6 Early quantum theory

6.1 Introduction: Cracks in the Classical Edifice

By the end of the 19th century, physics flexed like it had the final boss down. Newtonian mechanics ruled motion, Maxwell’s equations ruled fields, thermodynamics ruled heat. Yet precision experiments began posting L’s for classical theory. A new playbook—quantum—was loading.

What broke first? Three pillars:

  1. Radiation and heat. The spectrum of an ideal glowing cavity (a blackbody) refused to follow classical predictions. Equipartition and Maxwell’s waves led to infinite ultraviolet energy—an “ultraviolet catastrophe.”

  2. Light and electrons. Shining light on a clean metal knocked out electrons with sharp rules that defied wave-only intuition.

  3. Atoms and spectra. Gases glowed at razor-thin frequencies described by simple formulas, yet any classical electron orbiting a nucleus should radiate and spiral inward.

These cracks weren’t edge cases; they were mainstream physics behaving weirdly.


6.1.1 Characters entering the stage

Max Planck (1900). A thermodynamics purist backed into heresy. To fit blackbody data, he quantized energy exchange in chunks of hνh\nu. He thought it was a mathematical trick; the universe took him literally.

Albert Einstein (1905). The “patent-clerk energy.” He said the quiet part out loud: light itself comes in quanta (photons). With that, the photoelectric rules suddenly made sense, and E=hνE=h\nu stepped out of the blackbody and into matter.

Ernest Rutherford (1911). He yeeted Thomson’s “plum pudding” and installed a nuclear atom after alpha particles bounced at wild angles from gold foil. But a classical electron racing around a tiny nucleus should radiate energy and crash. Stable atoms? Not in classical land.

Niels Bohr (1913). He proposed that electrons live only on special orbits with quantized angular momentum and jump between them by absorbing or emitting light at specific frequencies. The hydrogen spectrum clicked into place—partial victory, but weird rules demanded deeper logic.

Louis de Broglie (1923–24). If light acts like particles, maybe matter acts like waves. He proposed λ=h/p\lambda=h/p, a hypothesis begging for an experiment.


6.1.2 The “classical no-go”s (with minimal math)

Blackbody catastrophe. Classical equipartition feeds every electromagnetic cavity mode an average energy kBTk_B T. Counting modes yields a spectral energy density that rises ν2\propto \nu^2, blowing up at high frequency. Data said “nope”—the spectrum peaks and falls.

We will show in §6.2 that introducing energy quanta leads to Planck’s distribution and cures the divergence

u(ν,T)=8πhν3c31exp ⁣(hνkBT)1u(\nu,T) = \frac{8\pi h \nu^{3}}{c^{3}}\,\frac{1}{\exp\!\left(\frac{h\nu}{k_B T}\right)-1}

No ultraviolet apocalypse appears in this curve


Photoelectric rules. Experiments found: (i) no electrons below a threshold frequency, however intense the light, (ii) above threshold, the maximum electron kinetic energy is linear in frequency, not intensity, (iii) emission is effectively instantaneous. Einstein’s one-liner,

Kmax=hνWK_{\max} = h\nu - W

nailed it, where WW is the work function. Intensity controls how many photons hit, not their individual energy


Atomic spectra. Balmer and Rydberg wrote beautiful empirical formulas for hydrogen lines. Rutherford’s nuclear atom raised the stakes: a tiny, charged nucleus should make orbiting electrons radiate continuously and smear the spectrum—contradiction. Bohr’s move was to quantize angular momentum and energy levels, producing discrete lines and a frequency rule that matched Rydberg’s data. But why those orbits? Old quantum theory worked, then stalled.


6.1.3 The vibe: discrete meets wave

The early quantum story is a series of “wait, both?” moments:

  • Light interferes like a wave but trades energy like a particle
  • Electrons behave like particles in tracks yet diffract like waves through crystals
  • Energy in atoms is not continuous real estate; it’s a floor plan of levels with allowed transitions

The common thread is discreteness sneaking into systems previously modeled as continuous. That discreteness shows up in spectra, specific heats, threshold effects, and stability.


6.1.4 Minimal mathematical promises (to be delivered in later sections)

  • Planck spectrum. From counting quanta with weight exp(nhν/kBT)\exp(-nh\nu/k_B T) we will derive the energy density above and recover the Stefan–Boltzmann law and Wien’s displacement
λpeakT=constant\lambda_{\text{peak}}\,T = \text{constant}
  • Einstein photoelectric equation. Straight line plots of KmaxK_{\max} vs ν\,\nu give slope hh and intercept WW—that is how early experiments measured Planck’s constant with electrons

  • Bohr energies. For hydrogen-like atoms we will get

En=mee4221n2E_n = -\,\frac{m_e e^4}{2\hbar^2}\,\frac{1}{n^2}

and the emitted photon frequency from a jump nnn\to n' will satisfy hν=EnEnh\nu = E_n - E_{n'} with Rydberg constant emerging naturally

  • de Broglie hypothesis. Assuming p=h/λp=h/\lambda and E=hνE=h\nu makes group velocity match particle velocity and predicts electron diffraction, later seen in labs

6.1.5 What “early quantum theory” is (scope note)

We will live mostly in 1900–1925: the old quantum theory (Planck → Einstein → Bohr–Sommerfeld → de Broglie) and its decisive experiments (Franck–Hertz, Davisson–Germer, Stern–Gerlach). Full quantum mechanics (Heisenberg’s matrices and Schrödinger’s waves, 1925–26) is the bridge to the next chapter, but we will aim the runway here: when classical intuition finally yields and a consistent quantum calculus becomes inevitable.


6.2 Blackbody Radiation and Planck’s Hypothesis

Blackbody radiation is where classical physics face-planted and quantum theory quietly spawned. A blackbody is an ideal cavity with a tiny hole: light that enters bounces around and is essentially absorbed; light that leaves carries the equilibrium spectrum of the cavity walls. The mission in the 1890s was to predict the energy density of this radiation as a function of frequency ν\nu and temperature TT and to explain why the measured curve rises, peaks, and then crashes instead of diverging. Spoiler: classical equipartition said “infinite UV energy,” nature said “lol, no,” and Max Planck invented quanta to get the curve right.


6.2.1 The experimental scene: precision burns theory

German labs (Lummer, Pringsheim, Rubens, Kurlbaum) built exquisitely calibrated cavities and detectors. At moderate and high frequencies, the data followed Wien’s law remarkably well; at low frequencies (long wavelengths), new measurements started to bend away from Wien’s exponential falloff, demanding a better formula. Planck followed these plots obsessively—enough that when Rubens showed him fresh long-wavelength data in October 1900, he went home and cooked up the curve that matched everything.

Two common ways to describe the spectrum are by frequency and by wavelength. They are related but not identical in shape because ν=c/λ\nu=c/\lambda changes the measure:

uν(ν,T)dν=uλ(λ,T)dλu_\nu(\nu,T)\,d\nu = u_\lambda(\lambda,T)\,d\lambda

You’ll see two standard forms. By frequency:

uν(ν,T)energy density per unit frequencyu_\nu(\nu,T) \equiv \text{energy density per unit frequency}

By wavelength:

uλ(λ,T)energy density per unit wavelengthu_\lambda(\lambda,T) \equiv \text{energy density per unit wavelength}

We’ll derive uνu_\nu first and then convert to uλu_\lambda.


6.2.2 Mode counting in a cavity: the 8πν2/c38\pi \nu^{2}/c^{3} factor

Take a cubic cavity with perfectly conducting walls. Electromagnetic standing waves fit only certain wavevectors k\boldsymbol k; each allowed mode is a harmonic oscillator. Counting modes with frequencies between ν\nu and ν+dν\nu+d\nu gives a density of states per unit volume

g(ν)dν=8πν2c3dνg(\nu)\,d\nu = \frac{8\pi \nu^{2}}{c^{3}}\,d\nu

The 8π8\pi comes from spherical shells in k\boldsymbol k-space and from two polarization states. This factor is pure geometry/kinematics and will be the same in classical and quantum derivations. The physics enters when we decide how much average energy to assign to each mode at equilibrium.


6.2.3 Classical attempt: equipartition and the ultraviolet catastrophe

In classical statistical mechanics, every quadratic degree of freedom contributes 12kBT\tfrac{1}{2}k_B T to the mean energy. A harmonic oscillator (one mode) has two quadratic pieces (potential and kinetic), so the mean energy per mode would be

εclassical=kBT\langle \varepsilon \rangle_{\text{classical}} = k_B T

Multiply by the mode density to get the spectral energy density

uνRJ(ν,T)=g(ν)ε=8πν2c3kBTu_\nu^{\text{RJ}}(\nu,T) = g(\nu)\,\langle \varepsilon \rangle = \frac{8\pi \nu^{2}}{c^{3}}\,k_B T

This is the Rayleigh–Jeans law. It works at low frequencies but skyrockets as ν2\nu^{2} and, when integrated over all frequencies, diverges. That divergence—the ultraviolet catastrophe—is not a cute nickname; it is classical physics predicting infinite energy in any warm room. Reality declines.


6.2.4 Wien’s law: a great half-truth

Empirically, Wien proposed

uνWien(ν,T)=Aν3exp ⁣(BνT)u_\nu^{\text{Wien}}(\nu,T) = A\,\nu^{3}\,\exp\!\left(-\,\frac{B\,\nu}{T}\right)

with constants A,BA,B fit to data. This law nails the high-frequency tail (it decays exponentially) and implies the correct displacement law (the peak slides linearly with TT), but it fails at low frequencies, where experiments show uνν2Tu_\nu\propto \nu^{2}T (Rayleigh–Jeans behavior), not ν3\nu^{3}. In short, Wien gets the UV, Rayleigh–Jeans gets the IR, and neither gets the full mixtape.


6.2.5 Planck’s move: discretize the energy exchange

Planck’s radical idea was to treat the cavity walls as a set of microscopic oscillators (resonators) that exchange energy with the radiation field in discrete packets of size hνh\nu:

εn=nhν,n=0,1,2,\varepsilon_n = n\,h\nu,\qquad n=0,1,2,\dots

Given Boltzmann weights, the partition function for one oscillator of frequency ν\nu is

Z=n=0enhν/(kBT)=11ehν/(kBT)Z = \sum_{n=0}^\infty e^{-n h\nu/(k_B T)} = \frac{1}{1-e^{-h\nu/(k_B T)}}

The mean energy per oscillator is then

εPlanck=βlnZ=hνehν/(kBT)1,β1kBT\langle \varepsilon \rangle_{\text{Planck}} = -\,\frac{\partial}{\partial \beta}\ln Z = \frac{h\nu}{e^{h\nu/(k_B T)}-1},\qquad \beta \equiv \frac{1}{k_B T}

Multiply by the density of modes g(ν)g(\nu) to get the full spectrum:

uν(ν,T)=8πhν3c3  1ehν/(kBT)1u_\nu(\nu,T) = \frac{8\pi h \nu^{3}}{c^{3}}\;\frac{1}{e^{h\nu/(k_B T)}-1}

This is Planck’s law, 1900. It reduces to Rayleigh–Jeans at low frequency (expand the exponential) and to Wien’s exponential at high frequency. One curve to rule them all.


6.2.6 Low- and high-frequency limits (sanity checks)

Let xhν/(kBT)x \equiv h\nu/(k_B T).

  • Low frequency (x1x\ll 1): ex1xe^{x}-1 \approx x, so
uν8πhν3c31x=8πν2c3kBTu_\nu \approx \frac{8\pi h \nu^{3}}{c^{3}}\,\frac{1}{x} = \frac{8\pi \nu^{2}}{c^{3}}\,k_B T

which is Rayleigh–Jeans

  • High frequency (x1x\gg 1): ex1exe^{x}-1 \approx e^{x}, so
uν8πhν3c3ehν/(kBT)u_\nu \approx \frac{8\pi h \nu^{3}}{c^{3}}\,e^{-\,h\nu/(k_B T)}

which is Wien’s law with A=8πh/c3A=8\pi h/c^{3} and B=h/kBB=h/k_B

So Planck’s formula gracefully bridges the two regimes that classical and empirical laws captured separately.

blackbody radiation


6.2.7 Total energy and Stefan–Boltzmann law

Integrate uνu_\nu over all frequencies to get the total energy density u(T)u(T):

u(T)=0uν(ν,T)dν=8πhc30ν3dνehν/(kBT)1u(T) = \int_{0}^{\infty} u_\nu(\nu,T)\,d\nu = \frac{8\pi h}{c^{3}}\int_{0}^{\infty}\frac{\nu^{3}\,d\nu}{e^{h\nu/(k_B T)}-1}

Change variables x=hν/(kBT)x=h\nu/(k_B T); then dν=(kBT/h)dxd\nu = (k_B T/h)\,dx and

u(T)=8π(kBT)4h3c30x3dxex1u(T) = \frac{8\pi (k_B T)^{4}}{h^{3} c^{3}} \int_{0}^{\infty}\frac{x^{3}\,dx}{e^{x}-1}

The integral is a famous one:

0x3dxex1=π415\int_{0}^{\infty}\frac{x^{3}\,dx}{e^{x}-1} = \frac{\pi^{4}}{15}

So

u(T)=aT4,a8π5kB415h3c3u(T) = a\,T^{4},\qquad a \equiv \frac{8\pi^{5} k_B^{4}}{15\,h^{3} c^{3}}

The radiated power per unit area from a black surface follows by I=c4uI=\tfrac{c}{4}u (kinematic factor for isotropic radiation):

I=σT4,σ=2π5kB415h3c2I = \sigma T^{4},\qquad \sigma = \frac{2\pi^{5} k_B^{4}}{15\,h^{3} c^{2}}

This is the Stefan–Boltzmann law with its constant σ\sigma expressed in terms of kB,h,ck_B,h,c.


6.2.8 Where is the peak? Wien’s displacement law

Maximize uν(ν,T)u_\nu(\nu,T) with respect to ν\nu at fixed TT. Setting x=hν/(kBT)x=h\nu/(k_B T), the condition reduces to

ddx(x3ex1)=03(ex1)xex=0\frac{d}{dx}\left(\frac{x^{3}}{e^{x}-1}\right)=0 \quad\Rightarrow\quad 3\,(e^{x}-1) - x\,e^{x} = 0

This transcendental equation has a root

x2.821439x_\ast \approx 2.821439\ldots

Hence the frequency of the peak scales as

νpeak=xkBhT\nu_{\text{peak}} = \frac{x_\ast k_B}{h}\,T

If you instead work with the wavelength form

uλ(λ,T)=8πhcλ5  1ehc/(λkBT)1u_\lambda(\lambda,T) = \frac{8\pi h c}{\lambda^{5}}\;\frac{1}{e^{hc/(\lambda k_B T)}-1}

and maximize with respect to λ\lambda, you get Wien’s displacement law

λpeakT=b\lambda_{\text{peak}}\,T = b

with a constant bb fixed by the root of a slightly different transcendental equation. Important PSA: the peak in uνu_\nu and the peak in uλu_\lambda are not at the same photon energy because the Jacobian dν/dλ|d\nu/d\lambda| reshapes the curve.


6.2.9 Converting between frequency and wavelength spectra

Use uλdλ=uνdνu_\lambda\,d\lambda = u_\nu\,d\nu with ν=c/λ\nu=c/\lambda and dν=c/λ2dλd\nu = -\,c/\lambda^{2}\,d\lambda. Then

uλ(λ,T)=uν ⁣(cλ,T)  cλ2u_\lambda(\lambda,T) = u_\nu\!\left(\frac{c}{\lambda},T\right)\;\frac{c}{\lambda^{2}}

Plugging Planck’s uνu_\nu gives the standard uλu_\lambda quoted above. The minus sign drops because decreasing λ\lambda increases ν\nu.


6.2.10 A statistical-mechanics view (why quantization matters)

Classically, energy is continuous and equipartition dumps kBTk_B T into each mode no matter how high its frequency. Quantum statistically, the occupation of a mode with energy spacing hνh\nu is Bose–Einstein with zero chemical potential:

nˉ(ν)=1ehν/(kBT)1\bar n(\nu) = \frac{1}{e^{h\nu/(k_B T)} - 1}

and the mean energy per mode is hνnˉ(ν)h\nu\,\bar n(\nu). High-frequency modes are exponentially hard to populate because you must pay hνh\nu upfront; low-frequency modes behave classically because hνkBTh\nu \ll k_B T makes nˉ\bar n large and εkBT\langle \varepsilon \rangle \to k_B T. The catastrophe disappears because discrete energy spacing throttles the UV.


6.2.11 Planck’s “I didn’t mean to be a revolutionary” arc

Planck was a dyed-in-the-wool thermodynamics guy who wanted a principled derivation of Wien’s empirical law. After the Rubens data contradicted Wien at low ν\nu, he sought an interpolation that matched both limits and then reverse-engineered a statistical story to justify it. He later called the quantization step “an act of desperation.” The act stuck. Within five years Einstein gave those energy packets—photons—independent life in the photoelectric effect. Planck’s constant hh went from curve-fitting fudge to a fundamental unit of nature.


6.2.12 Practical knobs: emissivity and real surfaces

Real materials are not perfect blackbodies. If a surface has emissivity ϵ(ν,T)\epsilon(\nu,T) between 0 and 1, its spectral radiance is

Iνreal(ν,T)=ϵ(ν,T)Iνblack(ν,T)I_\nu^{\text{real}}(\nu,T) = \epsilon(\nu,T)\,I_\nu^{\text{black}}(\nu,T)

Kirchhoff’s law of thermal radiation says emissivity equals absorptivity at each frequency, so cavities with a small aperture are effectively black even if the walls are not—rays bounce until they hit an absorptive patch. This is why laboratory blackbodies are cavities.


6.2.13 Worked mini-examples

(a) Power of a star. Approximating a star as a blackbody of radius RR and surface temperature TT, its luminosity is

L=4πR2σT4L = 4\pi R^{2}\,\sigma T^{4}

(b) Peak color of a filament. Using λpeakT=b\lambda_{\text{peak}}\,T=b, heating a filament so TT doubles halves the peak wavelength, shifting from red toward blue. Photographs feel the shift because camera sensors respond to uλu_\lambda weighting.

(c) Cavity photon number density. The photon number spectrum is nν=uν/(hν)n_\nu = u_\nu/(h\nu), hence the total number density is

n(T)=0uνhνdν=8πc30ν2dνehν/(kBT)1n(T) = \int_{0}^{\infty}\frac{u_\nu}{h\nu}\,d\nu = \frac{8\pi}{c^{3}}\int_{0}^{\infty}\frac{\nu^{2}\,d\nu}{e^{h\nu/(k_B T)} - 1}

Changing variables x=hν/(kBT)x=h\nu/(k_B T) yields n(T)T3n(T)\propto T^{3} with a proportionality constant involving ζ(3)\zeta(3).


6.2.14 What to remember (exam-tier TL;DR)

  • Geometry gives the density of modes g(ν)=8πν2/c3g(\nu) = 8\pi \nu^{2}/c^{3}; physics gives the average energy per mode
  • Classical equipartition \Rightarrow Rayleigh–Jeans ν2T\propto \nu^{2}T \Rightarrow ultraviolet catastrophe
  • Planck: energy exchange in quanta hνh\nu \Rightarrow ε=hν/(ehν/(kBT)1)\langle \varepsilon \rangle = h\nu/(e^{h\nu/(k_B T)}-1)
  • Multiply them to get Planck’s law; integrate to get Stefan–Boltzmann I=σT4I=\sigma T^{4}; maximize to get Wien’s displacement
  • Quantization is not optional decoration; it is the only way to make thermodynamics and electromagnetism coexist peacefully in a cavity

6.3 Photoelectric Effect and Einstein’s Light Quanta

Shine light on a clean metal and electrons pop out. Classical waves could explain “more light → more energy,” but experiments in the early 1900s said “frequency first, intensity second.” That plot twist pushed Einstein to a bold take: light energy arrives in discrete packets of size hνh\nu. The photoelectric effect is where photons stop being poetry and start being accounting.


6.3.1 The experimental rules (what needed explaining)

Laboratories found a crisp set of facts:

  • There exists a threshold frequency ν0\nu_0 below which no electrons are emitted, however intense the light
  • Above threshold, the maximum kinetic energy of emitted electrons depends linearly on frequency, not intensity
  • The emission is prompt on the scale of the light cycle; no measurable delay even at low intensities
  • The photoelectric current (number of electrons per second) scales with light intensity for fixed frequency, up to a saturation set by the apparatus

These rules are weird in a pure wave picture where energy is spread out over the wavefront and can accumulate continuously.


6.3.2 Einstein’s one-liner

Einstein’s 1905 hypothesis: light of frequency ν\nu behaves as a stream of quanta (photons), each carrying energy hνh\nu. An electron in the solid that absorbs one photon spends part of that energy to escape the material—this is the work function WW—and keeps the rest as kinetic energy. Thus

Kmax=hνWK_{\max} = h\nu - W

The threshold frequency is therefore

ν0=Wh\nu_0 = \frac{W}{h}

Intensity controls how many photons arrive per second, hence how many electrons can be emitted per second, but not the maximum energy of each electron. Frequency sets the budget per electron.


6.3.3 Measuring hh and WW with stopping potentials

Experimenters measure the maximum kinetic energy by applying a retarding voltage VsV_s to stop the most energetic electrons from reaching the collector. The definition is

eVs=Kmaxe\,V_s = K_{\max}

Combining with Einstein’s relation gives a straight-line plot

eVs=hνWe\,V_s = h\nu - W

So a graph of VsV_s versus ν\nu has slope h/eh/e and intercept W/e-W/e. That is how early experiments extracted Planck’s constant and the work function directly from electron kinetics.


6.3.4 Why the classical wave picture face-plants

A classical wave deposits energy continuously at a rate proportional to intensity. If that were the whole story, then at low intensities electrons should need a long time to accumulate enough energy to escape (a measurable delay), and for sufficiently intense low-frequency light the effect should happen anyway—no threshold. Neither occurs. The data scream “energy arrives in lumps,” with each electron grabbing at most one lump.


6.3.5 Millikan’s reluctant confirmation

Robert Millikan spent a decade trying to break Einstein’s interpretation by doing gold-standard measurements. He refined vacuum systems, surface preparation, and electronics. The result was the cleanest straight lines in physics: eVseV_s versus ν\nu with slope hh and a frequency intercept at W/hW/h. Millikan didn’t like the photon idea philosophically, but his data left no wiggle room. If your enemy measures your theory and it wins, that’s a W.


6.3.6 From intensity to flux: how many electrons?

Hold ν\nu fixed above threshold. Doubling intensity doubles the photon flux at the surface. If the surface is clean and space-charge effects are controlled, the saturation current IsatI_{\text{sat}} scales linearly with intensity because each absorbed photon can liberate at most one electron. The fraction that actually escape is the quantum efficiency η\eta; thus

IsatηPhνeI_{\text{sat}} \approx \eta\,\frac{P}{h\nu}\,e

where PP is the optical power absorbed on the active area. Real values of η\eta depend on the material, surface state, and whether photoelectrons are backscattered or trapped before reaching the collector.


6.3.7 Energy distribution and the role of the solid

KmaxK_{\max} is the upper edge. The distribution of kinetic energies depends on the band structure and scattering within the solid. In simple metals one can sketch an “initial state” between the Fermi level and a few kBTk_B T below it; an electron that absorbs a photon emerges with energy roughly hνWh\nu - W minus any losses to inelastic scattering on the way out. That is why measured spectra are broad, not delta functions, even though the top edge follows hνWh\nu - W cleanly.


6.3.8 Time-of-flight and “instantaneous” emission

Modern fast electronics confirm the old claim: emission is prompt within picoseconds or faster, consistent with a single-photon process instead of a long integration. In photon language, a single absorption event transfers hνh\nu in one go; there is no need to “wait” for energy to build up.


6.3.9 Surface physics and work function engineering

The work function WW is a surface property: it depends on the material, crystal face, cleanliness, and adsorbates. Alkali metals with low WW respond to red light; clean noble metals require blue/UV. Coatings and cesiation can lower WW, tuning ν0=W/h\nu_0=W/h downward. This is why phototubes and photocathodes care deeply about surface preparation.


6.3.10 Typical lab analysis workflow

  1. Illuminate a freshly prepared metal surface with monochromatic light at several frequencies
  2. For each ν\nu, sweep the collector voltage negative until the photocurrent drops to zero to find VsV_s
  3. Plot eVseV_s versus ν\nu; fit a line to extract slope hh and intercept W-W
  4. Independently, measure saturation current versus intensity at fixed ν\nu to confirm linear scaling of electron count with photon flux

The two experiments probe different pieces of Einstein’s model: the energy per photon and the counting of photons.


6.3.11 What about intensity at fixed frequency?

At fixed ν>ν0\nu>\nu_0, increasing intensity raises the number of emitted electrons (current) but leaves the maximum energy unchanged because hνh\nu is unchanged. Conversely, at fixed intensity, increasing frequency raises KmaxK_{\max} linearly. On the VsV_sν\nu plot, intensity just changes how quickly you reach saturation, not the slope or intercept.


6.3.12 Photons, momentum, and consistency checks

Photons also carry momentum p=h/λp=h/\lambda. In photoemission from a flat surface, most of the photon momentum is small compared with electron Fermi momenta, so pp is less central than hνh\nu for escape energetics, but radiation pressure and light-driven forces make sense in the same bookkeeping. Later phenomena—Compton scattering of X-rays—extend the particle-like picture by showing wavelength shifts that require both photon energy and photon momentum conservation. The photoelectric effect is the gentle on-ramp; Compton is the highway.


6.3.13 Why “one photon → one electron” has caveats

At high intensities and short pulses, multi-photon photoemission can occur: two (or more) photons combine to eject an electron even when hνh\nu is below WW. The signature is a nonlinear dependence of current on intensity. The original continuous-wave experiments lived in a regime where the single-photon process dominates, which is why Einstein’s linear law described them. Quantum mechanics contains both regimes—you just move along the intensity axis.


6.3.14 Quick derivations you can do on a napkin

Threshold. Set Kmax=0K_{\max}=0 in Einstein’s law to get ν0=W/h\nu_0=W/h. This is the cleanest way to define WW experimentally with light.

Slope extraction. From eVs=hνWeV_s=h\nu - W, the slope of VsV_s vs ν\nu is h/eh/e. Measure it and you have a value of hh independent of blackbody fits or spectroscopy.

Work function units. If WW is reported in electron-volts, then the threshold in hertz is

ν0=W(eV)h/e\nu_0 = \frac{W\,\text{(eV)}}{h/e}

and in wavelength

λ0=cν0\lambda_0 = \frac{c}{\nu_0}

It is common to quote λ0\lambda_0 for a photocathode because engineers think in wavelengths.


6.3.15 Photoelectric devices: from bench to tech

  • Vacuum photodiodes and photomultipliers. Use photoemission from a low-WW cathode to seed electrons, then multiply via dynodes. The VsV_s method turns into gain calibration and noise analysis
  • Solar cells (not the same effect). Photovoltaic junctions convert light to electricity via electron–hole pairs and built-in fields, not vacuum photoemission, but the hνh\nu threshold idea still sets which photons can be absorbed
  • Ultrafast electron sources. Femtosecond lasers drive single- or few-photon emission to produce short electron bunches for diffraction and microscopy

6.3.16 Conceptual audit

  • Particle-like energy transfer. hνh\nu per photon, one photon per primary electron escape event
  • Material gatekeeper. The work function WW sets the doorway; ν0=W/h\nu_0=W/h is the key frequency
  • Linear edge. KmaxK_{\max} tracks hνWh\nu-W regardless of intensity; stopping potential nails it as eVseV_s
  • Counting vs energy. Intensity changes how many electrons, frequency changes how energetic they can be

Quantum takes the wheel precisely where classical physics insists on the wrong dependencies.


6.4 Atomic Spectra and Bohr’s Model

Spectral lines were the OG QR codes of atoms—thin, repeatable, and screaming “there’s structure in here.” Long before quantum mechanics became a full theory, careful spectroscopy mapped those lines into simple integer patterns. The old quantum theory’s biggest W was Bohr’s 1913 model: a semi-classical atom with quantized orbits that nailed hydrogen’s spectrum and hinted at a deeper wave picture.


6.4.1 Empirical prelude: Balmer–Rydberg patterns

Hydrogen’s visible lines fit Balmer’s 1885 formula; Rydberg soon generalized it to a wavenumber series

ν~1λ=R(1nf21ni2),ni>nf\tilde\nu \equiv \frac{1}{\lambda} = R\,\left(\frac{1}{n_f^{2}} - \frac{1}{n_i^{2}}\right),\qquad n_i>n_f

Here RR is the Rydberg constant and (ni,nf)(n_i,n_f) are positive integers labeling initial and final “levels.” Famous series:

  • Lyman nf=1n_f=1 (ultraviolet)
  • Balmer nf=2n_f=2 (visible; Hα is ni=32n_i=3\to2 at 656\sim 656\,nm)
  • Paschen nf=3n_f=3 (infrared)

These were pure numerology—but spookily consistent across gases once you inserted effective constants.


6.4.2 Rutherford’s atom and the classical fail

Rutherford’s 1911 scattering showed a tiny, charged nucleus. Classical electrons orbiting such a nucleus are accelerated charges; Maxwell says they radiate and spiral in. That predicts continuous spectra and no stable atoms. Yet nature flexed discrete lines and stability. Enter Bohr.


6.4.3 Bohr’s postulates

Bohr’s model stapled three rules onto classical mechanics:

  1. Stationary states. Certain orbits are allowed and do not radiate despite acceleration
  2. Quantum of action. Angular momentum is quantized
mevr=n,n=1,2,3,m_e v r = n\,\hbar,\qquad n=1,2,3,\dots
  1. Frequency condition. Radiation occurs only when jumping between stationary states, with photon energy
hν=EiEfh\nu = E_i - E_f

These hacks were audacious—and shockingly predictive for one-electron atoms.


6.4.4 Coulomb plus quantization: solving for radii, speeds, energies

For a nucleus of charge +Ze+Ze and an electron in a circular orbit of radius rr and speed vv, centripetal force equals Coulomb attraction

mev2r=14πε0Ze2r2\frac{m_e v^{2}}{r} = \frac{1}{4\pi \varepsilon_0}\,\frac{Z e^{2}}{r^{2}}

Combine with the quantization mevr=nm_e v r = n\hbar to eliminate vv and solve for rnr_n:

rn=4πε02mee2n2Zr_n = \frac{4\pi \varepsilon_0\,\hbar^{2}}{m_e e^{2}}\,\frac{n^{2}}{Z}

The bracket is the Bohr radius

a04πε02mee2=mecαa_0 \equiv \frac{4\pi \varepsilon_0\,\hbar^{2}}{m_e e^{2}} = \frac{\hbar}{m_e c\,\alpha}

so rn=n2Za0r_n = \dfrac{n^{2}}{Z}\,a_0. The orbital speed follows as

vn=Zαcnv_n = \frac{Z\,\alpha\,c}{n}

where

αe24πε0c\alpha \equiv \frac{e^{2}}{4\pi \varepsilon_0 \hbar c}

is the fine-structure constant. The total energy is kinetic plus potential

En=12mevn214πε0Ze2rnE_n = \frac{1}{2} m_e v_n^{2} - \frac{1}{4\pi \varepsilon_0}\,\frac{Z e^{2}}{r_n}

which evaluates to

En=mee42(4πε0)22Z2n2=12mec2α2Z2n2E_n = -\,\frac{m_e e^{4}}{2 (4\pi \varepsilon_0)^{2} \hbar^{2}}\,\frac{Z^{2}}{n^{2}} = -\,\frac{1}{2}\,m_e c^{2}\,\alpha^{2}\,\frac{Z^{2}}{n^{2}}

For hydrogen (Z=1Z=1), the ground state is

E1=13.6eVE_1 = -\,13.6\,\text{eV}

and levels scale like 13.6eV/n2-13.6\,\text{eV}/n^{2}.


6.4.5 From Bohr levels to the Rydberg formula

Bohr’s frequency condition with EiE_i and EfE_f gives a photon wavenumber

ν~=νc=EiEfhc=mecα22hZ2(1nf21ni2)\tilde\nu = \frac{\nu}{c} = \frac{E_i - E_f}{h c} = \frac{m_e c\,\alpha^{2}}{2 h}\,Z^{2}\left(\frac{1}{n_f^{2}} - \frac{1}{n_i^{2}}\right)

Thus the Rydberg constant in the infinite nuclear mass limit is

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

For real hydrogen, the electron orbits the center of mass; replace mem_e by the reduced mass

μ=memNme+mN\mu = \frac{m_e m_N}{m_e + m_N}

to get

RH=RμmeR_{\text{H}} = R_\infty\,\frac{\mu}{m_e}

This tiny correction produces isotope shifts between H and D.


6.4.6 Why mvr=nm v r = n\hbar felt less random over time

A decade later, de Broglie proposed matter waves with λ=h/p\lambda = h/p. Demanding an integer number of wavelengths around a circular orbit,

2πrn=nλ=nhp2\pi r_n = n\,\lambda = n\,\frac{h}{p}

implies

prn=np r_n = n\,\hbar

i.e., Bohr’s rule. The standing wave picture explains why only certain radii are stable: destructive interference kicks out the rest. This wave logic is what Schrödinger will formalize in Chapter 7.


6.4.7 Correspondence principle and large nn

Bohr insisted that for large quantum numbers, quantum predictions must approach classical results. For n1n\gg1, the orbital frequency of the electron matches the emission frequency between neighboring levels, and the pattern of allowed transitions mimics the classical Fourier components of a Coulomb orbit. This “correspondence principle” guided the old quantum theory when full mechanics was still loading.


6.4.8 Selection rules and intensities (why some lines are strong)

Bohr’s model locates levels but doesn’t compute line intensities. Empirically, strong lines obey angular-momentum-like selection rules such as Δl=±1\Delta l=\pm 1 that later emerge from dipole transition matrix elements in wave mechanics. The old theory guessed at them via classical radiation patterns; quantum mechanics derives them cleanly.


6.4.9 Worked mini-derivations

(a) Lyman-α wavelength. From ni=2nf=1n_i=2\to n_f=1,

ν~=RH(114)=34RH\tilde\nu = R_{\text{H}}\left(1 - \frac{1}{4}\right) = \frac{3}{4} R_{\text{H}}

so

λLyα=43RH\lambda_{\text{Ly}\alpha} = \frac{4}{3 R_{\text{H}}}

Numerically this lands near 121.6121.6\,nm once μ\mu is included.

(b) Balmer Hα line. For ni=3nf=2n_i=3\to n_f=2,

ν~=RH(1419)=536RH\tilde\nu = R_{\text{H}}\left(\frac{1}{4} - \frac{1}{9}\right) = \frac{5}{36} R_{\text{H}}

so

λHα=365RH\lambda_{\text{H}\alpha} = \frac{36}{5 R_{\text{H}}}

which evaluates to about 656.3656.3\,nm.

(c) Ionization energy from level nn. Send nn\to\infty; the needed energy is

Eion(n)=En=13.6eVn2E_{\text{ion}}(n) = -E_n = \frac{13.6\,\text{eV}}{n^{2}}

for hydrogen.


6.4.10 Sommerfeld’s ellipses and early fine structure

To push further, Sommerfeld allowed elliptical orbits and quantized both the radial and angular actions. Adding a relativistic mass–velocity correction produced a partial account of fine structure (small level splittings) and the anomalous spacing of closely lying lines. It was clever but messy, and ran out of road on multi-electron atoms and precision details.


6.4.11 Where Bohr wins, where it breaks

Wins

  • Predicts En1/n2E_n \propto -1/n^{2} and the exact hydrogen and hydrogenic spectra when reduced mass is included
  • Explains series limits and ionization energies
  • Anticipates de Broglie and provides a clean bridge to wave mechanics

Breaks

  • Cannot explain line intensities and polarizations from first principles
  • Fumbles fine structure fully and the Zeeman effect anomalies
  • Faceplants on helium and other multi-electron atoms
  • Treats non-radiating “accelerated” electrons by fiat rather than mechanism

These failures are not bugs—they are breadcrumbs pointing to a wave equation and spin.


6.4.12 Practical constants and forms you will use

Equivalent and handy expressions:

a0=4πε02mee2=mecαa_0 = \frac{4\pi \varepsilon_0 \hbar^{2}}{m_e e^{2}} = \frac{\hbar}{m_e c\,\alpha} En=12mec2α2Z2n2E_n = -\,\frac{1}{2} m_e c^{2} \alpha^{2}\,\frac{Z^{2}}{n^{2}} R=α2mec2hR_\infty = \frac{\alpha^{2} m_e c}{2 h}

For a hydrogenic ion with nuclear charge ZZ, replace RHR_{\text{H}} by Z2RHZ^{2} R_{\text{H}} and EnE_n by 13.6Z2eV/n2-13.6\,Z^{2}\,\text{eV}/n^{2}.


6.4.13 Concept check: Bohr vs wave mechanics

Bohr’s radii rnr_n correspond to maxima in the radial probability density of the exact hydrogen wavefunctions. The quantum number nn remains, but ll and mm split levels and determine shapes; stationary states become standing waves in three dimensions. The photon frequency condition survives as energy eigenvalue differences, while selection rules arise from matrix elements of the dipole operator.


6.4.14 Minimal problem kit

  • Derive rnr_n and EnE_n from force balance plus mvr=nm v r=n\hbar and check a0a_0 numerically
  • Starting from EnE_n, recover the Rydberg formula and identify RR_\infty in terms of α\alpha
  • Include the reduced mass and estimate the H–D isotope shift of Hα
  • Use vn=Zαc/nv_n=Z\alpha c/n to check nonrelativistic consistency (v1αcv_1\approx \alpha c for H)

6.5 Rutherford Scattering and the Nuclear Atom

Sometimes one experiment deletes an entire worldview. The Geiger–Marsden gold-foil experiment did exactly that. They fired alpha particles (helium nuclei) at an ultra-thin gold foil and counted scintillations on a screen. Most alphas zipped through with tiny deflections, but a surprising few bounced at large angles, even backward. If positive charge were smeared throughout the atom (Thomson’s “plum pudding”), big-angle kicks would be essentially impossible. The data demanded a concentrated, tiny, positively charged nucleus.


6.5.1 What was measured and why it shocked people

  • Beam: alpha particles around a few MeV
  • Target: sub-micron gold foil, atomic number Z=79Z=79
  • Detector: zinc sulfide screen counting single scintillations
  • Surprise: an exceedingly small fraction scattered through tens of degrees or more, some near 180180^\circ

Classically, if positive charge were diffuse, only small, cumulative nudges add up. Large-angle events practically require a short-range, strong encounter with a compact center—i.e., a nucleus.


6.5.2 Coulomb repulsion and hyperbolic orbits

Treat the interaction as repulsive Coulomb between an alpha of charge zez e (with z=2z=2) and a nuclear charge +Ze+Ze:

V(r)=14πε0Zze2rV(r) = \frac{1}{4\pi\varepsilon_0}\,\frac{Z z e^2}{r}

It is convenient to define

kZze24πε0k \equiv \frac{Z z e^2}{4\pi\varepsilon_0}

Nonrelativistic trajectories are hyperbolae. The relation between scattering angle θ\theta and impact parameter bb for a projectile of kinetic energy EE is

θ=2arctan ⁣(k2Eb)\theta = 2\,\arctan\!\left(\frac{k}{2 E b}\right)

which inverts to

b(θ)=k2Ecot ⁣(θ2)b(\theta) = \frac{k}{2E}\,\cot\!\left(\frac{\theta}{2}\right)

Here E=12μv2E=\tfrac{1}{2}\mu v^2 with reduced mass μ\mu; for a very heavy target, μmα\mu\approx m_{\alpha} and lab \approx CM.


6.5.3 The Rutherford differential cross section

The measurable is the angular distribution, i.e. the differential cross section:

dσdΩ=bsinθdbdθ\frac{d\sigma}{d\Omega} = \frac{b}{\sin\theta}\,\left|\frac{db}{d\theta}\right|

Using b(θ)b(\theta) above and db/dθ=k4Ecsc2(θ/2)db/d\theta = -\tfrac{k}{4E}\csc^2(\theta/2) gives the famed Rutherford formula

dσdΩ=(k4E)2csc4 ⁣(θ2)\frac{d\sigma}{d\Omega} = \left(\frac{k}{4E}\right)^{2}\,\csc^{4}\!\left(\frac{\theta}{2}\right)

Key scalings:

  • Z2z2\propto Z^{2} z^{2}, so heavier nuclei punch harder
  • E2\propto E^{-2}, so higher beam energy suppresses scattering
  • Diverges at small angles like θ4\theta^{-4}

That 1/sin4(θ/2)1/\sin^{4}(\theta/2) falloff matches the data over a wide range—until other physics (screening, nuclear forces, relativity) kicks in.


6.5.4 Distance of closest approach and nuclear scale

For head-on approach (b=0b=0, θ=π\theta=\pi) the kinetic energy converts into potential energy at the turning point rminr_{\min}:

E=krminE = \frac{k}{r_{\min}}

so

rmin=kEr_{\min} = \frac{k}{E}

Numerically, the handy constant is e2/(4πε0)1.44 eVnme^{2}/(4\pi\varepsilon_{0}) \approx 1.44\ \text{eV}\cdot\text{nm}. For z=2z=2 alphas on gold (Z=79Z=79), k79×2×1.44 eVnm227.5 eVnmk \approx 79\times 2\times 1.44\ \text{eV}\cdot\text{nm} \approx 227.5\ \text{eV}\cdot\text{nm}. With E=5 MeVE=5\ \text{MeV},

rmin227.5 eVnm5×106 eV4.6×105 nm=4.6×1014 mr_{\min} \approx \frac{227.5\ \text{eV}\cdot\text{nm}}{5\times 10^{6}\ \text{eV}} \approx 4.6\times 10^{-5}\ \text{nm} = 4.6\times 10^{-14}\ \text{m}

That is tens of femtometers—orders of magnitude smaller than the atomic size (1010 m\sim 10^{-10}\ \text{m}). Pushing to higher energies drives rminr_{\min} down toward a few femtometers, the scale of nuclear radii, consistent with a tiny dense core.


6.5.5 From counts to cross sections in a foil

For a thin target with atomic number density nn and thickness tt, the expected rate into solid angle ΔΩ\Delta\Omega at angle θ\theta is

countsNbeam  nt  dσdΩ(θ)  ΔΩ\text{counts} \propto N_{\text{beam}}\; n t\; \frac{d\sigma}{d\Omega}(\theta)\; \Delta\Omega

“Thin” means single scattering dominates and attenuation is negligible. At larger tt or for softer beams, multiple scattering blurs the pure Rutherford shape at small angles.


6.5.6 Where the classical picture bends

  • Atomic screening (small θ\theta). At large bb the projectile sees a nearly neutral atom; bound electrons screen nuclear charge. Effective ZZ drops, taming the θ4\theta^{-4} divergence. A crude cutoff angle corresponds to impact parameters beyond the screening length a0/Z1/3\sim a_{0}/Z^{1/3}

  • Nuclear size and strong force (large θ\theta, high EE). When rminr_{\min} approaches a few fm, the finite nuclear radius and the strong interaction spoil the pure Coulomb potential. Deviations at the largest angles or highest energies carry nuclear-structure info

  • Relativity and spin. For faster beams and electrons as projectiles, one needs Mott scattering, i.e., Dirac theory corrections and spin–orbit effects. The qualitative csc4(θ/2)\csc^{4}(\theta/2) survives as a backbone but with spin-dependent factors

  • Quantum mechanics. The quantum calculation (partial waves or Born approximation) reproduces the Rutherford formula for Coulomb potentials—remarkable agreement between classical and quantum for this case


6.5.7 What the experiment proved

  1. Positive charge and most mass live in a tiny volume compared to the atom
  2. Electrons are peripheral; the atom is mostly empty space
  3. Long-range deflection is Coulombic and centered, not a diffuse push

Those facts killed the plum-pudding model and launched the nuclear atom, setting the stage for Bohr’s quantized orbits (§6.4) and, later, quantum mechanics proper.


6.5.8 Worked mini-derivations

(a) Differential cross section

Start from b(θ)=k2Ecot(θ/2)b(\theta)=\tfrac{k}{2E}\cot(\theta/2)

dσdΩ=bsinθdbdθ=(k2Ecot(θ/2)sinθ)(k4Ecsc2 ⁣θ2)=(k4E)2csc4 ⁣θ2\frac{d\sigma}{d\Omega} = \frac{b}{\sin\theta}\left|\frac{db}{d\theta}\right| = \left(\frac{k}{2E}\frac{\cot(\theta/2)}{\sin\theta}\right)\left(\frac{k}{4E}\csc^{2}\!\frac{\theta}{2}\right) = \left(\frac{k}{4E}\right)^{2} \csc^{4}\!\frac{\theta}{2}

(b) Integrated yield beyond a cut angle θ0\theta_{0}

Using dσ/dΩcsc4(θ/2)d\sigma/d\Omega \propto \csc^{4}(\theta/2) and dΩ=2πsinθdθd\Omega=2\pi\sin\theta\,d\theta,

σ(θ>θ0)=2πθ0π ⁣(k4E)2csc4 ⁣θ2sinθdθ=π(k2E)2cot2 ⁣θ02\sigma(\theta>\theta_{0}) = 2\pi \int_{\theta_{0}}^{\pi}\! \left(\frac{k}{4E}\right)^{2} \csc^{4}\!\frac{\theta}{2}\,\sin\theta\,d\theta = \pi \left(\frac{k}{2E}\right)^{2}\,\cot^{2}\!\frac{\theta_{0}}{2}

Handy for estimating count rates when your detector subtends large angles

(c) Closest approach at nonzero bb

Conserve energy and angular momentum L=μvbL=\mu v b; the turning point rminr_{\min} satisfies

E=L22μrmin2+krminE = \frac{L^{2}}{2\mu r_{\min}^{2}} + \frac{k}{r_{\min}}

Solve the quadratic for 1/rmin1/r_{\min} to link geometry (b)(b) to how near the alpha gets


6.5.9 Problem kit

  • Use dσ/dΩd\sigma/d\Omega to predict the ratio of counts at 3030^\circ and 6060^\circ for the same detector solid angle
  • For 5 MeV5\ \text{MeV} alphas on Au, compute rminr_{\min} at b=0b=0 and compare to an estimate of the nuclear radius Rr0A1/3R\approx r_{0} A^{1/3} with r01.2 fmr_{0}\sim 1.2\ \text{fm}
  • Include screening by imposing a maximum impact parameter equal to a screening length and show how it regularizes the small-angle divergence
  • Derive the lab–CM relation for EE and show why using lab EE is an excellent approximation for Au\text{Au} targets

In summary: The gold-foil experiment showed that atoms are mostly vacuum with a tiny, massive, positively charged nucleus at the center. The Coulomb scattering math predicts a sharp, parameter-free angular law,

dσdΩ=(Zze216πε0E)2csc4 ⁣(θ2)\frac{d\sigma}{d\Omega} = \left(\frac{Z z e^{2}}{16\pi\varepsilon_0 E}\right)^{2} \csc^{4}\!\left(\frac{\theta}{2}\right)

and the observed large-angle events made the nuclear atom unavoidable. Once you accept a compact nucleus, Bohr’s quantized orbits and discrete spectra stop being magic tricks and start being architecture.


6.6 Matter Waves: de Broglie to Davisson–Germer

If light can act like particles, maybe particles can act like waves. That was Louis de Broglie’s audacious 1923–24 symmetry flex. He guessed that every material particle with momentum pp carries a wavelength λ=h/p\lambda=h/p. Wild? Sure. But the prediction stuck landing after a few years: electrons diffracted from crystals exactly like X-rays, with the measured angles matching λ=h/p\lambda=h/p. Matter waves were not a metaphor; they were lab reality.


6.6.1 The hypothesis

de Broglie’s one-liner is

λ=hp\lambda = \frac{h}{p}

Together with photon E=hν=ωE=h\nu=\hbar \omega and p=kp=\hbar k, he suggested the wave–particle dictionary

E=ω,p=kE = \hbar \omega,\qquad p = \hbar k

The two equalities say: if something carries energy–momentum (E,p)(E,\boldsymbol p), there is a wave with angular frequency ω\omega and wavevector k\boldsymbol k marching in step.


6.6.2 Phase and group velocities

For a free particle with dispersion relation E(p)E(\boldsymbol p), define

ω=E,k=p\omega = \frac{E}{\hbar},\qquad \boldsymbol k = \frac{\boldsymbol p}{\hbar}

Then

vphaseωk=Ep,vgroupkω=pEv_{\text{phase}} \equiv \frac{\omega}{k} = \frac{E}{p},\qquad \boldsymbol v_{\text{group}} \equiv \nabla_{\boldsymbol k}\,\omega = \nabla_{\boldsymbol p} E
  • Relativistic E=γmc2E=\gamma m c^{2} and p=γmvp=\gamma m v give
vphase=c2v,vgroup=vv_{\text{phase}} = \frac{c^{2}}{v},\qquad v_{\text{group}} = v

Phase can be superluminal with no drama (it carries no information); group velocity equals the particle speed

  • Nonrelativistic E=p2/(2m)E=p^{2}/(2m) gives
ω=k22m,vgroup=km=pm=v\omega = \frac{\hbar k^{2}}{2m},\qquad v_{\text{group}} = \frac{\hbar k}{m} = \frac{p}{m} = v

This is why wave packets can track particles


6.6.3 Wave packets and uncertainty (sneak preview)

A localized particle corresponds to a superposition of plane waves—a wave packet. If the momentum spread is Δp\Delta p (i.e., spread in kk), then the spatial width obeys

ΔxΔp2\Delta x \,\Delta p \gtrsim \frac{\hbar}{2}

The packet moves at vgroupv_{\text{group}} but disperses in time for nonrelativistic Ek2E\propto k^{2}, because different kk components travel at slightly different speeds. This is the seed of the uncertainty principle and of Schrödinger dynamics.


6.6.4 Electrons from a potential: handy formulas

Accelerate an electron through a potential difference VV; it gains kinetic energy eVeV. The nonrelativistic momentum is

p=2meeVp = \sqrt{2 m_e e V}

so

λ=h2meeV\lambda = \frac{h}{\sqrt{2 m_e e V}}

Engineers love the quick constant (with λ\lambda in ångström, VV in eV):

λ(A˚)12.27V(eV)\lambda(\text{Å}) \approx \frac{12.27}{\sqrt{V(\text{eV})}}

For higher voltages, include relativity:

λ=hc(eV)(eV+2mec2)\lambda = \frac{h c}{\sqrt{(eV)\,(eV + 2 m_e c^{2})}}

which reduces to the nonrelativistic result when eVmec2eV \ll m_e c^{2}


6.6.5 Diffraction: crystals as electron gratings

A crystal is a 3D diffraction grating. For planes of spacing dd, constructive interference follows Bragg’s law

nλ=2dsinθ,n=1,2,n\,\lambda = 2 d \sin\theta,\qquad n=1,2,\dots

Given λ=h/p\lambda=h/p, electrons should diffract with angles set by dd—just like X-rays. This was the smoking gun de Broglie asked the world to look for.


6.6.6 Davisson–Germer (and G. P. Thomson): the receipts

Davisson & Germer (1927). Electrons of tens of eV scattered from a nickel single crystal produced sharp intensity peaks at angles consistent with Bragg’s law when you plug in λ=h/p\lambda=h/p. Varying the accelerating voltage shifted the peak exactly as λ1/V\lambda\propto 1/\sqrt{V} predicts.

G. P. Thomson (also 1927) sent keV electrons through thin polycrystalline foils and recorded ring patterns—the powder-diffraction version of Bragg. Two independent routes, one conclusion: matter waves are real.

Rule of thumb: V150 eVV\sim 150\ \text{eV} gives λ1 A˚\lambda\sim 1\ \text{Å} (interatomic distances), perfect for crystal diffraction.


6.6.7 From Bohr to de Broglie: standing waves

Bohr’s angular-momentum postulate mvr=nm v r = n\hbar stops feeling arbitrary once you impose standing waves on a circular orbit:

2πr=nλ=nhp2\pi r = n\,\lambda = n\,\frac{h}{p}

which gives

pr=np r = n\,\hbar

Same quantization, now with wave logic. Unstable radii are those where the phase does not fit—destructive interference kicks them out. de Broglie supplied the mechanism Bohr’s model lacked.


6.6.8 Schrödinger’s motivation (why a wave equation exists)

Take the free-particle dispersion E=p2/(2m)E=p^{2}/(2m) and the de Broglie dictionary EitE\to i\hbar\,\partial_t, pi\boldsymbol p\to -i\hbar\nabla. Demand a linear differential equation whose plane-wave solutions have ω=k2/(2m)\omega=\hbar k^{2}/(2m). You land on

iψt=22m2ψ+Vψi\hbar\,\frac{\partial \psi}{\partial t} = -\,\frac{\hbar^{2}}{2m}\,\nabla^{2}\psi + V\psi

the time-dependent Schrödinger equation. In other words, de Broglie’s E(p)E(\boldsymbol p) is the dispersion law that Schrödinger’s PDE realizes in real space.

A complementary viewpoint uses Hamilton–Jacobi theory: write ψexp(iS/)\psi \sim \exp(i S/\hbar); then S\nabla S behaves like momentum. Quantum mechanics is classical mechanics plus phase and interference.


6.6.9 Electron double-slit: the mic drop

Send single electrons through two slits; record impacts one by one. The screen builds an interference pattern over time—then vanishes when you measure which slit each electron took. The pattern wavelength matches λ=h/p\lambda=h/p. The message is not that electrons smear; it is that probabilities interfere with phases set by action differences.


6.6.10 Matter-wave tech: microscopes and metrology

  • Electron microscopes. Resolution scales roughly like λ\sim \lambda, so swapping visible light (500 nm\sim 500\ \text{nm}) for 100 keV electrons (λ0.004 nm\lambda\sim 0.004\ \text{nm} with relativistic correction) makes nanoworld imaging routine
  • LEED/RHEED. Low- and reflection-energy electron diffraction fingerprint surface structures and growth in real time
  • Atom interferometry. Neutral atoms also carry de Broglie waves; splitting and recombining them measures gravity, rotations, and fundamental constants with absurd precision

6.6.11 Worked mini-examples

(a) De Broglie wavelength at 54 eV54\ \text{eV}.

Using λ(A˚)12.27/V(eV)\lambda(\text{Å})\approx 12.27/\sqrt{V(\text{eV})},

λ12.2754 A˚1.67 A˚\lambda \approx \frac{12.27}{\sqrt{54}}\ \text{Å} \approx 1.67\ \text{Å}

Squarely in the range of crystal spacings—no wonder early peaks showed up near modest voltages

(b) Voltage for λ=1.0 A˚\lambda=1.0\ \text{Å}.

Invert the same rule:

V(12.27λ(A˚))2 eV150 eVV \approx \left(\frac{12.27}{\lambda(\text{Å})}\right)^{2}\ \text{eV} \approx 150\ \text{eV}

(c) Group velocity check (nonrelativistic).

With ω=k2/(2m)\omega=\hbar k^{2}/(2m),

vgroup=dωdk=km=pm=vv_{\text{group}} = \frac{d\omega}{dk} = \frac{\hbar k}{m} = \frac{p}{m} = v

The packet moves like the particle—consistency unlocked

(d) Bragg angle estimate.

For d=2.0 A˚d=2.0\ \text{Å}, λ=1.0 A˚\lambda=1.0\ \text{Å}, first order n=1n=1,

sinθ=λ2d=14θ14.5\sin\theta = \frac{\lambda}{2d} = \frac{1}{4}\quad\Rightarrow\quad \theta \approx 14.5^\circ

Real experiments include geometry factors (incidence vs detection), but the scaling is right on the nose


6.6.12 Subtleties and limits

  • Wave packet spread. Nonrelativistic packets broaden with time; real beams are continually collimated or detected before big dispersion
  • Scattering versus bound states. In crystals, elastic diffraction uses the crystal as a giant phase grating; in atoms, bound-state standing waves quantize energies
  • Heavy particles. λ=h/p\lambda=h/p shrinks fast with mass; neutron and atom interferometers compensate by running at ultra-low velocities to get usable wavelengths
  • Relativistic electrons. Use the corrected λ\lambda; at 100 keV, nonrelativistic formulas understate the momentum (and overstate λ\lambda)

6.6.13 Problem kit

  • Show that vphasevgroup=c2v_{\text{phase}}\,v_{\text{group}}=c^{2} for a free relativistic particle
  • For an electron beam at 200 keV200\ \text{keV}, compute λ\lambda with and without relativistic correction and compare
  • Given a cubic crystal with lattice constant aa, list dd-spacings for (100)(100), (110)(110), (111)(111) and predict first-order Bragg angles for a chosen λ\lambda
  • Build a Gaussian wave packet ψ(x,0)\psi(x,0), evolve it with the free Schrödinger kernel, and extract the time dependence of the rms width

6.7 Old Quantum Theory: Hits and Misses

Before full quantum mechanics landed in 1925–26, physicists hacked classical mechanics with a few bold patches. They quantized certain integrals, imposed selection rules, and stitched in Bohr’s frequency condition. This old quantum theory nailed hydrogen’s spectrum and hinted at deeper structure, but it stumbled on multi-electron atoms, intensities, and magnetic quirks. This section is the status report: what worked, what broke, and what carried over into modern QM.


6.7.1 What “old quantum” meant

The playbook had three pillars:

  1. Stationary states exist that do not radiate while the electron whirls in a Coulomb orbit
  2. Bohr–Sommerfeld quantum conditions select allowed motions
  3. Bohr frequency condition connects level differences to light

In symbols, energy levels EnE_n and jumps nnn\to n' emit or absorb photons with

hν=EnEnh\nu = E_n - E_{n'}

The deep math piece was the quantization of actions.


6.7.2 Bohr–Sommerfeld quantum conditions

For a system with separable coordinates qiq_i and conjugate momenta pip_i, allowed orbits satisfy

pidqi=nih,ni=0,1,2,\oint p_i\,dq_i = n_i\,h,\qquad n_i=0,1,2,\dots

The line integral runs over one full period of motion in coordinate qiq_i. Define the action variables

JipidqiJ_i \equiv \oint p_i\,dq_i

Then the rule is simply Ji=nihJ_i = n_i h. These JiJ_i are adiabatic invariants: under slow changes of parameters, they remain constant, a fact Ehrenfest elevated into the adiabatic principle. Where the dynamics is integrable and separable, the rules bite; where it is not, they flail.


6.7.3 Hydrogen, again but more general

Bohr’s circular orbits are a special case. With Sommerfeld’s ellipses, the Coulomb problem has two actions: radial JrJ_r and angular JϕJ_\phi. Quantization gives

Jϕ=h,Jr=nrh,n=nr+J_\phi = \ell\,h,\qquad J_r = n_r\,h,\qquad n = n_r + \ell

and the total energy depends only on nn:

En=mee42(4πε0)221n2E_n = -\,\frac{m_e e^{4}}{2 (4\pi\varepsilon_0)^2 \hbar^2}\,\frac{1}{n^{2}}

The principal quantum number nn counts nodes in the old theory too, but without a wavefunction to say it out loud. Sommerfeld also added a relativistic mass–velocity correction and extracted a qualitative fine structure splitting, a hint of what Dirac would do properly.


6.7.4 Harmonic oscillator: the “almost right” spectrum

For a 1D harmonic oscillator with frequency ω\omega, the action integral evaluates to

J=pdq=2πEωJ = \oint p\,dq = \frac{2\pi E}{\omega}

Bohr–Sommerfeld demands J=nhJ = n h, hence

En=nωE_n = n\,\hbar\omega

This is close to the truth; modern QM gives En=(n+12)ωE_n=(n+\tfrac12)\hbar\omega. The missing 12\tfrac12 is the vacuum’s zero-point energy, one of several places where old quantum theory is off by a half.


6.7.5 Correspondence principle and selection rules

Bohr’s correspondence principle guided guesses: in the limit of large quantum numbers n1n\gg 1, quantum predictions must morph into classical ones. For bound motion one can Fourier-expand a classical trajectory and show that the strongest allowed quantum transitions tend to frequencies present in the classical spectrum. This logic nudged the empirical dipole selection rule

Δ=±1\Delta \ell = \pm 1

and explained why transitions with wildly different frequencies are feeble. The principle later reappeared as actual mathematics in the Kramers–Heisenberg dispersion formula and sum rules.


6.7.6 Einstein’s AA and BB coefficients: blackbody from jumps

In 1916–17 Einstein wrote a kinetic model of radiative transitions between two levels 11 and 22 with degeneracies g1,g2g_1,g_2. In a radiation field of spectral energy density ρ(ν)\rho(\nu), the rates are

absorption: N1B12ρ(ν),stimulated emission: N2B21ρ(ν),spontaneous emission: N2A21\text{absorption: } N_1 B_{12}\,\rho(\nu),\qquad \text{stimulated emission: } N_2 B_{21}\,\rho(\nu),\qquad \text{spontaneous emission: } N_2 A_{21}

Thermal equilibrium plus Boltzmann factors N2/N1=(g2/g1)exp(hν/kBT)N_2/N_1=(g_2/g_1)\exp(-h\nu/k_B T) then enforce

A21B21=8πhν3c3,g1B12g2B21=1\frac{A_{21}}{B_{21}} = \frac{8\pi h \nu^{3}}{c^{3}},\qquad \frac{g_1\,B_{12}}{g_2\,B_{21}} = 1

and reproduce Planck’s law for ρ(ν)\rho(\nu). This was the first theoretical appearance of stimulated emission, which lasers would monetize decades later. Old quantum theory could balance the books for radiation.


6.7.7 Dispersion theory and the first sum rules

Kramers recast optical dispersion in terms of virtual jumps between levels and, using correspondence, derived frequency-dependent polarizabilities that matched classical Lorentz–Drude behavior at low frequency but respected quantum selection rules. Heisenberg and Born then took the hint and replaced classical amplitudes by arrays of transition amplitudes—Heisenberg’s bridge to matrix mechanics. Even before that leap, sum rules relating oscillator strengths to charges were early “no free lunch” constraints on spectra.


6.7.8 Thermodynamics wins: specific heats of solids

Classical equipartition predicts a constant molar heat capacity CV=3RC_V=3R (Dulong–Petit). Data at low TT tank that prediction. Einstein’s model (1907) treats each atom as an independent quantum oscillator of frequency ωE\omega_E:

CV=3R(θET)2eθE/T(eθE/T1)2,θEωEkBC_V = 3R\,\left(\frac{\theta_E}{T}\right)^2 \frac{e^{\theta_E/T}}{\left(e^{\theta_E/T}-1\right)^2},\qquad \theta_E \equiv \frac{\hbar\omega_E}{k_B}

It captures the drop but not the exact low-TT power. Debye improved it by quantizing acoustic modes up to a cutoff ωD\omega_D, yielding CVT3C_V\propto T^3 at low TT and 3R3R at high TT. While not part of orbital quantization, these successes showed that discreteness fixes thermodynamics too.


6.7.9 Where old quantum theory face-planted

  • Helium and many-electron atoms. The three-body Coulomb problem has no closed integrals of motion to quantize cleanly. Ad hoc quantum conditions gave conflicting energies and wrong term structures. Helium was the “here be dragons” sign until wave mechanics and spin arrived

  • Anomalous Zeeman effect. In weak magnetic fields, many spectral lines split into more components than classical or orbital angular momentum alone can explain. The fix needs electron spin and spin–orbit coupling, which the old theory lacked

  • Intensities and polarizations. Level positions were OK for hydrogenic systems; line strengths, polarizations, and lifetimes demanded matrix elements of operators between wavefunctions—tools old quantum did not possess

  • Radiation from accelerated charges. Old quantum declared stationary orbits non-radiating by fiat, while classical electrodynamics says accelerating charges radiate. The rule worked phenomenologically but had no mechanism until wave mechanics turned “orbit” into a stationary standing wave

  • Nonseparable systems and chaos. The action-quantization rule needs separable coordinates and closed invariant tori in phase space. Many motions (think coupled anharmonic oscillators, generic molecules) do not cooperate. Without a PDE for ψ\psi, the method stalls

  • Zero-point energy. The oscillator’s missing 12ω\tfrac12\hbar\omega showed up in experiments sensitive to low-temperature fluctuations and later in field quantization. Old quantum had no universal way to assign it


6.7.10 What survived into modern quantum mechanics

  • Actions and adiabatic invariants became the semiclassical WKB/EBK quantization rules
pdq=(n+μ4)h\oint p\,dq = \left(n + \frac{\mu}{4}\right)h

with Maslov indices μ\mu supplying the missing halves

  • Correspondence principle remains a sanity check and underlies sum rules, selection rules, and semiclassical approximations

  • Einstein AA/BB coefficients and oscillator strengths live on in atomic physics, lasers, and astrophysical diagnostics

  • Quantum numbers (n,,m)(n,\ell,m) are still the labels; they are now eigenvalue data of commuting operators rather than areas under curves

  • De Broglie’s standing waves turned into the full eigenfunction story, explaining stability and radiation probabilities from overlaps, not orbits


6.7.11 Worked mini-examples

(a) Action of the Kepler problem

For the Coulomb potential, the radial action can be computed from the periapsis–apoapsis bounds. One finds Jϕ=hJ_\phi=\ell h and

Jr=h(n)J_r = h\,(n - \ell)

so EE depends only on nn, reproducing En1/n2E_n\propto -1/n^2 and degeneracy in \ell that wave mechanics later lifts via spin–orbit coupling and Lamb shifts

(b) Oscillator’s missing half

Given J=2πE/ωJ=2\pi E/\omega, J=nhJ=n h yields En=nωE_n=n\hbar\omega. WKB with turning points adds a phase π/2\pi/2 at each boundary, shifting the rule to J=(n+12)hJ=(n+\tfrac12)h and fixing En=(n+12)ωE_n=(n+\tfrac12)\hbar\omega

(c) Dipole selection

Take a classical circular orbit and expand its dipole moment into Fourier modes at multiples of the orbital frequency. The strongest harmonic is the first, mirroring Δ=±1\Delta \ell=\pm 1 and nearest-neighbor nn transitions in the large-nn limit


6.7.12 Problem kit

  • Starting from pdq=nh\oint p\,dq = n h, derive EnE_n for the harmonic oscillator and compare to the WKB-corrected (n+12)ω(n+\tfrac12)\hbar\omega
  • For Coulomb motion, compute JϕJ_\phi and JrJ_r and show EE depends only on n=nr+n=n_r+\ell
  • Use Einstein’s rate equations to re-derive ρ(ν)\rho(\nu) and extract A21/B21A_{21}/B_{21} and B12/B21B_{12}/B_{21} in terms of (g1,g2)(g_1,g_2)
  • Show how Kramers’ dispersion reduces to the classical Lorentz model when ωkBT\hbar\omega \ll k_B T and transitions crowd densely
  • Identify one system where variables are not separable and explain why Bohr–Sommerfeld quantization stalls without wave mechanics

In summary: Old quantum theory was a brilliant stopgap. It guessed the right invariants to quantize, respected a correspondence to classical physics, and even anticipated stimulated emission. But without wavefunctions, operators, and spin, it could not compute intensities, multi-electron spectra, or magnetic fine print. The successes became the scaffolding for Heisenberg’s matrices and Schrödinger’s waves; the failures marked the exact spots where the full theory had to grow


6.8 Spin, Stern–Gerlach, and Exclusion

Atoms were shouting a secret in the 1910s–20s: angular momentum was coming in half-units and electron seats were limited. Two experiments and one principle sealed the deal—Stern–Gerlach split neutral beams into discrete tracks, spin explained “anomalous” Zeeman splittings, and Pauli’s exclusion principle organized the periodic table and built the stability of matter. This section introduces spin as a bona fide quantum degree of freedom, its magnetic moment, how fields act on it, and why “no two identical fermions in the same state” is the rule holding up chemistry and white dwarfs.


6.8.1 Stern–Gerlach: space quantization goes brrr

In 1922, Otto Stern and Walther Gerlach sent a beam of neutral silver atoms through a non-uniform magnetic field. Classically, the component of magnetic moment along the field gradient would vary continuously, giving a smear on the screen. Instead they saw two sharp spots. Conclusion: the zz-component of angular momentum is quantized.

Silver was clutch because its valence structure leaves a single unpaired electron dominating the atom’s magnetic moment while the atom remains neutral (no Lorentz deflection). The force is

Fz=μzBzzF_z = \mu_z\,\frac{\partial B_z}{\partial z}

and the observed two-way split screams that μz\mu_z only takes two values. That is spin shouting “ms=±12m_s=\pm \tfrac12.”


6.8.2 Spin-12\tfrac12 algebra and eigenvalues

Quantum mechanically, spin components do not commute:

[Si,Sj]=iϵijkSk[S_i,S_j]=i\hbar\,\epsilon_{ijk}\,S_k

For an electron, the total spin and its zz-projection are

S2=s(s+1)2,s=12S^2 = s(s+1)\hbar^2,\qquad s=\frac12 Sz=2,Sz=2S_z\,\ket{\uparrow}=\frac{\hbar}{2}\ket{\uparrow},\qquad S_z\,\ket{\downarrow}=-\frac{\hbar}{2}\ket{\downarrow}

A convenient matrix representation uses the Pauli matrices σ\boldsymbol{\sigma} with Si=2σiS_i=\tfrac{\hbar}{2}\sigma_i:

σx=(0110),σy=(0ii0),σz=(1001)\sigma_x=\begin{pmatrix}0&1\\[2pt]1&0\end{pmatrix},\quad \sigma_y=\begin{pmatrix}0&-i\\[2pt]i&0\end{pmatrix},\quad \sigma_z=\begin{pmatrix}1&0\\[2pt]0&-1\end{pmatrix}

Spin states are two-component spinors. Rotations act via these matrices, and measurement along any axis picks one of two outcomes.


6.8.3 Magnetic moment, Zeeman splitting, and Larmor precession

The electron’s spin carries a magnetic moment

μs=gse2meS\boldsymbol{\mu}_s = -\,g_s\,\frac{e}{2m_e}\,\boldsymbol{S}

where gs2g_s\approx 2 from Dirac theory (with small QED corrections). Define the Bohr magneton

μBe2me\mu_B \equiv \frac{e\hbar}{2m_e}

In a uniform magnetic field B=Bz^\boldsymbol{B}=B\hat{\boldsymbol{z}}, the interaction energy is

HZ=μB=gsμBmsBH_Z = -\,\boldsymbol{\mu}\cdot\boldsymbol{B} = g_s\,\mu_B\,m_s\,B

so a spin-12\tfrac12 splits into two levels separated by

ΔE=gsμBB\Delta E = g_s\,\mu_B\,B

Classically flavored but quantum-exact, spins also precess about B\boldsymbol{B} at the Larmor frequency

ωL=gsμBB\omega_L = \frac{g_s\,\mu_B\,B}{\hbar}

This underlies magnetic resonance (NMR/EPR) and shows up as beat frequencies in time-dependent spin dynamics.


6.8.4 Normal vs “anomalous” Zeeman: spin to the rescue

Old spectroscopy found that many lines in a magnetic field split into patterns beyond the normal Zeeman triplet. The fix is the total angular momentum

J=L+S\boldsymbol{J}=\boldsymbol{L}+\boldsymbol{S}

with a level-specific Landé factor gJg_J. The Zeeman energy shift is

ΔE=μBgJmJB\Delta E = \mu_B\,g_J\,m_J\,B

where mJ=J,J+1,,Jm_J=-J,-J+1,\dots,J. Spin explains the “anomalous” patterns immediately: gJg_J comes out of vector-coupling algebra and selection rules ΔmJ=0,±1\Delta m_J=0,\pm 1 handle line multiplicities and polarizations. No epicycles required.


6.8.5 Spin–orbit coupling and Thomas factor

In atoms, an electron moving in the nuclear field sees, in its rest frame, a magnetic field that couples L\boldsymbol{L} and S\boldsymbol{S}:

HSO=ξ(r)LSH_{SO} = \xi(r)\,\boldsymbol{L}\cdot\boldsymbol{S}

A careful relativistic treatment adds the Thomas precession factor 12\tfrac12, bringing predictions in line with observed fine structure. The result: terms split by JJ even when nn and \ell are fixed, e.g., the famous doublets in alkali spectra.


6.8.6 Pauli’s exclusion principle: the two-per-seat rule

In 1925 Wolfgang Pauli proposed: no two identical fermions can occupy the same quantum state. For electrons in atoms that means at most two per spatial orbital, with opposite spins. Label a one-electron state by (n,,m,ms)(n,\ell,m_\ell,m_s); exclusion forbids duplicate quadruples. Immediate wins:

  • Periodic table architecture. Shells and subshells fill in the observed order (with correlation and screening tweaks), explaining chemical valence. Two per orbital gives the 2,8,18,2,\,8,\,18,\dots shell capacities
  • Magnetism. Most electrons are spin-paired; unpaired spins produce paramagnetism. In metals, Pauli paramagnetism of conduction electrons is weak and temperature-independent compared to Curie’s law for localized moments
  • Stability of matter. Electrons cannot all collapse into the lowest orbital; packed fermions build degeneracy pressure that counters compression, from metals’ mechanical properties to white dwarf support

Under the hood, exclusion is a theorem in relativistic quantum field theory: spin–statistics. Particles with half-integer spin (fermions) have antisymmetric many-body wavefunctions and obey Fermi–Dirac statistics; integer-spin particles (bosons) are symmetric and obey Bose–Einstein statistics.


6.8.7 Fermi–Dirac distribution and the Fermi sea

For identical fermions with single-particle energies ϵ\epsilon and chemical potential μ\mu,

nˉ(ϵ)=1exp ⁣(ϵμkBT)+1\bar n(\epsilon) = \frac{1}{\exp\!\left(\frac{\epsilon-\mu}{k_B T}\right)+1}

At T=0T=0, states fill up to the Fermi energy ϵF\epsilon_F. The factor of 2 for spin-12\tfrac12 doubles the density of states per energy, shifting ϵF\epsilon_F and controlling metallic properties. Pauli paramagnetic susceptibility in a simple metal scales with the density of states at ϵF\epsilon_F—tiny but nonzero.


6.8.8 Sequential Stern–Gerlach (the measurement vibes)

Run a beam through three magnets: SGz_z (keep only \uparrow), then SGx_x (split into \rightarrow and \leftarrow; keep \rightarrow), then SGz_z again. The last stage shows both \uparrow and \downarrow. Lesson: measuring along xx re-prepares the spin state and erases previous zz information. Algebraically,

z=12 ⁣(x+x)\ket{\uparrow_z}=\frac{1}{\sqrt2}\!\left(\ket{\rightarrow_x}+\ket{\leftarrow_x}\right)

so a projective measurement in xx randomizes zz outcomes. This is the cleanest tabletop demo that incompatible observables do not share simultaneous sharp values.


6.8.9 Electron beams and why Stern–Gerlach used neutrals

You might ask: why not a spin split with electrons? Free electrons in inhomogeneous fields feel huge Lorentz forces qv×Bq\,\boldsymbol{v}\times\boldsymbol{B} and electric-field forces from fringe fields, which swamp the tiny (μB)\nabla(\boldsymbol{\mu}\cdot\boldsymbol{B}) splitting. Neutral atoms dodge that, so Stern–Gerlach reached the quantum without EM steering drama.


6.8.10 Tiny but real deviations: g2g-2

Dirac’s gs=2g_s=2 is insanely accurate but not perfect. Quantum electrodynamics adds a small anomalous magnetic moment

aegs22a_e \equiv \frac{g_s-2}{2}

whose leading term is α/(2π)\alpha/(2\pi) and higher orders keep theorists busy. Precision measurements of aea_e are stress tests for the Standard Model and help pin the fine-structure constant α\alpha.


6.8.11 Worked mini-examples

(a) Zeeman splitting size

For B=1 TB=1\ \text{T} and gs2g_s\approx 2,

ΔE=gsμBB2×5.788×105 eV/T×1 T1.16×104 eV\Delta E = g_s\,\mu_B\,B \approx 2\times 5.788\times 10^{-5}\ \text{eV/T}\times 1\ \text{T}\approx 1.16\times 10^{-4}\ \text{eV}

Corresponding frequency ν=ΔE/h28 GHz\nu=\Delta E/h\sim 28\ \text{GHz}, i.e., microwave EPR territory

(b) SG deflection estimate

With gradient Bz/z=10 T/m\partial B_z/\partial z = 10\ \text{T/m}, flight time τ=104 s\tau=10^{-4}\ \text{s}, and mass of Ag atom, the zz-kick is

ΔpzμBBzzτ\Delta p_z \approx \mu_B\,\frac{\partial B_z}{\partial z}\,\tau

Translate to a spot separation on a screen a few tens of cm away to see that mm-scale splits are reasonable

(c) White dwarf scale (order-of-magnitude)

Electron degeneracy pressure P2mene5/3P\sim \frac{\hbar^2}{m_e} n_e^{5/3} balances gravity; pushing nen_e up increases PP without heat—exclusion is literally structural support. Details live in astrophysics, but the scaling rests on spin-12\tfrac12 plus antisymmetry


6.8.12 Problem kit

  • Derive the two-outcome split of Stern–Gerlach using SzS_z eigenvalues and estimate spot separation for given field gradient and flight time
  • Starting from HZ=μBH_Z=-\boldsymbol{\mu}\cdot\boldsymbol{B} and Si=2σiS_i=\tfrac{\hbar}{2}\sigma_i, show that spinors precess with ωL=gsμBB/\omega_L=g_s\mu_B B/\hbar
  • Using vector addition rules, compute gJg_J for a given term symbol 2S+1LJ^{2S+1}L_J and predict Zeeman components and polarizations
  • Show how HSO=ξ(r)LSH_{SO}=\xi(r)\,\boldsymbol{L}\cdot\boldsymbol{S} splits a given n,n,\ell multiplet into J=±12J=\ell\pm \tfrac12 and sketch selection-rule-allowed lines
  • From the Fermi–Dirac distribution, derive the zero-temperature electron number density and Fermi energy of a free-electron metal including spin degeneracy

In summary: Spin is not a tiny orbit; it is an intrinsic two-level degree of freedom with algebra, matrices, and a magnetic moment. Stern–Gerlach revealed its discreteness, Zeeman patterns and fine structure confirmed its coupling to fields and orbital motion, and Pauli’s exclusion principle turned spin into architecture—filling shells, setting chemistry, and propping up stars. Add in Fermi–Dirac statistics and g2g\approx 2, and you have the modern, compact story behind a century of “anomalies” that were really just spin waiting to be named


6.9 From Old Quantum to Wave/Matrix Mechanics

By 1925, the “old quantum” patches were running out of duct tape. Selection rules worked, spectra mostly matched, but intensities, multi-electron messiness, and the logic behind the rules begged for a fresh OS. Two different apps booted almost simultaneously: Heisenberg’s matrix mechanics (with Born and Jordan) and Schrödinger’s wave mechanics. Same physics, wildly different vibes. By 1926, they were proved equivalent, and quantum theory finally had a consistent calculus.


6.9.1 From dispersion to Umdeutung

Heisenberg’s leap started with optical dispersion. Classical electrons driven by light oscillate and emit at combination frequencies. Old quantum theory rephrased this via transition frequencies between stationary states, ωmn=(EmEn)/\omega_{mn}=(E_m-E_n)/\hbar, and amplitudes linked to line strengths. Heisenberg’s “Umdeutung” (reinterpretation) was to ditch unobservable electron orbits and keep only observable transition data, organizing them into arrays XmnX_{mn} with rows and columns indexed by energy levels. Composition of motions became matrix multiplication. Out went trajectories; in came algebras of transitions.


6.9.2 Matrix mechanics in one page

Observables are matrices AA with elements AmnA_{mn} in the energy eigenbasis. Time dependence carries the Bohr frequencies:

Amn(t)=AmneiωmntA_{mn}(t) = A_{mn}\,e^{i\omega_{mn} t}

The product rule that respects composition of transitions is matrix multiplication. The Newtonian equation mx¨=V/xm\ddot x = -\partial V/\partial x becomes, after the dictionary xXx\to X, pPp\to P,

mX¨=V(X)Xm\,\ddot X = -\,\frac{\partial V(X)}{\partial X}

Time derivatives act elementwise via the phases, and one demands consistency with the correspondence principle at large quantum numbers. The crucial structure constant is the canonical commutator

[X,P]XPPX=iI[X,P] \equiv X P - P X = i\hbar\,\mathbb I

This is not an optional spice; it is the algebraic heartbeat that replaces Poisson brackets. With it, the harmonic oscillator, rotor, and hydrogen spectrum drop out in purely algebraic fashion.


6.9.3 Uncertainty from noncommutation

Once [X,P]=i[X,P]=i\hbar is law, spreads of incompatible observables cannot both be squeezed to zero. The general inequality reads

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

Take A=XA=X and B=PB=P to get the famous bound

ΔxΔp2\Delta x\,\Delta p \ge \frac{\hbar}{2}

This is not about measurement sloppiness; it is geometry in Hilbert space. The bound follows from Cauchy–Schwarz and the operator algebra, independent of any postulated “disturbance.”


6.9.4 Schrödinger’s wave mechanics

Schrödinger arrived from the opposite coast: waves. Starting with de Broglie’s E=ωE=\hbar\omega, p=kp=\hbar k, require a wave equation whose plane-wave solutions carry the correct dispersion. The result is the time-dependent Schrödinger equation

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

Looking for stationary states ψ(r,t)=ϕ(r)eiEt/\psi(\mathbf r,t)=\phi(\mathbf r)\,e^{-iEt/\hbar} yields the time-independent form

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

Boundary conditions discretize EE and generate the spectral lines Bohr had cataloged.


6.9.5 Born’s rule and the meaning of ψ\psi

Max Born supplied the interpretive key in 1926: ψ2|\psi|^{2} is a probability density. Expectation values are integrals with that weight. For a normalized state,

ψ(r,t)2d3r=1\int |\psi(\mathbf r,t)|^{2}\,d^{3}r = 1

and for an operator A^\hat A acting on wavefunctions,

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

Probability current and continuity follow from Schrödinger’s equation, making the probabilistic reading self-consistent.


6.9.6 Operators and the canonical dictionary

Wave mechanics uses differential operators that mirror the matrix commutators:

x^=x,p^=i\hat x = x,\qquad \hat p = -\,i\hbar\,\nabla

These satisfy

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

so the uncertainty relation and algebra of observables match matrix mechanics. Potentials become multiplication operators V(x^)V(\hat x); kinetic energy becomes 22/(2m)-\hbar^{2}\nabla^{2}/(2m). The correspondence principle is codified as “replace Poisson brackets by (1/i)(1/i\hbar) times commutators” in the classical–quantum dictionary.


6.9.7 Equivalence: showing the two apps are the same OS

Schrödinger himself showed that wave mechanics and matrix mechanics are unitarily equivalent. In the basis of energy eigenfunctions {ϕn}\{\phi_n\},

Amn=ϕm(r)(A^ϕn)(r)d3rA_{mn} = \int \phi_m^{\ast}(\mathbf r)\,(\hat A \phi_n)(\mathbf r)\,d^{3}r

That integral is exactly the matrix element used by Heisenberg–Born–Jordan. Conversely, given a matrix AmnA_{mn} one builds A^\hat A by its action on expansion coefficients. Same spectra, same transition amplitudes, same selection rules—different clothes. Dirac later packaged this as bra–ket notation and spectral theory, but the core equivalence is already here.


6.9.8 Worked snapshots: harmonic oscillator and hydrogen

Harmonic oscillator. In matrix mechanics, define ladder operators

a=12mω(mωX+iP),a=12mω(mωXiP)a = \frac{1}{\sqrt{2\hbar m\omega}}\left(m\omega X + i P\right),\qquad a^{\dagger} = \frac{1}{\sqrt{2\hbar m\omega}}\left(m\omega X - i P\right)

which 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)

giving eigenvalues

En=ω(n+12)E_n = \hbar\omega\left(n + \frac{1}{2}\right)

Wave mechanics solves the differential equation and gets the same ladder and energies; Hermite polynomials appear, but the spectrum is identical.

Hydrogen. In wave mechanics, separability in spherical coordinates produces radial equations with bound-state quantization

En=mee42(4πε0)221n2E_n = -\,\frac{m_e e^{4}}{2 (4\pi\varepsilon_0)^{2} \hbar^{2}}\,\frac{1}{n^{2}}

Degeneracies match the old nn-only dependence. In matrix mechanics, one uses hidden Runge–Lenz symmetry to reproduce the same 1/n21/n^{2} spectrum algebraically.


6.9.9 Ehrenfest’s theorem and the classical limit

Quantum expectation values obey equations that mirror classical motion:

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

This is Ehrenfest’s theorem. When the wave packet is narrow and the potential is smooth, V/xV(x)/x\langle \partial V/\partial x\rangle\approx \partial V(\langle x\rangle)/\partial x, and the centroid follows Newton. That is the correspondence principle rephrased as dynamics.


6.9.10 Stationary states, transitions, and selection rules

In wave language, stationary states ϕn\phi_n are energy eigenfunctions. Time dependence is a global phase eiEnt/e^{-iE_n t/\hbar}, so expectation values in a stationary state are time-independent. Driving the system with a weak periodic field H=d^E0cosωtH'=-\hat d\cdot \boldsymbol E_0\cos\omega t and using time-dependent perturbation theory yields transition rates proportional to matrix elements md^n\langle m|\hat d|n\rangle and a resonance at ω=EmEn\hbar\omega=E_m-E_n. The dipole operator enforces selection rules like Δ=±1\Delta \ell=\pm 1, matching old empirical rules while giving the intensities from first principles.


6.9.11 Commutators, Poisson brackets, and quantization rules

The formal bridge from classical mechanics uses

{A,B}PB1i[A^,B^]\{A,B\}_{\text{PB}} \longrightarrow \frac{1}{i\hbar}\,[\hat A,\hat B]

and promotes canonical variables (x,p)(x,p) to operators obeying [x,p]=i[x,p]=i\hbar. Ordering ambiguities appear when turning nonlinear classical expressions into operators; early prescriptions by Born–Jordan and Weyl tried to systematize this. In practice, symmetries and experimental checks decide the correct operator form. For quadratic Hamiltonians and central potentials, the naive promotion works cleanly.


6.9.12 Variational and spectral viewpoints

Schrödinger also wrote a variational principle: stationary states extremize the functional

E[ϕ]=ϕ[22m2+V]ϕd3rϕ2d3r\mathcal E[\phi] = \frac{\int \phi^{\ast}\left[-\frac{\hbar^{2}}{2m}\nabla^{2}+V\right]\phi\,d^{3}r}{\int |\phi|^{2}\,d^{3}r}

Minimizing E\mathcal E under normalization yields the lowest eigenvalue; constraining orthogonality gives excited states. This is the ancestor of modern Rayleigh–Ritz methods central to atomic, molecular, and solid-state calculations.


6.9.13 Continuum, scattering, and phase shifts

Not all spectra are discrete. For energies above binding, solutions are scattering states normalized to delta functions in energy or momentum. In 3D with short-range V(r)V(r), asymptotic wavefunctions look like plane waves plus outgoing spherical waves, and the physics hides in partial-wave phase shifts δ(E)\delta_\ell(E). Cross sections follow from them. Matrix mechanics encodes the same data in the SS-matrix, whose unitarity expresses probability conservation.


6.9.14 What changed conceptually

  • Observables as operators. No more hidden orbits; measurable quantities are spectra and expectation values of operators
  • States as vectors. A system is a vector in a complex inner-product space, evolved by a linear equation
  • Incompatibility via algebra. Noncommutation replaces classical simultaneity; uncertainty is structural, not instrumental
  • Probabilities from amplitudes. Interference is about complex phases of amplitudes, not only waves in space

The old rules become corollaries of a single linear framework.


6.9.15 Minimal problem kit

  • Show that [x,p]=i[x,p]=i\hbar implies ΔxΔp/2\Delta x\,\Delta p\ge \hbar/2 using Cauchy–Schwarz
  • Solve the 1D infinite square well with Schrödinger’s equation and extract energies and orthonormal eigenfunctions
  • Build the oscillator ladder operators and prove [a,a]=1[a,a^{\dagger}]=1 and En=ω(n+12)E_n=\hbar\omega(n+\tfrac12)
  • For hydrogen, separate variables in spherical coordinates and identify the three quantum numbers (n,,m)(n,\ell,m) and their degeneracies
  • Derive Fermi’s golden rule for a weak periodic perturbation and recover the dipole selection rule Δ=±1\Delta \ell=\pm 1

In summary: Heisenberg replaced paths by algebra of transitions; Schrödinger replaced orbits by waves whose squared magnitude is probability. Born provided the measurement rule; commutators supplied the logic; the two formalisms proved equivalent and complete. With this, quantum physics leveled up from clever hacks to a unified theory that predicts spectra and dynamics, computes intensities, and explains why “classical” is just a special zoomed-out limit. Chapter 7 will take this machinery for a spin, solving canonical systems and building the perturbation toolbox that powers modern physics and chemistry