnetic fields (Meissner effect). Microscopically, electrons form Cooper pairs that condense into a phase-coherent quantum state. This section builds the phenomenology (London, Meissner), the mesoscopic field theory (Ginzburg–Landau), the microscopic BCS results (gap, TcT_{c}, DOS), and the real-world signatures: Type I vs II, vortices, critical fields/currents, and Josephson effects.


10.9.1 London equations and the Meissner effect

Phenomenology first. The London equations for the supercurrent Js\boldsymbol J_{s} are

Jst=nse2mE\frac{\partial \boldsymbol J_{s}}{\partial t} = \frac{n_{s} e^{2}}{m}\,\boldsymbol E Js=1μ0λ2A,λ2mμ0nse2\boldsymbol J_{s} = -\frac{1}{\mu_{0}\lambda^{2}}\,\boldsymbol A,\qquad \lambda^{2} \equiv \frac{m}{\mu_{0} n_{s} e^{2}}

Combine with Maxwell to eliminate Js\boldsymbol J_{s} and get magnetic field screening

2B=Bλ2\nabla^{2}\boldsymbol B = \frac{\boldsymbol B}{\lambda^{2}}

so a static field decays inside the superconductor as B(x)=B0ex/λB(x)=B_{0} e^{-x/\lambda}. This Meissner effect distinguishes superconductors from ideal (non-superconducting) perfect conductors, which would merely freeze whatever BB they had when cooled.


10.9.2 Ginzburg–Landau (GL) theory

GL is a Landau free energy for a complex order parameter ψ(r)\psi(\boldsymbol r) (Cooper-pair wavefunction) minimally coupled to the vector potential:

F[ψ,A]=αψ2+β2ψ4+12m(i2eA)ψ2+B22μ0\mathcal F[\psi,\boldsymbol A] = \alpha |\psi|^{2} + \frac{\beta}{2} |\psi|^{4} + \frac{1}{2m^{\ast}} \left| \left(-i\hbar \nabla - 2e\,\boldsymbol A\right)\psi \right|^{2} + \frac{|\boldsymbol B|^{2}}{2\mu_{0}}

Minimizing yields the GL equations

αψ+βψ2ψ+12m(i2eA)2ψ=0\alpha \psi + \beta |\psi|^{2}\psi + \frac{1}{2m^{\ast}}\left(-i\hbar \nabla - 2e\boldsymbol A\right)^{2}\psi = 0 Js=2emRe{ψ(i2eA)ψ}\boldsymbol J_{s} = \frac{2e}{m^{\ast}}\,\mathrm{Re}\left\{\psi^{\ast}\left(-i\hbar \nabla - 2e\boldsymbol A\right)\psi\right\}

Key GL lengths:

ξ=22mα,λ=m4μ0e2ψ2\xi = \sqrt{\frac{\hbar^{2}}{2 m^{\ast} |\alpha|}},\qquad \lambda = \sqrt{\frac{m^{\ast}}{4 \mu_{0} e^{2} |\psi|^{2}}}

The dimensionless GL parameter κλ/ξ\kappa \equiv \lambda/\xi classifies superconductors:

  • Type I if κ<1/2\kappa < 1/\sqrt{2} (single critical field HcH_{c}, complete Meissner until abrupt breakdown)
  • Type II if κ>1/2\kappa > 1/\sqrt{2} (mixed state with vortices between Hc1H_{c1} and Hc2H_{c2})

10.9.3 BCS essentials: pairing, gap, and TcT_{c}

BCS shows that any weak, retarded attraction between electrons near ϵF\epsilon_{F} (e.g., phonon-mediated) causes a Cooper instability and a paired ground state with an energy gap Δ\Delta.

Zero-temperature gap and critical temperature (isotropic ss-wave, weak coupling):

Δ(0)=2ωDe1/[VN(0)]\Delta(0) = 2\hbar \omega_{D}\,e^{-1/[V N(0)]} kBTc=1.14ωDe1/[VN(0)]k_{B} T_{c} = 1.14\,\hbar \omega_{D}\,e^{-1/[V N(0)]}

Universal BCS ratio:

2Δ(0)kBTc3.53\frac{2\Delta(0)}{k_{B} T_{c}} \approx 3.53

The quasiparticle density of states (per spin) becomes

Ns(E)=N(0)ReEE2Δ2N_{s}(E) = N(0)\,\mathrm{Re}\,\frac{|E|}{\sqrt{E^{2}-\Delta^{2}}}

with coherence peaks at E=±ΔE=\pm \Delta. The specific heat jumps by ΔC/CnTc=1.43\Delta C/C_{n}|_{T_{c}}=1.43; the low-TT behavior is activated CeΔ/TC\sim e^{-\Delta/T} for fully gapped ss-wave.

BCS coherence length (clean limit):

ξ0=vFπΔ(0)\xi_{0} = \frac{\hbar v_{F}}{\pi \Delta(0)}

10.9.4 Critical fields, currents, and the mixed state

Type I has a thermodynamic critical field Hc(T)H_{c}(T) that satisfies Fs(Hc)=Fn(0)F_{s}(H_{c})=F_{n}(0).

Type II features three scales:

μ0Hc1Φ04πλ2lnκ\mu_{0} H_{c1} \approx \frac{\Phi_{0}}{4\pi \lambda^{2}} \ln \kappa μ0Hc2=Φ02πξ2\mu_{0} H_{c2} = \frac{\Phi_{0}}{2\pi \xi^{2}} Φ0=h2e\Phi_{0} = \frac{h}{2e}

For Hc1<H<Hc2H_{c1}<H<H_{c2}, magnetic field penetrates via Abrikosov vortices carrying one flux quantum Φ0\Phi_{0}. Vortex cores have radius ξ\sim \xi where ψ|\psi| is suppressed; circulating supercurrents decay over λ\lambda. Vortices form a triangular lattice in clean materials absent pinning.

Critical current density (GL depairing estimate near TcT_{c}):

JdHcμ0λJ_{d} \sim \frac{H_{c}}{\mu_{0}\lambda}

In practice, vortex motion (flux flow) limits current; strong pinning raises the critical current JcJdJ_{c}\ll J_{d} used in applications.


10.9.5 Flux quantization and phase stiffness

Single-valuedness of the condensate phase φ\varphi around a loop gives

(φ2eA)d=2πn\oint \left(\nabla \varphi - \frac{2e}{\hbar}\boldsymbol A\right)\cdot d\boldsymbol\ell = 2\pi n

which implies flux quantization Φ=nΦ0\Phi=n\Phi_{0} in a multiply connected superconductor. The phase rigidity is the origin of persistent currents and of the Josephson effects.


10.9.6 Josephson effects

Two superconductors separated by a thin barrier (SIS) support a supercurrent without voltage:

  • DC Josephson:
I=IcsinϕI = I_{c}\,\sin\phi

where ϕ\phi is the phase difference and IcI_{c} the critical current. With magnetic field, interference across a junction of width WW gives a Fraunhofer pattern Ic(B)sin(πΦ/Φ0)/(πΦ/Φ0)I_{c}(B)\propto |\sin(\pi \Phi/\Phi_{0})/(\pi \Phi/\Phi_{0})|.

  • AC Josephson: a DC voltage VV makes the phase run ϕ˙=2eV/\dot\phi = 2eV/\hbar, yielding an AC current at
f=2ehVf = \frac{2e}{h}\,V

Microwave irradiation produces Shapiro steps at voltages Vn=nhf/(2e)V_{n}=n\,h f/(2e). Two junctions in a loop form a SQUID with period Φ0\Phi_{0} vs applied flux, enabling ultrasensitive magnetometry.

In SNS junctions, transport below the gap proceeds via Andreev reflection, forming discrete Andreev bound states that carry the supercurrent.


10.9.7 Proximity and mesoscopic effects

A normal metal in good contact with a superconductor inherits pair correlations over a proximity length (dirty limit LϕD/kBTL_{\phi}\sim \sqrt{\hbar D/k_{B}T}). Subgap conductance of an NS interface is enhanced by Andreev reflection, approaching a factor of 2 in the ideal BTK limit at zero bias for a perfectly transparent contact.


10.9.8 Unconventional and high-TcT_{c} notes

Not all superconductors are ss-wave. dd-wave pairing (cuprates) features line nodes, giving power-law thermodynamics (CT2C\sim T^{2} in 3D) and a V-shaped tunneling DOS. pp-wave and other exotic states can host Majorana bound states in topological contexts. In many unconventional cases, the “glue” is not phonons but electron–electron interactions (spin fluctuations), and BCS ratios deviate from 3.53.


10.9.9 Experimental fingerprints

  • Transport: zero resistance below TcT_{c}; finite voltage only when current exceeds JcJ_{c} or vortices move
  • Magnetization: Meissner expulsion, irreversible hysteresis from vortex pinning in Type II
  • Thermodynamics: specific-heat jump at TcT_{c}; condensation energy from integrating C/TC/T
  • Spectroscopy: tunneling/STM coherence peaks at ±Δ\pm \Delta; ARPES gap symmetry; optical conductivity with missing area transferred to the zero-frequency delta (Ferrell–Glover–Tinkham sum rule)
  • Microwave/THz: AC conductivity shows a superfluid delta plus a finite-frequency quasiparticle response
  • Muon/NMR: penetration depth, vortex lattice, and gap symmetry from relaxation rates

10.9.10 Worked mini-examples

(a) Penetration depth from carrier density
Using λ2=m/(μ0nse2)\lambda^{2}=m/(\mu_{0} n_{s} e^{2}), estimate λ\lambda for ns=1028 m3n_{s}=10^{28}\ \mathrm{m^{-3}} and m=mem=m_{e}.

(b) Hc2H_{c2} estimate
A Type II superconductor has ξ=4 nm\xi=4\ \mathrm{nm}. Compute μ0Hc2=Φ0/(2πξ2)\mu_{0}H_{c2}=\Phi_{0}/(2\pi \xi^{2}).

(c) Coherence length from gap and vFv_{F}
With vF=2.0×105 m/sv_{F}=2.0\times 10^{5}\ \mathrm{m/s} and Δ(0)=1.5 meV\Delta(0)=1.5\ \mathrm{meV}, estimate ξ0=vF/(πΔ)\xi_{0}=\hbar v_{F}/(\pi \Delta) and compare to the GL ξ\xi near TcT_{c}.

(d) Josephson frequency
What microwave frequency results from V=1.00 mVV=1.00\ \mathrm{mV} across a junction Using f=(2e/h)Vf=(2e/h)V, show f483.6 GHz/mV×1.00 mVf\approx 483.6\ \mathrm{GHz/mV}\times 1.00\ \mathrm{mV}.

(e) Flux quantization in a ring
A superconducting loop of area A=20 μm2A=20\ \mu\mathrm{m}^{2} shows SQUID oscillations with period ΔB\Delta B. Verify ΔB=Φ0/A\Delta B=\Phi_{0}/A and compute the period.

(f) Depairing current (order of magnitude)
Given μ0Hc=0.2 T\mu_{0}H_{c}=0.2\ \mathrm{T} and λ=120 nm\lambda=120\ \mathrm{nm}, estimate JdHc/(μ0λ)J_{d}\sim H_{c}/(\mu_{0}\lambda) and compare with a measured JcJ_{c} limited by pinning.


10.9.11 Common pitfalls

  • Meissner vs perfect conductor: superconductors expel newly applied fields; perfect conductors merely lock in whatever field is present on cooling
  • Forgetting the 2e2e: Cooper pairs carry charge 2e2e; flux quantum is h/2eh/2e, not h/eh/e
  • Type I/II mix-ups: classification uses κ=λ/ξ\kappa=\lambda/\xi; it is not about “how strong the field is” but interfacial energy sign
  • Ignoring vortices: in Type II, finite resistance under current + field often comes from vortex motion, not a destroyed gap
  • Using clean-limit formulas in dirty samples: ξ\xi and λ\lambda renormalize with impurity scattering; check mean free path ll vs ξ0\xi_{0}
  • Assuming ss-wave: gap symmetry matters; power laws in C(T)C(T), λ(T)\lambda(T), and nodal quasiparticles betray unconventional pairing

10.9.12 Minimal problem kit

  • Starting from the London equations, derive 2B=B/λ2\nabla^{2}\boldsymbol B=\boldsymbol B/\lambda^{2} and show exponential decay of B(x)B(x) in a half-space
  • From GL, compute the surface energy between normal and superconducting regions and reproduce the Type I/II boundary at κ=1/2\kappa=1/\sqrt{2}
  • Show that flux quantization follows from single-valuedness of ψ\psi and derive Φ0=h/2e\Phi_{0}=h/2e
  • Use the weak-coupling BCS gap equation to obtain Δ(0)\Delta(0) and the ratio 2Δ(0)/kBTc2\Delta(0)/k_{B}T_{c}
  • Derive Hc2=Φ0/(2πξ2)H_{c2}=\Phi_{0}/(2\pi \xi^{2}) by linearizing the GL equation in a field and mapping to Landau levels for the order parameter
  • Compute the Fraunhofer Ic(B)I_{c}(B) envelope for a rectangular Josephson junction and extract the junction area from the period

10.10 Strong Correlation & Mott Physics

If weakly interacting electrons are chill, strongly correlated ones are messy drama: localization without band gaps, gigantic effective masses, bad-metal transport, and spectral weight teleporting across eV scales when you tweak doping by a few percent. The mascot is the Mott insulator—an odd creature that should conduct by band theory (half-filled band) but refuses because electrons won’t share a site when on-site repulsion UU is large. This section builds the Hubbard family, superexchange and the ttJJ model, the Mott transition (Brinkman–Rice vs. Slater), spectral weight transfer, multi-orbital/Hund physics, and a DMFT crash course, with experimental fingerprints you can spot at 10 paces.


10.10.1 Hubbard model: the minimalist drama

Put electrons on a lattice, let them hop with amplitude tt, and penalize double occupancy by UU:

H=tij,σ ⁣(ciσcjσ+h.c.)+UininiμiniH = -t \sum_{\langle i j \rangle,\sigma}\!\left(c^{\dagger}_{i\sigma} c_{j\sigma} + \text{h.c.}\right) + U \sum_{i} n_{i\uparrow} n_{i\downarrow} - \mu \sum_{i} n_{i}
  • At half filling and U/W1U/W \ll 1 (bandwidth WW), you get a metal.
  • For U/W1U/W \gg 1, charge motion freezes: each site wants one electron to dodge UU, producing a Mott insulator with local moments.
  • Doping away from half filling introduces mobile carriers in a strongly interacting background → strange transport and, in some materials, high-TcT_{c} superconductivity.

10.10.2 From Hubbard to Heisenberg: superexchange

When UtU\gg t at half filling, virtual hops create an effective antiferromagnetic exchange between neighboring spins:

Heff=JijSiSj,J=4t2UH_{\text{eff}} = J \sum_{\langle i j\rangle} \boldsymbol S_{i}\cdot \boldsymbol S_{j},\qquad J = \frac{4 t^{2}}{U}

This is superexchange. Intuition: parallel spins block exchange by Pauli, antiparallel can virtually double-occupy and return, lowering energy.


10.10.3 ttJJ model: doped Mott playbook

Doping a Mott insulator while keeping UU large gives the constrained model

HtJ=tij,σ ⁣c~iσc~jσ+Jij(SiSj14ninj)H_{tJ} = -t \sum_{\langle i j\rangle,\sigma}\! \tilde c^{\dagger}_{i\sigma} \tilde c_{j\sigma} + J \sum_{\langle i j\rangle} \left(\boldsymbol S_{i}\cdot \boldsymbol S_{j} - \frac{1}{4}\,n_{i} n_{j}\right)

where c~\tilde c forbids double occupancy. The no-double-occupancy constraint is the core difficulty; mean-field, slave boson/fermion, or numerical methods tackle it. In 2D, competition between kinetic energy and spin singlet formation seeds dd-wave pairing tendencies.


10.10.4 Slater vs Mott: why insulators aren’t all the same

Two routes to a gap at half filling:

  • Slater insulator: a weak-coupling antiferromagnet. Long-range order doubles the unit cell, folding bands and opening a mean-field gap m\propto m, which closes above the Néel temperature TNT_{N}.
  • Mott insulator: gap U\sim U from local physics; the material stays insulating without long-range order, even above TNT_{N}.

How to tell in the lab: if the high-TT phase is still insulating with a large activation energy and ARPES shows lower/upper Hubbard bands far from the chemical potential, you’re in Mott-land (e.g., NiO). If the gap collapses right at TNT_{N} with small spectral rearrangement, it’s Slater-ish.


10.10.5 Spectral function anatomy: Hubbard bands + quasiparticle

Define the one-electron spectral function A(k,ω)A(\boldsymbol k,\omega). In a correlated metal near a Mott transition:

  • A narrow quasiparticle (QP) peak of weight ZZ straddles the Fermi level.
  • Incoherent Hubbard bands sit at ω±U/2\omega \sim \pm U/2, carrying weight 1Z1-Z.

The quasiparticle weight is

Z=[1ReΣ(ω)ωω=0]1Z = \left[1 - \left.\frac{\partial \mathrm{Re}\,\Sigma(\omega)}{\partial \omega}\right|_{\omega=0}\right]^{-1}

with Σ\Sigma the self-energy. As correlations strengthen, ZZ \downarrow, effective mass m1/Zm^{\ast}\propto 1/Z \uparrow, and at the Mott point Z0Z\to 0—the QP dissolves, leaving only Hubbard bands.


10.10.6 Brinkman–Rice picture and the MIT

In a Fermi-liquid mean-field for correlations, the kinetic energy is renormalized by ZZ and the Mott metal–insulator transition (MIT) occurs when

Z0asUUcZ \to 0 \quad \text{as} \quad U \to U_{c}^{-}

The divergence of mm^{\ast} and collapse of Drude weight are hallmarks. In real 3D lattices, the transition at finite TT is often first order with hysteresis (two solutions coexist: metal and insulator), meeting at a critical endpoint—beautifully captured by DMFT.


10.10.7 Optical and ARPES fingerprints (how to spot Mottness)

  • Optics: Drude weight shrinks as ZZ, while a mid-infrared feature and Hubbard-band absorption grow. Spectral weight transfer violates simple band sum rules—weight moves over eV scales when you dope by a few percent.
  • ARPES: coherent QP peak near EFE_{F} plus broad incoherent background and high-energy waterfalls. Doping a Mott insulator pulls new low-energy states from the upper Hubbard band and reshapes the dispersion (strong kink structures if electron–phonon also matters).
  • Transport: resistivity above the Mott–Ioffe–Regel (MIR) limit (“bad metal”), huge thermopower asymmetries, Hall coefficient that changes sign with TT and doping.

10.10.8 Charge-transfer vs Mott insulators (Zaanen–Sawatzky–Allen)

In transition-metal oxides, the lowest unoccupied state might live on ligand pp orbitals rather than the metal dd. Compare UU (on dd) to the charge-transfer energy Δ\Delta (from O 2p2p to TM dd):

  • Mott–Hubbard insulator: Δ>U\Delta > U; gap mostly between lower and upper dd-Hubbard bands.
  • Charge-transfer insulator: Δ<U\Delta < U; gap is from O pp to upper dd band. Doping holes lands you on oxygen-like states (hello, cuprates).

This classification explains wildly different doping responses and optical edges across oxides with similar periodic tables.


10.10.9 Hund metals and multi-orbital spice

With several active orbitals, Hund’s coupling JHJ_{H} aligns spins on the same atom. Effects:

  • Suppresses interorbital charge fluctuations, reducing ZZ even when UU is moderate → “Hund metal” with heavy quasiparticles and large scattering.
  • Fosters orbital selectivity: some orbitals Mott-localize while others stay itinerant (orbital-selective Mott phase).
  • Competes/cooperates with crystal fields and SOC, key to ruthenates, iron-based superconductors, and 4d/5d oxides.

10.10.10 DMFT in one page: local self-energy, global consequences

Dynamical Mean-Field Theory (DMFT) maps a lattice model to a quantum impurity embedded in a self-consistent bath. The self-energy is taken local (momentum-independent): Σ(k,ω)Σ(ω)\Sigma(\boldsymbol k,\omega)\to \Sigma(\omega)—exact in infinite coordination, remarkably good in 3D.

Self-consistency (Bethe lattice with half bandwidth DD) reads

Δ(ω)=t2Gloc(ω),Gloc(ω)=dϵρ0(ϵ)ω+μΣ(ω)ϵ\Delta(\omega) = t^{2} G_{\text{loc}}(\omega),\qquad G_{\text{loc}}(\omega) = \int \frac{d\epsilon\,\rho_{0}(\epsilon)}{\omega+\mu-\Sigma(\omega)-\epsilon}

where Δ\Delta is the impurity hybridization, ρ0\rho_{0} the non-interacting DOS. Iterate: solve impurity → get Σ\Sigma → recompute GlocG_{\text{loc}} → update Δ\Delta.

What DMFT nails

  • Three-peak spectrum (QP + Hubbard bands), ZZ collapse at UcU_{c}, first-order MIT with coexistence.
  • Temperature evolution of resistivity including bad-metal behavior.
  • Doping asymmetries and spectral-weight transfer.
  • With LDA band inputs (LDA+DMFT): realistic materials, e.g., V2_2O3_3, SrVO3_3, iron pnictides.

10.10.11 1D twist: Luttinger liquids and spin–charge separation

In strictly 1D, Fermi liquids are unstable: correlations produce a Luttinger liquid with separate spin and charge modes. Signatures: power-law tunneling DOS, different velocities for spin vs charge excitations seen in ARPES or cold-atom analogues. The 1D Hubbard model is exactly solvable (Bethe Ansatz) and shows a Mott gap at half filling for any U>0U>0.


10.10.12 Kondo, heavy fermions, and the lattice cousin (one-paragraph cameo)

A local moment in a metal is screened by conduction electrons below the Kondo temperature TKT_{K}, forming a singlet with a narrow resonance at EFE_{F}. A periodic array of such moments (Kondo lattice) hybridizes at low TT into heavy quasiparticles with m ⁣mem^{\ast}\!\gg m_{e}; depending on filling and symmetry you can get heavy Fermi liquids, unconventional superconductivity, or Kondo insulators. (Deep dive would live in its own section; headline here: “correlations can also light up bands rather than only split them.”)


10.10.13 Worked mini-examples

(a) Superexchange sign and scale
For a two-site Hubbard dimer at half filling, do second-order perturbation in t/Ut/U and show J=4t2/UJ=4t^{2}/U and why it’s AFM.

(b) Brinkman–Rice mass blow-up
Assuming Z=1(U/Uc)2Z=1-(U/U_{c})^{2} near the MIT, show m/m=1/Zm^{\ast}/m=1/Z diverges as UUcU\to U_{c}^{-} and the Drude weight scales with ZZ.

(c) DMFT Bethe self-consistency
With semicircular DOS ρ0(ϵ)=2πD2D2ϵ2\rho_{0}(\epsilon)=\frac{2}{\pi D^{2}}\sqrt{D^{2}-\epsilon^{2}} and D=2tD=2t, derive Δ(ω)=t2Gloc(ω)\Delta(\omega)=t^{2}G_{\text{loc}}(\omega).

(d) Optical spectral weight transfer
Use the f-sum rule to argue why integrated low-energy spectral weight decreases with UU at fixed filling, and sketch how doping restores Drude weight while also creating mid-IR absorption.

(e) ZSA classification
Given U=6 eVU=6\ \mathrm{eV} and charge-transfer energy Δ=3 eV\Delta=3\ \mathrm{eV}, identify the insulator type and predict whether hole doping creates O-pp or TM-dd-like carriers.

(f) Hund metal toy
In a two-orbital Hubbard–Hund model with UU moderate but finite JHJ_{H}, explain qualitatively why interorbital charge fluctuations are suppressed and ZZ drops compared to the JH=0J_{H}=0 case.


10.10.14 Common pitfalls

  • Calling every AFM insulator “Mott”: if the gap collapses at TNT_{N}, it’s likely Slater. Check the paramagnetic phase.
  • Equating large UU with insulator at any filling: away from half filling you can have a correlated metal with small ZZ, not necessarily an insulator.
  • Forgetting spectral-weight transfer: simple band pictures conserve low-energy weight under doping; Mottness does not.
  • Using band effective masses in correlated fits: transport and thermodynamics see mm/Zm^{\ast}\sim m/Z; use the renormalized values.
  • Momentum-blind DMFT overreach: single-site DMFT can’t capture kk-dependent pseudogaps or short-range order well; cluster DMFT/DCA or diagrammatics are needed.
  • Mixing Mott and Anderson: interaction-driven localization (Mott) differs from disorder-driven (Anderson); both can coexist (Mott–Anderson), but diagnostics differ (e.g., interaction tuning vs. disorder tuning).

10.10.15 Minimal problem kit

  • Starting from the Hubbard model at half filling and UtU\gg t, derive the Heisenberg J=4t2/UJ=4t^{2}/U via Schrieffer–Wolff transformation
  • Implement a Hubbard-I self-energy and plot the emergence of upper/lower Hubbard bands with growing UU; compare to DMFT’s three-peak structure
  • For a semicircular DOS, solve the DMFT equations at T=0T=0 within iterated perturbation theory and chart Z(U)Z(U); locate UcU_{c}
  • Compute the doping evolution of the optical conductivity in a toy three-peak model (QP + two Hubbard bands); verify spectral-weight transfer
  • Classify a given oxide as Mott–Hubbard vs charge-transfer using tabulated UU and Δ\Delta and predict the sign/magnitude trend of the Seebeck coefficient upon light hole doping
  • Build a two-orbital Hubbard–Hund mean-field and map where an orbital-selective Mott phase appears as a function of UU, JHJ_{H}, and crystal-field splitting

10.11 Topological Band Theory

Some band insulators are “just” boring gaps. Others hide protected boundary states, quantized responses, and Berry-curvature superpowers. That’s topological band theory: use geometry in kk-space (Berry phase, curvature) plus symmetries (time reversal, inversion, chiral) to classify phases. Bulk invariants predict edge modes. Change the invariant? You must close the gap somewhere—cue phase transitions with Dirac or Weyl nodes. Let’s build the toolkit.


10.11.1 Berry phase, connection, and curvature

For a Bloch band unk|u_{n\boldsymbol k}\rangle smooth in k\boldsymbol k:

  • Berry connection
An(k)iunkkunk\boldsymbol A_{n}(\boldsymbol k) \equiv i \langle u_{n\boldsymbol k} | \nabla_{\boldsymbol k} u_{n\boldsymbol k} \rangle
  • Berry curvature
Ωn(k)k×An(k)\boldsymbol\Omega_{n}(\boldsymbol k) \equiv \nabla_{\boldsymbol k} \times \boldsymbol A_{n}(\boldsymbol k)
  • Berry phase along a closed loop C\mathcal C in the Brillouin zone (BZ)
γn[C]=CAndk\gamma_{n}[\mathcal C] = \oint_{\mathcal C} \boldsymbol A_{n}\cdot d\boldsymbol k

Gauge note: A\boldsymbol A depends on phase choice of u|u\rangle, Ω\boldsymbol\Omega and loop phases modulo 2π2\pi do not. In 1D, the Berry phase of occupied bands controls polarization (modern theory): PP is a Berry phase over the BZ (see §10.5.10).


10.11.2 Chern number and quantum Hall

In 2D, the Chern number of band nn is

Cn=12πBZd2k  Ωn,z(k)C_{n} = \frac{1}{2\pi} \int_{\mathrm{BZ}} d^{2}k\; \Omega_{n,z}(\boldsymbol k)

For a fully filled set of bands,

σxy=e2hnoccCn\sigma_{xy} = \frac{e^{2}}{h} \sum_{n\in \text{occ}} C_{n}

This is the TKNN integer quantum Hall result: a bulk topological invariant predicts a boundary chiral edge count. Change CC → the gap must close somewhere.


10.11.3 Anomalous Hall: Berry curvature without BB

Even at B=0B=0, broken time-reversal symmetry (TRS) yields an intrinsic anomalous Hall effect

σxyint=e2nd3k(2π)3fnkΩn,z(k)\sigma_{xy}^{\mathrm{int}} = -\frac{e^{2}}{\hbar} \sum_{n} \int \frac{d^{3}k}{(2\pi)^{3}} f_{n\boldsymbol k}\,\Omega_{n,z}(\boldsymbol k)

Here ff is the Fermi function. Think of Ω\boldsymbol\Omega as a “magnetic field in kk-space” deflecting wavepackets (§10.7 deferred Berry terms).


10.11.4 1D winding and the SSH warm-up

For a two-band, chiral-symmetric 1D model H(k)=d(k)σH(k)=\boldsymbol d(k)\cdot\boldsymbol\sigma with dz=0d_{z}=0, define ϕ(k)=arg[dx+idy]\phi(k)=\arg[d_{x}+id_{y}]. The winding number

ν=12ππ/aπ/adk  kϕ(k)\nu = \frac{1}{2\pi} \int_{-\pi/a}^{\pi/a} dk\; \partial_{k}\phi(k)

counts how many times d\boldsymbol d wraps the origin. Bulk ν=1\nu=1 gives edge zero modes in the SSH chain (§10.5.3).


10.11.5 2D Dirac mass flip → Chern insulator (Haldane vibe)

Consider a single massive Dirac cone

H(k)=v(kxσx+kyσy)+mσzH(\boldsymbol k)=v(k_{x}\sigma_{x}+k_{y}\sigma_{y}) + m\,\sigma_{z}

The valence-band Chern number is

C=12sgn(m)C = -\frac{1}{2}\,\mathrm{sgn}(m)

Two cones (as on honeycomb) contribute; if their masses have opposite signs, total C=±1C=\pm 1 → a Chern insulator with one chiral edge channel.


10.11.6 Time-reversal symmetry and Z2\mathbb{Z}_{2} topological insulators

With TRS (T2=1\mathcal T^{2}=-1 for electrons), bands come in Kramers pairs. In 2D and 3D, insulators are classified by a Z2\mathbb{Z}_{2} index ν\nu (2D) or (ν0;ν1ν2ν3)(\nu_{0};\nu_{1}\nu_{2}\nu_{3}) (3D). If inversion symmetry is present, use the Fu–Kane parity shortcut:

  • At each TRIM Γi\Gamma_{i}, multiply inversion eigenvalues of occupied Kramers pairs to get δi\delta_{i}
  • Then
(1)ν=iδi(-1)^{\nu} = \prod_{i}\delta_{i}

3D strong index is (1)ν0=iδi(-1)^{\nu_{0}}=\prod_{i}\delta_{i}; we omit the weak indices detail here. Strong TI (ν0=1\nu_{0}=1) hosts an odd number of Dirac cones on any surface, protected by TRS.

Without inversion, compute a Wilson loop of the Berry connection (a.k.a. Wannier charge-center flow). If the hybrid Wannier centers wind across the BZ, the phase is topological.


10.11.7 Bulk–boundary correspondence (in one graphic sentence)

Nontrivial bulk invariant ⇒ robust boundary states that cannot be gapped without either closing the bulk gap or breaking the protecting symmetry.

  • Chern insulator (CC): C|C| chiral edge modes (unidirectional).
  • 2D Z2\mathbb{Z}_{2} TI: helical edge (Kramers pair counterpropagating).
  • 3D strong TI: odd number of surface Dirac cones.
  • Crystalline topological phases (protected by mirror, rotation) host symmetry-selective boundary states.

10.11.8 Dirac and Weyl semimetals: topology without a gap

Gapless cousins:

  • Weyl nodes are monopoles of Berry curvature in 3D with integer chiral charge C=±1C=\pm 1. They require breaking either TRS or inversion to separate nodes in k\boldsymbol k-space. Consequences: Fermi arcs on surfaces and the chiral anomaly (longitudinal negative MR when EB\boldsymbol E\parallel \boldsymbol B). The anomaly equation reads
tρ5+j5=e22π22EB\partial_{t}\rho_{5} + \nabla\cdot \boldsymbol j_{5} = \frac{e^{2}}{2\pi^{2}\hbar^{2}}\,\boldsymbol E\cdot \boldsymbol B
  • Dirac semimetals superpose opposite-chirality Weyl nodes pinned together by symmetry; break the symmetry and they split into Weyl pairs.

10.11.9 Rashba, spin–momentum locking, and surface Dirac cones

Spin–orbit coupling (SOC) plus broken inversion gives Rashba splitting

HR=α(σ×k)z^H_{R} = \alpha (\boldsymbol\sigma \times \boldsymbol k)\cdot \hat{\boldsymbol z}

Topological surface states in TIs show spin–momentum locking (helical textures), suppressing backscattering from nonmagnetic disorder. Magnetic perturbations gap the surface Dirac cone, enabling quantized magneto-electric responses (axion physics below).


10.11.10 Axion electrodynamics and magneto-electric response

In 3D TIs, the effective EM action adds

Lθ=θe22πhEB\mathcal L_{\theta} = \frac{\theta e^{2}}{2\pi h}\,\boldsymbol E\cdot \boldsymbol B

with θ=0\theta=0 (trivial) or π\pi (strong TI) modulo 2π2\pi if TRS holds. At a surface where θ\theta jumps by π\pi, you get a half-quantized surface Hall conductance e2/(2h)e^{2}/(2h) when TRS is explicitly broken at the boundary.


10.11.11 Wilson loops and Wannier flow (how we compute invariants)

Discretize k\boldsymbol k along a closed path C\mathcal C and build the Wilson loop

W[C]=Pexp ⁣(iC ⁣A(k)dk)W[\mathcal C] = \mathcal P \exp\!\left(i\int_{\mathcal C}\! \mathsf A(\boldsymbol k)\cdot d\boldsymbol k\right)

where A\mathsf A is the non-Abelian Berry connection of the occupied subspace. Eigenphases of WW give Wannier centers; their flow vs transverse k\boldsymbol k reveals Chern (ν\nu winds by 2πC2\pi C per BZ) or Z2\mathbb{Z}_{2} (odd vs even partner-switching).


10.11.12 Orbital magnetization and modern polarization (Berry redux)

Two Berry-phase byproducts:

  • Polarization (insulators) is a Berry phase of occupied states across the BZ; only changes are physical and quantized under adiabatic cycles (Thouless pump).
  • Orbital magnetization in crystals can be expressed as BZ integrals involving Ω\boldsymbol\Omega and band energies; this links to the intrinsic AHE and to modern magneto-electric coefficients.

10.11.13 Minimal symmetry-class crib sheet (tiny slice of the tenfold way)

ClassSymmetry (TRS, PHS, chiral)2D invariant3D invariant
A (no symm)–, –, –Z\mathbb{Z} (Chern)0
AII (TRS with T2=1\mathcal T^{2}=-1)✓, –, –Z2\mathbb{Z}_{2}Z2\mathbb{Z}_{2} (strong ν0\nu_{0})
AIII (chiral only)–, –, ✓Z\mathbb{Z} (winding)0

The full tenfold way covers superconductors (BdG) and more invariants; we focus on the workhorses for electrons.


10.11.14 Experimental fingerprints

  • Transport: integer QHE plateaus; quantized edge conductance for Chern insulators; weak anti-localization from π\pi Berry phase; negative longitudinal MR in Weyls (careful: geometry artifacts exist).
  • ARPES: surface Dirac cones, Fermi arcs, Rashba splitting; spin-resolved ARPES confirms textures.
  • STM/STS: absence of 180180^{\circ} backscattering in QPI for helical surfaces; Landau-level ladders revealing Berry phase π\pi.
  • Optics: giant Kerr/Faraday angles with magnetized TI surfaces; cyclotron resonance with nontrivial phase offsets.
  • Quantum oscillations: phase shifts (Landau index plot intercept) tied to Berry phase, with spin-splitting caveats.

10.11.15 Worked mini-examples

(a) Chern of a two-band Dirac
For H=v(kxσx+kyσy)+mσzH=v(k_{x}\sigma_{x}+k_{y}\sigma_{y})+m\sigma_{z}, compute Ωz(k)\Omega_{z}(\boldsymbol k) and integrate to show C=12sgn(m)C=-\tfrac{1}{2}\mathrm{sgn}(m) per cone; two inequivalent cones with opposite mm give C=±1C=\pm 1.

(b) Parity test
Given inversion eigenvalues at four 2D TRIM as (+,+,+,)(+,+,+,-) for the occupied Kramers pairs, evaluate (1)ν=1(-1)^{\nu}=-1 ⇒ topological.

(c) Wilson loop wind
On a Chern band with C=1C=1, show the Wannier center angle θ(ky)\theta(k_{y}) advances by 2π2\pi as kyk_{y} goes across the BZ. Contrast with a trivial band.

(d) SSH end states
Compute the winding number for H(k)=(t1+t2coska)σx+t2sinkaσyH(k)=(t_{1}+t_{2}\cos ka)\sigma_{x}+t_{2}\sin ka\,\sigma_{y}. Show ν=1\nu=1 for t2>t1|t_{2}|>|t_{1}| and plot edge spectra for an open chain.

(e) Weyl monopole charge
Evaluate the flux of Ω\boldsymbol\Omega through a small sphere around a Weyl node and show it equals ±2π\pm 2\pi (i.e., charge ±1\pm 1).

(f) Surface Hall from axion
Assuming a TI with a magnetized surface so Δθ=π\Delta\theta=\pi, integrate Lθ\mathcal L_{\theta} across the boundary to obtain σxy=e2/(2h)\sigma_{xy}=e^{2}/(2h).


10.11.16 Common pitfalls

  • Gauge traps: A\boldsymbol A is gauge-dependent; only Ω\boldsymbol\Omega and loop phases mod 2π2\pi are physical. Use smooth gauges or Wilson loops for numerics.
  • Using Fu–Kane without inversion: the parity shortcut demands inversion symmetry; otherwise rely on Wilson loops or flow of Wannier centers.
  • Confusing Rashba with topology: Rashba splitting alone is not a TI; you need a bulk gap plus a nontrivial Z2\mathbb{Z}_{2} index.
  • Misreading oscillation phases: extracting Berry phase from Landau fan diagrams requires accounting for Zeeman, warping, and 3D geometry.
  • Edge ≠ surface every time: symmetry protects; break it at the boundary and you can gap the state (e.g., magnetic disorder gapping a TI surface).
  • Chern in TRS systems: with unbroken TRS and no magnetic order, the total Chern number of occupied bands must be zero; the Z2\mathbb{Z}_{2} index can still be nontrivial.

10.11.17 Minimal problem kit

  • Derive the TKNN formula starting from Kubo and show σxy=(e2/h)Cn\sigma_{xy}=(e^{2}/h)\sum C_{n} for a clean 2D insulator
  • Implement a discrete-BZ Chern number computation via link variables and validate on the two-band Dirac model
  • For a centrosymmetric SOC insulator, compute Fu–Kane Z2\mathbb{Z}_{2} from tabulated parities at TRIM; then confirm by Wilson-loop flow
  • Build a two-node Weyl model, compute surface Green’s function for a slab, and plot Fermi arcs
  • Add a surface Zeeman term to a 3D TI Dirac cone and show the opening of a gap and half-quantized Hall conductance
  • Simulate an adiabatic Thouless pump in a 1D lattice and show the quantized charge transport equals the Chern number of the (k,t)(k,t) torus

10.12 Mesoscopic Transport & Quantum Coherence

Shrink a conductor until its size rivals coherence and mean free paths, and transport flips from “traffic flow” to wave interference. This is the mesoscopic regime: electrons remain phase coherent across the device, conductance becomes sample-specific yet reproducible, and quantum noise reveals transmission statistics. The workhorses here are Landauer–Büttiker transport, Aharonov–Bohm interference, weak localization/antilocalization, and universal conductance fluctuations. We’ll keep one eye on experiments (QPCs, rings, nanowires) and one on formulas you can deploy.


10.12.1 Landauer: conductance as transmission

For a phase-coherent two-terminal conductor at low temperature and small bias, current is carried by independent quantum channels with transmissions {Tn}\{T_{n}\}:

G=2e2hnTnG = \frac{2e^{2}}{h}\sum_{n} T_{n}

Factor 2 is spin degeneracy (lifted by Zeeman/SOC). For a quantum point contact (QPC) in the ballistic limit, transverse modes open one by one so Tn1T_{n}\approx 1 and

GN2e2hG \approx N\,\frac{2e^{2}}{h}

giving quantized conductance steps vs gate voltage.

Multi-terminal (Büttiker) view. With contacts α,β\alpha,\beta and transmission probabilities TαβT_{\alpha\beta}, linear response reads

Iα=2e2hβ(TβαVαTαβVβ)I_{\alpha} = \frac{2e^{2}}{h}\sum_{\beta}\left( T_{\beta\alpha} V_{\alpha} - T_{\alpha\beta} V_{\beta} \right)

This enforces current conservation and automatically includes contact geometry.


10.12.2 Aharonov–Bohm (AB) and phase memory

In a ring, two paths enclose magnetic flux Φ\Phi. Wavefunctions pick up relative phase 2πΦ/Φ02\pi \Phi/\Phi_{0} with Φ0=h/e\Phi_{0}=h/e. Result:

  • AB oscillations in conductance periodic in h/eh/e
  • Altshuler–Aronov–Spivak (AAS) oscillations at h/2eh/2e from time-reversed loop pairs

Visibility decays with phase-coherence length LϕL_{\phi} set by dephasing (electron–electron, electron–phonon, external noise). Roughly, signals scale like exp(L/Lϕ)\exp(-L/L_{\phi}) along an interfering loop of length LL.


10.12.3 Shot noise: counting statistics under the hood

Even at T0T\to 0, current fluctuates because charge is discrete. For a phase-coherent conductor,

SI(0)=2eV2e2hnTn(1Tn)S_{I}(0) = 2 e V \frac{2e^{2}}{h}\sum_{n} T_{n}(1-T_{n})

where SI(0)S_{I}(0) is the zero-frequency noise spectral density and VV the bias. Define the Fano factor

FSI(0)2eI=nTn(1Tn)nTnF \equiv \frac{S_{I}(0)}{2 e I} = \frac{\sum_{n} T_{n}(1-T_{n})}{\sum_{n} T_{n}}

Key cases: ballistic plateau Tn1T_{n}\approx 1F0F\approx 0 (noise suppressed); diffusive wireF=1/3F=1/3; tunnel junction Tn1T_{n}\ll 1 ⇒ Poissonian F1F\approx 1.


10.12.4 Weak localization (WL) and antilocalization (WAL)

Quantum interference of time-reversed diffusive paths enhances backscattering:

  • With preserved time-reversal and weak spin–orbit, you get WL: a negative magnetoconductance dip at B=0B=0 (conductance reduced by quantum correction).
  • With strong SOC (or π\pi Berry phase), you get WAL: a positive cusp (antilocalization).

In 2D, the magnetoconductance near B=0B=0 follows the Hikami–Larkin–Nagaoka style correction (schematically)

Δσ(B)e22π2[f ⁣(BϕB)f ⁣(Bϕ+BsoB)]\Delta \sigma(B) \sim \frac{e^{2}}{2\pi^{2}\hbar} \left[ f\!\left(\frac{B_{\phi}}{B}\right) - f\!\left(\frac{B_{\phi}+B_{\mathrm{so}}}{B}\right) - \cdots \right]

where Bϕ1/Lϕ2B_{\phi}\propto 1/L_{\phi}^{2} and BsoB_{\mathrm{so}} encodes SOC. In 2D without SOC, the B=0B=0 correction scales like

Δσ(0)e2πhln ⁣(Lϕ)\Delta \sigma(0) \sim -\frac{e^{2}}{\pi h}\,\ln\!\left(\frac{L_{\phi}}{\ell}\right)

with mean free path \ell (sign flips for WAL).


10.12.5 Universal conductance fluctuations (UCF)

Mesoscopic conductance G(B)G(B) or G(Vg)G(V_{g}) exhibits aperiodic but reproducible wiggles from sample-specific interference. Their variance is universal (independent of material specifics), of order

var(G)(e2h)2×1β\mathrm{var}(G) \sim \left(\frac{e^{2}}{h}\right)^{2} \times \frac{1}{\beta}

where β=1,2,4\beta=1,2,4 labels orthogonal (TRS), unitary (broken TRS), and symplectic (strong SOC) symmetry classes. Precise coefficients depend on dimensionality and aspect ratio, but the order of magnitude is fixed—hence “universal.”


10.12.6 Dephasing and energy relaxation

Phase coherence dies due to inelastic processes:

  • Electron–electron scattering (dominant at low TT): often 1/τϕTp1/\tau_{\phi}\propto T^{p} with pp depending on dimension (sublinear in 1D, linear-ish in 2D diffusive)
  • Electron–phonon at higher TT: typically 1/τϕTq1/\tau_{\phi}\propto T^{q} with q>2q>2
  • Magnetic impurities: spin flips are coherence killers; Kondo physics may show up in TT trends

Extract Lϕ=DτϕL_{\phi}=\sqrt{D\tau_{\phi}} by fitting WL/WAL magnetoconductance or AB amplitude decay vs circumference.


10.12.7 Contact resistance and the four-probe cure

Landauer reminds us that even a perfect 1D wire has a contact-limited resistance

RQ=h2e212.9 kΩR_{\mathrm{Q}} = \frac{h}{2e^{2}} \approx 12.9\ \mathrm{k}\Omega

per occupied spin-degenerate channel. Multi-terminal geometries (Hall bars, four-probe) separate contact/interface effects from intrinsic device transmission.


10.12.8 Random-matrix universality

Fully chaotic/strongly disordered but phase-coherent cavities (“quantum dots”) have transmission eigenvalues whose statistics follow Wigner–Dyson ensembles:

  • GOE (β=1) with TRS, no SOC
  • GUE (β=2) broken TRS
  • GSE (β=4) strong SOC, TRS preserved

This governs distributions of GG, shot noise (e.g., F=1/4F=1/4 in chaotic cavities), and level spacings (Wigner surmise). Symmetry crossovers are tuned by magnetic field and SOC strength.


10.12.9 Mesoscopic superconductivity: Andreev in nanoscale

At a clean normal–superconductor (NS) interface, subgap transport occurs via Andreev reflection: an electron retroreflects as a hole, adding a Cooper pair to the condensate. Consequences:

  • Doubling of low-bias conductance for a perfectly transparent NS point contact (BTK limit)
  • Multiple Andreev reflections (MAR) in short SNS junctions, producing subharmonic gap features at eV=2Δ/neV=2\Delta/n
  • In proximitized nanowires with SOC and Zeeman field, Andreev bound states can morph into Majorana-like zero modes at tuned parameters (topological superconductivity cameo)

10.12.10 From mesoscopic to Anderson: the scaling bridge

As disorder grows or dimension sinks, interference corrections (WL/WAL, UCF) strengthen. In low dimensions:

  • 1D, 2D: arbitrarily weak disorder localizes noninteracting electrons at T=0T=0 in the orthogonal class (no SOC, no magnetic field)
  • 3D: a true metal–insulator transition can occur at critical disorder

The dimensionless conductance gG/(e2/h)g \equiv G/(e^{2}/h) obeys a scaling flow dlng/dlnL=β(g)d\ln g/d\ln L = \beta(g); the sign of β(g)\beta(g) decides metal vs insulator. We detail full Anderson localization and scaling theory in §10.13.


10.12.11 Worked mini-examples

(a) Landauer 101
A QPC shows plateaus at G=(2,4,6)×e2/hG=(2,4,6)\times e^{2}/h as the gate opens. Estimate the number of occupied modes and the average TnT_{n} on a sloped transition between plateaus.

(b) Shot-noise triad
Compute FF for (i) a tunnel junction with T1T\ll 1, (ii) a ballistic QPC on a plateau, (iii) a diffusive wire. Explain why FF distinguishes them even if GG is similar.

(c) AB vs AAS
A ring of circumference LL shows h/eh/e oscillations with amplitude A1eL/LϕA_{1}\propto e^{-L/L_{\phi}} and a weaker h/2eh/2e harmonic A2A_{2}. Argue why A2A_{2} decays as e2L/Lϕe^{-2L/L_{\phi}}.

(d) WL fit
Given a 2D magnetoconductance trace with a sharp cusp at B=0B=0 and curvature reversal after adding a top gate (stronger SOC), fit qualitatively for LϕL_{\phi} and LsoL_{\mathrm{so}} trends.

(e) UCF scaling
Show how thermal averaging over an energy window kBTk_{B}T reduces variance var(G)\mathrm{var}(G) when kBTk_{B}T exceeds the Thouless energy ET=D/L2E_{T}=\hbar D/L^{2}.

(f) Contact resistance
A ballistic single-mode nanowire of length LL\ll \ell shows RRQR\approx R_{\mathrm{Q}}. Explain why adding a fourth probe pair reports nearly zero internal resistance while two-probe still reads RQ\sim R_{\mathrm{Q}}.


10.12.12 Common pitfalls

  • Forgetting contacts: Landauer conductance includes contact resistance; don’t blame the “perfect wire” for a finite two-probe RR.
  • Mixing ensemble with sample noise: UCF are reproducible fingerprints vs BB or VgV_{g}, not random flicker; average over BB windows to expose WL on top.
  • WL sign confusion: WAL needs strong SOC or π\pi Berry phase (e.g., topological surfaces); otherwise expect WL.
  • Assuming Fano =1=1 always: Poissonian shot noise appears only in tunneling; ballistic and diffusive devices suppress noise predictably.
  • Phase vs energy relaxation: dephasing kills interference without necessarily equilibrating energy; don’t equate τϕ\tau_{\phi} with energy-relaxation time.
  • Overfitting HLN: the 2D WL/WAL formula has parameter correlations; report trends (LϕL_{\phi}, LsoL_{\mathrm{so}}) with care, not over-precise numbers.

10.13 Anderson Localization & Scaling Theory

Turn disorder up and interference stops being cute: wavefunctions localize, DC transport dies, and conductance becomes exponentially small. That’s Anderson localization—a single-particle quantum phase where disorder + interference trap eigenstates. This section builds the Anderson model, scaling theory and the β\beta-function, mobility edges and criticality in 3D, symmetry-class twists (orthogonal/unitary/symplectic), finite-size and multifractals, experimental fingerprints (VRH, Coulomb gap, WL/WAL magnetoresponse), and quick routes for calculations (transfer matrix, Thouless number, nonlinear sigma model cameo). We end with a postcard from many-body localization.


10.13.1 The minimal model

The tight-binding Anderson Hamiltonian on a lattice:

H=iϵiii+ijtijijH = \sum_{i} \epsilon_{i}\,|i\rangle\langle i| + \sum_{\langle ij\rangle} t_{ij}\,|i\rangle\langle j|

with random on-site energies ϵi\epsilon_{i} drawn from a distribution of width WW (e.g., box [W/2,W/2][-W/2,W/2]) and short-range hopping tij=tt_{ij}=t. Compete WW vs bandwidth WbandztW_{\mathrm{band}}\sim zt.

Heuristics. When elastic mean free path \ell drops to the Ioffe–Regel brink kF1k_{F}\ell \sim 1, waves can’t complete a phase-coherent mean-free-path without scrambling—interference runs the show.


10.13.2 From diffusion to localization: the scaling hypothesis

Define the dimensionless conductance gg of a hypercubic sample of size LL:

g(L)G(L)e2/hν(EF)DLd2ETΔg(L) \equiv \frac{G(L)}{e^{2}/h} \sim \nu(E_{F})\, D\, L^{d-2} \sim \frac{E_{T}}{\Delta}

with density of states ν\nu, diffusion constant DD, Thouless energy ET=D/L2E_{T}=\hbar D/L^{2}, and level spacing Δ[νLd]1\Delta\sim [\nu L^{d}]^{-1}. The one-parameter scaling postulate (Abrahams–Anderson–Licciardello–Ramakrishnan) says

β(g)dlngdlnL\beta(g) \equiv \frac{d\ln g}{d\ln L}

depends only on gg (and symmetry class). Consequences:

  • d=1d=1 and 2 (orthogonal class): β(g)<0\beta(g)<0 for all gg ⇒ no true metallic fixed point at T=0T=0; arbitrarily weak disorder localizes (2D logarithmically slowly).
  • d=3d=3: β(g)\beta(g) crosses zero at gg^{\ast}metal–insulator transition (MIT). Correlation/localization length
ξWWcν\xi \sim |W-W_{c}|^{-\nu}

diverges at critical disorder WcW_{c}; on the metallic side g(L ⁣ ⁣)(L/)d2g(L\!\gg\!\ell)\sim (L/\ell)^{d-2}.

Symmetry tweaks (unitary/symplectic) modify β\beta and can change the 2D verdict (see §10.13.6).


10.13.3 Mobility edge and 3D criticality

In 3D, eigenstates can be extended near band center and localized in the tails; the energy separating them is the mobility edge Ec(W)E_{c}(W). Tuning disorder moves EcE_{c}; at W=WcW=W_{c}, EcE_{c} hits EFE_{F} and DC transport vanishes.

Critical behavior (orthogonal class, noninteracting):

  • Localization length ξWWcν\xi\sim |W-W_{c}|^{-\nu} with universal ν\nu
  • Conductivity scaling on the metallic side
σe2hξ2dF ⁣(Lξ,Lϕξ)\sigma \sim \frac{e^{2}}{h}\, \xi^{2-d}\, \mathcal F\!\left(\frac{L}{\xi},\frac{L_{\phi}}{\xi}\right)

At T=0T=0 and LL\to\infty, σ(WcW)μ\sigma \propto (W_{c}-W)^{\mu} with an exponent μ\mu tied to ν\nu by scaling.


10.13.4 Quantum corrections: WL/WAL and lnT\ln T drifts

In the diffusive regime (kF1k_{F}\ell\gg 1), interference of time-reversed paths yields weak localization (WL) corrections. In 2D (orthogonal class), the zero-field correction behaves as

δσ(0)e2πhln ⁣(Lϕ)\delta \sigma(0) \sim -\frac{e^{2}}{\pi h}\,\ln\!\left(\frac{L_{\phi}}{\ell}\right)

A perpendicular magnetic field breaks time-reversal and suppresses WL, producing a negative magnetoresistance cusp around B=0B=0 (Hikami–Larkin–Nagaoka line shape). With strong spin–orbit coupling (symplectic class), the sign flips to weak antilocalization (WAL).

Interactions add Altshuler–Aronov corrections with lnT\ln T trends and zero-bias anomalies in tunneling DOS. Together, WL+AA often drive the observed σ(T)=σ0+alnT\sigma(T)=\sigma_{0}+a\ln T in 2D metals.


10.13.5 Insulating side: hopping and Coulomb gap

When states are localized with length ξ\xi, DC transport proceeds by phonon-assisted variable-range hopping (VRH).

  • Mott VRH (no long-range Coulomb gap):
ρ(T)exp ⁣[(T0T)1d+1]\rho(T) \sim \exp\!\left[\left(\frac{T_{0}}{T}\right)^{\frac{1}{d+1}}\right]
  • Efros–Shklovskii VRH (Coulomb interactions carve a soft gap in DOS):
ρ(T)exp ⁣[(T1T)1/2]\rho(T) \sim \exp\!\left[\left(\frac{T_{1}}{T}\right)^{1/2}\right]

Magnetoresponse can be large and positive in VRH (wavefunction shrinkage and interference of hopping paths).


10.13.6 Symmetry classes and 2D twists

The fate of 2D is not one-size-fits-all; it depends on global symmetries:

  • Orthogonal (TRS, no SOC): all states localized at T=0T=0; σ\sigma slides down logarithmically.
  • Unitary (TRS broken by BB): cooperon suppressed; localization still wins in the end in 2D, but β\beta is less negative.
  • Symplectic (strong SOC, TRS intact): WAL makes β(g)>0\beta(g)>0 at large gg and allows a genuine 2D MIT tuned by disorder or carrier density.

Other classes (chiral, Bogoliubov–de Gennes) add special cases: e.g., integer quantum Hall plateaus (class A) with topological transitions between localized Hall plateaus.


10.13.7 Finite-size scaling and the transfer-matrix route

Numerics extract ν\nu and WcW_{c} by computing a quasi-1D localization length λM(W)\lambda_{M}(W) for wide bars of width MM using transfer matrices, then collapsing

λMM=F ⁣[(WWc)M1/ν]\frac{\lambda_{M}}{M} = \mathcal F\!\left[(W-W_{c})\,M^{1/\nu}\right]

Curves for different MM cross at WcW_{c}; the slope yields ν\nu. This is the workhorse for noninteracting universality classes.


10.13.8 Multifractality at the mobility edge

Critical wavefunctions are neither plane waves nor exponentially localized—they are multifractal. Moments of amplitudes scale as

Pqiψ(i)2qLDq(q1)P_{q} \equiv \sum_{i} |\psi(i)|^{2q} \sim L^{-D_{q}(q-1)}

with a nontrivial spectrum of generalized dimensions DqD_{q} (not just the geometric dimension dd). Multifractality shows up in participation ratios, LDOS fluctuations, and critical conductance distributions.


10.13.9 Field theory cameo: nonlinear sigma model

Diagrammatics and disorder averaging lead to a nonlinear sigma model for diffusive modes (diffusons/cooperons). Coupling constants track gg; renormalization reproduces β(g)\beta(g), WL/WAL signs, and interaction corrections (Finkel’stein extension). This is the analytic backbone behind scaling and universality.


10.13.10 Optical and AC transport

On the metallic side, optical conductivity has a Drude peak with disorder-broadened width 1/τ1/\tau. Approaching criticality, Drude weight collapses. In an Anderson insulator (non-interacting), low-frequency AC conductivity rises as a power law (Mott ac hopping); interactions and Coulomb gap modify exponents. Finite-frequency probes thus access ξ\xi and hopping scales even when DC is immeasurably small.


10.13.11 Many-body localization (MBL): the sequel in interacting systems

Add interactions without a bath and localization can persist at finite energy density:

  • Emergent local integrals of motion (LIOMs), area-law entanglement in highly excited eigenstates
  • Logarithmic entanglement growth after quenches
  • Absence of DC transport and failure of eigenstate thermalization

MBL requires isolation; phonons or leads eventually delocalize. Still, it’s a striking “Anderson-but-for-many-body” phase for cold atoms and mesoscopic circuits.


10.13.12 Worked mini-examples

(a) Thouless number
Show gET/Δg\simeq E_{T}/\Delta by comparing energy sensitivity of levels under twisted boundary conditions (Thouless) to the mean level spacing.

(b) 2D WL slope
Starting from the Cooperon loop integral, derive δσ(0)=(e2/πh)ln(Lϕ/)\delta \sigma(0)=-(e^{2}/\pi h)\ln(L_{\phi}/\ell) in 2D and explain why a perpendicular BB cuts off the log via LB=/(2eB)L_{B}=\sqrt{\hbar/(2eB)}.

(c) VRH exponents
From a constant DOS in dd dimensions, optimize hopping distance and energy to obtain the Mott VRH exponent 1/(d+1)1/(d+1). Then include a Coulomb gap g(E)EEFd1g(E)\propto |E-E_{F}|^{d-1} to get the Efros–Shklovskii 1/21/2 law.

(d) Transfer-matrix crossing
Sketch λM/M\lambda_{M}/M vs WW for widths M=8,12,16M=8,12,16 and mark the single crossing at WcW_{c}. Explain how the slope ratio gives ν\nu.

(e) Multifractal moment
Given eigenstate amplitudes on a LdL^{d} lattice, compute P2P_{2} and extract an effective fractal dimension D2D_{2} from a log–log fit across sizes.

(f) Symmetry flip
Argue qualitatively how adding strong SOC flips the sign of the Cooperon contribution and yields WAL, and why a small BB restores WL-like behavior.


10.13.13 Common pitfalls

  • “All 2D localizes, period.” Only in the orthogonal class without strong SOC. Symplectic 2D can show a genuine MIT; QHE adds topological criticality.
  • Confusing VRH with activated gaps. VRH exponents are non-Arrhenius; check the (lnρ)(\ln \rho) vs TαT^{-\alpha} slope for dimension and Coulomb-gap effects.
  • Equating kF1k_{F}\ell\sim 1 with the transition. It’s a heuristic boundary of diffusive physics, not the precise WcW_{c}.
  • Ignoring interactions in 2D. Altshuler–Aronov terms are often as large as WL. If your lnT\ln T slope flips with gate-tuned SOC, that’s WAL, not an interaction sign change.
  • Over-reading finite samples. A rising ρ(L)\rho(L) does not prove insulating if LξL\ll \xi; do finite-size scaling.
  • MBL ≠ Anderson. MBL is an interacting, highly excited-state phenomenon; Anderson localization is single-particle (or noninteracting) at T=0T=0.

10.13.14 Minimal problem kit

  • Starting from Drude diffusion, build g(L)g(L) and the β\beta-function to first quantum-interference order in 2D; show why it’s negative in orthogonal and positive in symplectic at large gg
  • Compute the magnetoconductance Δσ(B)\Delta \sigma(B) near B=0B=0 using the Hikami–Larkin–Nagaoka formula; extract LϕL_{\phi} from a mock dataset
  • Fit synthetic ρ(T)\rho(T) to Mott vs Efros–Shklovskii VRH and decide the presence of a Coulomb gap
  • Implement a 3D Anderson model transfer-matrix code, produce λM/M\lambda_{M}/M crossings, and estimate ν\nu by data collapse
  • Evaluate PqP_{q} for q=2,3q=2,3 at WWcW\approx W_{c} and verify multifractal scaling across sizes
  • Show that g=ET/Δg=E_{T}/\Delta reduces to gσLd2/(e2/h)g\sim \sigma L^{d-2}/(e^{2}/h) by inserting D=σ/(e2ν)D=\sigma/(e^{2}\nu) and Δ1=νLd\Delta^{-1}=\nu L^{d}

10.14 Moiré Quantum Matter & 2D Materials

Stack atomically thin crystals with a tiny twist and the band structure goes full kaleidoscope. Van der Waals (vdW) heterostructures let us dial geometry (twist), electrostatics (dual gates), and symmetry (alignment, strain) with near-video-game control. The result: flat bands, correlated insulators, unconventional superconductivity, quantum anomalous Hall, and moiré excitons—all living in a designer superlattice you can reconfigure on a chip.


10.14.1 Why 2D, why moiré

  • 2D crystals (graphene, hBN, TMDs like MoS2_2, WSe2_2) can be peeled, stacked, rotated, and wired.
  • Weak vdW bonding between layers allows arbitrary heterostructures without nasty interfacial chemistry.
  • A small twist angle θ\theta between layers creates a long-wavelength moiré superlattice with period
Lm=a2sin(θ/2)L_{m} = \frac{a}{2\sin(\theta/2)}

where aa is the monolayer lattice constant. For θ1\theta \sim 1^{\circ}, LmL_{m} is tens of nanometers—huge compared to aa—so momentum space folds into mini Brillouin zones and bands flatten.


10.14.2 Continuum picture and “magic” flat bands

For twisted bilayer graphene (tBLG), the Bistritzer–MacDonald continuum model couples Dirac cones from the two layers with interlayer tunneling ww and relative rotation θ\theta. Define

kθ=2kDsin(θ/2),kD=4π3ak_{\theta} = 2 k_{D} \sin(\theta/2),\qquad k_{D} = \frac{4\pi}{3a}

and a dimensionless coupling

α=wvFkθ\alpha = \frac{w}{\hbar v_{F} k_{\theta}}

At magic angles (first one near θ1.1\theta \approx 1.1^{\circ}), α\alpha pushes the Dirac velocity to nearly zero; bandwidth WW collapses into a few meV. With interactions UU of similar size, correlation runs the show.

A cardboard back-of-the-envelope for interaction scale:

Ue24πε0εrLmU \sim \frac{e^{2}}{4\pi \varepsilon_{0} \varepsilon_{r} L_{m}}

which easily lands in the 10–50 meV ballpark at small θ\theta, i.e., U/W1U/W \gg 1.


10.14.3 Moiré Hubbard models and Wigner physics

The moiré potential defines effective lattices for Wannier functions:

  • tBLG: fragile topology complicates a strictly local tight-binding, but low-energy physics often projects to valley-resolved Chern bands near charge neutrality.
  • TMD heterobilayers (e.g., WSe2_2/WS2_2 with small twist): well-localized moiré triangular or honeycomb lattices; a bona fide Hubbard model with tunable tt by twist/field.

At low fillings and large rsr_{s}, kinetic energy loses and Wigner crystals or generalized charge order can appear. Competing orders (Mott, charge density waves, spin/valley polarization) are tuned by gates and displacement field.


10.14.4 Topology in moiré: valleys, Berry curvature, Chern

Monolayer TMDs carry valley-contrasting Berry curvature; stacking and moiré folding can engineer Chern bands even without net magnetic field. Valley-resolved Hall responses:

σxyv=e2hnoccCn(K)e2hnoccCn(K)\sigma_{xy}^{v} = \frac{e^{2}}{h} \sum_{n \in \text{occ}} C_{n}^{(K)} - \frac{e^{2}}{h} \sum_{n \in \text{occ}} C_{n}^{(K')}

Time-reversal demands total Chern Ctot=0C_{\text{tot}}=0 if both valleys are equally filled, but spontaneous valley polarization breaks that balance, yielding quantum anomalous Hall in zero BB.

Berry curvature dipole in non-centrosymmetric stacks drives nonlinear Hall signals at second order in EE.


10.14.5 Broken symmetries: correlated insulators, SC, QAHE, nematics

Dial carrier density per moiré cell ν\nu (via dual gates) and the phase diagram lights up:

  • Correlated insulators at integer ν\nu (Mott-like or flavor-polarized band insulators).
  • Unconventional superconductivity upon slight doping of correlated states; sensitivity to displacement field and Coulomb screening suggests pairing from electronic mechanisms in some platforms.
  • Quantum anomalous Hall effect (QAHE) when a single-valley Chern band is exchange-polarized.
  • Nematicity: rotational symmetry breaking evident in transport and STM, often tied to strain or interaction-driven Pomeranchuk-like instabilities.

A minimalist effective interaction for pairing in a flat band with DOS N(0)N(0):

TcWexp ⁣(1λeff),λeffVeffN(0)T_{c} \sim W\, \exp\!\left(-\frac{1}{\lambda_{\text{eff}}}\right),\qquad \lambda_{\text{eff}} \equiv V_{\text{eff}} N(0)

Even modest VeffV_{\text{eff}} gives sizable TcT_{c} if WW is tiny; details hinge on valley/spin structure and form factors.


10.14.6 TMD excitons, interlayer excitons, and moiré polaritons

Monolayer TMDs host tightly bound excitons (hundreds of meV) thanks to reduced screening and 2D kinematics. Stacking two different TMDs forms type-II band alignment and interlayer excitons with electron and hole on different layers:

  • Long lifetimes (ns–μ\mus), large electric dipoles, drift under in-plane fields.
  • With a twist, the moiré potential creates exciton minibands and trapped arrays (an “exciton lattice”).

A hydrogenic estimate for a 2D exciton (with effective masses and dielectric constant) still helps:

EXμme1εr2×13.6 eVE_{X} \approx \frac{\mu^{\ast}}{m_{e}} \frac{1}{\varepsilon_{r}^{2}} \times 13.6\ \mathrm{eV}

Real numbers require the Keldysh screening model, but the scaling with μ\mu^{\ast} and εr\varepsilon_{r} is the vibe.

Couple the excitons to a cavity: you get moiré exciton–polaritons with tunable dispersion and interactions.


10.14.7 Spin–valley locking and valleytronics

In TMD monolayers, SOC locks spin to valley. Circularly polarized light addresses KK vs KK' selectively, enabling valley polarization. With broken inversion in bilayers (or applied displacement field), you can steer valley Hall currents transverse to E\boldsymbol E without BB. The nonlinear Hall effect emerges from a finite Berry curvature dipole D\boldsymbol D:

J(2ω)τD×E(ω)2\boldsymbol J^{(2\omega)} \propto \tau\, \boldsymbol D \times \boldsymbol E(\omega)^{2}

measurable as a second-harmonic Hall voltage at zero magnetic field.


10.14.8 Experimental knobs and probes

  • Knobs: twist θ\theta, displacement field DD (dual gates), carrier density ν\nu per moiré cell, dielectric environment (hBN thickness, nearby metal gates), strain.
  • Transport: longitudinal/Hall resistivity vs ν\nu, Landau fans under BB, quantum oscillations reading off moiré FS pockets.
  • STM/STS: direct DOS, gap maps, visualization of moiré patterns and local order.
  • Optics: PL/reflectance for excitons, circular dichroism for valley, pump–probe for dynamics.
  • Capacitance/compressibility: thermodynamic DOS and symmetry breaking (spin/valley polarization plateaus).
  • Magnetometry: QAHE plateaus, orbital magnetization from Chern bands, SQUID-on-tip imaging of edge modes.

10.14.9 Minimal modeling routes

  • Continuum Dirac + periodic tunneling for graphene family; captures flat bands and topology with a few parameters (vF,w0/1,θ)(v_{F}, w_{0/1}, \theta).
  • Moiré Hubbard for TMD bilayers on triangular/honeycomb lattices with on-site UU, nearest-neighbor VV, and possible Hund/Ising anisotropies from SOC.
  • Hartree–Fock for flavor symmetry breaking; DMFT/ED for correlation and finite-TT spectra; Berry + Wilson loops (see §10.11) to track Chern.

10.14.10 Worked mini-examples

(a) Moiré length scale
Graphene with a=0.246 nma=0.246\ \mathrm{nm} and θ=1.1\theta=1.1^{\circ}. Compute Lm=a/[2sin(θ/2)]L_{m}=a/[2\sin(\theta/2)] and estimate Ue2/(4πε0εrLm)U \sim e^{2}/(4\pi \varepsilon_{0}\varepsilon_{r} L_{m}) for εr=5\varepsilon_{r}=5.

(b) Magic-angle parameter
With vF=1.0×106 m/sv_{F}=1.0\times 10^{6}\ \mathrm{m/s} and w=110 meVw=110\ \mathrm{meV}, compute α=w/(vFkθ)\alpha=w/(\hbar v_{F} k_{\theta}) at θ=1.1\theta=1.1^{\circ} and comment on flatness.

(c) Triangular-lattice Hubbard
Assume t=2 meVt=2\ \mathrm{meV}, U=20 meVU=20\ \mathrm{meV} for a TMD moiré. Estimate U/WU/W with W9tW\approx 9|t| and discuss Mott vs metal at half filling.

(d) Valley Hall sign
If a single-valley Chern band with C=1C=1 is fully polarized, what is σxy\sigma_{xy} in units of e2/he^{2}/h at zero BB Compare to the case where both valleys with opposite CC are filled.

(e) Interlayer exciton dipole
For layer spacing d=0.7 nmd=0.7\ \mathrm{nm}, estimate the dipole moment p=edp=ed and the Stark shift ΔE=pE\Delta E = -pE for E=0.1 V/nmE=0.1\ \mathrm{V/nm}.

(f) Nonlinear Hall estimate
Given a Berry curvature dipole D=0.1 A˚|D|=0.1\ \mathrm{\AA} and relaxation time τ=0.1 ps\tau=0.1\ \mathrm{ps}, estimate the order of magnitude of the second-harmonic Hall current under E(ω)=1 kV/cmE(\omega)=1\ \mathrm{kV/cm} (scaling answer is fine).


10.14.11 Common pitfalls

  • “Flat band = magic only.” Many moiré platforms (TMDs, graphene on hBN) host narrow bands without the exact magic-angle tuning; displacement field and dielectric screening reshape bandwidths dramatically.
  • Ignoring flavor: spin, valley, and layer give multiple “flavors”; broken-symmetry physics often picks one (or a combo). Models that compress flavors too soon miss real orders.
  • Topology blinders: total Chern can cancel across valleys; you still get valley Hall and orbital magnetization if flavors are imbalanced.
  • Wannierization traps: fragile topology in tBLG obstructs strictly local Wannier bases unless you enlarge the subspace—don’t force a too-minimal TB.
  • Overfitting mean-field: Hartree–Fock loves order; cross-check with thermodynamics, compressibility, and disorder robustness.
  • Exciton hydrogenic overuse: 2D screening is nonlocal (Keldysh); hydrogenic estimates are scale guides, not final numbers.