Quantum mechanics is a vibe only until you write the postulates. Then it’s a machine: crisp inputs, crisp outputs. Here is the compact rulebook we’ll use for the rest of the chapter, with enough Dirac notation to keep us fluent and enough coordinate formulas to stay grounded.
A closed quantum system lives in a complex Hilbert space H. A pure state is a unit vector (a ket) ∣ψ⟩∈H. Global phase is unphysical: ∣ψ⟩ and eiϕ∣ψ⟩ describe the same physical state. The dual ⟨ψ∣ is the complex-conjugate transpose (a bra). Inner products are ⟨ϕ⟩ψ.
An orthonormal basis {∣n⟩} resolves the identity
I=n∑∣n⟩⟨n∣
For continuous bases (like position), sums become integrals and Kronecker deltas become Dirac deltas
∫∣x⟩⟨x∣dx=I,⟨x⟩x′=δ(x−x′)
The wavefunction in position space is ψ(x)=⟨x⟩ψ, with normalization
Every observable A is a self-adjoint (Hermitian) operator A^ on H. Measurement outcomes are its eigenvalues; post-measurement states are the associated eigenvectors (modulo degeneracy). Spectral resolution:
A^=a∑aΠa
where Πa are orthogonal projectors onto the eigenspaces. For continuous spectra, replace sums by integrals A=∫adΠ(a).
Expectation values in state ∣ψ⟩:
⟨A⟩ψ=⟨ψ∣A^∣ψ⟩
Variance (ΔA)2=⟨A2⟩−⟨A⟩2 quantifies spread. Noncommuting observables cannot be simultaneously sharp; the general inequality is
ΔAΔB≥21⟨[A^,B^]⟩
Position and momentum satisfy the canonical commutation relation (CCR)
[x^,p^]=iℏ
implying ΔxΔp≥ℏ/2.
In the x-representation,
x^ψ(x)=xψ(x),p^ψ(x)=−iℏdxdψ
so the CCR becomes the familiar differentiation identity.
If you measure A in state ∣ψ⟩, the probability to get eigenvalue a is
P(a)=⟨ψ∣Πa∣ψ⟩
and the collapsed post-measurement state is
P(a)Πa∣ψ⟩
For degenerate a, Πa projects to the full eigenspace. More general measurements (POVMs) appear in §7.12; for most of this chapter, projective measurements suffice.
For subsystems A and B, states live in the tensor productHA⊗HB. If ∣ψA⟩ and ∣ψB⟩ are states, the product state is ∣ψA⟩⊗∣ψB⟩. But generic states are entangled, e.g.,
∣Ψ⟩=21(∣0⟩A∣1⟩B+∣1⟩A∣0⟩B)
which cannot be factored. Observables act as A⊗I or I⊗B. Partial information is described by the reduced density operator (see below).
Quantum expectation values evolve nearly classically when wave packets are narrow:
dtd⟨x^⟩=m⟨p^⟩,dtd⟨p^⟩=−⟨∂x∂V⟩
If V(x) is smooth over the packet width, ⟨∂V/∂x⟩≈∂V(⟨x⟩)/∂x, and the centroid follows Newton’s law. This is the correspondence principle in differential form.
For time-independent V, eigenfunctions ϕn(x) solve
[−2mℏ2dx2d2+V(x)]ϕn(x)=Enϕn(x)
Distinct eigenvalues are orthogonal
∫ϕm∗(x)ϕn(x)dx=0(m=n)
Real, confining potentials produce discrete spectra with increasing node counts as n rises. Bound vs scattering states will be treated carefully in §7.10.
If a unitary U leaves the Hamiltonian invariant, U†H^U=H^, then there is a conserved generator G (Noether-style). Infinitesimal unitaries U=exp(−iϵG/ℏ) imply
dtd⟨G⟩=ℏi⟨[H^,G]⟩
so [H,G]=0⇒⟨G⟩ is constant. Translational invariance conserves p, rotational invariance conserves L, and time invariance (no explicit t in H) conserves energy.
Quantum mechanics turns dynamics into linear wave evolution. The central object is the wavefunctionψ(r,t), whose squared magnitude gives probability density, and whose phase steers interference and current. The engine is the Schrödinger equation. This section lays out the equation, its conservation law, separation into stationary states, and the boundary conditions that quantize energies.
Build wave packets by superposing k values; packet center moves with group velocity vg=ℏk/m and spreads over time.
(b) Infinite square well of width L on 0<x<L:
ϕn(x)=L2sin(Lnπx),En=2mL2ℏ2π2n2,n=1,2,…
Real eigenfunctions ⇒ current j=0 in stationary states. Superpositions can carry current.
(c) Finite step and barriers. Impose continuity of ϕ and ϕ′ to get reflection R and transmission T; probability flux conservation enforces R+T=1 for real potentials. For barriers, tunneling gives T∼e−2κa with κ=2m(V0−E)/ℏ and width a.
For time-independent, real V, stationary bound eigenfunctions can be chosen real, so
j=0
The quantum number n counts nodes (zeros) of ϕn in 1D wells; higher n means more oscillations and higher energy. Relative phases only matter in superpositions, where interference redistributes probability and can create nonzero currents even in confining potentials.
The harmonic oscillator is quantum’s Swiss Army knife. It models vibrations, phonons, molecular stretches, the EM field’s normal modes, even the “zero-point jitter” of everything. It is also where the math becomes elegant: one potential, two complementary solutions—differential (Hermite polynomials) and algebraic (ladder operators)—and a gallery of physical insights.
With a weak drive H′(t)=−F(t)x, the Heisenberg equation for a is linear, producing
dtda=−iωa+2x0ℏiF(t)
A drive at frequency near ω displaces the coherent-state parameter α and yields resonant growth bounded by damping or finite interaction time. This is the quantum backbone of classical resonance and of light–matter coupling in cavity QED.
Zeros of Hn are real and bracket the classical turning points ξ≈±2n+1, foreshadowing the WKB picture where oscillatory regions live inside the turning points and exponential tails outside.
Because H=ℏω(N+21), each a† raises the energy by one quantumℏω. Promoting a,a† to mode operators of a field turns the EM field into a set of independent oscillators, with a† creating a photon. Your oscillator algebra is literally second-quantization’s ABCs.
Before tackling hydrogen, we need the geometry engine of 3D quantum mechanics: angular momentum. Rotational symmetry gives conserved L, spherical harmonics Yℓm carry the angles, and the radial Schrödinger equation sees an extra centrifugal barrier. This section builds the operator algebra, ladder machinery, spherical harmonics, and the central-potential separation that we will use on every spherically symmetric problem.
If V(r) is central, the Hamiltonian commutes with all rotations U(R)=exp(−ℏiθ⋅L). Noether’s theorem in the quantum language says
[H,L]=0
Hence ℓ and m label eigenstates. Energies depend on n and ℓ in general, but are m-degenerate by (2ℓ+1) because rotating a solution merely mixes m values inside the same ℓ multiplet.
Special symmetries can enlarge degeneracy. For hydrogen, a hidden Runge–Lenz symmetry makes E depend on n only; for the isotropic 3D oscillator, E depends on N=2nr+ℓ.
For a finite rotation R, the transformed state is U(R)ψ, with coordinates rotating oppositely. On the sphere, the Yℓm furnish irreducible representations of SO(3); under R, components mix within fixed ℓ via Wigner D-matrices Dm′m(ℓ)(R).
Electrons carry orbitalL and spinS angular momenta. The total
J=L+S
obeys the same algebra as any angular momentum. When two angular momenta couple, the allowed totals j follow triangle rules and states expand with Clebsch–Gordan coefficients. We will use these tools for fine structure, selection rules, and multi-particle systems later in the chapter.
(a) Small-r behavior of u(r).
For finite V(0), the dominant terms in the radial equation near r=0 are
−2mℏ2u′′(r)+2mr2ℏ2ℓ(ℓ+1)u(r)≈0
Try u∼rs to get s(s−1)=ℓ(ℓ+1), giving s=ℓ+1 (regular) and s=−ℓ (singular). Keep u∼rℓ+1.
(b) Expectation of L2 in a separable state.
For ψ=R(r)Yℓm(θ,ϕ),
⟨L2⟩=ℏ2ℓ(ℓ+1)
independent of R(r) because angles carry all the L structure.
(c) Degeneracy count at fixed ℓ.
Show that the set {Yℓm}m=−ℓℓ spans a (2ℓ+1)-dimensional irrep by applying L± repeatedly to Yℓℓ and Yℓ−ℓ.
(d) Effective potential sketch.
For V(r)=−k/r (Coulomb), plot Veff(r)=−k/r+ℏ2ℓ(ℓ+1)/(2mr2) to visualize classical-like turning points and bound-state regions for various ℓ.
Starting from ∇2 in spherical coordinates, derive the radial equation for u(r) and identify Veff(r)
Prove the ladder action of L± on ∣ℓm⟩ and normalize the coefficients by demanding ⟨ℓm∣ℓm⟩=1
Verify orthonormality of Yℓm by integrating two explicit low-ℓ examples over the sphere
For a central square well V(r)=−V0 for r<R and 0 otherwise, write the bound-state matching conditions for u(r) at r=R and discuss ℓ-dependence of the spectrum
Show that [H,L]=0 for any V(r) and use it to argue (2ℓ+1) degeneracy in m
This is the first full 3D victory lap of quantum mechanics. With rotational symmetry from §7.4 and the Coulomb potential, we separate variables, solve the radial equation, get exact energy levels
En=−2(4πε0)2ℏ2μe4n2Z2=−2μc2α2n2Z2
and explicit wavefunctions labeled by (n,ℓ,m). Hydrogen wins because the math secretly has extra symmetry; the degeneracy is n2 per n (ignoring spin).
Take a nucleus of charge +Ze and an electron of charge −e. Use the reduced mass
μ≡me+mNmemN
and the Coulomb constant
k≡4πε0Ze2
The time-independent Schrödinger equation with a central potential V(r)=−k/r separates as ψ(r,θ,ϕ)=Rnℓ(r)Yℓm(θ,ϕ). The radial function u(r)≡rR(r) obeys
which factor into angular pieces (Clebsch–Gordan algebra) and radial integrals built from the Rnℓ(r). This reproduces the Balmer/Lyman series patterns and polarization rules.
The Coulomb problem has a conserved Runge–Lenz vector (quantum-symmetrized)
A=2μ1(p×L−L×p)−rkr^
Together, L and A generate an SO(4) symmetry for bound states, explaining why En depends on n only, not on ℓ or m. Break that symmetry (e.g., by fields or relativity) and the degeneracy splits.
Hydrogen is not perfectly degenerate once we add small effects:
Reduced mass. Already included by μ; shifts RH slightly and creates isotope shifts
Fine structure. Relativistic kinetic correction, spin–orbit coupling, and Darwin term split levels by order α2 of the Rydberg; handled by perturbation theory in §7.6
Lamb shift. QED vacuum fluctuations separate 2S and 2P levels (famously nonzero)
Hyperfine. Proton–electron spin coupling yields the 21-cm line; scales with magnetic moments and wavefunction at the origin
Each effect has a clean perturbative footprint, making hydrogen a precision playground.
Real systems are rarely exactly solvable, but many are almost solvable. Perturbation theory is the art of starting from a Hamiltonian you can diagonalize, then adding a small extra term and tracking how energies and states shift. It powers fine structure in hydrogen, Zeeman/Stark effects, chemical shifts, and basically every “small correction” you care about.
Sign check: if H′ mixes n with higher states, denominators are negative so En typically drops.
Selection rules: many matrix elements vanish by symmetry (parity, angular momentum), so first-order shifts can be zero and second-order becomes the leading term.
If H(λ) depends smoothly on a parameter λ and ∣n(λ)⟩ is an exact eigenstate,
dλdEn=⟨n(λ)∂λ∂Hn(λ)⟩
This avoids differentiating messy wavefunctions: compute the expectation of the explicit derivative of H. Great for forces in molecules and for tracking how E slides with external fields.
If several H0 eigenstates share the same E(0), naive denominators blow up. Fix by diagonalizing H′within the degenerate subspace.
Let {∣ϕa⟩}a=1g be an orthonormal basis of a g-fold degenerate eigenspace of H0. Build the secular matrix
Wab≡⟨ϕa∣H′∣ϕb⟩
Diagonalize W to get eigenvalues {λα} and eigenvectors {c(α)}. Then
Eα(1)=λα∣ψα(0)⟩=a=1∑gca(α)∣ϕa⟩
These optimized zeroth-order states already include the dominant mixing; higher orders then proceed as in the nondegenerate case using ∣ψα(0)⟩.
Rule of thumb: choose a basis adapted to the symmetries of H′. You will often find that W block-diagonalizes (e.g., fixed m or fixed parity), making life easy.
For Ze2/4πε0r with Z=1, three relativistic corrections dress H0:
Relativistic kinetic energy
Hrel=−8me3c2p4
Spin–orbit coupling with Thomas factor
HSO=2me2c21r1drdVL⋅S
For Coulomb V(r)=−e2/4πε0r, this is ∝+L⋅S/r3
Darwin term (contact term from Zitterbewegung, nonzero only for ℓ=0)
HD=8me2c2ℏ2∇2V(r)
Combine expectation values using hydrogen eigenstates and angular momentum algebra. The sum depends on n and the totalj=ℓ±21 but not on m, giving the textbook fine-structure shift
ΔEn,j(fs)=2n3μc2(Zα)4(j+211−4n3)
Here μ is the reduced mass and α≡e2/4πε0ℏc. Fine structure lifts the n2 degeneracy down to the 2n degeneracy labeled by j and m.
Add a uniform magnetic field B=Bz^. The leading coupling is
HZ′=μB(Lz+gsSz)B
with gs≈2 and μB=eℏ/2me. In LS coupling, levels are labeled by ∣nℓsjmj⟩ and the shift is packaged as
ΔEZ=μBgJmJB
where the Landé factor
gJ=1+2j(j+1)j(j+1)+s(s+1)−ℓ(ℓ+1)
For s=21, this reproduces the familiar “anomalous” Zeeman patterns. The selection rule for dipole transitions remains ΔmJ=0,±1; polarizations track π and σ± components.
Strong-field (Paschen–Back) note: when μBB beats spin–orbit, L and S uncouple and you diagonalize Lz and Sz instead of J2, but that is beyond first-order weak-field PT.
Parity makes first-order shifts vanish in nondegenerate levels with definite parity (e.g., hydrogen 1s). Thus the leading shift is second order:
ΔE(2)=m=n∑En(0)−Em(0)∣⟨m(0)∣eEz∣n(0)⟩∣2
For hydrogen 1s this yields a quadratic Stark shift
ΔE1s=−21α1sE2,α1s=29a03
where α is the polarizability and a0 the Bohr radius.
Linear Stark in degenerate manifolds. When H0 has degeneracy of opposite parities (hydrogen n≥2), do degenerate PT in the n manifold using parabolic states. First-order shifts appear:
ΔEn,k,m=23nea0Ek
with k=n1−n2 the parabolic quantum number satisfying n1+n2+∣m∣+1=n. For n=2, k=±1 at m=0 split linearly, while ∣m∣=1 states stay unshifted at first order.
Exploit good quantum numbers. Work in a basis that diagonalizes all operators commuting with both H0 and H′
Use Wigner–Eckart to drop angles. For dipole-type perturbations rq, factor angular parts into Clebsch–Gordan coefficients and a reduced matrix element so you only carry one radial integral
Mind selection rules. If first order vanishes by symmetry, skip straight to second order (saves algebra and surprises)
Estimate sizes early. Compare typical matrix elements of H′ to level spacings of H0. If the ratio is not ≪1, perturbation theory will not slay this dragon
Sketch: only ℓ=1 intermediate states contribute because z∝Y10 changes ℓ by ±1. Evaluate the radial sum or use known hydrogenic polarizability.
(b) Zeeman splitting of a 2P3/2 level
With j=23, gJ=34. Allowed mJ={−23,−21,21,23} produce four equally spaced lines with separation μBgJB.
(c) Fine-structure order of magnitude
For n=2, the scale is
∣ΔE(fs)∣∼μc2α4/16
which for hydrogen lands in the 10−4eV range, i.e., tens of GHz—microwave domain.
(d) Two-level avoided crossing via second order
For a pair {∣1⟩,∣2⟩} with E2(0)−E1(0)=Δ and off-diagonal ⟨1∣H′∣2⟩=V, diagonalize
(E1(0)VVE2(0))
to get exact E±=21(E1(0)+E2(0))±21Δ2+4V2. For ∣V∣≪∣Δ∣, recover second-order shifts ±V2/Δ. This is the algebraic skeleton behind many “near-degenerate” perturbations.
Forgetting to orthogonalize in degenerate spaces. Always diagonalize H′ first; otherwise denominators lie to you
Using unperturbed labels after mixing. Once you diagonalize W, the “good” zeroth-order states are the W eigenvectors, not the original basis
Boundary terms in Hellmann–Feynman. HF needs exact eigenstates; variational or truncated bases require adding Pulay-type corrections when the basis depends on λ
Order counting. If first order vanishes, second order can be dominant but still “small.” Check scales before trusting the series
Derive En(1), ∣n(1)⟩, and En(2) from the series ansatz with intermediate normalization
For a 3-fold degenerate level {∣ϕ1⟩,∣ϕ2⟩,∣ϕ3⟩} and perturbation H′, build W, solve the secular equation det(W−λI)=0, and construct the mixed eigenkets
Compute the Landé gJ formula from HZ′ using vector coupling and verify special cases S (ℓ=0) and L-only (s=0) limits
Evaluate the quadratic Stark shift for 1s by explicit summation over np states or by using known closure relations
Starting from Hrel,HSO,HD, show that the combined fine-structure correction depends only on n and j and reproduce ΔEn,j(fs) above