You are on page 1of 51

Late 19th-Early 20th Century: Crisis in Physics

(“Everything You Know is Wrong” *)


• f • inability to explain
Rutherford atom
• inability to explain blackbody
radiation (light emission
from heated object)
• inability to explain light
emission from energetically
excited H atoms
• failure to explain other
experiments

* For a perverse take on this same line, see a book with this title from a past “great” Tulane
graduate: http://www.lloydpye.com/
Resolving the Problems Caused When Applying
Classical Physics to Atoms and Molecules: Solution

Step 1: Discard classical physics

Step 2: Develop a new theory. It will have to depend


on new fundamental principles, that through ab
initio application (not assumption) bring about the
necessary quantized phenomena. This will be a
wave theory of matter.

Older textbooks tended to call this


approach wave mechanics. Books of
more recent vintage call it quantum
mechanics. The old terminology has
some advantages: That’s basically what
the approach is. Since standing wave
solutions (whether quantum or classical)
involve quantization, the newer
terminology eventually predominated.
Key Observations That Doomed
Classical Physical Theory

1. Rutherford scattering (1911)


2. Blackbody radiation (Planck solution,1900)
3. Photoelectric effect (Einstein, 1905)
4. Franck-Hertz (1914)
5. Compton scattering (1923)
6. H atom emission (Bohr, 1913)
7. Wave character of matter (de Broglie, 1926)
Plum Pudding Atom Model

“Plum pudding” atom model


(J.J. Thomson—discoverer
of electron, 1897)

First meaningful model:


• Consistent with charge neutrality
• Consistent with a stably bound atom
• Testable in future experiments
1. Rutherford Scattering (1911)*
Schematic of the α
scattering experiment
of Rutherford (student
of J.J. Thomson)

Scattering of an alpha particle by


the central repulsive Coulomb force
leads to a hyperbolic trajectory.
From the scattering angle and
momentum, one can calculate the
impact parameter and closest
approach to the target nucleus.
Extreme angle backscattering =>
positive charge was instead highly
compact, with a spatial extent ~10-5 of
the spatial extent of the negative
charge (electrons).
* http://hyperphysics.phy-astr.gsu.edu/hbase/rutsca.html
2. Blackbody Radiation
When objects are heated, they emit a characteristic distribution of
electromagnetic energy as a function of frequency. At lower
temperatures the emission is mainly in the infrared. As T is raised,
the distribution peaks at higher frequencies (shorter, bluer,
wavelengths): first an object becomes red hot, then white hot.

Now-common device based on


blackbody radiation curve:
remote T sensors. They measure
the shape of the blackbody curve
in the infrared and fit it to a best-
fit blackbody temperature. (You
can buy a cheap one on sale at
Harbor Freight Tools for about
$10.)
Electromagnetic Radiation

region of spectrum wavelengths typical sources


radio more than 30 cm radio, television
microwave 3 mm to 30 cm radar, microwave oven
infrared 750 nm to 3 mm hot objects; atomic/molecular emission

visible (“light”) 400 nm to 750 nm very hot objects; atomic/molecular emission

ultraviolet 20-400 nm sun; black lights; atomic/molecular emission


x-rays 3 pm to 20 nm cathode ray tubes
gamma rays less than 3 pm radioactive decay
Implementation to use in Experiment and Theory

• For both theory and measurement, consider a box (cubic


convenient for theory) that can be heated to high T. A
small measurement hole is in the box, for experimentally
measuring energy as a function of frequency inside the
box.
Important Digression: Traveling Wave Properties
• Waves have amplitudes. When
two waves meet, their amplitudes
combine. At the combination
points the result can be either a
larger amplitude (crest adds to
crest) or a smaller amplitude (crest
adds to trough).
• Common examples are water
waves, waves on strings, sound
waves.

The repeat length is the wavelength, λ. The time to execute one oscillation is the
period, T. The frequency, ν (Greek nu), is given by the number of full oscillations
per time; ν = 1/T. Suppose the wave is traveling at speed v. In one cycle it had
to have advanced a distance λ in a time T. So, for any wave that travels at some
speed v, it has to be the case that v = λ/T = νλ. ! Beware the often subtle
difference between Roman “v”
! and Greek nu.
Blackbody Radiation—classical derivation 1
• Compute number of standing wave electromagnetic radiation modes in
a 3d cubic box. Start in 1d.
1 2

{
  c, so c    , or c   c  k  2  
wave  
properties 2
That is, ck  . k  is the 1D wavevector;   2 is angular frequency.

{
Standing wave 1D solutions in 1D box, x = 0 to x = L:

mode Amplitude must disappear at ends, so solutions of form u ( x)  c  sin( kx) : 1d mode
counting x = 0: u (0)  const  sin(0)  0, as needed spacing is
n 2π/L
x  L: u ( L)  const  sin( kL)  0  kL   n , or k   , n  0,1, 2,3,
L

2d

This box in 2d k space shows the 2d


“volume” taken by one allowed mode,
(2π/L)2.
Blackbody Radiation—classical derivation 2

• => volume of mode in 3D k space: (2π/L)3 = 8π3/V

• volume in 3d k space with magnitude ≤ some value k: (4/3)πk3


=> total number of standing wave EM states therein is
N = (k space volume)/(states per volume) =
[(4/3)πk3 ]/[(2π)3/V]

• correction for electromagnetic radiation: Two polarizations of EM


waves means N value should be doubled.
Blackbody Radiation—classical derivation 3
• Number of modes N in terms of frequency
V 4 3
N  2 k
(2 )3 3
3
2 2 4  2  8 3
Now use k  : N V   V  3
c (2 )3 3  c  3 c

• Differentiate above expression

2
dN  V 8 3 d # modes between ν and ν + dν
c

• Energy in the blackbody box at thermal equilibrium at temperature T

from statistical mechanics: average energy in each EM mode is kT


total energy per volume V in frequency range  to  + d   ( , T )
8 v 2 kT
 ( , T )  3
(Rayleigh-Jeans law) Called ρυ(T) in
c
McQuarrie
Ultraviolet Catastrophe!
Planck’s Fix Using a Quantization Kludge*
1. Key kludge*: Possible energies in an EM mode with frequency ν
described by ε ∝ nν (n = 0,1,2,..); i.e. ε = h∙nν = nhν, where h is some
proportionality constant.

2. Boltzmann statistics: Probability that the mode in thermal


equilibrium will have a particular n (quantum) energy value nhν ∝
exp(-nhν/kT).

3. Find mean energy  averaged over all n: Details on next slide.

Now use  ( , T )  8 v 2
4.
3

c

* “A ‘solution’ to a problem, doing a task, or fixing a system (whether hardware or software)


that is inefficient, inelegant, or even unfathomable, but which nevertheless (more or less)
works” [from Wikipedia]
Step 3 (above): Statistical Details
So, then, pn  K  exp(  nh / kT )  K  exp(  n h ), where   1/ kT
K is a constant that normalizes the total probability to 1:

 pn  K  n 0 exp( n h )  1
n  n 
n 0
1
Proceed using the useful function Z   n 0 exp( n h ); so KZ  1  K 
n 

Z
Z  1  exp(  h )  exp(2 h )  ...  1  x  x 2  x 3  ... where x  exp(  h )
1 1
note: 1  x  x 2  x 3  ...  for x  1  Z 
1 x 1  exp(   h )

dZ 1 dZ 1 d  1 
   n 0  n  pn   n 0 (nh )  pn   K
n  n 
   
d Z d Z d   1  exp(  h ) 
2
1 1  h
    exp(   h ) h 
Z  1  exp(   h )  exp(  h )  1
Step 4: Ψ(ν,T)

8 v 2 8 v 2
• Rayleigh-Jeans:  ( , T )  3    3  kT
c c

• Planck: 8 v 2 8 v 2 h
 ( , T )  3    3 
c c exp(  h )  1

Can be precisely fitted to experimental


curve for h = 6.63·10-34 J·s
Footnote: Einstein’s Treatment of Heat Capacity in
Crystals*: Problem Without It
Law of Dulong and Petit: cV = 3R for a solid
Handwave** rationalization for above:
*This was not one of the compelling
Consider a one mole chunk of the solid, NA
arguments for a quantum theory,
atoms. Each has 3 quadratic degrees of though we can see its reasonableness.
translational KE and 3 of PE. At thermal Einstein’s result, assuming just one
equilibrium at temperature T, each degree vibrational frequency, couldn’t seriously
has on average an energy of kT/2. So, the model real crystals.
total energy in one mole of solid is
E(molar) = NA∙(6∙kT/2) = 3NAkT = 3RT.
cV = ∂E(molar)/∂T = 3R = 24.94 J/mol∙K **Describes either the technique of failing to
address an argument rigorously, in an
attempt to bypass the argument altogether,
or an admission that one is intentionally
D-P okay at high glossing over detail for the sake of time or
T; bad to awful at clarity. It can be meant as an accusation or
lower T. in a more positive light, depending on the
context.
Rationalization of Low T Results with a
Simple Quantization Model Due to Einstein*
• Assume that all atoms in the solid vibrate with the same frequency ν;
assume quantum vibrational levels with spacing hν as in the Planck
development.

• Consider just the vibrational contribution to energy at thermal equilibrium at


temperature T, using the result we got for one such mode in the earlier
Planck development:   h
1  exp(h / kT )

• 3NA such modes in a mole; look at the form of their heat capacity
contribution:
   h  h 
cV (vibr contrib)        exp(h / kT )
T T 1  exp( h / kT )   1  exp(  h / kT )  T
2

h 2 2 k exp( h / kT )
   0 as T  0
 1  exp(h / kT ) 
2
T2

*McQuarrie has further discussion on pp 200-201


3. Photoelectric effect
• If the light frequency ν is below ν0, no electrons are ejected from the
surface, no matter how intense the light is made.

• Light directed onto a metal surface causes electrons to be ejected if the


frequency of the light is above a certain threshold value frequency
value, ν0.

• If the light frequency ν is above ν0, the kinetic energy of the ejected
electrons is proportional to ν - ν0.
Einstein’s Explanation (1905)
• The quantity hν from the Planck hypothesis is not
just some mathematical artifact; light is in fact
quantized, and hν is the energy of one quantum
of light.
• An electron absorbs the energy hν of the photon:
1. If the energy hν is below the threshold energy
for ejecting an electron (work function) Φ, no
electron is ejected.
2. If the energy hν exceeds Φ, an electron is
ejected with excess KE = hν – Φ.
4. Franck-Hertz (1914)

This original Franck-Hertz data shows electrons losing 4.9 eV


per collision with mercury atoms. It is possible to observe ten
sequential bumps at intervals of 4.9 volts.
5. Compton Effect (1923)
• x-ray (high energy photon) scattering via
interaction with electrons
• We can model the interaction as the
collision with a light particle (photon) with
an electron, while conserving energy and
momentum
6. H atom emission
problems
• Why should emission from
energetically excited H
atoms only occur at certain
discrete wavelengths?

• Similar discrete wavelength


emission was observed for
other atoms as well.
Puzzling observation (before quantum
theory) is that the emission is at discrete
wavelengths.
Numerical fitting of the observed H
emission lines

1  1 1 
Rydberg formula:  109680   2  2  (units of cm 1 )
  n1 n2 

series n1 n2
Lyman 1 2,3,4,
Balmer 2 3,4,5,
Paschen 3 4,5,6
Bohr H Atom: a new quantum mechanical kludge

1. Use planetary orbital model: An electron orbits a


proton, with a coulombic central force.

would not work: electron would collapse into nucleus

2. Bohr hypothesis (kludge): By assumption, constrain


the angular momentum to quantized values:
nh/2π, n = 1,2,3,
• angular momentum always nonzero (n > 0) means no
electron collapse to nucleus
• quantized angular momenta => quantized energies
Bohr H atom derivation*
• Circular orbits: Coulomb force = centrifugal force
e2/r2 = mv2/r, leading to mv2 = e2/r
For later use, v2 = e2/rm
• Energy of orbiting system = KE + PE
E = ½ mv2 - e2/r = ½ e2/r - e2/r = - ½ e2/r
• Bohr postulate: mvr = n h/2π
Use v2 = n2h2/(2πmr)2 = e2/rm; solve for r
r = n2h2/me2(2π)2 ≡ n2a0 (a0 = 0.53 Á – Bohr radius)
So, r is quantized!
E = En = - ½ e2/r = -2π2me4/n2h2 = -RH/n2 (RH = 2.179 x 10-18 J)
So, E is quantized!

* This is really a two body problem, involving a proton and an electron.


Since mp >> me = m, we can treat it as a one body (electron orbit) problem
with negligible error. Note: Gaussian (non-SI) electrical units are used here.
Bohr theory problems/critique
• The approach “works” (i.e. replicates Rydberg formula).

• There was the assumption of angular momentum/energy


quantization. A comprehensive theory would need to show,
rather than assume, quantization.

• It’s interesting that the size of the quantization intervals again


involves h. Evidently h is a deeply fundamental quantity that
specifies the “graininess” of quantization present in nature.
Brief Digression: How come we don’t directly see
and experience angular momentum quantization?

Evidently angular momentum must come in quantized units of


h/2π = ħ = 1.05∙10-34 J∙s. Compare this to the angular
momentum arising from swinging a rock of mass 0.1 kg on a
string of r = 1 m with a circumferential speed of 10 m/s:

J(rock system) = mvr = 1 kg∙m2/s = 1 J∙s

The ratio of J(rock system): h/2π ≈ 1034! (i.e. ≈ 1034 quanta)

So the loss or gain of a quantum of angular momentum would make a


change of around one part in 1034—kind of easy not to notice!
Application to Excited H Atom Emission
• An excited H atom is in one
of the higher (n>1) quantum
levels. Following emission,
the atom is in a lower n
energy level. The difference
in energy is the energy of an
emitted photon.

E(photon) = hν
= En(initial) - En(final)

This explanation accounts for


the discrete wavelength or
frequency character of the
emission and also quantitatively
accounts for the emission
frequencies observed.
H atom practice problem
Calculate the wavelength of the following
transition in the hydrogen atom
n=5  n=1. Useful information:
 RH
En  2
n
RH = 2.179 x 10-18 J
n – the quantum number

E = hu
h = 6.63 x 10-34 J-s
Solution
The photon energy is given by E = hu = hc/ λ
(To convert frequency to wavelength, use λu = c)
Then λ = hc/ E

E = (E5 - E1) = - RH(1/52) – (- RH(1/12) )


= 0.96 RH

λ = hc/(0.96 RH)
= (6.63 x 10-34 J-s)(3.00 x 108 m/s)/(0.96)(2.179 x 10-18 J)
= 9.51 x 10-8 m = 95.1 x 10-9 m = 95.1 nm
Correct numerical calculations have required quantized energy levels in H
atoms and have also required that the electromagnetic radiation is in
quantized units, photons.
7. Wave-particle duality
Prelude to wave-particle duality: Early
Observations of Light as a Wave
Is light a stream of particles or a wave?
• Thomas Young, 1801: light passed through two tiny
adjacent slits
– If light consisted of classical particles:
• target would be brightest where light passing
through the slits overlapped
• target would darken steadily moving away from
the overlap region
This was not observed.
– A pattern of light and dark stripes was observed
instead.
• Young explained the stripes as a combination of
diffraction and interference.
• These interference fringes are a sure sign of
wave behavior.
•                         
– White areas are peak-peak or trough-trough overlaps
(constructive interference)
– black areas are peak-trough overlaps (destructive
interference)
– width of interference bands suggested the
wavelength of visible light was less than a millionth
of a meter!
Dual Wave-Particle Behavior in Light
• This experiment involves passing photons one at a time through 3 slits
side by side.
• You can clearly make out 5 regions where photons were detected--
indicating that individual photons must be diffracted, even when they don’t
have any other waves to interfere with.

White dots
Detector

arise from
detection of
individual
photons.

Each individual photon must


also have wave like properties.
Wave-Particle Duality:
De Broglie Hypothesis (1926)
• classical EM radiation (waves) ↔ photons (“particles”)

E 2  p 2 c 2  m0 2c 4 (from special relativity)  E  pc for a photon (m0  0)


h h recipe for obtaining
E  h and h  pc  p  
c  p in terms of λ

• classical particles ↔ matter waves

We have seen that for a classically “λ-like” phenomenon (EM radiation) that
the wave/matter dual p is given in terms of λ by p = h/λ.

For a classically “p-like” phenomenon (particles) assume now that the


wave/matter dual λ is given in terms of p by turning around the same relation:
de Broglie
λ = h/p
relation
Davisson-Germer Experiment (1927)
• Electrons from a heated
filament were accelerated
by a voltage and allowed to
strike the surface of nickel
metal as a function of angle
relative to the lattice.

• Results were consistent


with electron waves
satisfying the Bragg
condition, with λ given by
the de Broglie relation
De Broglie Relation ↔ Wave Mechanics

• For matter on a sufficiently small size scale, problem must be


approached on a matter wave basis (Schrodinger equation)

• The de Broglie relation can tell us when a wave mechanical


approach is necessary and when a classical physical approach can
be used.

For matter characterized by a de Broglie value


Application of these
λ, where the interaction are over a distance d:
simple criteria tell us
whether or not
λ ≈ d, λ > d: wave mechanical problem
quantum effects are
important in a given
d >> λ: classical limit; no need for wave
situation.
mechanics
End of Historical Introduction

Introduction to a “handwave” wave theory of matter


Toward a Matter Wave Theory

Better Rationalization of Bohr H atom


quantization hypothesis

In what follows, we’ll use standing wave concepts and


then specialize the arguments to matter waves, using
the de Broglie relation.
Electrons as “Particle Waves” with
time-stable (standing wave) conditions
• Wavelike electrons
act as standing
waves.
• Standing wave
boundary conditions
limit the permitted
wavelengths and
results in quantization
of the allowed energy
levels.
Bohr Hypothesis in Wave Terms
Consider the electrons to be orbiting
waves. The only stable solutions will
be ones where the circular wave
closes on itself (standing wave). For
this to happen, there must be an
integral number n of wavelengths in
the circular circumference.

• For this to happen, 2πr = nλ.


• Using λ = h/mv (de Broglie), 2πr = nh/mv. Rearrangement
gives
mvr = n(h/2π), the original form of the Bohr hypothesis.
• Evidently, then, the Bohr hypothesis works because electrons
in atoms are matter waves with standing wave solutions.
Elements of a Wave Mechanics: Kinds of
details that will have to be developed
• An acceptable theory for electrons in atoms has to be a wave
theory that somehow incorporates the de Broglie relation.

• For stable atoms, standing wave solutions will result; there will be
sets of possible quantized energies for the electronic system.

• Three dimensional wave amplitude solutions of the form Ψ(x,y,z)


will result for an electron. The wave amplitudes themselves have
no direct physical interpretation. The absolute value squared of
Ψ(x,y,z) gives the probability of “finding” (localizing) the electron
at that (x,y,z) point.
Waves Versus Particles

Waves can …and fan out Particles effuse


bend from pinholes. from pinholes.
around small
obstacles…

A wave can't bend around obstacles much larger than its


wavelength wave behavior particle behavior
waves interfere particles collide
waves diffract particles effuse
waves are particles are
delocalized localized
“Everything you know is wrong”: Individual electrons
(“particles”, you say?) are wave-like!

                                    
                   100 electrons • Electrons passing one
at a time through a
double slit. Each spot
shows an electron
impact on the
                                     detector.
                   3000 electrons
• Interference fringes
are clearly evident, so
electrons must
behave as a waves.
70000
                                    
electrons: These experimental results
                   
interference are analogous to the ones
fringes
seen earlier for the wave
interference of individual
photons.
Then how come baseballs aren’t wavy when
they’re thrown?
We need to consider numerical examples that show that objects like
electrons can act like waves (e.g. crystal diffraction) but that no
“waviness” should be observed for familiar objects—consistent with
1.experience.
0.10 kg ball thrown at 30 m/s:
h h 6.63 10 34 J  s
    2.2 10 34 m -- about 10 21 of a nuclear dimension!
p mv  0.1 kg   30 m / s 

2. an electron, m e  9.11 1031 kg , moving at v = 107 m/s:


h h 6.63 1034 J  s 1A
    .73  10 10
m   0.73 A
p mv  9.11 1031 kg   107 m / s  1010 m
This  may seem at first to be small, but it is on the order of a typical atomic dimension.
So, the electron is entirely wavelike if confined to regions of atomic dimension!

For any matter on our scale of dimensions, any “waviness” is utterly


negligible. That’s because, on our scale, the degree of quantum
“graininess”, h, is utterly negligible.
Heisenberg Uncertainty
• 1d: Δx∙Δpx ≥ ħ/2 – a “Heisenberg uncertainty relation”

• It can be developed in analytical form; it is not a


philosophical* dictum concerning “uncertainty”.

• It is deeply linked to the “waviness” of matter on


sufficiently small size scales.

* or, as may frequently happen when it is invoked in this way, “philosophical”


Fourier analysis for waves with a finite
repeat length
• Any wave can be described in terms
of contributions from a “basis set” of
sine waves of increasingly high
frequency.

• A sine wave of frequency ν would be


described by a single sin(2πνt) term
of proper amplitude.

• The decomposition of a square


wave of frequency ν would have a
sin(2πνt) term of relatively large
amplitude, but there would be many
other higher frequency terms with
smaller relative amplitudes.
Fourier Analysis allowing for Infinite Repeat Distance
(“distance” domain) (“wavevector” domain)

Note the trend from bottom up: As the wave to be described (LHS) becomes
confined to a smaller interval in time, the number of frequency components
needed to describe it (RHS) gets larger. A similar set of plots would hold for
waveforms as a function of position in space (LHS). In this case we would
need to specify the range of contributing sine wavelength components on the
RHS, but the two set of plots would be unchanged.
Classical Fourier Analysis: Description of Number of
Pure Wave Components Needed for Decomposition

1. Time domain-frequency domain


t    1, where t is the time duration and 
the range of frequency components needed to describe
 h  t    t  h  t  E  h
2. Position domain-wavelength domain
 2 
Position dependent terms will be of form sin   x   sin(k  x)
  
for different values of  (i.e. k ).
The analogous version of the previous relation is
x  k  1    x  k  x  k  x  p
Movement of a Spatial Wave Packet in Time
• The spatial width is Δx. At propagation speed v, the time
needed for this width to pass a given point is Δt ≈ Δx/v.

• The particle wave has an uncertainty in energy that is


given formally by E   E   p  v  p  p  E / v
 p  p  p0

• Use the two expressions above to rewrite Δx∙Δp:

x  p   v  t    E / v   t  E  h
Conversion of the Classical Fourier
Relations to Quantum Relations

1. photon time-energy: h∙(Δt∙Δν ≈ 1) → Δt∙Δ(hν) ≈ h → Δt∙ΔE


≈ h, where E is the energy of a photon. Evidently a photon
can have an uncertainty in its energy E by an amount given
by this relation over a short time Δt.

2. position-momentum: ħ∙(Δx∙Δk ≈ 1) → Δx∙Δ(ħk) ≈ ħ →


Δx∙Δp≈ ħ

You might also like