7 Quantum mechanics

7.1 Foundations and Postulates

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.


7.1.1 State space and Dirac essentials

A closed quantum system lives in a complex Hilbert space H\mathcal H. A pure state is a unit vector (a ket) ψH\ket{\psi}\in\mathcal H. Global phase is unphysical: ψ\ket{\psi} and eiϕψe^{i\phi}\ket{\psi} describe the same physical state. The dual ψ\bra{\psi} is the complex-conjugate transpose (a bra). Inner products are ϕψ\braket{\phi}{\psi}.

An orthonormal basis {n}\{\ket{n}\} resolves the identity

I=nnn\mathbb I = \sum_n \ket{n}\bra{n}

For continuous bases (like position), sums become integrals and Kronecker deltas become Dirac deltas

xxdx=I,xx=δ(xx)\int \ket{x}\bra{x}\,dx = \mathbb I,\qquad \braket{x}{x'}=\delta(x-x')

The wavefunction in position space is ψ(x)=xψ\psi(x)=\braket{x}{\psi}, with normalization

ψ(x)2dx=1\int |\psi(x)|^{2}\,dx = 1

Momentum-space amplitudes are Fourier partners

ψ~(p)=pψ=12πeipx/ψ(x)dx\tilde\psi(p)=\braket{p}{\psi}=\frac{1}{\sqrt{2\pi\hbar}}\int e^{-ipx/\hbar}\,\psi(x)\,dx

and invert with the conjugate transform.


7.1.2 Observables and spectra

Every observable AA is a self-adjoint (Hermitian) operator A^\hat A on H\mathcal H. Measurement outcomes are its eigenvalues; post-measurement states are the associated eigenvectors (modulo degeneracy). Spectral resolution:

A^=aaΠa\hat A = \sum_a a\,\Pi_a

where Πa\Pi_a are orthogonal projectors onto the eigenspaces. For continuous spectra, replace sums by integrals A=adΠ(a)A=\int a\,d\Pi(a).

Expectation values in state ψ\ket{\psi}:

Aψ=ψA^ψ\langle A \rangle_\psi = \bra{\psi}\,\hat A\,\ket{\psi}

Variance (ΔA)2=A2A2(\Delta A)^2=\langle A^2\rangle-\langle A\rangle^2 quantifies spread. Noncommuting observables cannot be simultaneously sharp; the general inequality is

ΔAΔB12[A^,B^]\Delta A\,\Delta B \ge \frac{1}{2}\,\big|\langle[\hat A,\hat B]\rangle\big|

Position and momentum satisfy the canonical commutation relation (CCR)

[x^,p^]=i[\hat x,\hat p] = i\hbar

implying ΔxΔp/2\Delta x\,\Delta p\ge \hbar/2.

In the xx-representation,

x^ψ(x)=xψ(x),p^ψ(x)=idψdx\hat x\,\psi(x)=x\,\psi(x),\qquad \hat p\,\psi(x) = -\,i\hbar\,\frac{d\psi}{dx}

so the CCR becomes the familiar differentiation identity.


7.1.3 Measurement postulate (projective version)

If you measure AA in state ψ\ket{\psi}, the probability to get eigenvalue aa is

P(a)=ψΠaψP(a) = \bra{\psi}\,\Pi_a\,\ket{\psi}

and the collapsed post-measurement state is

ΠaψP(a)\frac{\Pi_a\ket{\psi}}{\sqrt{P(a)}}

For degenerate aa, Πa\Pi_a projects to the full eigenspace. More general measurements (POVMs) appear in §7.12; for most of this chapter, projective measurements suffice.


7.1.4 Time evolution (unitary dynamics)

A closed system evolves unitarily. There exists a one-parameter family of unitaries U(t)U(t) with

ψ(t)=U(t)ψ(0),U(t1)U(t2)=U(t1+t2),U(0)=I\ket{\psi(t)} = U(t)\,\ket{\psi(0)},\qquad U(t_1)U(t_2)=U(t_1+t_2),\quad U(0)=\mathbb I

Stone’s theorem says U(t)=exp ⁣(iH^t)U(t)=\exp\!\left(-\tfrac{i}{\hbar}\hat H t\right) for a self-adjoint Hamiltonian H^\hat H. Differentiating gives the time-dependent Schrödinger equation

iddtψ(t)=H^ψ(t)i\hbar\,\frac{d}{dt}\ket{\psi(t)}=\hat H\,\ket{\psi(t)}

In the xx-representation with a standard kinetic-plus-potential Hamiltonian,

iψ(x,t)t=[22m2x2+V(x,t)]ψ(x,t)i\hbar\,\frac{\partial \psi(x,t)}{\partial t} = \left[-\,\frac{\hbar^{2}}{2m}\,\frac{\partial^{2}}{\partial x^{2}} + V(x,t)\right]\psi(x,t)

If HH is time-independent, the stationary states solve

H^n=Enn\hat H\,\ket{n}=E_n\ket{n}

and evolve by phases n(t)=eiEnt/n\ket{n(t)}=e^{-iE_n t/\hbar}\ket{n}. General solutions expand as ψ(t)=ncneiEnt/n\ket{\psi(t)}=\sum_n c_n e^{-iE_n t/\hbar}\ket{n}.


7.1.5 Composite systems and entanglement

For subsystems AA and BB, states live in the tensor product HAHB\mathcal H_A\otimes\mathcal H_B. If ψA\ket{\psi_A} and ψB\ket{\psi_B} are states, the product state is ψAψB\ket{\psi_A}\otimes\ket{\psi_B}. But generic states are entangled, e.g.,

Ψ=12(0A1B+1A0B)\ket{\Psi} = \frac{1}{\sqrt2}\left(\ket{0}_A\ket{1}_B + \ket{1}_A\ket{0}_B\right)

which cannot be factored. Observables act as AIA\otimes \mathbb I or IB\mathbb I\otimes B. Partial information is described by the reduced density operator (see below).


7.1.6 Density operators (mixed states, quick intro)

Not all states are pure; ignorance or decoherence yields mixed states described by positive, unit-trace operators ρ\rho with expectation values

A=Tr(ρA^)\langle A\rangle = \mathrm{Tr}(\rho\,\hat A)

Pure states are projectors ρ=ψψ\rho=\ket{\psi}\bra{\psi} with ρ2=ρ\rho^{2}=\rho. For a composite ρAB\rho_{AB}, the state of AA alone is the partial trace

ρA=TrBρAB\rho_A = \mathrm{Tr}_B\,\rho_{AB}

Unitary evolution is ρ(t)=U(t)ρ(0)U(t)\rho(t)=U(t)\rho(0)U^\dagger(t). Projective measurement updates are ρΠaρΠa/Tr(Πaρ)\rho\mapsto \Pi_a\rho\,\Pi_a/\mathrm{Tr}(\Pi_a\rho) conditioned on outcome aa.


7.1.7 Probability current and continuity

With ψ(x,t)\psi(x,t) normalized, define the probability density ρ(x,t)=ψ(x,t)2\rho(x,t)=|\psi(x,t)|^{2} and current

j(x,t)=2mi[ψψx(ψx)ψ]j(x,t)=\frac{\hbar}{2mi}\left[\psi^{\ast}\frac{\partial \psi}{\partial x}-\left(\frac{\partial \psi^{\ast}}{\partial x}\right)\psi\right]

They obey a continuity equation

ρt+jx=0\frac{\partial \rho}{\partial t} + \frac{\partial j}{\partial x} = 0

so total probability is conserved. In 3D, replace derivatives by gradients and divergences.


Proof

We assume a real potential V(x,t)V(x,t) (no gain/loss). Define

ρ(x,t)=ψ(x,t)2=ψ(x,t)ψ(x,t),j(x,t)=2mi[ψxψ(xψ)ψ].\rho(x,t)=|\psi(x,t)|^{2}=\psi^{\ast}(x,t)\,\psi(x,t),\qquad j(x,t)=\frac{\hbar}{2mi}\Big[\psi^{\ast}\partial_x\psi-(\partial_x\psi^{\ast})\psi\Big].

The time-dependent Schrödinger equation and its complex conjugate are

itψ=22mx2ψ+Vψ,itψ=22mx2ψ+Vψ.i\hbar\,\partial_t\psi =-\frac{\hbar^{2}}{2m}\,\partial_x^{2}\psi+V\psi, \qquad -\,i\hbar\,\partial_t\psi^{\ast} =-\frac{\hbar^{2}}{2m}\,\partial_x^{2}\psi^{\ast}+V\psi^{\ast}.

Step 1: Differentiate ρ\rho in time

tρ=t(ψψ)=(tψ)ψ+ψ(tψ).\partial_t\rho=\partial_t(\psi^{\ast}\psi) =(\partial_t\psi^{\ast})\psi+\psi^{\ast}(\partial_t\psi).

Use the equations to substitute tψ\partial_t\psi and tψ\partial_t\psi^{\ast}:

tψ=1i ⁣(22mx2ψ+Vψ),tψ=1i ⁣(22mx2ψ+Vψ).\partial_t\psi=\frac{1}{i\hbar}\!\left(-\frac{\hbar^{2}}{2m}\partial_x^{2}\psi+V\psi\right), \qquad \partial_t\psi^{\ast}=-\frac{1}{i\hbar}\!\left(-\frac{\hbar^{2}}{2m}\partial_x^{2}\psi^{\ast}+V\psi^{\ast}\right).

Therefore

tρ=[1i ⁣(22mx2ψ+Vψ)]ψ+ψ[1i ⁣(22mx2ψ+Vψ)]=1i[22m(ψx2ψψx2ψ)+VψψVψψ].\begin{aligned} \partial_t\rho &=\left[-\frac{1}{i\hbar}\!\left(-\frac{\hbar^{2}}{2m}\partial_x^{2}\psi^{\ast}+V\psi^{\ast}\right)\right]\psi +\psi^{\ast}\left[\frac{1}{i\hbar}\!\left(-\frac{\hbar^{2}}{2m}\partial_x^{2}\psi+V\psi\right)\right] \\ &=\frac{1}{i\hbar}\left[ -\frac{\hbar^{2}}{2m}\big(\psi^{\ast}\partial_x^{2}\psi-\psi\,\partial_x^{2}\psi^{\ast}\big) +\cancel{V\psi^{\ast}\psi}-\cancel{V\psi^{\ast}\psi} \right]. \end{aligned}

Since the potential cancels,

tρ=2mi(ψx2ψψx2ψ).\partial_t\rho =-\,\frac{\hbar}{2mi}\Big(\psi^{\ast}\partial_x^{2}\psi-\psi\,\partial_x^{2}\psi^{\ast}\Big).

Step 2: Rewrite as divergence form

Using product rule identities:

x(ψxψ)=(xψ)(xψ)+ψx2ψ,\partial_x\big(\psi^{\ast}\partial_x\psi\big) =(\partial_x\psi^{\ast})(\partial_x\psi)+\psi^{\ast}\partial_x^{2}\psi, x(ψxψ)=(xψ)(xψ)+ψx2ψ.\partial_x\big(\psi\,\partial_x\psi^{\ast}\big) =(\partial_x\psi)(\partial_x\psi^{\ast})+\psi\,\partial_x^{2}\psi^{\ast}.

Subtracting gives

x(ψxψψxψ)=ψx2ψψx2ψ.\partial_x\big(\psi^{\ast}\partial_x\psi-\psi\,\partial_x\psi^{\ast}\big) =\psi^{\ast}\partial_x^{2}\psi-\psi\,\partial_x^{2}\psi^{\ast}.

Thus

tρ=2mix ⁣(ψxψψxψ)=x ⁣[2mi(ψxψ(xψ)ψ)].\partial_t\rho =-\,\frac{\hbar}{2mi}\,\partial_x\!\left(\psi^{\ast}\partial_x\psi-\psi\,\partial_x\psi^{\ast}\right) =-\,\partial_x\!\left[\frac{\hbar}{2mi}\Big(\psi^{\ast}\partial_x\psi-(\partial_x\psi^{\ast})\psi\Big)\right].

Recognizing the bracket as j(x,t)j(x,t):

  tρ+xj=0  \boxed{\;\partial_t\rho+\partial_x j=0\;}

This is the continuity equation ensuring probability conservation.

Remarks

  • With boundary condition j0j\to 0 at ±\pm\infty:
ddtρ(x,t)dx=0,\frac{d}{dt}\int_{-\infty}^{\infty}\rho(x,t)\,dx=0,

so the total probability remains 1.

  • In dd-dimensions:
j(r,t)=2mi(ψψ(ψ)ψ),tρ+ ⁣j=0.\mathbf j(\boldsymbol r,t)=\frac{\hbar}{2mi}\big(\psi^{\ast}\nabla\psi-(\nabla\psi^{\ast})\psi\big),\qquad \partial_t\rho+\nabla\!\cdot\boldsymbol j=0.
  • If VV is complex (absorbing or amplifying medium), the continuity equation gains a source/sink term.

7.1.8 Ehrenfest’s theorem (classical limit peeks through)

Quantum expectation values evolve nearly classically when wave packets are narrow:

ddtx^=p^m,ddtp^=Vx\frac{d}{dt}\langle \hat x \rangle = \frac{\langle \hat p \rangle}{m},\qquad \frac{d}{dt}\langle \hat p \rangle = -\,\left\langle \frac{\partial V}{\partial x} \right\rangle

If V(x)V(x) is smooth over the packet width, V/xV(x)/x\langle \partial V/\partial x\rangle \approx \partial V(\langle x\rangle)/\partial x, and the centroid follows Newton’s law. This is the correspondence principle in differential form.


7.1.9 Stationary states, nodes, and orthogonality

For time-independent VV, eigenfunctions ϕn(x)\phi_n(x) solve

[22md2dx2+V(x)]ϕn(x)=Enϕn(x)\left[-\,\frac{\hbar^{2}}{2m}\frac{d^{2}}{dx^{2}} + V(x)\right]\phi_n(x) = E_n\,\phi_n(x)

Distinct eigenvalues are orthogonal

ϕm(x)ϕn(x)dx=0(mn)\int \phi_m^{\ast}(x)\,\phi_n(x)\,dx=0\quad (m\ne n)

Real, confining potentials produce discrete spectra with increasing node counts as nn rises. Bound vs scattering states will be treated carefully in §7.10.


7.1.10 Symmetries and conserved quantities

If a unitary UU leaves the Hamiltonian invariant, UH^U=H^U^\dagger \hat H U=\hat H, then there is a conserved generator GG (Noether-style). Infinitesimal unitaries U=exp(iϵG/)U=\exp(-i\epsilon G/\hbar) imply

ddtG=i[H^,G]\frac{d}{dt}\langle G\rangle = \frac{i}{\hbar}\langle [\hat H,G]\rangle

so [H,G]=0G[H,G]=0 \Rightarrow \langle G\rangle is constant. Translational invariance conserves pp, rotational invariance conserves L\boldsymbol L, and time invariance (no explicit tt in HH) conserves energy.


7.1.11 Quick dictionary: pictures and operators

  • Schrödinger picture: states evolve, operators are fixed (unless explicitly time-dependent)
  • Heisenberg picture: operators carry time, states are fixed. The equation of motion reads
dA^Hdt=i[H^,A^H]+(A^t)H\frac{d\hat A_H}{dt} = \frac{i}{\hbar}\,[\hat H,\hat A_H] + \left(\frac{\partial \hat A}{\partial t}\right)_H

Both pictures are unitarily equivalent. Choose the one that makes the algebra short.


7.1.12 Worked mini-examples

(a) Minimum-uncertainty Gaussian

Take

ψ(x)=(1πd2)1/4exp ⁣(x22d2)\psi(x) = \left(\frac{1}{\pi d^{2}}\right)^{1/4}\exp\!\left(-\frac{x^{2}}{2d^{2}}\right)

Then Δx=d/2\Delta x=d/\sqrt{2} and Δp=/(2d)\Delta p=\hbar/(\sqrt{2}d), hence ΔxΔp=/2\Delta x\,\Delta p=\hbar/2. Time-evolving under a free Hamiltonian broadens d(t)d(t) while keeping the product fixed.

(b) Free-particle propagator in 1D (form)

The kernel K(x,t;x,0)K(x,t;x',0) solving itψ=(2/2m)x2ψi\hbar\partial_t\psi=-(\hbar^2/2m)\partial_x^2\psi with ψ(x,0)=ψ0(x)\psi(x,0)=\psi_0(x) is

K(x,t;x,0)=m2πit  exp ⁣(im(xx)22t)K(x,t;x',0)=\sqrt{\frac{m}{2\pi i\hbar t}}\;\exp\!\left(\frac{i m (x-x')^{2}}{2\hbar t}\right)

and ψ(x,t)=K(x,t;x,0)ψ0(x)dx\psi(x,t)=\int K(x,t;x',0)\,\psi_0(x')\,dx'. We will use this in §7.9.

(c) Commutator algebra warmup

With [x^,p^]=i[\hat x,\hat p]=i\hbar, show

[x^,p^2]=2ip^[\hat x,\hat p^{2}] = 2i\hbar\,\hat p

by [x^,AB]=A[x^,B]+[x^,A]B[\hat x,AB]=A[\hat x,B]+[\hat x,A]B. This identity drives many quick manipulations in later sections.


7.1.13 What to keep in RAM

  • States live in a Hilbert space; global phase is fluff
  • Observables are Hermitian operators; outcomes are eigenvalues; expectation A=ψAψ\langle A\rangle=\bra{\psi}A\ket{\psi}
  • Measurement is probabilistic with projectors, and post-measurement states are projected and renormalized
  • Closed systems evolve unitarily by U(t)=eiHt/U(t)=e^{-iHt/\hbar}; Schrödinger vs Heisenberg are just different UIs
  • CCR [x^,p^]=i[\hat x,\hat p]=i\hbar \Rightarrow uncertainty and Fourier duality of xx and pp
  • Composites live in tensor products; entanglement is the default, not the exception
  • Mixed states need density operators; partial traces describe subsystems

7.2 The Schrödinger Equation: Waves, Probability, and Stationary States

Quantum mechanics turns dynamics into linear wave evolution. The central object is the wavefunction ψ(r,t)\psi(\mathbf 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.


7.2.1 Time-dependent Schrödinger equation and normalization

For a particle of mass mm in potential V(r,t)V(\mathbf r,t),

iψ(r,t)t=[22m2+V(r,t)]ψ(r,t)i\hbar\,\frac{\partial \psi(\mathbf r,t)}{\partial t} = \left[-\,\frac{\hbar^2}{2m}\,\nabla^2 + V(\mathbf r,t)\right]\psi(\mathbf r,t)

The probabilistic reading demands unit normalization

R3ψ(r,t)2d3r=1\int_{\mathbb R^3} |\psi(\mathbf r,t)|^2\,d^3r = 1

Dimension check: in 3D, ψ2|\psi|^2 is a probability density, so [ψ]=length3/2[\psi]=\text{length}^{-3/2}.


7.2.2 Continuity equation and probability current

Schrödinger’s equation implies local probability conservation. Define the probability current

j(r,t)=2mi[ψψψψ]\boldsymbol j(\boldsymbol r,t) = \frac{\hbar}{2mi}\left[\psi^{\ast}\,\nabla\psi - \psi\,\nabla\psi^{\ast}\right]

Then

tψ2+j=0\frac{\partial}{\partial t}|\psi|^{2} + \nabla\cdot \boldsymbol j = 0

With electromagnetic fields via minimal coupling p^p^qA\hat{\boldsymbol p}\to \hat{\boldsymbol p}-q\boldsymbol A and VV+qϕV\to V+q\phi, the current becomes

j=2mi[ψψψψ]qmAψ2\boldsymbol j = \frac{\hbar}{2mi}\left[\psi^{\ast}\,\nabla\psi - \psi\,\nabla\psi^{\ast}\right] - \frac{q}{m}\,\boldsymbol A\,|\psi|^{2}

showing gauge-covariant flow.


7.2.3 Observables and expectation values

Promote classical quantities to operators acting on ψ\psi

x^=r,p^=i\hat{\boldsymbol x}=\boldsymbol r,\qquad \hat{\boldsymbol p}=-\,i\hbar\,\nabla

The expectation of an observable A^\hat A is

A(t)=ψ(r,t)A^ψ(r,t)d3r\langle A \rangle(t) = \int \psi^{\ast}(\boldsymbol r,t)\,\hat A\,\psi(\boldsymbol r,t)\,d^{3}r

Physical observables correspond to self-adjoint operators so that A\langle A\rangle is real and the spectral theorem applies. Canonical commutation

[x^i,p^j]=iδij[\hat x_i,\hat p_j]=i\hbar\,\delta_{ij}

implies the uncertainty bound ΔxΔp/2\Delta x\,\Delta p \ge \hbar/2.


7.2.4 Stationary states and the time-independent equation

If V(r)V(\boldsymbol r) is time-independent, seek separable solutions

ψ(r,t)=ϕ(r)eiEt/\psi(\boldsymbol r,t) = \phi(\boldsymbol r)\,e^{-iEt/\hbar}

Plugging in gives the time-independent Schrödinger equation

[22m2+V(r)]ϕ(r)=Eϕ(r)\left[-\,\frac{\hbar^{2}}{2m}\,\nabla^{2} + V(\boldsymbol r)\right]\phi(\boldsymbol r) = E\,\phi(\boldsymbol r)

Bound-state eigenvalues {En}\{E_n\} are discrete under confining VV, and the corresponding {ϕn}\{\phi_n\} can be chosen orthonormal

ϕmϕnd3r=δmn\int \phi_m^{\ast}\phi_n\,d^{3}r = \delta_{mn}

Any solution evolves as a superposition ψ(r,t)=ncnϕn(r)eiEnt/\psi(\boldsymbol r,t)=\sum_n c_n \phi_n(\boldsymbol r)\,e^{-iE_n t/\hbar} with ncn2=1\sum_n |c_n|^{2}=1.


7.2.5 Boundary conditions and self-adjointness

The Hamiltonian is self-adjoint only for states obeying appropriate boundary conditions. In practice:

  • For finite, piecewise-smooth VV, both ϕ\phi and its first derivative are continuous across finite steps
  • At infinite walls (e.g., ϕ=0\phi=0 at the boundary), wavefunctions vanish
  • For δ\delta-like spikes, ϕ\phi is continuous but its derivative has a prescribed jump proportional to the spike strength
  • Normalizable bound states must satisfy ϕ0\phi\to 0 fast enough at infinity

These conditions kill boundary terms so that ϕ(H^ψ)=(H^ϕ)ψ\int \phi^{\ast}(\hat H\psi)=\int (\hat H\phi)^{\ast}\psi.


7.2.6 Canonical 1D examples (fast but useful)

(a) Free particle. V=0V=0 gives plane waves

ψk(x,t)=12πei(kxωt),ω=k22m\psi_k(x,t)=\frac{1}{\sqrt{2\pi}}\,e^{i(kx-\omega t)},\qquad \omega=\frac{\hbar k^{2}}{2m}

Build wave packets by superposing kk values; packet center moves with group velocity vg=k/mv_g=\hbar k/m and spreads over time.

(b) Infinite square well of width LL on 0<x<L0<x<L:

ϕn(x)=2Lsin ⁣(nπxL),En=2π2n22mL2,n=1,2,\phi_n(x)=\sqrt{\frac{2}{L}}\,\sin\!\left(\frac{n\pi x}{L}\right),\qquad E_n=\frac{\hbar^{2}\pi^{2}n^{2}}{2mL^{2}},\quad n=1,2,\dots

Real eigenfunctions \Rightarrow current j=0j=0 in stationary states. Superpositions can carry current.

(c) Finite step and barriers. Impose continuity of ϕ\phi and ϕ\phi' to get reflection RR and transmission TT; probability flux conservation enforces R+T=1R+T=1 for real potentials. For barriers, tunneling gives Te2κaT\sim e^{-2\kappa a} with κ=2m(V0E)/\kappa=\sqrt{2m(V_0-E)}/\hbar and width aa.


7.2.7 Probability current in 1D scattering

For a stationary scattering state ψ(x)=Aineikx+Areeikx\psi(x)=A_{\text{in}}e^{ikx}+A_{\text{re}}e^{-ikx} on the left and ψ(x)=Atreikx\psi(x)=A_{\text{tr}}e^{ik'x} on the right, the fluxes

jin=kmAin2,jre=kmAre2,jtr=kmAtr2j_{\text{in}}=\frac{\hbar k}{m}|A_{\text{in}}|^{2},\quad j_{\text{re}}=\frac{\hbar k}{m}|A_{\text{re}}|^{2},\quad j_{\text{tr}}=\frac{\hbar k'}{m}|A_{\text{tr}}|^{2}

define reflection and transmission

R=jrejin,T=jtrjinR=\frac{j_{\text{re}}}{j_{\text{in}}},\qquad T=\frac{j_{\text{tr}}}{j_{\text{in}}}

For real VV, R+T=1R+T=1 follows from the continuity equation.

Detailed Explanation

Setup (incoming + reflected on the left, transmitted on the right):

For xx \to -\infty:

ψL(x)=Aineikx+Areeikx,k=2m(EVL)\psi_L(x) = A_{\text{in}} e^{ikx} + A_{\text{re}} e^{-ikx}, \qquad k = \frac{\sqrt{2m(E - V_L)}}{\hbar}

For x+x \to +\infty:

ψR(x)=Atreikx,k=2m(EVR)\psi_R(x) = A_{\text{tr}} e^{ik'x}, \qquad k' = \frac{\sqrt{2m(E - V_R)}}{\hbar}

Here VL,VRV_L, V_R are constants, with E>max{VL,VR}E > \max\{V_L, V_R\}.

1) Continuity equation and probability flux

The time-dependent Schrödinger equation is

iψt=22m2ψx2+V(x)ψ.i\hbar \frac{\partial \psi}{\partial t} = -\frac{\hbar^2}{2m}\frac{\partial^2 \psi}{\partial x^2} + V(x)\psi.

Its complex conjugate is

iψt=22m2ψx2+V(x)ψ.-\,i\hbar \frac{\partial \psi^\ast}{\partial t} = -\frac{\hbar^2}{2m}\frac{\partial^2 \psi^\ast}{\partial x^2} + V(x)\psi^\ast.

Multiplying the first by ψ\psi^\ast and the second by ψ\psi, then subtracting, gives

i(ψψt+ψψt)=22m[ψ2ψx2ψ2ψx2]+(VV)ψ2.i\hbar \Bigl(\psi^\ast \frac{\partial \psi}{\partial t} + \psi \frac{\partial \psi^\ast}{\partial t}\Bigr) = -\frac{\hbar^2}{2m}\Bigl[\psi^\ast \frac{\partial^2 \psi}{\partial x^2} - \psi \frac{\partial^2 \psi^\ast}{\partial x^2}\Bigr] + \cancel{(V - V)|\psi|^2}.

The left-hand side becomes

itψ2,i\hbar \frac{\partial}{\partial t}|\psi|^2,

and the right-hand side can be written as a total derivative:

22mx[ψψx(ψx)ψ].-\frac{\hbar^2}{2m}\frac{\partial}{\partial x} \Bigl[\psi^\ast \frac{\partial \psi}{\partial x} - \Bigl(\frac{\partial \psi^\ast}{\partial x}\Bigr)\psi\Bigr].

Thus, for real V(x)V(x),

tψ2+xj(x,t)=0,\frac{\partial}{\partial t}|\psi|^2 + \frac{\partial}{\partial x}\,j(x,t) = 0,

with the probability flux

j(x,t)=2mi[ψψx(ψx)ψ].j(x,t) = \frac{\hbar}{2mi}\Bigl[\psi^\ast \frac{\partial \psi}{\partial x} - \Bigl(\frac{\partial \psi^\ast}{\partial x}\Bigr)\psi\Bigr].

For a stationary state ψ(x,t)=ψ(x)eiEt/\psi(x,t) = \psi(x)e^{-iEt/\hbar},
tψ2=0\partial_t |\psi|^2 = 0, so

ddxj(x)=0j(x)=constant.\frac{d}{dx}j(x) = 0 \quad \Rightarrow \quad j(x) = \text{constant}.

2) Flux on the left side

On the left,

ψL(x)=Aineikx+Areeikx.\psi_L(x) = A_{\text{in}} e^{ikx} + A_{\text{re}} e^{-ikx}.

Derivative and conjugate:

dψLdx=ikAineikxikAreeikx,ψL=Aineikx+Areeikx.\frac{d\psi_L}{dx} = ikA_{\text{in}} e^{ikx} - ikA_{\text{re}} e^{-ikx}, \qquad \psi_L^\ast = A_{\text{in}}^\ast e^{-ikx} + A_{\text{re}}^\ast e^{ikx}.

Compute

ψLdψLdx=(Aineikx+Areeikx)(ikAineikxikAreeikx).\psi_L^\ast \frac{d\psi_L}{dx} = (A_{\text{in}}^\ast e^{-ikx} + A_{\text{re}}^\ast e^{ikx}) (ikA_{\text{in}} e^{ikx} - ikA_{\text{re}} e^{-ikx}).

Expansion gives

=ikAin2ikAre2+ik(AinAree2ikx+AreAine2ikx).= ik|A_{\text{in}}|^2 - ik|A_{\text{re}}|^2 + ik\bigl(-A_{\text{in}}^\ast A_{\text{re}} e^{-2ikx} + A_{\text{re}}^\ast A_{\text{in}} e^{2ikx}\bigr).

Similarly,

(dψLdx)ψL=(ikAineikx+ikAreeikx)(Aineikx+Areeikx),\Bigl(\frac{d\psi_L^\ast}{dx}\Bigr)\psi_L = (-ikA_{\text{in}}^\ast e^{-ikx} + ikA_{\text{re}}^\ast e^{ikx}) (A_{\text{in}} e^{ikx} + A_{\text{re}} e^{-ikx}),

which expands to

=ikAin2+ikAre2+ik(AinAree2ikx+AreAine2ikx).= -ik|A_{\text{in}}|^2 + ik|A_{\text{re}}|^2 + ik\bigl(-A_{\text{in}}^\ast A_{\text{re}} e^{-2ikx} + A_{\text{re}}^\ast A_{\text{in}} e^{2ikx}\bigr).

Subtracting cancels the interference terms:

ψLdψLdx(dψLdx)ψL=2ik(Ain2Are2).\psi_L^\ast \frac{d\psi_L}{dx} - \Bigl(\frac{d\psi_L^\ast}{dx}\Bigr)\psi_L = 2ik\bigl(|A_{\text{in}}|^2 - |A_{\text{re}}|^2\bigr).

Therefore

jL=2mi2ik(Ain2Are2)=km(Ain2Are2).j_L = \frac{\hbar}{2mi}\cdot 2ik\bigl(|A_{\text{in}}|^2 - |A_{\text{re}}|^2\bigr) = \frac{\hbar k}{m}\bigl(|A_{\text{in}}|^2 - |A_{\text{re}}|^2\bigr).

Defining

jin=kmAin2,jre=kmAre2,j_{\text{in}} = \frac{\hbar k}{m}|A_{\text{in}}|^2, \qquad j_{\text{re}} = \frac{\hbar k}{m}|A_{\text{re}}|^2,

we find

jL=jinjre.j_L = j_{\text{in}} - j_{\text{re}}.

3) Flux on the right side

On the right,

ψR(x)=Atreikx.\psi_R(x) = A_{\text{tr}} e^{ik'x}.

Derivative and conjugate:

dψRdx=ikAtreikx,ψR=Atreikx.\frac{d\psi_R}{dx} = ik'A_{\text{tr}} e^{ik'x}, \qquad \psi_R^\ast = A_{\text{tr}}^\ast e^{-ik'x}.

Thus,

ψRdψRdx(dψRdx)ψR=2ikAtr2.\psi_R^\ast \frac{d\psi_R}{dx} - \Bigl(\frac{d\psi_R^\ast}{dx}\Bigr)\psi_R = 2ik'|A_{\text{tr}}|^2.

Hence

jR=2mi2ikAtr2=kmAtr2jtr.j_R = \frac{\hbar}{2mi}\cdot 2ik'|A_{\text{tr}}|^2 = \frac{\hbar k'}{m}|A_{\text{tr}}|^2 \equiv j_{\text{tr}}.

4) Conservation of flux R+T=1\Rightarrow R+T=1

Stationarity and real V(x)V(x) imply j(x)j(x) is constant:

jL=jR.j_L = j_R.

Therefore,

jinjre=jtrjin=jre+jtr.j_{\text{in}} - j_{\text{re}} = j_{\text{tr}} \quad \Rightarrow \quad j_{\text{in}} = j_{\text{re}} + j_{\text{tr}}.

Define reflection and transmission:

R=jrejin,T=jtrjin.R = \frac{j_{\text{re}}}{j_{\text{in}}}, \qquad T = \frac{j_{\text{tr}}}{j_{\text{in}}}.

Finally,

1=jrejin+jtrjinR+T=1.1 = \frac{j_{\text{re}}}{j_{\text{in}}} + \frac{j_{\text{tr}}}{j_{\text{in}}} \quad \Rightarrow \quad \boxed{R + T = 1}.

Remark:
If V(x)V(x) is complex (absorptive), the continuity equation gains a source term

jx=2ImV(x)ψ(x)2,\frac{\partial j}{\partial x} = \frac{2}{\hbar}\,\operatorname{Im}V(x)\,|\psi(x)|^2,

so in general R+T<1R+T<1.


7.2.8 Ehrenfest’s theorem (centers move classically)

Even in full quantum dynamics,

ddtx^=p^m,ddtp^=Vx\frac{d}{dt}\langle \hat x \rangle = \frac{\langle \hat p \rangle}{m},\qquad \frac{d}{dt}\langle \hat p \rangle = -\,\left\langle \frac{\partial V}{\partial x} \right\rangle

When ψ\psi is narrow and VV is smooth, V/xV(x)/x\langle \partial V/\partial x\rangle\approx \partial V(\langle x\rangle)/\partial x, so the packet centroid obeys Newton’s law while spread and interference track the departures.


7.2.9 Nodes, phases, and currents in bound states

For time-independent, real VV, stationary bound eigenfunctions can be chosen real, so

j=0\boldsymbol j = \boldsymbol 0

The quantum number nn counts nodes (zeros) of ϕn\phi_n in 1D wells; higher nn 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.


7.2.10 Units and rescalings you will actually use

Set natural or atomic units to declutter:

  • Natural: =c=1\hbar=c=1 for relativistic work
  • Atomic (nonrelativistic electrons): =me=e=4πε0=1\hbar=m_e=e=4\pi\varepsilon_0=1, so energies in Hartrees and lengths in Bohr radii

Dimensionless rescaling often turns H^\hat H into a minimal form with one control parameter, improving numerics and insight.


7.2.11 Worked mini-examples

(a) Normalization of a Gaussian packet

Take ψ(x,0)=Ce(xx0)2/(4σ2)eik0x\psi(x,0)=C\,e^{-(x-x_0)^2/(4\sigma^2)}\,e^{ik_0 x}. Enforce ψ2dx=1\int|\psi|^2 dx=1 to get

C=1(2πσ2)1/4C=\frac{1}{(2\pi\sigma^{2})^{1/4}}

Time evolution preserves normalization.

(b) Expectation values in the well

For ϕn\phi_n in the infinite well, symmetry gives x=L/2\langle x\rangle = L/2 and

x2=L23L22π2n2\langle x^{2}\rangle = \frac{L^{2}}{3} - \frac{L^{2}}{2\pi^{2}n^{2}}

so the spatial spread shrinks slowly with nn.

(c) Jump across a delta spike

For V(x)=λδ(x)V(x)=\lambda\,\delta(x), continuity ϕ(0+)=ϕ(0)\phi(0^{+})=\phi(0^{-}) and derivative jump

ϕ(0+)ϕ(0)=2mλ2ϕ(0)\phi'(0^{+})-\phi'(0^{-})=\frac{2m\lambda}{\hbar^{2}}\,\phi(0)

lead to one bound state at E=mλ2/(22)E=-m\lambda^{2}/(2\hbar^{2}) for λ<0\lambda<0.


7.3 The Quantum Harmonic Oscillator: Algebra, Waves, and Coherent Motion

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.


7.3.1 Setup and natural scales

Consider a 1D particle of mass mm with

V(x)=12mω2x2V(x) = \frac{1}{2} m\omega^{2} x^{2}

Two natural scales declutter the algebra:

x0mω,p0mωx_{0} \equiv \sqrt{\frac{\hbar}{m\omega}},\qquad p_{0} \equiv \sqrt{m\hbar\omega}

Introduce a dimensionless coordinate ξx/x0\xi \equiv x/x_{0}. We will hop between xx and ξ\xi as convenient.


7.3.2 Differential solution: Hermite world

The time-independent Schrödinger equation

22md2ϕdx2+12mω2x2ϕ=Eϕ-\frac{\hbar^{2}}{2m}\,\frac{d^{2}\phi}{dx^{2}} + \frac{1}{2} m\omega^{2} x^{2}\,\phi = E\,\phi

becomes, in ξ\xi,

d2ϕdξ2+(2ϵξ2)ϕ=0,ϵEω\frac{d^{2}\phi}{d\xi^{2}} + \left(2\epsilon - \xi^{2}\right)\phi = 0,\qquad \epsilon \equiv \frac{E}{\hbar\omega}

Normalizable solutions exist only when the series terminates, yielding Hermite polynomials Hn(ξ)H_{n}(\xi) and the quantized spectrum

En=ω(n+12),n=0,1,2,E_{n} = \hbar\omega\left(n+\frac{1}{2}\right),\qquad n=0,1,2,\dots

The normalized eigenfunctions are

ϕn(x)=1π1/42nn!x0  Hn ⁣(xx0)exp ⁣(x22x02)\phi_{n}(x) = \frac{1}{\pi^{1/4}\,\sqrt{2^{n} n!\,x_{0}}}\;H_{n}\!\left(\frac{x}{x_{0}}\right)\,\exp\!\left(-\frac{x^{2}}{2x_{0}^{2}}\right)

Orthogonality and completeness follow from Hermite properties; nodes increase with nn and sit where HnH_{n} changes sign.


7.3.3 Algebraic solution: ladder operators in 3 lines

Define

a12(xx0+ipp0),a12(xx0ipp0)a \equiv \frac{1}{\sqrt{2}}\left(\frac{x}{x_{0}} + i\,\frac{p}{p_{0}}\right),\qquad a^{\dagger} \equiv \frac{1}{\sqrt{2}}\left(\frac{x}{x_{0}} - i\,\frac{p}{p_{0}}\right)

These satisfy

[a,a]=1[a,a^{\dagger}] = 1

The Hamiltonian becomes

H=ω(aa+12)H = \hbar\omega\left(a^{\dagger}a + \frac{1}{2}\right)

with number operator NaaN\equiv a^{\dagger}a and eigenstates n\ket{n} such that

Nn=nn,an=nn1,an=n+1n+1N\ket{n}=n\ket{n},\quad a\ket{n}=\sqrt{n}\,\ket{n-1},\quad a^{\dagger}\ket{n}=\sqrt{n+1}\,\ket{n+1}

Algebra alone gives En=ω(n+12)E_{n}=\hbar\omega(n+\tfrac12), no differential equations needed.


7.3.4 Ground state and zero-point energy

The ground state obeys a0=0a\ket{0}=0. In xx-space this reads

(xx0+ipp0)ϕ0(x)=0\left(\frac{x}{x_{0}} + i\,\frac{p}{p_{0}}\right)\phi_{0}(x) = 0

With pixp\to -i\hbar\,\partial_{x} you recover the Gaussian

ϕ0(x)=1π1/4x0exp ⁣(x22x02)\phi_{0}(x) = \frac{1}{\pi^{1/4}\,\sqrt{x_{0}}}\,\exp\!\left(-\frac{x^{2}}{2x_{0}^{2}}\right)

Unavoidably, E0=12ωE_{0}=\tfrac12\hbar\omega—the zero-point energy. No potential minimum is truly at rest in quantum mechanics.


7.3.5 Position and momentum matrix elements

Using x=x02(a+a)x = \tfrac{x_{0}}{\sqrt{2}}(a+a^{\dagger}) and p=p02(aa)/ip = \tfrac{p_{0}}{\sqrt{2}}(a^{\dagger}-a)/i,

nxn±1=x02n+1δn,n+1+x02nδn,n1\langle n|x|n\pm 1\rangle = \frac{x_{0}}{\sqrt{2}}\sqrt{n+1}\,\delta_{n',\,n+1} + \frac{x_{0}}{\sqrt{2}}\sqrt{n}\,\delta_{n',\,n-1} npn±1=p0i2n+1δn,n+1p0i2nδn,n1\langle n|p|n\pm 1\rangle = \frac{p_{0}}{i\sqrt{2}}\sqrt{n+1}\,\delta_{n',\,n+1} - \frac{p_{0}}{i\sqrt{2}}\sqrt{n}\,\delta_{n',\,n-1}

Selection rule: only Δn=±1\Delta n=\pm 1 connect via xx or pp, the oscillator’s dipole pattern.


7.3.6 Uncertainty and virial: clean checks

For the ground state,

Δx2=x022=2mω,Δp2=p022=mω2,ΔxΔp=2\Delta x^{2} = \frac{x_{0}^{2}}{2} = \frac{\hbar}{2m\omega},\qquad \Delta p^{2} = \frac{p_{0}^{2}}{2} = \frac{m\hbar\omega}{2},\qquad \Delta x\,\Delta p = \frac{\hbar}{2}

A minimal-uncertainty state. For any stationary n\ket{n} the quantum virial theorem gives

T=V=En2\langle T\rangle = \langle V\rangle = \frac{E_{n}}{2}

so half the energy sits in kinetic, half in potential—exactly as the classical time average.


7.3.7 Coherent states: quantum that cosplays classical

Define coherent states as eigenstates of aa:

aα=ααa\ket{\alpha} = \alpha\ket{\alpha}

Expand in number states to get

α=eα2/2n=0αnn!n\ket{\alpha} = e^{-|\alpha|^{2}/2}\sum_{n=0}^{\infty}\frac{\alpha^{n}}{\sqrt{n!}}\ket{n}

Time evolution is shape-preserving:

α(t)=eiωt/2αeiωt\ket{\alpha(t)} = e^{-i\omega t/2}\ket{\alpha e^{-i\omega t}}

Expectation values trace classical motion,

x(t)=x02 ⁣[αeiωt],p(t)=p02 ⁣[αeiωt]\langle x(t)\rangle = x_{0}\sqrt{2}\,\Re\!\left[\alpha e^{-i\omega t}\right],\qquad \langle p(t)\rangle = p_{0}\sqrt{2}\,\Im\!\left[\alpha e^{-i\omega t}\right]

while uncertainties remain constant and minimal, ΔxΔp=/2\Delta x\,\Delta p=\hbar/2. These states are the closest quantum analogs to phase-space points.


7.3.8 Forced oscillator and driven response

With a weak drive H(t)=F(t)xH'(t) = -\,F(t)\,x, the Heisenberg equation for aa is linear, producing

dadt=iωa+i2x0F(t)\frac{d a}{dt} = -\,i\omega a + \frac{i}{\sqrt{2}\,x_{0}\hbar}\,F(t)

A drive at frequency near ω\omega displaces the coherent-state parameter α\alpha 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.


7.3.9 Hermite toolbox: recursions and nodes

Hermite polynomials satisfy

Hn+1(ξ)=2ξHn(ξ)2nHn1(ξ)H_{n+1}(\xi) = 2\xi\,H_{n}(\xi) - 2n\,H_{n-1}(\xi)

and

dHndξ=2nHn1(ξ)\frac{d H_{n}}{d\xi} = 2n\,H_{n-1}(\xi)

Zeros of HnH_{n} are real and bracket the classical turning points ξ±2n+1\xi \approx \pm\sqrt{2n+1}, foreshadowing the WKB picture where oscillatory regions live inside the turning points and exponential tails outside.


7.3.10 Creation of quanta: field-theory teaser

Because H=ω(N+12)H=\hbar\omega(N+\tfrac12), each aa^{\dagger} raises the energy by one quantum ω\hbar\omega. Promoting a,aa,a^{\dagger} to mode operators of a field turns the EM field into a set of independent oscillators, with aa^{\dagger} creating a photon. Your oscillator algebra is literally second-quantization’s ABCs.


7.3.11 Worked mini-examples

(a) Expectation values in n\ket{n}.

Using xa+ax\propto a+a^{\dagger} and Nn=nnN\ket{n}=n\ket{n},

nxn=0,nx2n=(n+12)x02\langle n|x|n\rangle = 0,\qquad \langle n|x^{2}|n\rangle = \left(n+\frac{1}{2}\right)x_{0}^{2}

Similarly,

np2n=(n+12)p02\langle n|p^{2}|n\rangle = \left(n+\frac{1}{2}\right)p_{0}^{2}

Check that T=p2/(2m)=12En\langle T\rangle = \langle p^{2}\rangle/(2m) = \tfrac12 E_{n}.

(b) Transition matrix element n+1xn\langle n+1|x|n\rangle.

From x=x02(a+a)x=\tfrac{x_{0}}{\sqrt{2}}(a+a^{\dagger}),

n+1xn=x02n+1\langle n+1|x|n\rangle = \frac{x_{0}}{\sqrt{2}}\sqrt{n+1}

This controls dipole selection rules and line strengths for an oscillator-like transition.

(c) Overlap of coherent and number states.

The Poisson weights are

nα2=eα2α2nn!|\langle n|\alpha\rangle|^{2} = e^{-|\alpha|^{2}}\frac{|\alpha|^{2n}}{n!}

Mean quanta nˉ=α2\bar n = |\alpha|^{2} and variance Δn2=α2\Delta n^{2}=|\alpha|^{2}.


7.3.12 Problem kit

  • Derive the Hermite solution by power series, show termination, and recover EnE_{n}
  • Starting from a,aa,a^{\dagger}, build HH and prove En=ω(n+12)E_{n}=\hbar\omega(n+\tfrac12) without solving a differential equation
  • Compute Δx\Delta x and Δp\Delta p for ϕ0\phi_{0} and verify the minimal-uncertainty product /2\hbar/2
  • Show that x\langle x\rangle for a coherent state follows xcl(t)x_{\text{cl}}(t), while ψ(x,t)2|\psi(x,t)|^{2} keeps a fixed Gaussian width
  • Add a linear drive FcosΩtF\cos\Omega t and solve for a(t)a(t); identify the resonant response and phase lag
  • Use x=x02(a+a)x=\tfrac{x_{0}}{\sqrt{2}}(a+a^{\dagger}) to prove that nxn=0\langle n'|x|n\rangle=0 unless n=n±1n'=n\pm 1

7.4 Angular Momentum and Central Potentials

Before tackling hydrogen, we need the geometry engine of 3D quantum mechanics: angular momentum. Rotational symmetry gives conserved L\mathbf L, spherical harmonics YmY_\ell^m 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.


7.4.1 Angular momentum operators and algebra

Define orbital angular momentum

L=r×p,Li=iϵijkxjk\boldsymbol L = \boldsymbol r \times \boldsymbol p,\qquad L_i = -\,i\hbar\,\epsilon_{ijk}\,x_j\,\partial_k

The commutators encode the rotation group

[Li,Lj]=iϵijkLk[L_i,L_j] = i\hbar\,\epsilon_{ijk}\,L_k

The Casimir and components obey

L2Lx2+Ly2+Lz2,[L2,Li]=0L^2 \equiv L_x^2+L_y^2+L_z^2,\qquad [L^2,L_i]=0

Because [L2,Li]=0[L^2,L_i]=0, we can simultaneously diagonalize L2L^2 and one component, conventionally LzL_z.


7.4.2 Ladder operators and eigenvalue structure

Introduce

L±Lx±iLyL_{\pm} \equiv L_x \pm i L_y

They satisfy

[Lz,L±]=±L±,[L+,L]=2Lz[L_z,L_{\pm}] = \pm \hbar\,L_{\pm},\qquad [L_+,L_-] = 2\hbar\,L_z

Let m\ket{\ell m} be simultaneous eigenstates of L2L^2 and LzL_z

L2m=2(+1)m,Lzm=mmL^2\ket{\ell m} = \hbar^2 \ell(\ell+1)\ket{\ell m},\qquad L_z\ket{\ell m} = \hbar m \ket{\ell m}

Ladders step mm within a fixed \ell

L±m=(+1)m(m±1),m±1L_{\pm}\ket{\ell m} = \hbar\,\sqrt{\ell(\ell+1)-m(m\pm 1)}\,\ket{\ell,m\pm 1}

Allowable quantum numbers are

=0,1,2,,m=,+1,,\ell = 0,1,2,\dots,\qquad m=-\ell,-\ell+1,\dots,\ell

giving a (2+1)(2\ell+1)-fold degeneracy for each \ell.


7.4.3 Spherical harmonics: the angular wavefunctions

In position representation the simultaneous eigenfunctions are the spherical harmonics Ym(θ,ϕ)Y_\ell^m(\theta,\phi), obeying

L2Ym=2(+1)Ym,LzYm=mYmL^2\,Y_\ell^m = \hbar^2 \ell(\ell+1)\,Y_\ell^m,\qquad L_z\,Y_\ell^m = \hbar m\,Y_\ell^m

with orthonormality and completeness on the unit sphere

02π ⁣ ⁣0πYm(θ,ϕ)Ym(θ,ϕ)sinθdθdϕ=δδmm\int_{0}^{2\pi}\!\!\int_{0}^{\pi} Y_{\ell}^{m}(\theta,\phi)^{\ast} Y_{\ell'}^{m'}(\theta,\phi)\,\sin\theta\,d\theta\,d\phi = \delta_{\ell\ell'}\delta_{mm'} =0m=Ym(Ω)Ym(Ω)=δ(cosθcosθ)δ(ϕϕ)\sum_{\ell=0}^{\infty}\sum_{m=-\ell}^{\ell} Y_{\ell}^{m}(\Omega)\,Y_{\ell}^{m}(\Omega')^{\ast} = \delta(\cos\theta-\cos\theta')\,\delta(\phi-\phi')

Parity is simple

Ym(πθ,ϕ+π)=(1)Ym(θ,ϕ)Y_{\ell}^{m}(\pi-\theta,\phi+\pi) = (-1)^{\ell}\,Y_{\ell}^{m}(\theta,\phi)

Explicitly, with associated Legendre functions PmP_{\ell}^{m},

Ym(θ,ϕ)=NmPm(cosθ)eimϕY_{\ell}^{m}(\theta,\phi) = N_{\ell m}\,P_{\ell}^{m}(\cos\theta)\,e^{i m \phi}

where NmN_{\ell m} is the standard normalization constant.


7.4.4 Laplacian in spherical coordinates and separation of variables

The 3D Laplacian is

2=1r2r ⁣(r2r)L22r2\nabla^{2} = \frac{1}{r^{2}}\frac{\partial}{\partial r}\!\left(r^{2}\frac{\partial}{\partial r}\right) - \frac{L^{2}}{\hbar^{2} r^{2}}

For a central potential V(r)V(r) the time-independent Schrödinger equation

[22m2+V(r)]ψ(r)=Eψ(r)\left[-\,\frac{\hbar^{2}}{2m}\nabla^{2} + V(r)\right]\psi(\mathbf r) = E\,\psi(\mathbf r)

separates with the ansatz ψ(r,θ,ϕ)=Rn(r)Ym(θ,ϕ)\psi(r,\theta,\phi)=R_{n\ell}(r)\,Y_{\ell}^{m}(\theta,\phi). The angular part yields the spherical harmonics above; the radial part obeys

22m[1r2ddr ⁣(r2dRdr)(+1)r2R]+V(r)R=ER-\,\frac{\hbar^{2}}{2m}\left[\frac{1}{r^{2}}\frac{d}{dr}\!\left(r^{2}\frac{dR}{dr}\right) - \frac{\ell(\ell+1)}{r^{2}} R\right] + V(r)\,R = E\,R

It is cleaner to remove the first derivative by u(r)rR(r)u(r)\equiv r\,R(r), giving the radial Schrödinger equation

22md2udr2+[V(r)+2(+1)2mr2]u=Eu-\,\frac{\hbar^{2}}{2m}\,\frac{d^{2}u}{dr^{2}} + \left[V(r) + \frac{\hbar^{2}\ell(\ell+1)}{2m r^{2}}\right]u = E\,u

The bracketed term is the effective potential

Veff(r)=V(r)+2(+1)2mr2V_{\text{eff}}(r) = V(r) + \frac{\hbar^{2}\ell(\ell+1)}{2m r^{2}}

with a quantum centrifugal barrier 1/r2\propto 1/r^{2}.


7.4.5 Radial normalization and boundary conditions

Assuming angular functions are normalized, the full normalization reduces to

0u(r)2dr=1\int_{0}^{\infty} |u(r)|^{2}\,dr = 1

Physical solutions satisfy

  • Regularity at the origin: u(r)r+1u(r)\sim r^{\ell+1} as r0r\to 0 for finite V(0)V(0)
  • Normalizability for bound states: u(r)0u(r)\to 0 as rr\to \infty
  • Appropriate scattering asymptotics for E>0E>0

These conditions quantize EE for confining V(r)V(r).


7.4.6 Rotational symmetry, Noether, and degeneracy

If V(r)V(r) is central, the Hamiltonian commutes with all rotations U(R)=exp ⁣(iθL)U(R)=\exp\!\left(-\frac{i}{\hbar}\,\boldsymbol{\theta}\cdot\boldsymbol L\right). Noether’s theorem in the quantum language says

[H,L]=0[H,\boldsymbol L]=\boldsymbol 0

Hence \ell and mm label eigenstates. Energies depend on nn and \ell in general, but are mm-degenerate by (2+1)(2\ell+1) because rotating a solution merely mixes mm values inside the same \ell multiplet.

Special symmetries can enlarge degeneracy. For hydrogen, a hidden Runge–Lenz symmetry makes EE depend on nn only; for the isotropic 3D oscillator, EE depends on N=2nr+N=2n_{r}+\ell.


7.4.7 Angular momentum as the generator of rotations

Infinitesimal rotations act as

δψ(r)=iδθLψ(r)\delta \psi(\boldsymbol r) = -\,\frac{i}{\hbar}\,\delta\boldsymbol{\theta}\cdot\boldsymbol L\,\psi(\boldsymbol r)

For a finite rotation RR, the transformed state is U(R)ψU(R)\psi, with coordinates rotating oppositely. On the sphere, the YmY_{\ell}^{m} furnish irreducible representations of SO(3)SO(3); under RR, components mix within fixed \ell via Wigner DD-matrices Dmm()(R)D^{(\ell)}_{m'm}(R).


7.4.8 Orbital vs spin, and addition (teaser)

Electrons carry orbital L\boldsymbol L and spin S\boldsymbol S angular momenta. The total

J=L+S\boldsymbol J = \boldsymbol L + \boldsymbol S

obeys the same algebra as any angular momentum. When two angular momenta couple, the allowed totals jj 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.


7.4.9 Worked mini-examples

(a) Small-rr behavior of u(r)u(r).
For finite V(0)V(0), the dominant terms in the radial equation near r=0r=0 are

22mu(r)+2(+1)2mr2u(r)0-\,\frac{\hbar^{2}}{2m}\,u''(r) + \frac{\hbar^{2}\ell(\ell+1)}{2m r^{2}}\,u(r) \approx 0

Try ursu\sim r^{s} to get s(s1)=(+1)s(s-1)=\ell(\ell+1), giving s=+1s=\ell+1 (regular) and s=s=-\ell (singular). Keep ur+1u\sim r^{\ell+1}.

(b) Expectation of L2L^{2} in a separable state.
For ψ=R(r)Ym(θ,ϕ)\psi=R(r)Y_{\ell}^{m}(\theta,\phi),

L2=2(+1)\langle L^{2}\rangle = \hbar^{2}\ell(\ell+1)

independent of R(r)R(r) because angles carry all the LL structure.

(c) Degeneracy count at fixed \ell.
Show that the set {Ym}m=\{Y_{\ell}^{m}\}_{m=-\ell}^{\ell} spans a (2+1)(2\ell+1)-dimensional irrep by applying L±L_{\pm} repeatedly to YY_{\ell}^{\ell} and YY_{\ell}^{-\ell}.

(d) Effective potential sketch.
For V(r)=k/rV(r)=-k/r (Coulomb), plot Veff(r)=k/r+2(+1)/(2mr2)V_{\text{eff}}(r)=-k/r+\hbar^{2}\ell(\ell+1)/(2mr^{2}) to visualize classical-like turning points and bound-state regions for various \ell.


7.4.10 Minimal problem kit

  • Starting from 2\nabla^{2} in spherical coordinates, derive the radial equation for u(r)u(r) and identify Veff(r)V_{\text{eff}}(r)
  • Prove the ladder action of L±L_{\pm} on m\ket{\ell m} and normalize the coefficients by demanding mm=1\langle \ell m|\ell m\rangle=1
  • Verify orthonormality of YmY_{\ell}^{m} by integrating two explicit low-\ell examples over the sphere
  • For a central square well V(r)=V0V(r)=-V_{0} for r<Rr<R and 00 otherwise, write the bound-state matching conditions for u(r)u(r) at r=Rr=R and discuss \ell-dependence of the spectrum
  • Show that [H,L]=0[H,\boldsymbol L]=\boldsymbol 0 for any V(r)V(r) and use it to argue (2+1)(2\ell+1) degeneracy in mm

7.5 The Hydrogen Atom: Solving the Coulomb Problem

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=μe42(4πε0)22Z2n2=μc2α22Z2n2E_n = -\,\frac{\mu e^{4}}{2 (4\pi\varepsilon_0)^{2} \hbar^{2}}\,\frac{Z^{2}}{n^{2}} = -\,\frac{\mu c^{2}\alpha^{2}}{2}\,\frac{Z^{2}}{n^{2}}

and explicit wavefunctions labeled by (n,,m)(n,\ell,m). Hydrogen wins because the math secretly has extra symmetry; the degeneracy is n2n^{2} per nn (ignoring spin).


7.5.1 Setup and separation

Take a nucleus of charge +Ze+Ze and an electron of charge e-e. Use the reduced mass

μmemNme+mN\mu \equiv \frac{m_e m_N}{m_e + m_N}

and the Coulomb constant

kZe24πε0k \equiv \frac{Z e^{2}}{4\pi\varepsilon_0}

The time-independent Schrödinger equation with a central potential V(r)=k/rV(r)=-k/r separates as ψ(r,θ,ϕ)=Rn(r)Ym(θ,ϕ)\psi(r,\theta,\phi)=R_{n\ell}(r)\,Y_{\ell}^{m}(\theta,\phi). The radial function u(r)rR(r)u(r)\equiv r\,R(r) obeys

22μd2udr2+[2(+1)2μr2kr]u=Eu-\,\frac{\hbar^{2}}{2\mu}\,\frac{d^{2}u}{dr^{2}} + \left[\frac{\hbar^{2}\ell(\ell+1)}{2\mu r^{2}} - \frac{k}{r}\right]u = E\,u

Bound states have E<0E<0.


7.5.2 Non-dimensionalization and solution shape

Define

κ2μE,ρ2κr\kappa \equiv \frac{\sqrt{-2\mu E}}{\hbar},\qquad \rho \equiv 2\kappa r

Look for solutions of the form

u(ρ)=ρ+1eρ/2v(ρ)u(\rho) = \rho^{\ell+1}\,e^{-\rho/2}\,v(\rho)

Regularity and normalizability force v(ρ)v(\rho) to be a polynomial, specifically an associated Laguerre

v(ρ)Ln12+1(ρ)v(\rho) \propto L_{n-\ell-1}^{\,2\ell+1}(\rho)

and quantization pops out as

κ=μk2  1n,n=1,2,3,\kappa = \frac{\mu k}{\hbar^{2}}\;\frac{1}{n},\qquad n=1,2,3,\dots

which yields the spectrum quoted above.


7.5.3 Bohr radius, length scale, and quantum numbers

Introduce the reduced-mass Bohr radius

aμ4πε02μe2a_{\mu} \equiv \frac{4\pi\varepsilon_0\,\hbar^{2}}{\mu e^{2}}

and the hydrogenic length

aaμZa \equiv \frac{a_{\mu}}{Z}

Quantum numbers:

  • Principal n=1,2,n=1,2,\dots
  • Orbital =0,1,,n1\ell=0,1,\dots,n-1
  • Magnetic m=,+1,,m=-\ell,-\ell+1,\dots,\ell

For each nn, the total degeneracy is

=0n1(2+1)=n2\sum_{\ell=0}^{n-1} (2\ell+1) = n^{2}

Doubling by electron spin gives 2n22n^{2}.


7.5.4 Radial wavefunctions and normalization

With a=aμ/Za=a_{\mu}/Z, the normalized hydrogenic radial functions are

Rn(r)=2n2a3/2(n1)!(n+)!  exp ⁣(rna)(2rna)Ln12+1 ⁣(2rna)R_{n\ell}(r) = \frac{2}{n^{2}\,a^{3/2}}\, \sqrt{\frac{(n-\ell-1)!}{(n+\ell)!}}\; \exp\!\left(-\frac{r}{n a}\right)\, \left(\frac{2r}{n a}\right)^{\ell}\, L_{n-\ell-1}^{\,2\ell+1}\!\left(\frac{2r}{n a}\right)

so that

0Rn(r)2r2dr=1\int_{0}^{\infty} |R_{n\ell}(r)|^{2}\,r^{2}\,dr = 1

Sanity checks:

  • 1s1s (n=1,=0)(n{=}1,\ell{=}0) gives R10(r)=2a3/2er/aR_{10}(r) = 2 a^{-3/2} e^{-r/a}
  • 2p2p (n=2,=1)(n{=}2,\ell{=}1) gives R21(r)rer/(2a)R_{21}(r) \propto r\,e^{-r/(2a)}

7.5.5 Probability, nodes, and orbital shapes

The radial probability density is

Pn(r)=Rn(r)2r2P_{n\ell}(r) = |R_{n\ell}(r)|^{2}\,r^{2}

Node counting:

  • Angular: \ell nodal planes/surfaces from YmY_{\ell}^{m}
  • Radial: n1n-\ell-1 nodes from RnR_{n\ell}

Peaks move outward with nn; for 1s1s, PP peaks at r=ar=a.


7.5.6 Useful expectation values

In terms of a=aμ/Za=a_{\mu}/Z,

1rn=1n2a\big\langle \frac{1}{r} \big\rangle_{n\ell} = \frac{1}{n^{2} a} rn=a2[3n2(+1)]\langle r \rangle_{n\ell} = \frac{a}{2}\,\left[3n^{2} - \ell(\ell+1)\right] r2n=a2n22[5n2+13(+1)]\langle r^{2} \rangle_{n\ell} = a^{2}\,\frac{n^{2}}{2}\,\left[5n^{2} + 1 - 3\ell(\ell+1)\right]

These nail orders of magnitude for atom size and screening estimates.


7.5.7 Dipole selection rules and lines

For electric-dipole radiation with d^=er\hat{\boldsymbol d}=-e\boldsymbol r,

Δ=±1,Δm=0,±1\Delta \ell = \pm 1,\qquad \Delta m = 0,\pm 1

Line strengths follow from matrix elements

nmrnm\langle n'\ell'm' | \boldsymbol r | n\ell m \rangle

which factor into angular pieces (Clebsch–Gordan algebra) and radial integrals built from the Rn(r)R_{n\ell}(r). This reproduces the Balmer/Lyman series patterns and polarization rules.


7.5.8 Runge–Lenz symmetry and nn-only energy

The Coulomb problem has a conserved Runge–Lenz vector (quantum-symmetrized)

A=12μ(p×LL×p)krr^\boldsymbol A = \frac{1}{2\mu}\,(\boldsymbol p \times \boldsymbol L - \boldsymbol L \times \boldsymbol p) - \frac{k}{r}\,\hat{\boldsymbol r}

Together, L\boldsymbol L and A\boldsymbol A generate an SO(4)SO(4) symmetry for bound states, explaining why EnE_n depends on nn only, not on \ell or mm. Break that symmetry (e.g., by fields or relativity) and the degeneracy splits.


7.5.9 Real-world splittings: what lifts the degeneracy

Hydrogen is not perfectly degenerate once we add small effects:

  • Reduced mass. Already included by μ\mu; shifts RHR_{\text H} slightly and creates isotope shifts
  • Fine structure. Relativistic kinetic correction, spin–orbit coupling, and Darwin term split levels by order α2\alpha^{2} of the Rydberg; handled by perturbation theory in §7.6
  • Lamb shift. QED vacuum fluctuations separate 2S2S and 2P2P 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.


7.5.10 Worked mini-examples

(a) Most probable radius for 1s1s.

Maximize P10(r)=4r2a3e2r/aP_{10}(r)=4 r^{2} a^{-3} e^{-2r/a}. The derivative zero gives r=ar=a.

(b) Rydberg constant from the spectrum.

The photon wavenumber between nnn\to n' is

ν~=EnEnhc=RμmeZ2(1n21n2)\tilde\nu = \frac{E_n - E_{n'}}{hc} = R_{\infty}\,\frac{\mu}{m_e}\,Z^{2}\left(\frac{1}{n'^{2}} - \frac{1}{n^{2}}\right)

with

R=α2mec2hR_{\infty} = \frac{\alpha^{2} m_e c}{2h}

(c) Radial integral for 2p1s2p\to 1s.

The dipole matrix element reduces to

1sr2p=0R10(r)rR21(r)r2dr\langle 1s||r||2p\rangle = \int_{0}^{\infty} R_{10}(r)\,r\,R_{21}(r)\,r^{2}\,dr

Plugging the explicit RR’s gives a nonzero value, consistent with Δ=1\Delta \ell=1.

(d) Expectation of 1/r1/r from Virial.

For V1/rV\propto 1/r, the virial theorem yields T=E\langle T\rangle = -E and V=2E\langle V\rangle = 2E, hence

1r=2Ek=1n2a\left\langle \frac{1}{r} \right\rangle = \frac{-2E}{k} = \frac{1}{n^{2} a}

matching the formula above.


7.5.11 Minimal problem kit

  • Starting from the radial equation, perform the ρ=2κr\rho=2\kappa r substitution and show that termination of the Laguerre series gives n=+1+nrn=\ell+1+n_{r} and En1/n2E_n\propto -1/n^{2}
  • Derive and normalize R10(r)R_{10}(r) and R21(r)R_{21}(r) explicitly
  • Prove the degeneracy count =0n1(2+1)=n2\sum_{\ell=0}^{n-1}(2\ell+1)=n^{2} and extend to 2n22n^{2} with spin
  • Compute rn\langle r \rangle_{n\ell} using orthogonality of associated Laguerres
  • Evaluate the angular selection rules for dipole transitions via Wigner–Eckart or Clebsch–Gordan coefficients
  • Estimate the fine-structure splitting for n=2n=2 levels using leading relativistic and spin–orbit terms as a teaser for §7.6

7.6 Time-Independent Perturbation Theory: Nondegenerate, Degenerate, and First Applications

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.


7.6.1 Setup and the expansion game

Let

H(λ)=H0+λHH(\lambda) = H_{0} + \lambda H'

where H0H_{0} is solvable with eigenpairs {En(0),n(0)}\{E_{n}^{(0)},\ket{n^{(0)}}\} and HH' is the perturbation. Seek series

En=En(0)+λEn(1)+λ2En(2)+E_{n} = E_{n}^{(0)} + \lambda E_{n}^{(1)} + \lambda^{2} E_{n}^{(2)} + \cdots n=n(0)+λn(1)+λ2n(2)+\ket{n} = \ket{n^{(0)}} + \lambda \ket{n^{(1)}} + \lambda^{2}\ket{n^{(2)}} + \cdots

Adopt intermediate normalization n(0)n=1\langle n^{(0)}|n\rangle = 1, which implies n(0)n(k)=0\langle n^{(0)}|n^{(k)}\rangle=0 for k1k\ge 1 and keeps formulas tidy.


7.6.2 Nondegenerate perturbation theory

Assume En(0)E_{n}^{(0)} is nondegenerate.

First-order energy

En(1)=n(0)Hn(0)E_{n}^{(1)} = \langle n^{(0)}| H' | n^{(0)} \rangle

First-order state

n(1)=mnm(0)m(0)Hn(0)En(0)Em(0)\ket{n^{(1)}} = \sum_{m\neq n} \frac{\ket{m^{(0)}}\,\langle m^{(0)}|H'|n^{(0)}\rangle}{E_{n}^{(0)} - E_{m}^{(0)}}

Second-order energy

En(2)=mnm(0)Hn(0)2En(0)Em(0)E_{n}^{(2)} = \sum_{m\neq n} \frac{|\langle m^{(0)}|H'|n^{(0)}\rangle|^{2}}{E_{n}^{(0)} - E_{m}^{(0)}}

Sign check: if HH' mixes nn with higher states, denominators are negative so EnE_{n} 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.


7.6.3 Hellmann–Feynman and quick derivatives

If H(λ)H(\lambda) depends smoothly on a parameter λ\lambda and n(λ)\ket{n(\lambda)} is an exact eigenstate,

dEndλ=n(λ)Hλn(λ)\frac{dE_{n}}{d\lambda} = \left\langle n(\lambda)\left|\frac{\partial H}{\partial \lambda}\right|n(\lambda)\right\rangle

This avoids differentiating messy wavefunctions: compute the expectation of the explicit derivative of HH. Great for forces in molecules and for tracking how EE slides with external fields.


7.6.4 Degenerate perturbation theory

If several H0H_{0} eigenstates share the same E(0)E^{(0)}, naive denominators blow up. Fix by diagonalizing HH' within the degenerate subspace.

Let {ϕa}a=1g\{\ket{\phi_{a}}\}_{a=1}^{g} be an orthonormal basis of a gg-fold degenerate eigenspace of H0H_{0}. Build the secular matrix

WabϕaHϕbW_{ab} \equiv \langle \phi_{a} | H' | \phi_{b} \rangle

Diagonalize WW to get eigenvalues {λα}\{\lambda_{\alpha}\} and eigenvectors {c(α)}\{\mathbf c^{(\alpha)}\}. Then

Eα(1)=λαE_{\alpha}^{(1)} = \lambda_{\alpha} ψα(0)=a=1gca(α)ϕa\ket{\psi_{\alpha}^{(0)}} = \sum_{a=1}^{g} c^{(\alpha)}_{a}\,\ket{\phi_{a}}

These optimized zeroth-order states already include the dominant mixing; higher orders then proceed as in the nondegenerate case using ψα(0)\ket{\psi_{\alpha}^{(0)}}.

Rule of thumb: choose a basis adapted to the symmetries of HH'. You will often find that WW block-diagonalizes (e.g., fixed mm or fixed parity), making life easy.


7.6.5 First applications: hydrogen fine structure (order α4\alpha^{4})

For Ze2/4πε0rZ e^{2}/4\pi\varepsilon_{0}r with Z=1Z=1, three relativistic corrections dress H0H_{0}:

  1. Relativistic kinetic energy
Hrel=p48me3c2H_{\mathrm{rel}} = -\,\frac{p^{4}}{8 m_{e}^{3} c^{2}}
  1. Spin–orbit coupling with Thomas factor
HSO=12me2c21rdVdrLSH_{\mathrm{SO}} = \frac{1}{2 m_{e}^{2} c^{2}}\,\frac{1}{r}\frac{dV}{dr}\,\boldsymbol L\cdot \boldsymbol S

For Coulomb V(r)=e2/4πε0rV(r)=-e^{2}/4\pi\varepsilon_{0}r, this is +LS/r3\propto +\,\boldsymbol L\cdot\boldsymbol S/r^{3}

  1. Darwin term (contact term from Zitterbewegung, nonzero only for =0\ell=0)
HD=28me2c22V(r)H_{\mathrm{D}} = \frac{\hbar^{2}}{8 m_{e}^{2} c^{2}}\,\nabla^{2}V(r)

Combine expectation values using hydrogen eigenstates and angular momentum algebra. The sum depends on nn and the total j=±12j=\ell\pm \tfrac12 but not on mm, giving the textbook fine-structure shift

ΔEn,j(fs)=μc2(Zα)42n3(1j+1234n)\Delta E_{n,j}^{(\mathrm{fs})} = \frac{\mu c^{2}\,(Z\alpha)^{4}}{2 n^{3}}\left(\frac{1}{j+\tfrac12} - \frac{3}{4n}\right)

Here μ\mu is the reduced mass and αe2/4πε0c\alpha\equiv e^{2}/4\pi\varepsilon_{0}\hbar c. Fine structure lifts the n2n^{2} degeneracy down to the 2n2n degeneracy labeled by jj and mm.


7.6.6 Zeeman effect: weak magnetic fields

Add a uniform magnetic field B=Bz^\boldsymbol B = B\,\hat{\boldsymbol z}. The leading coupling is

HZ=μB(Lz+gsSz)BH'_{\mathrm{Z}} = \mu_{B}\,(L_{z} + g_{s} S_{z})\,B

with gs2g_{s}\approx 2 and μB=e/2me\mu_{B}=e\hbar/2m_{e}. In LS coupling, levels are labeled by nsjmj\ket{n\ell s j m_{j}} and the shift is packaged as

ΔEZ=μBgJmJB\Delta E_{\mathrm{Z}} = \mu_{B}\,g_{J}\,m_{J}\,B

where the Landé factor

gJ=1+j(j+1)+s(s+1)(+1)2j(j+1)g_{J} = 1 + \frac{j(j+1) + s(s+1) - \ell(\ell+1)}{2j(j+1)}

For s=12s=\tfrac12, this reproduces the familiar “anomalous” Zeeman patterns. The selection rule for dipole transitions remains ΔmJ=0,±1\Delta m_{J}=0,\pm 1; polarizations track π\pi and σ±\sigma^{\pm} components.

Strong-field (Paschen–Back) note: when μBB\mu_{B}B beats spin–orbit, LL and SS uncouple and you diagonalize LzL_{z} and SzS_{z} instead of J2J^{2}, but that is beyond first-order weak-field PT.


7.6.7 Stark effect: electric fields and parity trouble

For a static electric field E=Ez^\boldsymbol E=E\,\hat{\boldsymbol z},

HS=eEzH'_{\mathrm{S}} = e E z

Parity makes first-order shifts vanish in nondegenerate levels with definite parity (e.g., hydrogen 1s1s). Thus the leading shift is second order:

ΔE(2)=mnm(0)eEzn(0)2En(0)Em(0)\Delta E^{(2)} = \sum_{m\neq n} \frac{|\langle m^{(0)}|eEz|n^{(0)}\rangle|^{2}}{E_{n}^{(0)} - E_{m}^{(0)}}

For hydrogen 1s1s this yields a quadratic Stark shift

ΔE1s=12α1sE2,α1s=92a03\Delta E_{1s} = -\,\frac{1}{2}\,\alpha_{1s}\,E^{2},\qquad \alpha_{1s} = \frac{9}{2}\,a_{0}^{3}

where α\alpha is the polarizability and a0a_{0} the Bohr radius.

Linear Stark in degenerate manifolds. When H0H_{0} has degeneracy of opposite parities (hydrogen n2n\ge 2), do degenerate PT in the nn manifold using parabolic states. First-order shifts appear:

ΔEn,k,m=32nea0Ek\Delta E_{n,k,m} = \frac{3}{2}\,n\,e\,a_{0}\,E\,k

with k=n1n2k=n_{1}-n_{2} the parabolic quantum number satisfying n1+n2+m+1=nn_{1}+n_{2}+|m|+1=n. For n=2n=2, k=±1k=\pm 1 at m=0m=0 split linearly, while m=1|m|=1 states stay unshifted at first order.


7.6.8 Practical workflow and symmetry hacks

  • Exploit good quantum numbers. Work in a basis that diagonalizes all operators commuting with both H0H_{0} and HH'
  • Use Wigner–Eckart to drop angles. For dipole-type perturbations rqr_{q}, 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 HH' to level spacings of H0H_{0}. If the ratio is not 1\ll 1, perturbation theory will not slay this dragon

7.6.9 Worked mini-examples

(a) Quadratic Stark for hydrogen ground state

Use closure over npnp states to show

ΔE1s(2)=12α1sE2,α1s=92a03\Delta E_{1s}^{(2)} = -\,\frac{1}{2} \alpha_{1s} E^{2},\qquad \alpha_{1s} = \frac{9}{2}\,a_{0}^{3}

Sketch: only =1\ell=1 intermediate states contribute because zY10z\propto Y_{1}^{0} changes \ell by ±1\pm 1. Evaluate the radial sum or use known hydrogenic polarizability.

(b) Zeeman splitting of a 2P3/22P_{3/2} level

With j=32j=\tfrac32, gJ=43g_{J}=\tfrac{4}{3}. Allowed mJ={32,12,12,32}m_{J}=\{-\tfrac32,-\tfrac12,\tfrac12,\tfrac32\} produce four equally spaced lines with separation μBgJB\mu_{B}g_{J}B.

(c) Fine-structure order of magnitude

For n=2n=2, the scale is

ΔE(fs)μc2α4/16|\Delta E^{(\mathrm{fs})}| \sim \mu c^{2}\,\alpha^{4}/16

which for hydrogen lands in the 104 eV10^{-4}\ \text{eV} range, i.e., tens of GHz—microwave domain.

(d) Two-level avoided crossing via second order

For a pair {1,2}\{\ket{1},\ket{2}\} with E2(0)E1(0)=ΔE_{2}^{(0)}-E_{1}^{(0)}=\Delta and off-diagonal 1H2=V\langle 1|H'|2\rangle=V, diagonalize

(E1(0)VVE2(0))\begin{pmatrix} E_{1}^{(0)} & V \\ V & E_{2}^{(0)} \end{pmatrix}

to get exact E±=12(E1(0)+E2(0))±12Δ2+4V2E_{\pm} = \tfrac{1}{2}(E_{1}^{(0)}+E_{2}^{(0)}) \pm \tfrac{1}{2}\sqrt{\Delta^{2}+4V^{2}}. For VΔ|V|\ll |\Delta|, recover second-order shifts ±V2/Δ\pm V^{2}/\Delta. This is the algebraic skeleton behind many “near-degenerate” perturbations.


7.6.10 Common pitfalls and how not to trip

  • Forgetting to orthogonalize in degenerate spaces. Always diagonalize HH' first; otherwise denominators lie to you
  • Using unperturbed labels after mixing. Once you diagonalize WW, the “good” zeroth-order states are the WW 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\lambda
  • Order counting. If first order vanishes, second order can be dominant but still “small.” Check scales before trusting the series

7.6.11 Minimal problem kit

  • Derive En(1)E_{n}^{(1)}, n(1)\ket{n^{(1)}}, and En(2)E_{n}^{(2)} from the series ansatz with intermediate normalization
  • For a 3-fold degenerate level {ϕ1,ϕ2,ϕ3}\{\ket{\phi_{1}},\ket{\phi_{2}},\ket{\phi_{3}}\} and perturbation HH', build WW, solve the secular equation det(WλI)=0\det(W-\lambda I)=0, and construct the mixed eigenkets
  • Compute the Landé gJg_{J} formula from HZH'_{\mathrm{Z}} using vector coupling and verify special cases SS (=0\ell=0) and LL-only (s=0s=0) limits
  • Evaluate the quadratic Stark shift for 1s1s by explicit summation over npnp states or by using known closure relations
  • Starting from Hrel,HSO,HDH_{\mathrm{rel}}, H_{\mathrm{SO}}, H_{\mathrm{D}}, show that the combined fine-structure correction depends only on nn and jj and reproduce ΔEn,j(fs)\Delta E_{n,j}^{(\mathrm{fs})} above