Sie sind auf Seite 1von 227

MIT Open Courseware Compillation

Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #1 page 1

Quantum Mechanics – Historical Background


Physics in the Late 19 th Century (prior to quantum mechanics (QM))

• Atoms are basic constituents of matter

• Newton’s Laws apply universally

• The world is deterministic

According to classical mechanics (CM):


  
Given initial positions r0 and velocities v 0 , and given all forces F t () 
all the future can be predicted!

 


 F   dv 
() 
v t
v t = 
v0
dv  =  t0 m
dt  F = ma = m dt 
 
    dr  
() 
r t
r t = 
r0
dr = 
t0
vdt  dt = v 

Physics was complete except for a few decimal places !

• Newtonian mechanics explained macroscopic behavior of


matter -- planetary motion, fluid flow, elasticity, etc.

• Thermodynamics had its first two laws and most of their


consequences

• Basic statistical mechanics had been applied to chemical


systems

• Light was explained as an electromagnetic wave

Prepared By Kevin Boudreaux


1 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #1 page 2

 However there were several experiments that could not be explained


by classical physics and the accepted dogma !

• Blackbody radiation

• Photoelectric effect

• Discrete atomic spectra

• The electron as a subatomic particle

 Inescapable conclusions would result from these problems

• Atoms are not the most microscopic objects

• Newton’s laws do not apply to the microscopic world of the


electron

OUTCOME  New Rules!!!

Quantum Mechanics!

• Describes rules that apply to electrons in atoms and molecules

• Non-deterministic, probabilistic ! A new philosophy of nature

 Explains unsolved problems of late 19th century physics

 Explains bonding, structure, and reactivity in chemistry

Prepared By Kevin Boudreaux


2 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #2 page 1

The DEMISE of CLASSICAL PHYSICS

(a) Discovery of the Electron


In 1897 J.J. Thomson discovers the electron and measures e me ( )
(and inadvertently invents the cathode ray (TV) tube)

Faraday (1860’s – 1870’s) had already shown using electrochemistry that


amounts of electric current proportional to amounts of some substances could
be liberated in an electrolytic cell. The term “electron” was suggested as a
natural “unit” of electricity.

But Thomson experimentally observes electrons as particles with charge & mass.

y = displacement
induced by
deflector voltage 

Deflector plates Phosphor


Cathode
screen
Anode

Thomson found that results are independent of (1) cathode material


(2) residual gas composition

 “electron” is a distinct particle, present in all materials!

Classical mechanics  force on electron due to deflector voltage:

Fy =  e (force starts at time t = 0 when electron enters region between plates)


dv y dv y  e 
=m
dt
( F = ma ) 
dt
= 
 me 

 e 
Integrating  v y =  t
 me 
(
[Note v y t = 0 = 0] )

Prepared By Kevin Boudreaux


3 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #2 page 2

 e  t f
2

Integrating again since v y =


dy
dt
and y t = 0 = 0  y =  (
 me  2
)
t f = total time electron is between the plates (easily calculated)
 e 
Set voltage  , calculate time t f , measure displacement y   m = 1x10 C/kg
11

 e
 e 
Modern day value is   = 1.758x1011 C/kg
 me 

(b) 1909 Milliken oil drop experiment determines e, me separately

mist of micron-size
oil droplets in air

Gravitational force downward: Fg


Fg =  Mg M = mass of droplet, g = gravitational constant

Frictional force upward due to air: Ff


F f = 6 rv r = radius of droplet,  = air viscosity, v = droplet velocity

Since F f  v , terminal velocity vt is reached when forces balance


6 rvt = Mg  get droplet mass M = 6 rvt g

Now use x-rays or -rays to add some charge ne to the droplets ( )


Voltage  across plates exerts Coulomb force Fc =  ne on the charged droplet

x-rays n-
Fc Fg Ff

Prepared By Kevin Boudreaux


4 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #2 page 3

Adjust voltage until drop stops falling: v = 0  F f = 0, Fc = Fg


 ne = Mg Determine ne = Mg 

Mulliken did this for lots of droplets i = 1,2,3,...


( )
They all had different charges ni e but all integer multiples of charge e ()
19
Determined elementary charge as e = 1.59x10 C

19
(very close to today’s value e = 1.602x10 C)

Combining values for e me ( ) and ( e)  me = 9.11x10


31
kg

27
Hydrogen mass was known: mH = 1.66x10 kg  electron is subatomic!!

(c) Where are the electrons? What’s the structure of the atom?

Angstrom (10 -10 m) atomic size scale already inferred from gas kinetics
First “jellium” model didn’t last long

n+
(jelly)

e

Rutherford backscattering experiment Au foil


2+ 2+
He He

(no electrons) 98%


undeflected
2%
scatter back

 (1) He2+ nucleus very small, << 10 -10 m (Rutherford estimated 10 -14 m)
(2) Au atoms are mostly empty!

Prepared By Kevin Boudreaux


5 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #2 page 4

Rutherford planetary model: classical mechanical model of atomic structure


Coulomb attraction plays the role of gravity
me v 2
centripetal force Fc =
r
v Ze2
r Coulomb force FC =
e 4 0 r 2
+Z
me v 2 Ze2 Ze2
for stable orbit =  r=
r 4 0 r 2 4 0 me v 2

This is stable compared to separated electron & nucleus

1  Ze2  1 Ze2
E = K.E. + P.E. = me v +  
2
 = <0
2 4 0 r
2 4 0 r

BUT model not consistent with classical electrodynamics:


Accelerating charge emits radiation! (centripetal acceleration = v2/r)
And since light has energy, E must be getting more negative with time

 r must be getting smaller with time!


 Electron spirals into nucleus in ~ 10-10 s !

Also, as r decreases, v should increase


Frequency  of emitted light = frequency of rotation

v (m/s)
 (Hz = cycles/s) =
2 r (m/cycle)

circumference of orbit

 atom should emit light at all frequencies – that is it should produce a


continuous spectrum

Prepared By Kevin Boudreaux


6 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #2 page 5

BUT emission from atoms was known to be discrete, not continuous!

For H:
n1 = 3 n1 = 2 n1 = 1
n2 = 3 4 5 6… n2 = 2 3 4…

10,000 30,000 50,000 100,000 (cm-1)

For the H atom, Rydberg showed that the spectrum was consistent with the
simple formula:

 1 1
 (cm -1 ) = R  2  2
 n1 n2 
with n1 = 1, 2, 3, ... and n2 = n1 + 1, n1 + 2, n1 + 3, ...
R = 1.097x105 cm -1 (Rydberg constant)

n1 = 1 Lyman series
n1 = 2 Balmer series
n1 = 3 Paschen series

visible & UV lines well known

Summary: Rutherford’s model of the atom

(1) Is not stable relative to collapse of electron into nucleus


(2) Does not yield discrete emission lines,
(3) Does not explain the Rydberg formula

Prepared By Kevin Boudreaux


7 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #3 page 1

The DEMISE of CLASSICAL PHYSICS (cont’d)

(a) Blackbody radiation – When heated all objects emit light!

Classically: (1) Radiation from a blackbody is the result of electrons


oscillating with frequency 
Oscillating charged particle  antennae

(2) The electrons can oscillate (& radiate) equally well at any
frequency

 Rayleigh-Jeans Law for spectral density (), where intensity of emitted


light in frequency range from  to  + d is I() ~ ()d

8 kT 2
( )
d  =   ,T d =
c 3
 d   2

where d(,T) = density of radiative energy in frequency range


from  to  + d at temperature T

k = Boltzmann’s constant [= R/NA (gas constant per molecule)]


c = speed of light

Rayleigh- ()
    2  divergence at high freq.
Jeans Law  lots of emission in UV, x-ray regions!
“UV catastrophe” – not observed in exp’t
( )

Exp’t T3
T2 T1 < T2 < T3
T1

Prepared By Kevin Boudreaux


8 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #3 page 2

Planck (~1900)  First “quantum” ideas

(1) The energy of the oscillator  frequency

E
(2) The energy  an integral multiple of 
(the # of oscillators n)
E  n

or E = nh

constant

h becomes a “quantum” of energy

Planck used statistical mechanics (5.62) to derive the expression for black body
radiation
8 h 3  
( )
d  =   ,T d =
1
c3  eh kT  1

d

 3

( )
 1 
 eh kT  1 

Fitting exp’t to model  h = 6.626 x 10-34 J-sec

Planck’s constant

Prepared By Kevin Boudreaux


9 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #3 page 3

(b) Photoelectric effect

UV light e-
Exp’t:
K.E. = _mv2

metal

Classically K.E. K.E.


expect:

I 

Exp’t: K.E. K.E. linear


Opposite!

I 0 

threshold

Einstein (1905) proposed:

(1) Light is made up of energy “packets: “photons”

(2) The energy of a photon is proportional to the light frequency

E = h h Planck’s constant

Prepared By Kevin Boudreaux


10 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #3 page 4

New model of photoelectric effect:

E K.E. energy of
of e- photon = h
0
 = work function
of metal

 K.E. = h –  = h – h0 = h( – 0)

K.E.
 K.E. = h( – 0)

0 = / h 

Comparing to exp’t, value of “h” matches the one found by Planck!

This was an extraordinary result !

Summary:

(1) Structure of atom can’t be explained classically


(2) Discrete atomic spectra and Rydberg’s formula can’t be explained
(3) Blackbody radiation can be “explained” by quantifying energy of
oscillators E = h
(4) Photoelectric effect can be “explained” by quantifying energy of light
E = h

Prepared By Kevin Boudreaux


11 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #4 page 1

The ATOM of NIELS BOHR


Niels Bohr, a Danish physicist who established the Copenhagen school.

(a) Assumptions underlying the Bohr atom

(1) Atoms can exist in stable “states” without radiating. The states have
discrete energies En, n = 1, 2, 3,..., where n = 1 is the lowest energy
state (the most negative, relative to the dissociated atom at zero
energy), n = 2 is the next lowest energy state, etc. The number “n” is
an integer, a quantum number, that labels the state.

(2) Transitions between states can be made with the absorption or


E
emission of a photon of frequency  where  = .
h
En1
h h
or
Absorption Emission
En2

These two assumptions “explain” the discrete spectrum of atomic vapor


emission. Each line in the spectrum corresponds to a transition between two
particular levels. This is the birth of modern spectroscopy.

h
(3) Angular momentum is quantized:  = n where  =
2
Angular momentum

   
L=rp = L

For circular motion:
L
  
L is constant if r and p are constant 
r
l = mrv is a constant of the motion  
p = mv
Other useful properties

1
Prepared By Kevin Boudreaux
12 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #4 page 2

v

( )
= 2 r   rot = r  rot
 
velocity
circumference frequency angular
(m/s)
(m/cycle) (cycles/s) frequency
(rad/s)

  = mvr = mr 2 rot

Recall the moment of inertia I =  mi ri2


i
2
 For our system I = mr

  = I  rot

Note: Linear motion vs. Circular motion

mass m  I moment of inertia


velocity v   rot angular velocity
momentum p = mv   = I angular momentum

Kinetic energy is often written in terms of momentum:

1 2 p2 1 m2 r 2 v 2  2
K.E. = mv = K.E. = =
2 2m 2 mr 2 2I

Introduce Bohr‛s quantization into the Rutherford‛s planetary model.

For a 1-electron atom with


r e-
a nucleus of charge +Ze
+Ze

Ze2 n2 2
r=
4 0 mv 2
 r=
Z
(
4 0
me2
) The radius is quantized!!

2
( 4 ) 0
me2
 a0 the Bohr radius

2
Prepared By Kevin Boudreaux
13 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #4 page 3

For H atom with n = 1, r = a 0 = 5.29x10 -11 m = 0.529 Å (1 Å = 10 -10 m)

Take Rutherford‛s energy and put in r,

1 Ze2 1 Z 2 me4
E=  En =  2 Energies are quantized!!!
2 4 0 r n 8 02 h2

For H atom, emission spectrum

me4  1 1
(
 cm -1
) =
hc
En2

En1
= 2 3  2  2
hc 8 0 h c  n1 n2

me4
Rydberg formula ! with R = 2 3 = 109,737 cm -1
8 0 h c

Measured value is 109,678 cm-1 (Slight difference due to model that gives
nucleus no motion at all, i.e. infinite mass.)

3
Prepared By Kevin Boudreaux
14 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #5 page 1

WAVE-PARTICLE DUALITY of LIGHT and MATTER

(A) Light (electromagnetic radiation)

Light as a wave

For now we neglect the polarization vector orientation


Propagating in x-direction:

0

 0
x (const. t) or
t (const. x)

 2 
( )
 x,t = 0 sin  x  ct
( )


2
Define k= (“wavevector” magnitude)


At some fixed time, say t = 0, can write simply  x,t = 0 = 0 sin kx ( ) ( )


Superposition principle

sin(kx)

Constructive
+ = 2sin(kx) interference

sin(kx)
(in phase)

Prepared By Kevin Boudreaux


15 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #5 page 2

sin(kx)

Destructive
+ =0 interference

sin(kx + )
(out of phase by /2)

This leads to many interference phenomena

Young‛s 2-slit experiment

D
l
s
Interference
pattern
light,  screen observed

D  
=  D=
 s s

Light as a particle

Compton exp‛t
e-
x-rays (  1Å)

If just a wave, expect light to scatter off electron

e-
 
Experimentally:
e-
 
Prepared By Kevin Boudreaux
16 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #5 page 3

The backscattered wave is red-shifted (   >  ), i.e. less energy/photon.

hc hc
E = < =E
 

Energy (and momentum) transferred to the electron.


Need relativistic mechanics to solve

h  h 
p= = for the light
c   

Light is a particle with energy


h
momentum p=
c

Light can behave both as a wave and as a particle!!


Which aspect is observed depends on what is measured.

(B) Matter

Matter as particles  obvious from everyday experience

Matter as waves (deBroglie, 1929, Nobel Prize for his Ph.D. thesis!)

Same relationship between momentum and wavelength for light and for matter

h h
p=  =  de Broglie wavelength
 p

Amazing notion! But wavelength only observable for microscopic momentum

Diffraction
e- crystal pattern

Prepared By Kevin Boudreaux


17 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #5 page 4

Consequences (I)

(1) on Bohr atom

+ +

If e- wave does not close on If e- wave does close on


itself, eventually destructive itself, then constructive
interference will kill it! interference preserves it.

Criterion for stability:


nh nh
2 r = n = =
p mv
nh
or mvr = = n
2
  = n

As Bohr had assumed angular momentum is quantized!

Prepared By Kevin Boudreaux


18 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #6 page 1

WAVE-PARTICLE DUALITY of MATTER


Consequences (II)


Heisenberg Uncertainty Principle xpx 
2

Consider diffraction through a single slit

x
peak-null
D
distance
s
l
light, 
D  
=  D=
 s s

Now consider a beam of electrons with de Broglie wavelength . The slit


restricts the possible positions of the electrons in the x direction: at the slit,
the uncertainty in the electron x-position is
xs=

x Coming out of the slit, the electrons


D spread out to form a diffraction pattern
with width D.

This means the electrons must go through the slit with some range of
velocity components vx

v
x xv
v

Prepared By Kevin Boudreaux


19 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #6 page 2

vx D  

= =
v  s x
xvx =  v or xpx =  p

h
But from deBroglie =
p

 xpx = h

So the position and momentum of a particle cannot both be determined with


arbitrary position! Knowing one quantity with high precision means that the
other must necessarily be imprecise!

The conventional statement of the Heisenberg Uncertainty Principle is


xpx 
2

(depends on how “uncertainty” is defined: 1/e half-width, FWHM, etc.)

Uncertainty can always be larger than  2 , but not smaller.

Note that this sort of uncertainty is standard in classical wave mechanics. If


you focus a light beam or a water wavelet to a small spot size, at the focus
there is a wide range of propagation directions. What is new is the idea that
particles inherently show wavelike behavior, with similar consequences.

Implications for atomic structure

Apply Uncertainty Principle to e- in H atom

me = 9.11 x 10-31 kg x  10-10 m (1 Å)

Prepared By Kevin Boudreaux


20 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #6 page 3

  1
px   vx   x106 m s
2x 2mx 2

Basically, if we know the e- is in the atom, then we can’t know its velocity at all!

Bohr had assumed the electron was a particle with a known position and velocity.
To complete the picture of atomic structure, the wavelike properties of the
electron had to be included.

So how do we properly represent where the particle is??

Schrödinger (1933 Nobel Prize)

A particle in a “stable” or time-independent state can be represented


mathematically as a wave, by a “wavefunction” (x ) (in 1-D) which is a solution to
the differential equation

Time-independent
2  2

2m x 2
() ()
+ V x  x = E x () Schrödinger
equation

potential energy total energy

We cannot prove the Schrödinger equation. But we can motivate why it might be
reasonable.

A v
( ) ( )
1 x,t = Asin kx   t is a right-traveling wave.

2
k=  = 2  = v


Similarly, a left-traveling wave can be represented as

( )
2 x,t = Asin kx +  t .( )
Both are solutions to the wave equation

Prepared By Kevin Boudreaux


21 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #6 page 4

( )=
 2 x,t 1   x,t
2
( )
x 2 v 2 t 2

( )
Further, the sum  x,t = 1 x,t + 2 x,t ( ) ( ) of left and right traveling waves is
also a solution.

( ) ( ) ( )
 x,t = A sin kx   t + sin kx +  t  = 2 Asin kx cos  t ( ) ( )
This is a stationary wave or standing wave. Its peaks and nulls remain stationary.

At various times during a full cycle (2/):

( )
 x,0 = 2 Asin kx ( ) t=0
 1 2  2
  x,  = 2 A 

8 
sin kx t = /4 ( )

2
 1 2 t = /2
  x,  = 0

4 
t = 3/4
 3 2   2
  x,  = 2 A 

8 
sin kx ( )

2 nodes
t = /
 1 2
( ) ( )
  x,  = 2 A 1 sin kx

2 

As in a vibrating violin string, the node positions are independent of time. Only
the amplitude of the fixed waveform oscillates with time.

More generally, we can write wave equation solutions in the form

( ) () ( )
 x,t =  x cos  t

()
In the particular case above,  x = 2 Asin kx . ( )

Prepared By Kevin Boudreaux


22 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #6 page 5

For the general case,

( )=
 2  x,t 1   x,t
2
( )
 2
x 2 v 2
t 2
= 2  x,t = k 2  x,t
v
( ) ( ) ( k=v )

( )
Plugging in  x,t =  x cos  t () ( )
 2 x ( ) = k   2 
2


x 2
2
()
x = 
 x


()
 2 x( ) =  p 2

de Broglie relation =
h
p

x 2

  x ()

 2 x ( )= p
 2

x 2
2
( x)

( )
But p = 2m K.E. = 2m  E  V x 
2
 ( ) (assuming t-independent potential)

2   x
2
()
 
2m x 2
+ V x  x = E x () () ()
time-independent Schrödinger equation in one dimension

We now have the outline of:

• a physical picture involving wave and particle duality of light and matter !

• a quantitative theory allowing calculations of stable states and their


properties !

Prepared By Kevin Boudreaux


23 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #7 page 1

SOLUTIONS TO THE SCHRÖDINGER EQUATION

Free particle and the particle in a box

Schrödinger equation is a 2nd-order diff. eq.

2   x
2
()

2m x 2
+ V x  x = E x () () ()
We can find two independent solutions 1 x and 2 x () ()
The general solution is a linear combination

()
 x = A1 x + B2 x() ()
A and B are then determined by boundary conditions on  x and   x . () ()
Additionally, for physically reasonable solutions we require that  x and ()
 x () be continuous function.

(I) Free particle V( x ) = 0

2   x
2
()

2m x 2
= E x ()
2mE 2 k 2
Define k = 2
2
or E =
 2m

p2
()
V x = 0, E =
2m
 p 2 = 2 k 2  p = k

h 2
de Broglie p=  k=
 

Prepared By Kevin Boudreaux


24 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #7 page 2

 2 x ( ) = k 
The wave eq. becomes
x 2
2
( x)
with solutions ()
 x = Acos kx + B sin kx( ) ( )
Free particle  no boundary conditions

2 k 2
 any A and B values are possible, any E = possible
2m

So any wavelike solution (traveling wave or standing wave) with any wavelength,
wavevector, momentum, and energy is possible.

(II) Particle in a box


 

()
V x = ( x < 0, x > a )
V ( x) = 0 (0  x  a ) V( x )
0
0 x a

Particle can’t be anywhere with V x =  ()


 (
 x < 0, x > a = 0 )
For 0  x  a , Schrödinger equation is like that for free particle.

2   x
2
()

2m x 2
= E x ()
 2 x ( ) = k 
x 2
2
( x) with same definition

2mE 2 k 2
k2 = or E =
2 2m

Prepared By Kevin Boudreaux


25 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #7 page 3

again with solutions () ( )


 x = Acos kx + B sin kx ( )
But this time there are boundary conditions!

Continuity of  x ()  ()
 0 = a = 0 ()
(i) () ()
 0 = Acos 0 + B sin 0 = 0 ()  A=0

(ii) ()
 a = B sin ka = 0 ( )
Can’t take B = 0 (no particle anywhere!)

Must have ( )
sin ka = 0  ka = n n = 1, 2, 3,...
n
 k is not continuous but takes on discrete values k =
a
Thus integer evolves naturally !!

So solutions to the Schrödinger equation are

 n x
( )
 0  x  a = B sin 
 a 

n = 1,2,3,...

These solutions describe different stable (time-independent or


“stationary”) states with energies

2 k 2 n2 h2
E=  En =
2m 8ma 2

Energy is quantized!! And the states are labeled by a quantum number n


which is an integer.

Properties of the stationary states

Prepared By Kevin Boudreaux


26 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #7 page 4

(a) The energy spacing between successive states gets progressively


larger as n increases

 
E5 = 25E1
2

( 2 h
)
2
En+1  En = n + 1  n
  8ma 2
E4 = 16E1

E3 = 9E1
(
En+1  En = 2n + 1 E1 )
E2 = 4E1
E1 = h2 8ma 2

(b) ()
The wavefunction  x is sinusoidal, with the number of nodes
increased by one for each successive state

4 = a 2 (
 4 = B sin 4 x a ) (3 nodes)

3 = 2a 3 (
 3 = B sin 3 x a ) ( 2 nodes)
n = 2a n # nodes = n - 1
a=
( ) (1 node )
2
 2 = B sin 2 x a

1 = 2a 1 = B sin ( x a ) (0 nodes)
0 x a

(c) The energy spacings increase as the box size decreases.


1
E
a2

Prepared By Kevin Boudreaux


27 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #7 page 5

We’ve solved some simple quantum mechanics problems! The P-I-B model is a
good approximation for some important cases, e.g. pi-bonding electrons on
aromatics.

Electronic transitions shift to lower energies as molecular size increases !

Prepared By Kevin Boudreaux


28 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #8 page 1

QUANTUM MECHANICAL PARTICLE IN A BOX

Summary so far:

( ) ( )
 
V x < 0, x > a =   x < 0, x > a = 0
 n x
V (0  x  a ) = 0  ( 0  x  a ) = B sin 
V( x )
n
a

0 n2 h2 n 2a
0 a En = k= = n = 1,2,3,...
x 8ma 2 a n

What is the “wavefunction”  x ? ()


Max Born interpretation:

() () ()
2
 x = * x  x is a probability distribution or probability density
for the particle

()
2
  x dx is the probability of finding the particle in the interval
between x and x + dx

This is a profound change in the way we view nature!! We can only know the
probability of the result of a measurement – we can’t always know it with
certainty! Makes us re-think what is “deterministic” in nature.

Easy implication: Normalization of the wavefunction

()
x2 2
  x1
 x dx = probability of finding particle in interval

The total probability of finding the particle somewhere must be 1.

For a single particle in a box,

Prepared By Kevin Boudreaux


29 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #8 page 2

()  ( )
 2 a 2
 
 x dx =
0
 x dx = 1 Normalization condition

a  n x  2
0
B 2 sin 2 
 a

dx = 1  B=
a

 n x 
()
n x =
2
a
sin 
 a 
n = 1,2,3,... Normalized wavefunction

()
2

( ) ( ) Interpretation of  x
2
 4 = 2 a sin 4 x a 2

based on measurement

Each measurement of the


( ) ( )
2
 = 2 a sin 3 x a 2
position gives one result.
()()
3
Many measurements give
a probability distribution
of outcomes.
( ) ( )
2
 2 = 2 a sin 2 2 x a

( ) ( )
2
 1 = 2 a sin 2  x a

Expectation values or average values

For a discrete probability distribution


0.4

0.35

0.3

0.25
e.g.
0.2

0.15

0.1

0.05

0
2 4 6 8 10 12 14

Prepared By Kevin Boudreaux


30 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #8 page 3

<x > = average value of x


= 4(0.1) + 6(0.1) + 8(0.2) + 10(0.4) + 12(0.2)
= 4(P4) + 6(P6) + 8(P8) + 10(P10) + 12(P12)

where Px is probability that measurement yields value “ x ”

 x =  xPx

Now switch to continuous probability distribution

()
2
Px   x dx

  

()
 2
 x =  
x  x dx

Similarly

()
 2
x2 =  
x 2  x dx

Often written in “sandwich” form

() ()

x =  
 * x x x dx

=   * ( x ) x  ( x )dx

x2 2


For particle in a box

2 a  n x 
x = 
a 0
x sin 2 
 a 
dx

Prepared By Kevin Boudreaux


31 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #8 page 4

a
Integrate by parts  x =
2

The average particle position is in the middle of the box.

Prepared By Kevin Boudreaux


32 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #9 page 1

VARIANCE, ROOT-MEAN SQUARE, OPERATORS,


EIGENFUNCTIONS, EIGENVALUES

xi  x  Deviation of ith measurement from average value <x >

xi  x  Average deviation from average value <x >

But for particle in a box, xi  x =0

(x  x )
2
i
 Square of deviation of ith measurement from average
value <x >

(x  x )
2
i
  x2  the Variance in x

(x  x )
2 2
Note i
= x2  x =  x2

The Root Mean Square (rms) or Standard Deviation is then

12
 x =  x2  x 
2

 

The uncertainty in the measurement of x, x , is then defined as

x =  x

 x for particle in a box

() () () ()
a 
 x2 = 0
 * x x 2 x dx   * x x x dx

2
 2 a  n x   2  a 2  n x 

=
 x 2 sin 2
dx 
 0 x sin dx 
a 0 a  a 

a 


Prepared By Kevin Boudreaux


33 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #9 page 2

Evaluate integral by parts

 2   2
a a2   a 
  =
2

3 2 n
( )  4

x 2

( )
 n 
2
2
a 
 = 2
 2

( )  3

x 2
4 n
12

( ) 
2
a  n
x =  x =  2

2 n  3

( )

Note that deviation increases with a, and depends weakly on n.

Now suppose we want to test the Heisenberg Uncertainty Principle for the
particle in a box.

12
p and p 2 to get p =  p =  p 2  p 
2
We need


() ()

But do we write p =  
 * x p x dx ?

what do we put in here??

We need the concept of an OPERATOR

ˆ x =g x
Af () ()
operator acts on function to get a new function

Prepared By Kevin Boudreaux


34 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #9 page 3

d  x2   2x 
e.g.   =  
dx  3   3

operator function new function

Prepared By Kevin Boudreaux


35 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #9 page 4

Special Case

If ()
ˆ x =af x
Af ()
number (constant)

then f x () is called an eigenfunction of the operator


and is the eigenvalue.
This is called an eigenvalue problem (as in linear algebra).

Quantum mechanics is full of operators and eigenvalue problems!!

e.g. Schrödinger’s equation:

 2 d 2 
 2
+ V x  x () () = E x ()
 2m dx 

Ĥ operator Eigenfunction constant


(Hamiltonian)

2 d 2
or Ĥ = E with ()
Ĥ x = 
2m dx 2
+V x () (in 1D)

The Hamiltonian operator, acting on an eigenfunction, gives the energy.


i.e. the Hamiltonian is the energy operator

p2
()
If V x = 0 , then E = K.E. =
2m

( p̂ )
2
d2
( p̂ )
2
 Ĥ =  = 2

2m dx 2

( p̂ ) ( pˆ ) ( p̂ )
2
means i.e. the operator acts sequentially on the function

Prepared By Kevin Boudreaux


36 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #9 page 5

( p̂ ) f ( x ) = ( pˆ ) ( pˆ ) f ( x ) = pˆ  pfˆ ( x ) = pˆ  g ( x )


2

 d  d d2
 ( )( )
p̂ p̂ =  i   i  = 2 2
 dx   dx  dx

d
 p̂ = i Momentum operator (in 1D)
dx
p
for Particle in a Box

2 2
 d
  d
() () () ()
2 
 2p = p 2  p =   * x  i  x dx     * x  i  x dx 
  dx 
  dx 

Note order is now very important! Operator acts only on the function to its
right.

 d
() ()

p =   * x  i
dx
 x dx

a
 2 1 2  n x  d
 2
12
 n x
= 
0
 sin 
  a  a



i
dx

 a

sin  a dx

=0

a  2 1 2  n x  d  d  2
12
 n x
=
 sin    i dx   i dx   a  sin  a  dx
 
2
p
0
  a   a   
2 2
22 a  n x  n  n x 22  n a
2  n x
=
a
0 sin  a   a  sin  a  dx = a  a 
0  a  dx
sin

n2 2 2
=
a2

Prepared By Kevin Boudreaux


37 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #9 page 6

n2 h2 n2 h2
Note p 2
= = 2m = 2mE as expected
4a 2 8ma 2
E = K.E. since V(x) = 0

n2 2 2
( )
2
 =2
p
= p
a2

12 12

( )  
( ) 
2 2
a n n   n
 xp =  2 =  2
2 n 3

( ) 

a 2 3



always > 1


 xp  as expected from Heisenberg Uncertainty Principle
2

Prepared By Kevin Boudreaux


38 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #10 page 1

THE POSTULATES OF QUANTUM MECHANICS


(time-independent)

Postulate 1: The state of a system is completely described by a


wavefunction  r,t . ( )
Postulate 2: All measurable quantities (observables) are described by
Hermitian linear operators.

Postulate 3: The only values that are obtained in a measurement of an


observable “A” are the eigenvalues “an” of the corresponding operator “ Â ”. The
measurement changes the state of the system to the eigenfunction of  with
eigenvalue an.

Postulate 4: If a system is described by a normalized wavefunction ,


then the average value of an observable corresponding to  is

a =   *Â d

Implications and elaborations on Postulates

2
#1] (a) The physically relevant quantity is 

( ) ( ) ( )
2
 * r,t  r,t =  r,t  probability density at time t
and position r

(b) ( )
 r,t must be normalized

  * d = 1

(c) ( )
 r,t must be well behaved

Prepared By Kevin Boudreaux


39 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #10 page 2

(i) Single valued


(ii)  and   continuous
(iii) Finite

#2] (a) Example: Particle in a box eigenfunctions of Ĥ

12
 2  n x 
() ()
Ĥ x  n x = En n x () n ()
x =
 a
sin 
 a 

But if  is not an eigenfunction of the operator, then the statement is not


true.

e.g. ()
 n x above with momentum operator

d  2   n x 
12

() d
dx
()
p̂n n x = i  n x = i
 sin
dx
a a  

 
 2  1 2  n x 
 pn
 sin  

 a a 

(b) In order to create a Q.M. operator from a classical observable, use


d
x̂ = x and p̂x = i and replace in classical expression.
dx
e.g.
2 d 2
K.E. =
1 2
2m
p̂ =
1
2m
p̂ p̂ =  ( )( )
2m dx 2
(1D)

2  2 2 2 
= + + (3D)
2m  x 2 y 2 z 2 

Another 3D example: Angular momentum L=rp

Prepared By Kevin Boudreaux


40 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #10 page 3

 d d
lx = ypz  zp y = i  y  z 
 dz dy 
 d d
l y = zpx  xpz = i  z x 
 dx dz 
 d d
lz = xp y  ypx = i  x  y 
 dy dx 

(c) Linear means

() () ()
  f x + g x  = Af ˆ x + Ag
ˆ x () and

() ()
  cf x  = c  f x 

(d) Hermitian means that

( )

  1 Â 2 d =   2 Â 1 d


and implies that the eigenvalues of  are real. This is important!!


Observables should be represented as real numbers.

Proof: Take  = a

( ) ( ) d

  Â d =   Â


( )

 a d =   a

d
 a = a*

true only if a is real

(e) Eigenfunctions of Hermitian operators are orthogonal

i.e. if  m = am m and  n = an n

Prepared By Kevin Boudreaux


41 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #10 page 4



then m
 n d = 0 if m  n

Proof:

( )

  m  n d =   n  m d


an   m  n d = am   n m d
 (a n
 am )  
 n d = 0
m

(a n
 am )  
m
 n d = 0

= 0 if = 0 if n  m
n=m

Example: Particle in a box

4
 1 4 As much + as - area

3
Eigenfunctions  1 3 
of
2
 1 2
1
0 x a 0 x a

In addition, if eigenfunctions of  are normalized, then they are orthonormal



m
 n d =  mn

Krönecker delta

Prepared By Kevin Boudreaux


42 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #10 page 5

1 if m = n (normalization)
 mn = 
0 if m  n (orthogonality)

#3] If  is an eigenfunction of the operator, then it’s easy, e.g.

Ĥ n = En n measurement of energy yields value

But what if  is not an eigenfunction of the operator?

e.g.  could be a superposition of eigenfunctions

 = c11 + c22

where Â1 = a11 and Â2 = a22

2 2
Then a measurement of A returns either a1 or a2 , with probability c1 or c2
respectively, and making the measurement changes the state to either 1 or 2 .

a1 1 (probability c12 )

a2 2 (probability c22 )
measure

#4] This connects to the expectation value

(i) If  n is an eigenfunction of  , then  n = an n

a =   n  n d = an   n n d = an

a = an only value possible

Prepared By Kevin Boudreaux


43 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #10 page 6

(ii) If  = c11 + c22 as above

 (c  ) ( )

a =    Aˆ d = 1 1
+ c22 Aˆ c11 + c22 d = c12 a1 + c22 a2

c12 is the probability of measuring a1

<a > = average of possible values weighted by their probabilities

Prepared By Kevin Boudreaux


44 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #11 page 1

PRINCIPLES OF QUANTUM MECHANICS (cont’d)


,
COMMUTATORS

Order counts when applying multiple operators!

e.g. Â = p̂x̂ = ? and can we write p̂x̂ = x̂p̂ ?

 operate on function to obtain .

 
()
ˆ x =g x
Af ()  ( p̂x̂ ) f ( x ) =  i dxd 
( x ) f ( x )
 
= ix
d
dx
() ()
d
f x  if x = ix  i f x
 dx

()
 


d
 = ix  i = p̂x̂
dx

( )

Now try B̂ = x̂p̂

ˆ x = x  i d  f x =  ix d  f x
Bf () () () ()
 dx 
 dx 

d
 B̂ = ix = x̂p̂
dx
 x̂p̂  p̂x̂

Define commutator

For two operators  and B̂ ,

 Â, B̂  = ÂB̂  B̂Â = Ĉ


 
need not be zero!
e.g.
 x̂, p̂ = x̂p̂  p̂x̂ = i  0 !

Prepared By Kevin Boudreaux


45 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #11 page 2

Important general statements about commutators:

1) For operators that commute

 Â, B̂  = 0
 

• it is possible to find a set of wavefunctions that are eigenfunctions


of both operators simultaneously.

e.g. can find wavefunctions  n such that

 n = an n and B̂ n = bn n

• This means that we can know the exact values of both observables
A and B simultaneously (no uncertainty limitation).

2) For operators that do not commute

 Â, B̂   0
 

• it is not possible to find a set of wavefunctions that are


simultaneous eigenfunctions of both operators.

• This means that we cannot know the exact values of both


observables A and B simultaneously  uncertainty !


e.g.  x̂, p̂ = i  0  xp 
2

Prepared By Kevin Boudreaux


46 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 1

THE HARMONIC OSCILLATOR

• Nearly any system near equilibrium can be approximated as a H.O.

• One of a handful of problems that can be solved exactly in quantum


mechanics

examples

m1 m2 B (magnetic
field)
A diatomic molecule μ
(spin
magnetic
moment)
E (electric
field)

Classical H.O.

m
k

X0 X

Hooke’s Law: (
f = k X  X 0  kx )
(restoring force)
d2x d2x  k 
f = ma = m 2 = kx  + x=0
dt dt 2  m

Prepared By Kevin Boudreaux


47 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 2

Solve diff. eq.: General solutions are sin and cos functions

() ( )
x t = Asin  t + B cos  t ( ) =
k
m
or can also write as

() (
x t = C sin  t +  )
where A and B or C and  are determined by the initial conditions.

e.g. ()
x 0 = x0 ()
v 0 =0
spring is stretched to position x 0 and released at time t = 0.

Then

() ()
x 0 = A sin 0 + B cos 0 = x0 ()  B = x0

()
v 0 =
dx
dt
()
=  cos 0   sin 0 = 0 ()  A=0
x=0

So ()
x t = x0 cos  t ( )
k
Mass and spring oscillate with frequency:  =
m
and maximum displacement x0 from equilibrium when cos(t)= ±1

Energy of H.O.
Kinetic energy  K

2
1  dx 
1 1 1
( ) ( )
2
K = mv 2 = m  = m
  x0 sin  t = kx02 sin 2  t
2 2  dt 2 2

Potential energy  U

()
f x =
dU
dx
 ()
U =   f x dx =  ( kx )dx =
1
2
kx 2
=
1 2
2
( )
kx0 cos 2  t

Prepared By Kevin Boudreaux


48 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 3

Total energy = K + U = E

E=
1 2
2
( )
kx0 sin 2  t + cos 2  t  ( ) E=
1 2
kx
2 0

x (t )
x 0(t )

0 t

-x0(t)

U K
1 2
kx E
2 0

0 t

Most real systems near equilibrium can be approximated as H.O.

e.g. Diatomic molecular bond A B

X
U

X
X0 A + B separated atoms

equilibrium bond length

Prepared By Kevin Boudreaux


49 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 4

1 d 2U 1 d 3U
( ) ( ) dU
(X  X ) (X  X ) (X  X )
2 3
U X = U X0 + 0
+ 0
+ 0
+
dX X = X0
2 dX 2 X = X0
3! dX 3 X = X0

Redefine x = X  X 0 and ( )
U X = X0 = U x = 0 = 0 ( )
1 d 2U 1 d 3U
()
U x =
dU
dx
x+
2 dx 2
x +
2

3! dx 3
x3 + 
x=0 x=0 x=0

real potential

H.O. approximation

dU
At eq. =0
dx x=0

For small deviations from eq. x 3 << x 2

1 d 2U
 U x  () 2 dx 2
x2 
1 2
2
kx
x=0

Prepared By Kevin Boudreaux


50 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 5

Total energy of molecule in 1D


m1 m2

X1 XCOM X2 X
M = m1 + m2 total mass
xrel
m1 m2
μ= reduced mass
m1 + m2
m1 X 1 + m2 X 2
X COM = COM position
m1 + m2
xrel = X 2  X 1  x relative position

2 2 2 2
1  dX  1  dX  1  dX  1  dx 
K = m1  1  + m2  2  = M  COM  + μ  
2  dt  2  dt  2  dt  2  dt 

1 2
U= kx
2

2 2
1  dX  1  dx  1
E = K + U = M  COM  + μ   + kx 2
2  dt  2  dt  2

COM coordinate describes translational motion of the molecule

2
1  dX 
Etrans = M  COM 
2  dt 

QM description would be free particle or PIB with mass M

We’ll concentrate on relative motion (describes vibration)

2
1  dx  1
Evib = μ   + kx 2
2  dt  2

and solve this problem quantum mechanically.

Prepared By Kevin Boudreaux


51 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 6

THE QUANTUM MECHANICAL HARMONIC OSCILLATOR

 2 d 2 1 2 
()
Ĥ x =   2
+ kx   x = E x () ()
 2m dx 2

K U

Note: replace m with μ (reduced mass) if m1 m2

Goal: Find eigenvalues En and eigenfunctions n(x )

Rewrite as:

d 2 x ( ) + 2m  E  1 kx 
dx 2

2 
2
()
 x = 0
2

This is not a constant, as it was for P-I-B,


so sin and cos functions won’t work.

()
f x = e x
2
2
TRY: (gaussian function)

d2 f x ( ) =  e  x 2 2
+  2 x 2 e x () ()
2

2
2
=  f x +  2 x 2 f x
dx

d2 f x ( ) + f
or rewriting,
dx 2 ( x)   2
()
x 2 f x = 0w

which matches our original diff. eq. if

2mE mk
= and 2 =
2 2

 k
 E=
2 m

Prepared By Kevin Boudreaux


52 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 7

We have found one eigenvalue and eigenfunction

k 1 k
Recall = or =
m 2 m

1 1
 E=  = h
2 2

This turns out to be the lowest energy: the “ground” state

For the wavefunction, we need to normalize:

()
 x = Nf x = Ne x ()
2
2
where N is the normalization constant

14
2

()
 
 
 x2
 x dx = 1  N 2
e =1  N =

  

 

14

() e x
2
0 x =
2

  0 x ()
1 1 1
E0 = 
= h E0 = 
2 2 2
x

()
Note  0 x is symmetric. It is an even function:  0 x =  0 x () ( )
There are no nodes, & the most likely value for the oscillator displacement is 0.

So far we have just one eigenvalue and eigenstate. What about the others?

Prepared By Kevin Boudreaux


53 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 8

14

() e x
1
2
0 x =
2
E0 = h
 2
14
1 
()
1 x =

2 
( 2 x ) e
12  x 2 2
E1 =
3
2
h
14
1 
2 ()x =

8 
( 4 x 2
)
 2 e x
2
2
E2 =
5
2
h

14
1 
3 ()x =

48 
(8 32
x 3  12 1 2 x e x ) 2
2
E3 =
7
2
h

 
12
 km 
with  = 2 
 

These have the general form

14

()
n x =
1
1/ 2
 
(
H n  1 2 x e x ) 2
2
n = 0,1,2,...
( 2 n!)
n

Normalization Gaussian
Hermite polynomial (pronounced “air-MEET”)

( )
H0 y = 1 even ( n = 0)
H ( y) = 2 y
1
odd ( n = 1)
H ( y) = 4 y  2
2
2
even ( n = 2 )
H ( y ) = 8y  12 y
3
3
odd ( n = 3)
H ( y ) = 16 y  48y
4
4 2
+ 12 even ( n = 4 )
 

Prepared By Kevin Boudreaux


54 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 9

() ()
2
n x n x

7
3 x () E3 =
2

5
2 x () E2 =
2

3
1 x () E1 =
2

0 x () E0 =
2

 1
Energies are En =  n +  h
 2

Note E increases linearly with n.


Energy levels are evenly spaced

 1  1

( 2
)
En+1  En =  n + 1 + h   n + h = h
 2
regardless of n

1
There is a “zero-point” energy E0 = h
2

E = 0 is not allowed by the Heisenberg Uncertainty Principle.

Prepared By Kevin Boudreaux


55 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 10

Symmetry properties of ’s

 0,2,4,6,.... are even functions ( ) ()


 x =  x
 1,3,5,7,.... are odd functions  ( x ) =  ( x )

Useful properties: (even) (even) = even


(odd) (odd) = even
(odd) (even) = odd

( )=
d odd (
d even )=
dx
(even ) dx
( odd )
 ( odd ) dx = 0  (even ) dx = 2  (even ) dx
  

  0

Just from symmetry:

 d
() () ()
 
x n
= 
 n x x n x dx = 0 p n
=  n  dx
 n x dx = 0
 ih

odd odd

Average displacement & average momentum = 0

IR spectroscopy
H.O. selection rules

Intensity of vibrational absorption features + -

n’ = 1

Vibrational transition h

n=0

Prepared By Kevin Boudreaux


56 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 11

2
dμ  
dx  n n '
Intensity I nn   x dx

1) Dipole moment of molecule must change as molecule vibrates


HCl can absorb IR radiation, but N 2, O2, H2 cannot.

2) Only transitions with n = n ± 1 allowed (selection rule).


(Prove for homework.)

QUANTUM MECHANICAL HARMONIC OSCILLATOR &


TUNNELING

Classical turning points E

1 2
Classical H.O.: Total energy ET = kx
2 0
oscillates between K and U.
ET

Maximum displacement x0 occurs when


all the energy is potential.
x
-x0 x0
2ET
x0 = is the “classical turning point”
k

The classical oscillator with energy ET can never exceed this displacement, since
if it did it would have more potential energy than the total energy.

Prepared By Kevin Boudreaux


57 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 12

Quantum Mechanical Harmonic Oscillator.

1 2
kx
()
2
()
2 2
3 x  12 x
()

()
2
2 x

()
2
1 x At high n, probability
density begins to look
classical, peaking at turning

() points.
2
0 x

Non-zero probability at x > x 0!

()
2
Prob. of (x > x 0, x < -x 0):
3 x 1

()
 2 
2

e x dx
2
 02 x dx = 2


1 2 1 2
 

()
2
2 x 2
()


2
= e y dy = erfc 1
1 2 1

“Complementary error function”


()
2
1 x tabulated or calculated
numerically

()
2
Prob. of (x > x 0, x < -x 0) = erfc(1)
0 x
= 0.16
Significant probability!
x
Prepared By Kevin Boudreaux
58 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 13

The oscillator is “tunneling” into the classically forbidden region. This is a purely
QM phenomenon!

Tunneling is a general feature of QM systems, especially those with very low


mass like e- and H.

()
 x ~ sin kx ( ) ()
 x ~ e  x
Finite

V0
barrier
()
 x ~ sin kx +  ( )

x
Even though the energy is less than the barrier height, the wavefunction is
nonzero within the barrier! So a particle on the left may escape or “tunnel” into
the right hand side.

 2 d 2 
Inside barrier: 

2m 2
+ V0
 x = E x () ()
 dx

d 2 x ( ) =  2m (V E )
or
dx 2 
0
2 ()
  x   2 x ()
 
1

(
 2m V0
E )  2

Solutions are of the form ()


 x = Be
 x with  =
2


( )
12
Note   V0
E and   m1 2

If barrier is not too much higher then the energy and if the mass is light, then
tunneling is significant.
Important for protons (e.g. H-bond fluctuations, tautomerization)

Prepared By Kevin Boudreaux


59 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 14

Important for electrons (e.g. scanning tunneling microscopy)

Nonstationary states of the QM H.O.

System may be in a state other than an eigenstate, e.g.

2 2 1
 = c0 0 + c1 1 with c0 + c1 = 1 (normalization), e.g. c0 = c1 =
2
Full time-dependent eigenstates can be written as

( )
 0 x,t =  0 x e ()
i 0 t
( ) ()
1 x,t =  1 x e

i1t

where
1 1 3 3
 0 = E0 =  vib   0 =  vib  1 = E1 =  vib   1 =  vib
2 2 2 2

System is then time-dependent:

( )
 x,t =
1
e

i 0 t
()
0 x +
1
e

i1t
() () ( )
 1 x = c0 t  0 x + c1 t  1 x () ( )
2 2

where ()
c0 t =
1
e

i 0 t
c1 t = () 1
e

i1t

2 2

What is probability density?

( ) ( )
  x,t  x,t =
1 
2

0
x e ()
i 0 t
+

1
x e
i1t
 ()
 
0 x e i 0 t +
1 x e i1t 
() ()
= 
0
0 +
1
1 +
1
0 e ( 1 0 ) +
0
1e ( 1 0 )  = 
02 +
12 + 2
0
1 cos  vib t 
1
2
i   t  i   t 1
2 ( )

Probability density oscillates at the vibrational frequency!

Prepared By Kevin Boudreaux


60 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 15

 t = 0,  4,  2, 3 4, 
( )
2
 x,t

() ( )
2
1 x

()
2
0 x
()

x
2 0 1 cos ( t )
 t = 0,  4,  2, 3 4, 

What happens to the expectation value <x>?

Prepared By Kevin Boudreaux


61 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 16

( ) ( )

x = 
  x,t x̂ x,t dx

=
1  
2 
() () ()

0 x ei 0 t +
1 x ei1t x 

0 x e i 0 t +
1 x e i1t dx ()
= 
0 x
0 dx +
1 x
1 dx +
1 x
0 e ( 1 0 ) dx +
0 x
1e ( 1 0 ) dx
1    i   t   i   t

2
    
( )

<x >0 = 0 <x >1 = 0 = cos  vib t
0 x
1 dx


<x>(t) oscillates at the vibrational frequency, like the classical H.O.!



Vibrational amplitude is 

 0 x 1 dx
1 1
 1 
() () ( 2 x ) e
4 4
 0 x =
e
 x
 x 2 2
2
2
1 x =

12


 2

1

() ( ) ()
4
1 2
x 0 x =
xe
 x 2 = 2
2
 1 x



( )  ( ) () ( ) ( )

1 2 
1 2
1 2
 

 0 x 1 dx = 2


 02 dx = 2 x t = 2 cos  vib t

Relations among Hermite polynomials

Recall H.O. wavefunctions


1
4
()
n x =
1
1/ 2
 
H  12
x (
e
 x 2
) 2
n = 0,1,2,...
( 2 n!)
n
n

Normalization Gaussian
Hermite polynomial

Prepared By Kevin Boudreaux


62 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 17

( )
H0 y = 1 even ( n = 0)
H ( y) = 2 y
1
odd ( n = 1)
H ( y) = 4 y
2
2
2
even ( n = 2 )
H ( y ) = 8y
12 y
3
3
odd ( n = 3)
H ( y ) = 16 y
48y
4
4 2
+ 12 even ( n = 4 )
 

Generating formula for all the Hn:


d n
y2
( ) ( )
n y2
Hn y =
1 e e
dy n

A useful derivative formula is:


( )=
dH n y d n
y2 n+1

( ) ( ) ( ) ( )
n y2 n y2 d
y2

1 2 ye e +
1 e e = 2 yH n y
H n+1 y
dy dy n dy n+1

Another useful relation among the Hn’s is the recursion formula:

( ) ( )
H n+1 y
2 yH n y + 2nH n
1 y = 0 ( )
Substituting ( ) ( )
2 yH n y = H n+1 y + 2nH n
1 y above gives ( )
dH n y ( ) = 2nH
dy n
1 ( y)
Use these relations to solve for momentum <p>(t)

( ) ( )

p = 
  x,t p̂ x,t dx

=
1  

2 
() () ()

0 x ei 0 t +
1 x ei1t pˆ 

0 x e i 0 t +
1 x e i1t dx ()
= 
0 pˆ
0 dx +
1 pˆ
1 dx +
1 pˆ
0 e ( 1 0 ) dx +
0 pˆ
1e ( 1 0 ) dx
1    i   t   i   t

2
    
<p >0 = 0 <p >1 = 0

Prepared By Kevin Boudreaux


63 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 18

1 1
 
d
() (
 x ) e ()
4 2

 x 2 2
0 x = =
1 x
dx   2
1 1
  i   t    i t

1 p̂
0 e ( )
dx = i  e ( 1 0 )
1
1 dx = i  e vib
 i 1   0 t
2  2
  2  2

 0 p̂ 1e ( 1 0 ) dx use relations among Hn’s


  i   t
To solve integral  

d
dx
1 x =
d 
dx 
()
2

( ) d
N 1 H1  1 2 x e
 x 2  =  1 2 N 1  H1 y e
y 2 
dy 
2

( )
d d
with y   1 2 x dy =  1 2 dx dx = 
1 2 dy = 1 2
dx dy
d 
d
() ( ) ( )
2 2
 1 x =  1 2 N1 H1 y e
y 2
yH1 y e
y 2
dx  dy
d
( )
H y = 2nH 0 y = 2H 0 y
dy 1
( ) ( )
( )
yH1 y =
1
2
( )
1
( )
 2nH 0 y + H 2 y  = H 0 y + H 2 y
2
( ) ( )
   1 
d
() ( )
1
( ) 1
() ()
2 2
 1 x =  1 2 N1 H 0 y e
y 2
H 2 y e
y 2 =  1 2 N1  0 x
2 x
dx  2  N0 2N 2


0 p̂
1e ( 1 0 ) dx = e ( 1 0 ) i ( ) d
  i   t  i   t 
 

0
dx
dx 1
1 
( )
= e ( 1 0 ) i  1 2 N1
1
 i   t  

0
0 dx   0 2 



dx

 N0 2N 2 
1
N1     i t
= e ( 1 0 ) i  1 2 ( )
2
 i   t
= i  e vib
N0 2

Finally

Prepared By Kevin Boudreaux


64 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lectures #12-15 page 19

 1
 1

()
  2 i t
p t = i e vib
e vib
1
2  2

i t
( )  =
   2 sin  t
  2 vib( )



Average momentum also oscillates at the vibrational frequency.

Prepared By Kevin Boudreaux


65 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 12-15 Lecture Supplement Page 1

Harmonic Oscillator Energies and Wavefunctions


via Raising and Lowering Operators

We can rearrange the Schrödinger equation for the HO into an interesting form ...
1   d 2
2
1

+ ( m x )   =  p 2 + ( m x )2  = E
2m   i dx  2m 

with
1
H=  p 2 + ( m x )2 
2m  
which has the same form as
u 2 + v 2 = ( iu + v )(
iu + v ) .

We now define two operators

1
a±  ( ip + m x )
2m

that operate on the test function f(x) to yield

( a a ) f ( x ) =  ( ip + m x )(
ip + m x ) f ( x )
1

+
2m 

1
 p 2 + ( m x )
im (xp
px)  f (x)
2
=
2m  

= { 1
2m
( i
)
p 2 + ( m x )
[ x, p ] f ( x )
2
2 }
a
a+ =
1
2m
( 1
p 2 + ( m x ) + =
2 1
2 
H+ )
1
2

Which leads to a new form of the Schrödinger equation in terms of a+ and a- …

 1
H  =  a
a+
 
 2

If we reverse the order of the operators-- a


a+  a+ a
-- we obtain …

Prepared By Kevin Boudreaux


66 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 12-15 Lecture Supplement Page 2

 1
H  =  a+ a
+  
 2

or
 1
  a± a ±  = E
 2

and the interesting relation


a
a+
a+ a = [ a a ] = 1


+

A CLAIM: If  satisfies the Schrödinger equation with energy E, then a+


satisfies it with energy (E+ ) !
 1  1 
H ( a +  ) =   a + a
+  ( a +  ) =  a + a
a + + a +  
 2
 2

 1   1   1
=  a+ a
a+ +  = a+  a+ a
+ 1 +   = a+  a+ a
+ +  
 2
 2 
 2 
= a+ ( H +  ) = ( E +  )( a+ )

H ( a+ ) = (E +  ) ( a+ )

Likewise, a- satisfies the Schrödinger equation with energy (E- ) …


 1  1   1
H ( a
 ) =  a
a+
 ( a
 ) =  a
a+ a

a
  = a  a+ a

 
 2
 2
 2

  1
= a
 a
a+
1
  = a
( H
 ) = a
( E
 )

 2 

H ( a
 ) = ( E
 ) ( a
 )

So, these are operators connecting states and if we can find one state then we can
use them to generate other wavefunctions and energies. In the parlance of the
trade the a± are known as LADDER operators or

a+ = RAISING and a- = LOWERING operators.

We know there is a bottom rung on the ladder 0 so that

a
 0 = 0

Prepared By Kevin Boudreaux


67 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 12-15 Lecture Supplement Page 3

1  d 
  + m x   0 = 0
2m dx

Integrating this equation yields


d 0 m
=
x 0
dx 

d m m 2
  00 =
 
xdx  ln 0 =

2h
x + A0


m x2 1
 ( x) = A e
0 0
2
and E0 = 
2

E0 comes from plugging  0 into H = E . We will perform the normalization


below.

Now that we are firmly planted on the bottom rung of the ladder, we can
utilize a+ repeatedly to obtain other wavefunctions, n , and energies, En. That is,

m
n
2  x2  1
 n ( x ) = An ( a+ ) e , with En =  n +  
 2
Thus, for 1 we obtain

1 1
 m  4  2m  2
m2 x2
 1 ( x ) = A1
    
xe

where you still have to determine the normalization constant A1.

Prepared By Kevin Boudreaux


68 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 12-15 Lecture Supplement Page 4

Algebraic Normalization of the Wavefunctions: We can perform the normalization


algebraically. We know that

a+ n = cn n+1 a


 n = dn n
1

What are the proportionality factors cn and dn ? For any functions f (x) and g(x)
 

f  (a± )gdx =  
(a f ) g dx

here a is the Hermitian conjugate of a±

Proof:
 1   d

f  (a± g) dx =
(2m )1 2

f   + m x
g dx
 dx

recall that
1 1   d  1  d
12 [
a± = ip + mwx ] =  i  i dx + mwx  = (2m )1 2  dx + mwx 

(2m ) (2m )1 2 

Integrate by parts

1   d   

 ± dx + mwx   (a f ) g dx
 
f (a± g) dx = dx = f  gdx =
(2mw)  

So we can write
 


(a± n ) *(a± n )dx =  (a a±


n ) *  n dx

We now use
 1  1
  a± a ±  n = En n and En =  n +  w
 2  2
 1  1
 a+ a
±   n =  n +   n
2 2
and therefore
a+ a
 n = n n
And
 1  1
 a
a+
  n =  n +   n
 2
 2

Prepared By Kevin Boudreaux


69 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 12-15 Lecture Supplement Page 5

a
a+ n = ( n + 1) n
We can now calculate cn :

   
 (a+ n ) * (a+ n )dx = cn 
 n+1 n +1dx = 
 ( a
a+ n ) * n dx = (n + 1) 
 n n dx
2 * *



cn = n +1

The calculation for dn proceeds in a similar manner :


   
 
 n
1 n
1dx = 
 ( a+ a
 n )  n dx = n 
 n n dx
*
(a
 n ) * (a
 n )dx = dn 2 * *



dn = n

Thus we obtain the two normalization constants for the a±

a+ n = ( n + 1)  n+1
12
a
 n = n1 2  n
1

HO Wavefunctions: Rearranging the equations to a slightly more useful form yields

1 1
 n +1 = a+ n  n
1 = a
 n
( n + 1)12
n1 2

We can now use these equations to generate other wavefunctions. Thus, if we start
with  0 we obtain:

1
n=0 1 = a+ 0 = a+ 0
( 0 + 1)1 2
1 1
n =1 2 = a+ 1 = ( a+ )2  0
2 2
1 1
n=2 3 = a+ 2 = a+3 0
2 +1 3 2
1 1
n=3 4 = a+ 3 = a+4 0
3+1 4  3 2
So that  n is
1
( a+ )  0
n
n =
n!

Prepared By Kevin Boudreaux


70 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 12-15 Lecture Supplement Page 6

Orthogonality of the HO Wavefunctions: Recall the orthogonality condition for


two wave functions is


 n mdx =  nm
*

Using the a± operators we can show this condition also holds for the HO
wavefunctions. The proof is as follows.

  ( a+ a
) dx = n   m*  n dx
*
m n

 ( a ) ( a ) dx =  ( a+ a ) 


 
m m m n dx = m   m*  n dx

( n
m )   m*  n dx = 0

The trivial case occurs when n = m ; but when n  m then

  n dx = 0
*
m

Potential Energy of the Harmonic Oscillator: We can now use the a± operators to
perform some illustrative calculations. Consider the potential energy associated with
the HO.

1 2 1
V= kx = m 2 x 2
2 2
and therefore
1 1
V = m 2 x 2 = m 2   n* x̂ 2 n dx
2 2

First, we express x̂ and p̂ in terms of a± operators …

12 12
    m 
x̂ = 
 2m 
( a+ + a
) p̂ = i 
 2 
( a+
a
)
and
     
x2 = 
 2m 
( a+ + a
) = 
2
( a a + a+ a
+ a
a+ + a
a
)
 2m  + +

Prepared By Kevin Boudreaux


71 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 12-15 Lecture Supplement Page 7

Dirac Notation: Before we evaluate this expression let’s introduce some new
notation that will make life simpler for us on future occasions. Instead of writing
the integral between ± we use brackets to denote this integral. The first
half is called a “bra” and the second a “ket”. That is, “bra”-c-“ket” notation is

bra = and ket =


and for the probability density we would have an expression such as




 m*  n dx = mn
where the present of  m* is understood. Using this notation, integrals for
x , p , and x 2 assume the form

 


 m* x̂ n dx = m x̂ n and 

 m* p̂ n dx = m p̂ n

1 1 1
V = m 2 x 2 = m 2   n* x̂ 2 n dx = m 2 n x̂ 2 n
2 2 2

  1
V = m 2  n a+ a+ n + n a+ a
n + n a
a+ n + n a
a
n

 2m 2
yielding
h   1
V =  n + n + 1  =  n+ 
4 2  2

It is important to not get totally embroiled in the equations and neglect the
chemistry and physics. Accordingly, we should ask the question as to the physical
significance of this formula ?

Prepared By Kevin Boudreaux


72 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture Summary 12-15, Supplement page 1

Ehrenfest’s Theroem

In the lecture notes for the harmonic oscillator we derived the


expressions for x̂ (t) and p̂x (t) using standard approaches – integrals
involving Hermite polynomials (see pages 17 and 18, Lecture Summary 12-15).
The calculations are algebraically intensive, but showed that x̂ (t) and
p̂x (t) oscillate at the vibrational frequency. The results were as follows:

12
  
x (t) = ( 2 ) cos ( vibt ) = cos ( vibt )

1 2

 2 μ 

and
1   
12 12
 μ
p (t) = i
2
 2
(e i vib t

e
i vib t
)  =

 2
sin ( vibt )


The issue considered here is an approach to calculate x (t) and p (t) in a


more straightforward manner.

Classically, (we use m instead of μ since we are dealing with a free


particle)
dx
p = mv = m
dt
So, quantum mechanically we might expect
d x (t)
p (t) = m .
dt
But, is this expression valid ? We can show that in fact it is with the
following argument.

d x (t)
For our original expression was …
dt
d x (t) d   *
 d * 
d
=   x̂ dx =  x̂ dx +  * x̂ dx
dt dt     dt 
dt
Recall the time dependent Schrödinger equation is

  1
i = H or = H
t t i
Inserting these results into the expression above yields

Prepared By Kevin Boudreaux


73 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture Summary 12-15, Supplement page 2

 
d x (t) 1
( ) 1
 x̂ ( Ĥ ) dx
*
=   Ĥ x̂ dx + *

dt i  i 

 

=
1

i 
( ) i
(
* x̂Ĥ  Ĥx̂ dx =  * Ĥx̂  x̂Ĥ dx
 
)
Evaluating the commutator (assuming that H is the HO Hamiltonian) we find

 2 d 2   2 d 2 
( )
Ĥx̂  x̂Ĥ f (x) =  
 2 μ dx 2
+ U(x)

ˆ
x f (x)  x̂  
 2 μ dx 2
+ U(x) f (x)


2 df 2  i  i
= 2 =   p̂x  =  p̂x
2 μ dx μ   μ
And therefore
d x (t) i  * 

dt
( 1
)
=  Ĥx̂  x̂Ĥ dx =  * p̂x dx
  m 

Which is the result that is desired


d x (t)
p̂x (t) = μ
dt

Thus, we can now obtain p̂x (t) without the lengthy calculation contained in
the HO lecture notes.

d x (t) d 
 
12

 μ 
12

p̂x (t) = μ = μ  cos ( t )  =   sin ( t )


dt dt
 2 μ 
2

12
 μ  
p̂x (t) =   sin ( t )
 2 

which is the result with which we started initially.

The equations above are a specific illustration of a more general


result due to Paul Ehrenfest (an Austrian physicst who later resided in
Leiden, The Netherlands) and known as Ehrenfest’s Theorem. In particular,
for any dynamical variable F

Prepared By Kevin Boudreaux


74 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture Summary 12-15, Supplement page 3

d F (t) i  *
dt
(
=  ĤF̂  F̂Ĥ dx
 
)
For further information see McQuarrie Problems 4-43 and 4-44, p 187-188.

Prepared By Kevin Boudreaux


75 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Separable Systems page 1

48$1780 ,1  6(3$5$%/( 6<67(06

' 6\VWHPV ' 6\VWHPV


K K K
x̂ r̂ = ( x̂ ŷ ẑ ) = i x̂ + j ŷ + k ẑ
= d K= ∂ K= ∂ K= ∂
p̂ =
i dx (
p̂ = p̂x p̂y p̂z = i )
i ∂x
+j
i ∂y
+k
i ∂y
ª¬ x̂, p̂¼º = i= ª¬ x̂, p̂x º¼ = i= ª yˆ, pˆ y º = i= ª¬ ẑ, p̂z º¼ = i=
¬ ¼
p̂2 −=2 d 2 p̂2 −=2 ∂ 2 −=2 ∂ 2 −=2 ∂ 2
Tˆ = = Tˆ = = + +
2m 2m dx 2 2m 2m ∂x 2 2m ∂y2 2m ∂y 2
ψ (x) ψ ( x, y, z )
Ô = ³ψ ( x ) Ô ψ ( x ) dx Ô = ³ψ ( x, y, z ) Ô ψ ( x, y, z ) dx dy dz
* *

%\ ILDW RSHUDWRUV FRUUHVSRQGLQJ WR GLIIHUHQW D[HV FRPPXWH ZLWK RQH DQRWKHU


ˆ ˆ = yx
xy ˆˆ pˆ z yˆ = yp
ˆˆz pˆ z pˆ x = pˆ x pˆ z etc.
)XUWKHU RSHUDWRUV LQ RQH YDULDEOH KDYH QR HIIHFW RQ IXQFWLRQV RI DQRWKHU
ˆ ( y ) = f ( y ) xˆ
xf pˆ z f ( x ) = f ( x ) pˆ z f * ( z ) pˆ x = pˆ x f * ( z ) etc.

7KH 7LPH ,QGHSHQGHQW 6FKU|GLQJHU (TXDWLRQ EHFRPHV


ª =2 § ∂ 2 ∂2 ∂2 · º
« − ¨ 2 + + 2 ¸
+ V ( x̂, ŷ, ẑ ) »ψ ( x, y, z ) = Eψ ( x, y, z )
¬ 2m © ∂x ∂y 2
∂z ¹ ¼

∇ 2 WKH /DSODFLDQ

ª =2 2 º
Ÿ « − 2m ∇ +V ( x̂, ŷ, ẑ ) »ψ ( x, y, z ) = Eψ ( x, y, z )
¬ ¼

=2 2
Ĥ = − ∇ +V ( x̂, yˆ, zˆ ) +DPLOWRQLDQ RSHUDWRU LQ '
2m

Hˆψ ( x, y , z ) = Eψ ( x, y , z ) ' 6FKU|GLQJHU HTXDWLRQ


7LPH ,QGHSHQGHQW
6HSDUDWLRQ RI YDULDEOHV

Prepared By Kevin Boudreaux


76 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Separable Systems page 2

,) V ( x̂, ŷ, ẑ ) = Vx ( x̂ ) +V y ( ŷ ) +Vz ( ẑ )

ª =2 ∂ 2 º ª =2 ∂ 2 º ª =2 ∂ 2 º
H ( x, y, z ) = « −
ˆ +V ( ˆ
x ) + −
» « 2m ∂y 2 +V ( ˆ
y ) + −
» « 2m ∂z 2 +V ( ˆ
z ) »
¬ 2m ∂x
2 x y z
WKHQ ¼ ¬ ¼ ¬ ¼
= Ĥ x + Ĥ y + Ĥ z
Ÿ 6FKU|GLQJHU·V (T EHFRPHV
ª Ĥ x + Ĥ y + Ĥ z ºψ ( x, y, z ) = Eψ ( x, y, z )
¬ ¼

7KHQ WU\ VROXWLRQ RI IRUP ψ (x, y, z )= ψ x (x )ψ y (y )ψ z (z )


VHSDUDWLRQ RI YDULDEOHV
:KHUH ZH DVVXPH WKDW WKH ' IXQFWLRQV VDWLVI\ WKH DSSURSULDWH ' 7,6(
Ĥ xψ x ( x ) = E xψ x ( x )
Ĥ yψ y ( y ) = E yψ y ( y )
Ĥ zψ z ( z ) = E zψ z ( z )

)LUVW WHUP
Ĥ xψ x ( x )ψ y ( y )ψ z ( z ) = ψ y ( y )ψ z ( z ) Ĥ xψ x ( x ) = ψ y ( y )ψ z ( z ) E xψ x ( x )
= E xψ x ( x )ψ y ( y )ψ z ( z )

6DPH IRU Ĥ y DQG Ĥ z Ÿ

Ĥ ψ = E ψ
ª Hˆ x + Hˆ y + Hˆ z º ª¬ψ x ( x )ψ y ( y )ψ z ( z )º¼ = ( E x + E y + E z ) ª¬ψ x ( x )ψ y ( y )ψ z ( z )º¼
¬ ¼

E = Ex + E y + Ez
7KXV LI WKH +DPLOWRQLDQ KDV WKLV VSHFLDO IRUP WKH HLJHQIXQFWLRQV RI WKH '
+DPLOWRQLDQ DUH MXVW SURGXFWV RI WKH HLJHQIXQFWLRQV RI WKH ' +DPLOWRQLDQ DQG
WKH VLWXDWLRQ LV HTXLYDOHQW WR GRLQJ WKUHH VHSDUDWH ' SUREOHPV

&RQFOXVLRQ :DYHIXQFWLRQV PXOWLSO\ DQG WKH HQHUJLHV DGG LI Ĥ LV VHSDUDEOH LQWR

Prepared By Kevin Boudreaux


77 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Separable Systems page 3

Ĥ = Ĥ x + Ĥ y + Ĥ z 

7KH ' SURGXFW VWDWHV DUH QDWXUDOO\ QRUPDOL]HG LI WKH ' ZDYHIXQFWLRQV DUH
QRUPDOL]HG

³³³ψ ( x )ψ ( y )ψ ( z )ψ ( x )ψ ( y )ψ ( z ) dx dy dz =
* * *
x y z x y z

³ψ ( x )ψ ( x ) dx ³ψ ( y )ψ ( y ) dy ³ψ ( z )ψ ( z ) dz = 1
* * *
x x y y z z

 [  [ 
1RWLFH WKDW IRU ψ ZH DUH IUHH WR FKRRVH DQ\ HLJHQIXQFWLRQ ψ nx  RI Ĥ x

WRJHWKHU ZLWK DQ\ HLJHQIXQFWLRQ ψ ny RI Ĥ y DQG DQ\ HLJHQIXQFWLRQ ψ nz  RI Ĥ z 


7KXV ZKLOH LQ ' ZH XVXDOO\ KDG RQH TXDQWXP QXPEHU LQ ' ZH ZLOO KDYH WKUHH
Q[Q\Q]  )XUWKHU IRU WZR SURGXFW VWDWHV WR EH RUWKRJRQDO ZH GR QRW KDYH WR
KDYH DOO WKUHH ' IXQFWLRQV EH GLIIHUHQW ,I DQ\ RQH RI WKH WKUHH '
ZDYHIXQFWLRQV [\ RU ] LV RUWKRJRQDO WR LWV FRXQWHUSDUW WKHQ WKH WZR '
ZDYHIXQFWLRQV DUH DOVR RUWKRJRQDO )RU H[DPSOH FRQVLGHU WKH WZR ZDYHIXQFWLRQV
ψ 111 ( x, y, z ) = ψ x1 ( x )ψ y1 ( y )ψ z1 ( z ) DQG ψ 311 ( x, y, z ) = ψ x 3 ( x )ψ y1 ( y )ψ z1 ( z ) 

³³³ψ 111 ( x, y, z )ψ 311 ( x, y, z ) dx dy dz


*

Ÿ ³³³ψ 1* ( x )ψ 1* ( y )ψ 1* ( z )ψ 3 ( x )ψ 1 ( y )ψ 1 ( z ) dx dy dz
x y z x y z

Ÿ ³ψ 1* ( x )ψ 3 ( x ) dx ³ψ 1* ( y )ψ 1 ( y ) dy ³ψ 1* ( z )ψ 1 ( z ) dz = 0
x x y y z z

 [  [ 
%HFDXVH WKH SURGXFW VWDWHV DUH RUWKRJRQDO DQG QRUPDOL]HG WKH\·UH RUWKRQRUPDO
DQG ZH VXPPDUL]H WKLV E\ ZULWLQJ

³³³ ( ) (
ψ nxnynz * x, y, z ψ mxmymz x, y, z dx dy dz = δ nx ,mxδ ny ,myδ nz ,mz )
([DPSOH ' +DUPRQLF 2VFLOODWRU

/HW·V FRQVLGHU D SDUWLFOH LQ ' VXEMHFW WR D +DUPRQLF SRWHQWLDO LQ [\ DQG ]


)XUWKHU DVVXPH WKH IRUFH FRQVWDQWV LQ HDFK GLUHFWLRQ DUH GLIIHUHQW 7KLV PLJKW
EH WUXH IRU H[DPSOH IRU D SDUWLFOH WUDSSHG LQVLGH D SURWHLQ WKH UHVRUWLQJ IRUFH
IRU PRYLQJ LW LQ WKH [ GLUHFWLRQ ZLOO EH GLIIHUHQW IURP \ RU ] EHFDXVH WKH SURWHLQ
KDV D GLIIHUHQW VKDSH DORQJ [ WKDQ \ RU ] VHH EHORZ ULJKW  0XOWLGLPHQVLRQDO

Prepared By Kevin Boudreaux


78 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Separable Systems page 4

+DUPRQLF SRWHQWLDOV DUH DOVR LPSRUWDQW IRU GHVFULELQJ WKH YLEUDWLRQV RI


SRO\DWRPLF PROHFXOHV )RU H[DPSOH LQ +&2 [ PLJKW FRUUHVSRQG WR WKH &+
VWUHWFK \ WR WKH &2 VWUHWFK DQG ] WR WKH +&2 EHQG

,Q HLWKHU FDVH WKH JHQHUDO SRWHQWLDO LV JLYHQ E\


1 1 1
V ( x, y , z ) = k x x 2 + k y y 2 + k z z 2 = V x ( x ) + V y ( y ) + Vz ( z )
2 2 2
1RZ EHFDXVH WKH SRWHQWLDO LV D
VXP RI DQ [ SRWHQWLDO D \
SRWHQWLDO DQG D ] SRWHQWLDO ZH
FDQ HDVLO\ ZULWH WKH
+DPLOWRQLDQ GRZQ DV D VXP
Ĥ = Ĥ x + Ĥ y + Ĥ z
p̂ x 2 1
Ĥ x = + k x x̂ 2
2m 2
ˆ p̂ y 2 1
Hy = + k y ŷ 2
2m 2
p̂ 2 1
Ĥ z = z + k z zˆ 2
2m 2
ZKHUH HDFK ' +DPLOWRQLDQ GHVFULEHV D SDUWLFOH VXEMHFW WR D +DUPRQLF SRWHQWLDO
ZLWK WKH DSSURSULDWH VSULQJ FRQVWDQW N[ N\ RU N]  %DVHG RQ WKH GLVFXVVLRQ
DERYH ZH FDQ LPPHGLDWHO\ ZULWH GRZQ DOO WKH HLJHQIXQFWLRQV DQG HLJHQYDOXHV
1
α x x2 α y y2 α z z2
§ αz ·
ψ n n n ( x, y , z ) = N x H n (α x x ) e N y H n y (α y y ) e ( )
− − 4 −
1/2 2 1/2 2
¨ ¸ N H α 1/2
z e 2

©π ¹
x y z x z nz z

ª§ 1· º ª§ 1· º ª§ 1· º
Enxn y nz = «¨ n x + ¸ =ωx » + «¨ n y + ¸ =ω y » + «¨ n x + ¸ =ωz »
¬© 2¹ ¼ ¬© 2¹ ¼ ¬© 2¹ ¼
( mk x )
12
kx
αx = ωx = etc.
= m

1RWLFH DJDLQ WKDW IRU WKH ' SUREOHP ZH KDYH  TXDQWXP QXPEHUV DQG WKH
HQHUJ\ DQG ZDYHIXQFWLRQ GHSHQG RQ DOO WKUHH VLPXOWDQHRXVO\

'HJHQHUDFLHV
,Q ' WKHUH DUH D QXPEHU RI LQWHUHVWLQJ WKLQJV WKDW FDQ KDSSHQ WKDW ZH GLGQ·W
VHH LQ ' 2QH H[DPSOH RI WKLV LV WKDW LQ ' LW LV SRVVLEOH IRU WZR GLIIHUHQW

Prepared By Kevin Boudreaux


79 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Separable Systems page 5

HLJHQIXQFWLRQV RI WKH +DPLOWRQLDQ WR KDYH WKH VDPH HQHUJ\ :KHQ WKLV KDSSHQV
WKHVH WZR VWDWHV DUH FDOOHG GHJHQHUDWH ,I WKHUH DUH WKUHH IRXU« VWDWHV ZLWK
WKH VDPH HQHUJ\ WKH\ DUH VDLG WR EH WKUHHIROG IRXUIROG « GHJHQHUDWH 7KLV
QHYHU KDSSHQHG IRU WKH 3DUWLFOH LQ D %R[ RU WKH +DUPRQLF 2VFLOODWRU ,Q IDFW
RQH FDQ VKRZ WKDW IRU ERXQG VWDWHV LQ ' RQH QHYHU KDV GHJHQHUDF\ HYHU\ VWDWH
KDV LWV RZQ HQHUJ\ +RZHYHU LQ ' ZH FDQ ILQG GHJHQHUDF\ YHU\ HDVLO\ )RU
H[DPSOH LI RXU VSULQJ FRQVWDQWV DUH DOO WKH VDPH
kx = k y = kz ≡ k Ÿ ωx = ω y = ωz ≡ ω
§ 3·
Ÿ Enxn y nz = =ω ¨ n x + n y + n z + ¸
© 2¹
7KH JURXQG VWDWH KDV DQ HQHUJ\ E000 = 23 =ω  7KHUH LV RQO\ RQH ZD\ , FDQ JHW WKLV
HQHUJ\ n x = 0, n y = 0, nz = 0  VR LW LV QRW GHJHQHUDWH +RZHYHU WKHUH DUH WKUHH
ZD\V , FDQ JHW WKH ILUVW H[FLWHG VWDWH HQHUJ\ 5
2
=ω  n x = 1, ny = 0, nz = 0 
n x = 0, ny = 1, nz = 0 RU n x = 0, ny = 0, nz = 1  6R ZH ILQG WKDW LI ZH FKRRVH DOO WKH
VSULQJ FRQVWDQWV WR EH HTXDO
3=ω
E000 = nondegenerate level
2
5=ω
E100 = E010 = E001 = 3-fold degenerate level
2
HWF
:H ZLOO W\SLFDOO\ GUDZ WKLV ZLWK D SLFWXUH OLNH

(
«

( ( ( ( ( (


IROG GHJHQHUDWH

( ( ( IROG GHJHQHUDWH

( 1RQGHJHQHUDWH

1RWH WKH ZDYHIXQFWLRQV DUH GLVWLQFW


ψ 100 ( x, y, z ) ≠ ψ 010 ( x, y , z ) ≠ ψ 001 ( x, y, z )
7KLV OHDGV WR DQ LQWHUHVWLQJ HIIHFW 6XSSRVH ZH PDNH XS D ZDYHIXQFWLRQ WKDW LV
D VXP RI WZR GHJHQHUDWH VWDWHV VD\
ψ ( x, y, z ) = a ψ 010 ( x, y, z ) + b ψ 001 ( x, y, z )

Prepared By Kevin Boudreaux


80 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Separable Systems page 6

ZKHUH D DQG E DUH FRQVWDQWV 7KHQ LW WXUQV RXW WKDW ψ LV DOVR DQ HLJHQVWDWH RI
WKH +DPLOWRQLDQ 7R VHH WKLV
Ĥψ = Ĥ ( a ψ 010 + b ψ 001 ) = a Hˆ ψ 010 + b Hˆψ 001
=a 5=ω
2 ψ 010 + b 5=ω
2 ψ 001 = 5=2ω ( a ψ 010 + b ψ 001 )
= 5=2ω ψ
7KLV LOOXVWUDWHV WKH LPSRUWDQW SRLQW WKDW DQ\ VXP RI GHJHQHUDWH HLJHQVWDWHV LV
DOVR DQ HLJHQVWDWH RI WKH +DPLOWRQLDQ ZLWK WKH VDPH HLJHQYDOXH

6HWWLQJ WKH VSULQJ FRQVWDQWV HTXDO DPRXQWV WR DVVXPLQJ WKH ZHOO LV V\PPHWULF
ZLWK UHVSHFW WR [ \ DQG ] ,I WKH V\PPHWU\ LV ´EURNHQµ LH k x ≠ k y ≠ k z WKHQ
WKH   DQG  VWDWHV ZLOO EHFRPH QRQGHJHQHUDWH ,Q WKLV FDVH ZH ZLOO
XVXDOO\ VD\ WKDW WKH GHJHQHUDF\ KDV EHHQ ´OLIWHGµ RU WKDW WKH GHJHQHUDWH VWDWHV
KDYH EHHQ VSOLW 7KH ODWWHU ODQJXDJH FRPHV IURP WKH SLFWRULDO YLHZ LI WKH VSULQJ
FRQVWDQWV DUH RQO\ VOLJKWO\ GLIIHUHQW WKHQ WKH HQHUJ\ OHYHOV PLJKW ORRN OLNH

( (
( (
( (

( ( (

(

+HUH \RX FDQ VHH WKDW WKH Q  OHYHOV DUH DOPRVW GHJHQHUDWH WKH\·YH EHHQ
´VSOLWµ DQG WKH Q  OHYHOV DUH DOPRVW GHJHQHUDWH EXW QRW TXLWH EHFDXVH WKH
IRUFH FRQVWDQWV DUH VOLJKWO\ GLIIHUHQW 1RWLFH WKDW LW LV SRVVLEOH WR EUHDN VRPH
GHJHQHUDFLHV EXW NHHS RWKHUV )RU H[DPSOH  LI ZH FKRRVH
kx = k y ≡ k ≠ kz Ÿ ωx = ω y ≡ ω ≠ ωz
7KHQ WKH HQHUJLHV EHFRPH
§ 1·
Enxn y nz = =ω ( n x + n y + 1) + =ωz ¨ nz + ¸
© 2¹
DQG ZH ILQG

Prepared By Kevin Boudreaux


81 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Separable Systems page 7

1
E000 = =ω + =ωz nondegenerate level
2
1
E100 = E010 = 2=ω + =ωz 2-fold degenerate level
2
3
E001 = =ω + =ωz nondegenerate level
2

Prepared By Kevin Boudreaux


82 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Separable Systems page 8

([DPSOH  3DUWLFOH LQ ' ER[

F V ( x, y, z ) = Vx ( x ) +V y ( y ) +Vz ( z )
Vx ( x ) = 0 0≤ x≤a
Vy ( y ) = 0 0≤ y≤b
Vz ( z ) = 0 0≤ z≤c

D \ Vx ( x ) ,V y ( y ) ,Vz ( z ) = ∞ otherwise
E
[
,QVLGH WKH ER[ 2XWVLGH WKH ER[
=2 § ∂ 2 ∂2 ∂2 ·
− + + ψ ( x, y, z ) = Eψ ( x, y, z ) ψ (x, y, z )= 0
2m ¨© ∂x 2 ∂y 2 ∂z 2 ¸¹

:H FDQ DJDLQ DSSO\ VHSDUDWLRQ RI YDULDEOHV VLQFH Ĥ = Ĥ x + Ĥ y + Ĥ z ZKHUH


Ĥ x  Ĥ y  Ĥ z DUH HDFK ' SDUWLFOH LQ D ER[ +DPLOWRQLDQV 6R WKH VROXWLRQV WR WKH
' HTXDWLRQ DUH SURGXFWV RI WKH ' VROXWLRQV

Ÿ ψ ( x , y , z ) = ψ n ( x )ψ n
x y
( y )ψ n ( z )
z

ZKHUH IURP WKH ' SUREOHP ZH KDYH WKH VROXWLRQV


1
§2· §n πx·
2 = 2 n x2
ψ nx ( x ) = ¨ ¸ sin ¨ x ¸ n x = 1, 2,3,... En x =
©a¹ © a ¹ 8m a 2
1
§n πy·
2
§ 2 ·2 =2 n y
ψ n y ( y ) = ¨ ¸ sin ¨ y ¸ n y = 1, 2,3,... En y =
©b¹ © b ¹ 8m b2
1
§2· §n πz·
2 = 2 nz2
ψ nz ( z ) = ¨ ¸ sin ¨ z ¸ nz = 1, 2,3,... En z =
©c¹ © c ¹ 8m c 2
:KHUH WKH HQHUJ\ LV QRZ D IXQFWLRQ RI DOO WKUHH 4XDQWXP QXPEHUV

Prepared By Kevin Boudreaux


83 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Separable Systems page 9

= 2 § n x2 n y nz2 ·
2

En x n y n z = E n x + E n y + E nz = ¨ + + ¸
8m ¨© a 2 b 2 c 2 ¸¹

'HJHQHUDFLHV

'HJHQHUDFLHV RFFXU LQ D VLPLODU IDVKLRQ IRU WKH 3L% HJ LI D E F LQ RXU '
ER[

h2
2 ( x
Ÿ En x n y n z = n 2 + n 2y + nz2 )
8ma

3h 2
E111 = LV QRQGHJHQHUDWH
8ma 2

1
§ 8 ·2 §π x · §π y · §π z ·
ψ 111 ( x, y, z ) = ψ 1 ( x )ψ 1 ( y )ψ 1 ( z ) = ¨ 3 ¸ sin ¨ ¸ sin ¨ ¸ sin ¨ ¸
©a ¹ © a ¹ © a ¹ © a ¹

%XW«

h2
2 (
E211 = E121 = E112 = 22 +12 + 12 ) IROG GHJHQHUDF\
8ma

1RWH WKH ZDYHIXQFWLRQV DUH DJDLQ GLVWLQFW

ψ 211 (x, y, z )≠ ψ 121 (x, y, z )≠ ψ 112 (x, y, z )


(

( 1RQGHJHQHUDWH
( ( (
IROG GHJHQHUDWH
( ( (
1RQGHJHQHUDWH
(

Prepared By Kevin Boudreaux


84 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Separable Systems page 10

,I WKH V\PPHWU\ LV ´EURNHQµ LH a≠ b≠c WKHQ WKH GHJHQHUDF\ LV OLIWHG

h2 § 1 1 4 · h2 § 1 4 1·
E112 = ¨ 2 + 2 + 2 ¸ ≠ E121 = ¨ 2+ 2+ 2¸
8m © a b c ¹ 8m © a b c ¹

6XPPDU\
' ER[
1
§ 8 ·2 § n xπ x · § n yπ y · § n zπ z ·
ψ nxn ynz ( x, y , z ) = ¨ ¸ sin ¨ ¸ sin ¨ ¸ sin ¨ ¸
© abc ¹ © a ¹ © b ¹ © c ¹
h 2 § n x2 n y nz2 ·
2

En x n y n z = ¨ + + ¸ n x = 1, 2,3,... n y = 1, 2,3,... nz = 1, 2,3,...


8m ©¨ a 2 b2 c 2 ¸¹

Prepared By Kevin Boudreaux


85 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Rigid Rotor page 1

5LJLG 5RWDWLRQV
&RQVLGHU WKH URWDWLRQ RI WZR SDUWLFOHV DW D IL[HG GLVWDQFH R IURP RQH DQRWKHU

ω
r1 + r2 ≡ R
P
U m1r1 = m2 r2 FHQWHU RI PDVV &20
&20 U P

7KHVH WZR SDUWLFOHV FRXOG EH DQ HOHFWURQ DQG D SURWRQ LQ ZKLFK FDVH ZH·G EH
ORRNLQJ DW D K\GURJHQ DWRP RU WZR QXFOHL LQ ZKLFK FDVH ZH·G EH ORRNLQJ DW D
GLDWRPLF PROHFXOH &ODVVLFDOO\ HDFK RI WKHVH URWDWLQJ ERGLHV KDV DQ DQJXODU
PRPHQWXP Li = I iω ZKHUH ω LV WKH DQJXODU YHORFLW\ DQG ,L LV WKH PRPHQW RI
LQHUWLD Ii = mri 2 IRU WKH SDUWLFOH 1RWH WKDW LQ WKH &20 WKH WZR ERGLHV PXVW
KDYH WKH VDPH DQJXODU IUHTXHQF\ 7KH FODVVLFDO +DPLOWRQLDQ IRU WKH SDUWLFOHV LV
L12 L2 2 1
= m1r12ω 2 + m2 r22ω 2 = ( m1r12 + m2 r22 ) ω 2
1 1
H= +
2I1 2I 2 2 2 2
,QVWHDG RI WKLQNLQJ RI WKLV DV WZR URWDWLQJ SDUWLFOHV LW ZRXOG EH UHDOO\ QLFH LI ZH
FRXOG WKLQN RI LW DV RQH HIIHFWLYH SDUWLFOH URWDWLQJ DURXQG WKH RULJLQ :H FDQ GR
WKLV LI ZH GHILQH WKH HIIHFWLYH PRPHQW RI LQHUWLD DV
m1 m2
I = m1r12 + m2 r22 = μ r02 μ=
m1 + m2
ZKHUH LQ WKH VHFRQG HTXDOLW\ ZH KDYH QRWHG WKDW WKLV WZR SDUWLFOH V\VWHP
EHKDYHV DV D VLQJOH SDUWLFOH ZLWK D UHGXFHG PDVV μ URWDWLQJ DW D GLVWDQFH R
IURP WKH RULJLQ 7KXV ZH KDYH
]
1 2 L2
H = Iω =
2 2I
ZKHUH LQ WKH VHFRQG HTXDOLW\
ZH KDYH GHILQHG WKH DQJXODU μ
U
PRPHQWXP IRU WKLV HIIHFWLYH θ
SDUWLFOH L = Iω  7KH SUREOHP
LV QRZ FRPSOHWHO\ UHGXFHG WR D
\
ERG\ SUREOHP ZLWK PDVV μ φ

6LPLODUO\ LI ZH KDYH D JURXS RI


REMHFWV WKDW DUH KHOG LQ ULJLG [

Prepared By Kevin Boudreaux


86 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Rigid Rotor page 2

SRVLWLRQV UHODWLYH WR RQH DQRWKHU ² VD\ WKH DWRPV LQ D FU\VWDO ² DQG ZH URWDWH WKH
ZKROH DVVHPEO\ ZLWK DQ DQJXODU YHORFLW\ ω DERXW D JLYHQ D[LV r WKHQ E\ D VLPLODU
PHWKRG ZH FDQ UHGXFH WKH FROOHFWLYH URWDWLRQ RI DOO RI WKH REMHFWV WR WKH
URWDWLRQ RI D VLQJOH ´HIIHFWLYHµ REMHFW ZLWK D PRPHQW RI LQHUWLD ,U ,Q WKLV
PDQQHU ZH FDQ WDON DERXW URWDWLRQV RI D PROHFXOH RU D ERRN RU D SHQFLO ZLWKRXW
KDYLQJ WR WKLQN DERXW WKH PRYHPHQW RI HYHU\ VLQJOH HOHFWURQ DQG TXDUN
LQGLYLGXDOO\

,W LV LPSRUWDQW WR UHDOL]H KRZHYHU WKDW HYHQ IRU D FODVVLFDO V\VWHP URWDWLRQV


DERXW GLIIHUHQW D[HV GR QRW FRPPXWH ZLWK HDFK RWKHU )RU H[DPSOH

5RWDWH Ü 5RWDWH Ü


DERXW [ DERXW \
5[ 5\
]
\
[
5RWDWH Ü 5RWDWH Ü
DERXW \ DERXW [
5\ 5[

+HQFHRQH JHWV GLIIHUHQW DQVZHUV GHSHQGLQJ RQ ZKDW RUGHU WKH URWDWLRQV DUH
SHUIRUPHG LQ *LYHQ RXU H[SHULHQFH ZLWK TXDQWXP PHFKDQLFV ZH PLJKW GHILQH DQ
RSHUDWRU R̂x R̂y WKDW URWDWHV DURXQG [ \  7KHQ ZH ZRXOG ZULWH WKH DERYH
H[SHULPHQW VXFFLQFWO\ DV Rˆ Rˆ ≠ Rˆ Rˆ  7KLV UDWKHU SURIRXQG UHVXOW KDV QRWKLQJ
x y y x

WR GR ZLWK TXDQWXP PHFKDQLFV ² DIWHU DOO WKHUH LV QRWKLQJ TXDQWXP PHFKDQLFDO


DERXW WKH ER[ GUDZQ DERYH ² EXW KDV HYHU\WKLQJ WR GR ZLWK JHRPHWU\ 7KXV ZH
ZLOO ILQG WKDW ZKLOH OLQHDU PRPHQWXP RSHUDWRUV FRPPXWH ZLWK RQH DQRWKHU
pˆ x pˆ y = pˆ y pˆ x WKH VDPH ZLOO QRW EH WUXH IRU DQJXODU PRPHQWD EHFDXVH WKH\ UHODWH
WR URWDWLRQV Lˆ Lˆ ≠ Lˆ Lˆ 
x y y x

&ODVVLFDOO\ DQJXODU PRPHQWXP LV JLYHQ E\ L = r × p  7KLV PHDQV WKH FRUUHVSRQGLQJ


TXDQWXP RSHUDWRU VKRXOG EH Lˆ = rˆ × pˆ  7KLV YHFWRU RSHUDWRU KDV WKUHH
FRPSRQHQWV ZKLFK ZH LGHQWLI\ DV WKH DQJXODU PRPHQWXP RSHUDWRUV DURXQG HDFK
RI WKH WKUHH &DUWHVLDQ D[HV [\ DQG ]

Prepared By Kevin Boudreaux


87 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Rigid Rotor page 3

i j k
Lˆ = rˆ × p̂ = ( x̂, yˆ, zˆ ) × ( p̂ x , p̂ y , pˆ z ) = x̂ ŷ ẑ
p̂ x p̂ y p̂z
= ( yp ˆ ˆ y ) i + ( zp
ˆ ˆ z − zp ˆ ˆ z ) j + ( xp
ˆ ˆ x − xp ˆˆx )k
ˆ ˆ y − yp

∴ Lˆx = yp
ˆ ˆ z − zp
ˆˆ y Lˆ y = zp
ˆˆ x − xp
ˆˆz Lˆ y = xp
ˆ ˆ y − yp
ˆˆx

1RZ LW WXUQV RXW WKDW WKHVH DQJXODU PRPHQWXP RSHUDWRUV GR QRW FRPPXWH )RU
H[DPSOH
+ L̂x L̂y = + ( ŷp̂ z − ẑp̂ y )(ẑp̂ x − x̂p̂ z ) = + ŷp̂ z ẑp̂ x − ẑp̂ y ẑp̂ x − yˆ pˆ z xˆpˆ z + ẑp̂ y x̂p̂ z

− L̂y L̂x = −(ẑp̂ x − x̂p̂ z )( ŷp̂ z − ẑp̂ y ) = −ẑp̂ x ŷp̂ z + ẑp̂ x ẑp̂ y + xˆpˆ z yˆ pˆ z − x̂p̂ z ẑp̂ y

[ ]
Ÿ Lˆ x , Lˆ y = ŷp̂ x [ p̂ z , ẑ ] + x̂p̂ y [ẑ, p̂ z ]

= −i =ŷp̂ x + i=x̂p̂ y = i=L̂z ∴


1RZ ZH FRXOG SURFHHG WR ZRUN RXW WKH UHODWLRQVKLSV IRU ª¬ L̂ y , L̂z º¼ DQG ª¬ L̂z , L̂ x º¼ 
+RZHYHU ZH QRWH WKDW RXU DQVZHU PXVW DOZD\V EH LQYDULDQW WR D F\FOLF
SHUPXWDWLRQ x → y, y → z, z → x  7KLV PXVW EH WKH FDVH EHFDXVH RXU ODEHOLQJ RI
WKH [ \ DQG ] D[HV LV WRWDOO\ DUELWUDU\ LI ZH FKRVH WR UHODEHO RXU D[HV VR WKDW
x → y, y → z, z → x  WKHQ DOO WKH PDWKHPDWLFV ZRXOG ZRUN RXW H[DFWO\ WKH VDPH
ZLWK WKH OHWWHUV VKXIIOHG DURXQG LQ WKH FRUUHVSRQGLQJ PDQQHU 7KH RQO\ WKLQJ ZH
PXVW EH FDUHIXO RI LV WKDW RXU UHODEHOLQJ SUHVHUYHV WKH ULJKWKDQGHGQHVV RU
FKLUDOLW\ RI RXU FRRUGLQDWHV &\FOLF SHUPXWDWLRQV SUHVHUYH WKH KDQGHGQHVV
ZKLOH D VLPSOH LQWHUFKDQJH RI WZR D[HV LH x ↔ y ZLOO UHYHUVH WKH KDQGHGQHVV
RI RXU FRRUGLQDWHV DQG JLYH XV WKH ZURQJ DQVZHU WU\ LW DQG VHH  7KLV F\FOLF
LQYDULDQFH LV YHU\ LPSRUWDQW EHFDXVH LW UHGXFHV WKH ZRUN ZH QHHG WR GR E\ D
IDFWRU RI  EXW ZH PXVW EH FDUHIXO WR DSSO\ LW FRUUHFWO\ ,Q WKH IXWXUH ZH FDQ
WKHUHIRUH VWDWH WKH UHVXOW IRU WKH ] D[LV DQG WKHQ LQIHU WKH UHVXOWV IRU [ DQG \
E\ F\FOLF SHUPXWDWLRQV ,Q WKLV FDVH ZH LQIHU
ª º ˆ
¬ L̂ x , L̂ y ¼ = i=Lz
⎯⎯⎯
x→y
→ ª¬ L̂ y , Lˆ z º¼ = i=Lˆ x
y→z
z→x

Prepared By Kevin Boudreaux


88 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Rigid Rotor page 4

⎯⎯⎯
x→y
→ ª¬ L̂z , Lˆ x º¼ = i=Lˆ y
y→z
z→x

%HFDXVH WKH DQJXODU PRPHQWXP RSHUDWRUV GR QRW FRPPXWH WKH GR QRW VKDUH D
FRPPRQ VHW RI HLJHQIXQFWLRQV 7KXV ZH DUULYH DW WKH LPSRUWDQW FRQFOXVLRQ WKDW
D V\VWHP FDQQRW VLPXOWDQHRXVO\ KDYH ZHOOGHILQHG DQJXODU PRPHQWXP DURXQG
WKH [ \ DQG ] D[HV 7KH EHVW ZH FDQ KRSH IRU LV ZHOOGHILQHG DQJXODU
PRPHQWXP DURXQG  D[LV

5HWXUQLQJ WR WKH SUREOHP RI ULJLG URWDWLRQV ZH DUH LQWHUHVWHG LQ WKH


HLJHQVWDWHV RI WKH +DPLOWRQLDQ
L̂2
= ( Lx 2 + Ly 2 + Lz 2 ) 
1
Ĥ =
2I 2I
,W LV UHODWLYHO\ HDV\ WR VKRZ WKDW ZKLOH WKH [ \ DQG ] DQJXODU PRPHQWXP
RSHUDWRUV GR QRW FRPPXWH ZLWK HDFK RWKHU WKH\ GR FRPPXWH ZLWK L̂2 

ª L , L2 º = ª L , L2 º + ª L , L2 º + ª L , L2 º
¬ z ¼ ¬ z x¼ ¬ z y¼ ¬ z z¼
= ª Lz , L2x º + ª Lz , L2y º = ¬ª Lz ,Lx º¼ Lx + Lx ª¬ Lz ,Lx º¼ + ¬ª Lz ,Ly º¼ Ly + Ly ª¬ Lz ,Ly º¼
¬ ¼ ¬ ¼
( ) ( )
= −i=Ly Lx + Lx −i=Ly + ( i=Lx ) Ly + Ly (i=Lx )
=0
$W WKLV SRLQW ZH PDNH XVH RI F\FOLF SHUPXWDWLRQV WR DVVHUW WKDW ª¬ L̂ y , L̂2 º¼ = 0 DQG
ª 2º
¬ L̂ x , L̂ ¼ = 0 DV ZHOO 6LQFH L̂z FRPPXWHV ZLWK L̂  WKH WZR RSHUDWRUV VKDUH
2

FRPPRQ HLJHQIXQFWLRQV VR ZH FDQ WDON DERXW D SDUWLFOH ZLWK D ZHOOGHILQHG ]


FRPSRQHQW RI DQJXODU PRPHQWXP FDOO LW m DQG D ZHOOGHILQHG WRWDO DQJXODU
PRPHQWXP FDOO LW l  7KXV ZKHQ ZH DUH ORRNLQJ IRU WKH HLJHQIXQFWLRQV RI WKH
ULJLG URWRU +DPLOWRQLDQ ZH DUH ORRNLQJ IRU VWDWHV LQGH[HG E\ WZR TXDQWXP
QXPEHUV l DQG m 7KLV PDNHV VHQVH EHFDXVH ZH VWDUWHG ZLWK D ' V\VWHP WKDW
ZRXOG KDYH KDG WKUHH TXDQWXP QXPEHUV EXW ZH·YH QRZ UHVWULFWHG WKH PRWLRQ WR
WKH VXUIDFH RI D VSKHUH ZKLFK LV WZR GLPHQVLRQDO )RU VXFK D ' V\VWHP ZH
H[SHFW  TXDQWXP QXPEHUV %DVHG RQ WKH DERYH DQDO\VLV ZH ZLOO GHQRWH WKH
DQJXODU PRPHQWXP HLJHQVWDWHV E\ Ylm  ZLWK m DVVRFLDWHG ZLWK WKH HLJHQYDOXH RI
L̂z DQG l DVVRFLDWHG ZLWK L̂2 

:H DUH QRZ OHIW ZLWK WKH IDLUO\ GLIILFXOW SUREOHP RI VROYLQJ IRU WKH HLJHQYDOXHV
(O DQG HLJHQVWDWHV Ylm  RI WKH ULJLG URWRU 1RWLFH WKDW WKH HLJHQYDOXHV RI WKH

Prepared By Kevin Boudreaux


89 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Rigid Rotor page 5

ULJLG URWRU +DPLOWRQLDQ ZLOO RQO\ GHSHQG RQ l EHFDXVH WKH +DPLOWRQLDQ LV


SURSRUWLRQDO WR L̂2  ,Q RUGHU WR IXOO\ VROYH WKLV GLIIHUHQWLDO HTXDWLRQ LW LV PRVW
FRQYHQLHQW WR ZRUN LQ VSKHULFDO SRODU FRRUGLQDWHV ,Q PRVW PDWK WH[WERRNV φ LV
GHILQHG WR EH WKH DQJOH UHODWLYH WR WKH z
]
D[LV ZKLOH WKH YDVW PDMRULW\ RI TXDQWXP
PHFKDQLFV WH[WV XVH θ LQ WKLV FDSDFLW\ :H θ
ZLOO XVH WKH ODWWHU GHILQLWLRQ EXW EH FDUHIXO
WKDW DQ\ HTXDWLRQV WDNHQ IURP RWKHU VRXUFHV
XVH WKLV VDPH FRQYHQWLRQ +HUH DUH VRPH U
XVHIXO UHODWLRQV LQ VSKHULFDO SRODU \
FRRUGLQDWHV
x ≡ r cos φ sin θ
y ≡ r sin φ sin θ
φ
z ≡ r cos θ
1 ∂ ∂ 1 ∂2 1 ∂ ∂ [
∇2 = 2 r 2 + 2 2 + 2 sin θ
r ∂r ∂r r sin θ ∂φ 2
r sin θ ∂θ ∂θ

,Q RUGHU WR PDNH SURJUHVV ZH QHHG WR H[SUHVV WKH DQJXODU PRPHQWXP RSHUDWRUV


LQ VSKHULFDO SRODU FRRUGLQDWHV DV ZHOO 7KLV UHDUUDQJHPHQW WXUQV RXW WR EH D
IDLUO\ WHGLRXV DSSOLFDWLRQ RI WKH FKDLQ UXOH DQG ZH ZLOO PHUHO\ VWDWH WKH UHVXOWV
§ ∂ ∂ ·
L̂x = −i= ¨ − sin φ − cot θ cos φ ¸
© ∂θ ∂φ ¹
§ ∂ ∂ ·
L̂y = −i= ¨ cos φ − cot θ sin φ
© ∂θ ∂φ ¸¹

L̂z = −i=
∂φ

ª 1 ∂ § ∂ · 1 ∂2 º
L̂2 = L̂2x + L̂2Y + L̂2z Ÿ L̂2 = − = 2 « ¨ sin θ ¸+
¬ sin θ ∂θ © ∂θ ¹ sin 2 θ ∂φ 2 »¼
$V D UHVXOW WKH 6FKU|GLQJHU HTXDWLRQ IRU WKH ULJLG URWRU EHFRPHV

−= 2 ª 1 ∂ § ∂ · 1 ∂2 º m
2» l (
« sin θ ∂θ ¨ sin θ ¸ + Y θ , φ ) = ElYl m (θ , φ )
2I ¬ © ∂θ ¹ sin θ ∂φ ¼
2

Prepared By Kevin Boudreaux


90 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Rigid Rotor page 6

ZKHUH ZH KDYH QRWHG WKDW EHFDXVH WKH URWDWLQJ SDUWLFOH LV FRQVWUDLQHG WR EH RQ


WKH VXUIDFH RI D VSKHUH RI UDGLXV R WKH DVVRFLDWHG ZDYHIXQFWLRQ PXVW EH D
IXQFWLRQ RI θ DQG φ RQO\

$V LW WXUQV RXW WKHUH DUH VRPH YHU\ QLFH RSHUDWRU DOJHEUD WULFNV WKDW DOORZ XV
WR JHW WKH HLJHQYDOXHV (O TXLFNO\ 7KHVH WULFNV DUH DQDORJRXV WR WKH UDLVLQJ DQG
ORZHULQJ RSHUDWRU PDQLSXODWLRQV ZH XVHG WR VROYH WKH +DUPRQLF RVFLOODWRU DQG ZH
ZLOO XVH WKHP WR GHULYH WKH HLJHQYDOXHV LQ WKH QH[W OHFWXUH 7KH UHVXOW LV WKDW
WKH HQHUJLHV RI WKH ULJLG URWRU REH\
=2
El = l ( l +1) l = 0, 1, 2,...
2I
2QH FRQIXVLQJ SRLQW DERXW DQJXODU PRPHQWXP LV WKDW IRU GLIIHUHQW NLQGV RI
DQJXODU PRPHQWXP ZH XVH GLIIHUHQW OHWWHUV 6R ZKLOH O LV WKH OHWWHU FKRVHQ IRU
DQ HOHFWURQ PRYLQJ DURXQG WKH QXFOHXV - LV WKH FKRVHQ OHWWHU IRU WKH URWDWLRQ RI
D GLDWRPLF PROHFXOH
=2
EJ = J ( J +1) J = 0, 1, 2,...
2I
7KXV WKH VSDFLQJ EHWZHHQ WKH HQHUJ\ OHYHOV LQFUHDVHV ZLWK LQFUHDVLQJ J XQOLNH
IRU WKH KDUPRQLF RVFLOODWRU ZKHUH WKH\ ZHUH HTXDOO\ VSDFHG
=2 =2
E J +1 − E J = ª( J +1)( J + 2 ) (
− J J +1)º¼ = ( J + 1)
2I ¬ I
)XUWKHU IRU HDFK J ZH KDYH PXOWLSOH SRVVLEOH YDOXHV RI m DOVR NQRZQ DV K LQ
VRPH FRQWH[WV :
m = 0, ± 1, ± 2,..., ±J
7KXV HDFK HQHUJ\ OHYHO LV J IROG GHJHQHUDWH 3K\VLFDOO\ m UHIOHFWV WKH
FRPSRQHQW RI WKH DQJXODU PRPHQWXP DORQJ WKH ]GLUHFWLRQ )RU IL[HG J WKH
GLIIHUHQW YDOXHV RI m UHIOHFW WKH GLIIHUHQW GLUHFWLRQV WKH DQJXODU PRPHQWXP
YHFWRU FRXOG EH SRLQWLQJ ² IRU ODUJH
SRVLWLYH m WKH DQJXODU PRPHQWXP LV ]
PRVWO\ DORQJ ] LI m LV ]HUR WKH m = +J
DQJXODU PRPHQWXP LV RUWKRJRQDO WR ]
3K\VLFDOO\ ZH NQRZ WKDW WKH HQHUJ\
RI WKH URWDWLRQ GRHVQ·W GHSHQG RQ
\
WKH GLUHFWLRQ 7KLV LV UHIOHFWHG LQ m=0
WKH IDFW WKDW WKH HQHUJ\ GHSHQGV [
RQO\ RQ J ZKLFK PHDVXUHV WKH OHQJWK m = −J
RI WKH YHFWRU QRW LWV GLUHFWLRQ

Prepared By Kevin Boudreaux


91 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Angular Momentum Page 1

$QJXODU 0RPHQWXP
6LQFH L̂2 DQG L̂z FRPPXWH WKH\ VKDUH FRPPRQ HLJHQIXQFWLRQV 7KHVH IXQFWLRQV DUH
H[WUHPHO\ LPSRUWDQW IRU WKH GHVFULSWLRQ RI DQJXODU PRPHQWXP SUREOHPV ² WKH\
GHWHUPLQH WKH DOORZHG YDOXHV RI DQJXODU PRPHQWXP DQG IRU V\VWHPV OLNH WKH 5LJLG
5RWRU WKH HQHUJLHV DYDLODEOH WR WKH V\VWHP 7KH ILUVW WKLQJV ZH ZRXOG OLNH WR NQRZ
DUH WKH HLJHQYDOXHV DVVRFLDWHG ZLWK WKHVH HLJHQIXQFWLRQV :H ZLOO GHQRWH WKH
HLJHQYDOXHV RI L̂2 DQG L̂z E\ α DQG β UHVSHFWLYHO\ VR WKDW
L̂2Y β (θ , φ ) = α Y β (θ , φ )
α αLˆ Y β (θ , φ ) = β Y β (θ , φ )
z α α

)RU EUHYLW\ LQ ZKDW IROORZV ZH ZLOO RPLW WKH GHSHQGHQFH RI WKH HLJHQVWDWHV RQ θ
DQG φ VR WKDW WKH DERYH HTXDWLRQV EHFRPH
Lˆ 2Yαβ = α Yαβ Lˆ zYαβ = β Yαβ
,W LV FRQYHQLHQW WR GHILQH WKH UDLVLQJ DQG ORZHULQJ RSHUDWRUV QRWH WKH VLPLODULW\ WR
WKH +DUPRQLF RVFLOODWRU 
L̂± ≡ L̂ x ± iL̂ y
:KLFK VDWLVI\ WKH FRPPXWDWLRQ UHODWLRQV
ª L̂+ , L̂− º = 2=L̂z ª L̂z , L̂± º = ± =L̂± ª L̂± , L̂2 º = 0
¬ ¼ ¬ ¼ ¬ ¼
7KHVH UHODWLRQV DUH UHODWLYHO\ HDV\ WR SURYH XVLQJ WKH FRPPXWDWLRQ UHODWLRQV ZH·YH
DOUHDG\ GHULYHG
ªˆ ˆ º ˆ ªˆ ˆ º ˆ ªˆ ˆ º ˆ ª ˆ2 ˆ º
¬ L x , L y ¼ = i=Lz ¬ L y , Lz ¼ = i=L x ¬ Lz , L x ¼ = i=Ly ¬ L , Lz ¼ = 0
)RU H[DPSOH
ª L̂z , L̂± º = ª L̂z , L̂ x º ± i ª L̂z , L̂ y º
¬ ¼ ¬ ¼ ¬ ¼
(
= i=L y ± i ( −i=L x ) = ± = L x ± iLy )
= ± =L̂±
7KH UDLVLQJ DQG ORZHULQJ RSHUDWRUV KDYH D SHFXOLDU HIIHFW RQ WKH HLJHQYDOXH RI L̂z 
Lˆ z (Lˆ ±Yαβ ) = ( ª¬ Lˆ z , Lˆ ± º¼ + Lˆ ± Lˆ z )Yαβ = ( ± =Lˆ ± + Lˆ ± β )Yαβ = ( β ± = ) (Lˆ ±Yαβ )
7KXV L̂+ L̂− UDLVHV ORZHUV WKH HLJHQYDOXH RI L̂z E\ =  KHQFH WKH QDPHV 6LQFH
WKH UDLVLQJ DQG ORZHULQJ RSHUDWRUV FRPPXWH ZLWK L̂2 WKH\ GR QRW FKDQJH WKH YDOXH
RI α DQG VR ZH FDQ ZULWH
Lˆ ±Yαβ ∝ Yαβ ±=
DQG VR WKH HLJHQYDOXHV RI L̂z DUH HYHQO\ VSDFHG

:KDW DUH WKH OLPLWV RQ WKLV ODGGHU RI HLJHQYDOXHV" 5HFDOO WKDW IRU WKH KDUPRQLF
RVFLOODWRU ZH IRXQG WKDW WKHUH ZDV D PLQLPXP HLJHQYDOXH DQG WKH HLJHQVWDWHV FRXOG

Prepared By Kevin Boudreaux


92 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Angular Momentum Page 2

EH FUHDWHG E\ VXFFHVVLYH DSSOLFDWLRQV RI WKH UDLVLQJ RSHUDWRU WR WKH ORZHVW VWDWH


7KHUH LV DOVR D PLQLPXP HLJHQYDOXH LQ WKLV FDVH 7R VHH WKLV QRWH WKDW
Lˆ2 + Lˆ2 = Lˆ2 + Lˆ2 ≥ 0 x y x y

7KLV UHVXOW VLPSO\ UHIOHFWV WKH IDFW WKDW LI \RXWDNH DQ\ REVHUYDEOH RSHUDWRU DQG
VTXDUH LW \RXPXVW JHW EDFN D SRVLWLYH QXPEHU 7R JHW D QHJDWLYH YDOXH IRU WKH
DYHUDJH YDOXH RI L̂2x RU L̂2y ZRXOG LPSO\ DQ LPDJLQDU\ HLJHQYDOXH RI L̂ x RU L̂ y  ZKLFK LV
LPSRVVLEOH VLQFH WKHVH RSHUDWRUV DUH +HUPLWLDQ %HVLGHV ZKDW ZRXOG DQ LPDJLQDU\
DQJXODU PRPHQWXP PHDQ" :H QRZ DSSO\ WKH DERYH HTXDWLRQ IRU WKH VSHFLILF
ZDYHIXQFWLRQ Yαβ 

³ ( ) ³ (
0 ≤ Yαβ * L̂2x + L̂2y Yαβ = Yαβ * L̂2 − L̂2z Yαβ )
= ³ Yαβ * (α − β 2 ) Yαβ

=α −β2
+HQFH β 2 ≤ α DQG WKHUHIRUH − α ≤ β ≤ α  :KLFK PHDQV WKDW WKHUH DUH ERWK
PD[LPXP DQG PLQLPXP YDOXHV WKDW β FDQ WDNH RQ IRU D JLYHQ α ,I ZH GHQRWH WKHVH
YDOXHV E\ βPD[ DQG βPLQ UHVSHFWLYHO\ WKHQ LW LV FOHDU WKDW
Lˆ +Yαβmax = 0 Lˆ −Yαβmin = 0 
:H FDQ WKHQ XVH WKLV NQRZOHGJH DQG VRPH DOJHEUD WULFNV WULFN WR GHWHUPLQH WKH
UHODWLRQVKLS EHWZHHQ α DQG βPD[ RU βPLQ  )LUVW QRWH WKDW
Ÿ Lˆ − Lˆ +Yαβ max = 0 Lˆ + Lˆ −Yαβ min = 0
:H FDQ H[SDQG WKLV H[SOLFLWO\ LQ WHUPV RI L̂ x DQG L̂ x 
( )
Ÿ Lˆ2x + Lˆ2y − i( Lˆ y Lˆ x − Lˆ x Lˆ y ) Yαβ max = 0 (
Lˆ2x + Lˆ2y + i( Lˆ y Lˆ x − Lˆ x Lˆ y ) Yαβ min = 0 )
+RZHYHU WKLV LV QRW WKH PRVW FRQYHQLHQW IRUP IRU WKH RSHUDWRUV EHFDXVH ZH GRQ·W
NQRZ ZKDW L̂ x RU L̂ y JLYHV ZKHQ DFWLQJ RQ Yαβ  +RZHYHU ZH FDQ UHZULWH WKH VDPH
H[SUHVVLRQ LQ WHUPV RI L̂2 DQG L̂z 
(
Lˆ2x + Lˆ2y ± i( Lˆ y Lˆ x − Lˆ x Lˆ y ) )
L̂2 − L̂2z −i=Lˆ z
6R WKHQ ZH KDYH
( )
Ÿ L̂2 − L̂2z − =L̂z Yαβ max = 0 ( L̂ − L̂ + =L̂ ) Yαβ = 0
2 2
z z
min

Ÿ (α − β 2
max − =β max = 0 ) ( α − β + =β ) = 0
2
min min

Ÿ α = β max ( β max + =) = β min ( β min − =)


Ÿ β max = − β min ≡ =l

Prepared By Kevin Boudreaux


93 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Angular Momentum Page 3

ZKHUH LQ WKH ODVW OLQH ZH KDYH VLPSO\ GHILQHG D QHZ YDULDEOH l WKDW LV GLPHQVLRQOHVV
QRWLFH WKDW = KDV WKH XQLWV RI DQJXODU PRPHQWXP  6R FRPELQLQJ WKHVH PLQLPXP
DQG PD[LPXP YDOXHV ZH KDYH WKDW − =l ≤ β ≤ =l  )XUWKHU VLQFH ZH FDQ JHW IURP WKH
ORZHVW WR WKH KLJKHVW HLJHQYDOXH LQ LQFUHPHQWV RI = E\ VXFFHVVLYH DSSOLFDWLRQV RI
WKH UDLVLQJ RSHUDWRU LW LV FOHDU WKDW WKH GLIIHUHQFH EHWZHHQ WKH KLJKHVW DQG
ORZHVW YDOXHV > =j − ( −=j ) = 2=l @ PXVW EH DQ LQWHJHU PXOWLSOH RI =  7KXV l LWVHOI PXVW
HLWKHU EH DQ LQWHJHU RU D KDOILQWHJHU

3XWWLQJ DOO WKHVH IDFWV WRJHWKHU ZH FRQFOXGH 'HILQH m ≡ β / = 

Lˆ 2Ylm = =2l(l + 1)Ylm l = 0, 12 ,1, 23 ,2...


DQG
Lˆ zYlm = m=Ylm m = −l, −l + 1...l − 1, l

ZKHUH ZH KDYH UHSODFHG α ZLWK l DQG β ZLWK m VR WKDW Yαβ EHFRPHV Ylm  $OVR LQ WKH
ILUVW HTXDWLRQ ZH KDYH QRWHG WKDW 0 ≤ L̂2 = = 2l ( l + 1) LPSOLHV l ≥ 0  7KHVH DUH WKH
IXQGDPHQWDO HLJHQYDOXH HTXDWLRQV IRU DOO IRUPV RI DQJXODU PRPHQWXP

1RWLFH WKDW WKHUH LV D GLIIHUHQFH KHUH IURP ZKDW ZH VDZ IRU WKH ULJLG URWRU
7KHUH ZH KDG
=2
EJ = J ( J +1) J = 0, 1, 2,...
2I
ZKHUH DV D UHPLQGHU WKH TXDQWXP QXPEHU J IRU WKH ULJLG URWRU LV HTXLYDOHQW WR WKH
TXDQWXP QXPEHU l GHILQH DERYH +HUH WKH GHSHQGHQFH RI WKH HQHUJ\ RQ J ²
E ∝ J ( J + 1)  LV WKH VDPH DV ZH IRXQG LQ RXU GHULYDWLRQ IRU L̂2  7KH IDFWRU RI 1 / 2I
VLPSO\ DULVHV IURP WKH IDFW WKDW WKH ULJLG URWRU +DPLOWRQLDQ LV L̂2 / 2I UDWKHU WKDQ
L̂2  7KH UHDO GLIIHUHQFH LV WKDW KDOI LQWHJHU YDOXHV RI J GR QRW DSSHDU IRU WKH
ULJLG URWRU $W ILUVW \RXPLJKW WKLQN WKLV PHDQV ZH PDGH D PLVWDNH LQ RXU
GHULYDWLRQ DERYH DQG WKDW l VKRXOG RQO\ EH DQ LQWHJHU DQG QRW D KDOI LQWHJHU
+RZHYHU WKHUH LV QR HUURU 7KH GLIIHUHQFH DULVHV EHFDXVH RXU GHULYDWLRQ DERYH LV
YDOLG IRU DQ\ NLQG RI DQJXODU PRPHQWXP 7KXV ZKLOH FHUWDLQ YDOXHV RI l PD\ QRW
DSSHDU IRU FHUWDLQ W\SHV RI DQJXODU PRPHQWXP HJ WKH\ GRQ·W RFFXU IRU WKH ULJLG
URWRU ZH ZLOO VHH ODWHU RQ WKDW WKH\ FDQ DSSHDU IRU RWKHU W\SHV RI DQJXODU
PRPHQWXP 0RVW QRWDEO\ HOHFWURQV KDYH DQ LQWULQVLF VSLQ DQJXODU PRPHQWXP ZLWK
l = 12  7KXV ZKLOH LQGLYLGXDO V\VWHPV PD\ KDYH DGGLWLRQDO UHVWULFWLRQV RQ WKH

Prepared By Kevin Boudreaux


94 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Angular Momentum Page 4

DOORZHG YDOXHV RI l DQJXODU PRPHQWXP VWDWHV DOZD\V REH\ WKH DERYH HLJHQYDOXH
UHODWLRQV

Prepared By Kevin Boudreaux


95 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Spherical Harmonics page 1

$1*8/$5 020(1780
1RZ WKDW ZH KDYH REWDLQHG WKH JHQHUDO HLJHQYDOXH UHODWLRQV IRU DQJXODU
PRPHQWXP GLUHFWO\ IURP WKH RSHUDWRUV ZH ZDQW WR OHDUQ DERXW WKH DVVRFLDWHG
ZDYH IXQFWLRQV 5HWXUQLQJ WR VSKHULFDO SRODU FRRUGLQDWHV ZH UHFDOO WKDW WKH
DQJXODU PRPHQWXP RSHUDWRUV DUH JLYHQ E\
§ ∂ ∂ ·
L̂x = −i= ¨ − sin φ − cot θ cos φ ¸
© ∂θ ∂φ ¹
§ ∂ ∂ ·
L̂y = −i= ¨ cos φ − cot θ sin φ
© ∂θ ∂φ ¸¹

L̂z = −i=
∂φ

2 ª 1 ∂ § ∂ · 1 ∂2 º
Lˆ2 = Lˆ2x + Lˆ2Y + Lˆ2z ˆ
Ÿ L = −= «
2
¨ sin θ ¸+
¬ sin θ ∂θ © ∂θ ¹ sin 2 θ ∂φ 2 »¼
,Q WHUPV RI WKHVH RXU RULJLQDO 6FKU|GLQJHU (TXDWLRQ IRU ULJLG URWDWLRQV ZDV

ˆ L̂2 m
HYl = Yl = ElYl m
m

2I
−= ª 1 ∂ §
2
∂ · 1 ∂2 º m
2» l (
Ÿ « ¨ sin θ ¸+ 2 Y θ , φ ) = ElYl m (θ , φ )
2I ¬ sin θ ∂θ © ∂θ ¹ sin θ ∂φ ¼
2
ZKHUH l ZDV WKH TXDQWXP QXPEHU IRU L̂ DQG m ZDV WKH TXDQWXP QXPEHU IRU L̂z 
2
7DNLQJ ZKDW ZH OHDUQHG LQ WKH ODVW VHFWLRQ DERXW WKH HLJHQYDOXHV RI L̂ DQG L̂z
ZH FDQ VD\ WKDW DW PRVW ZH FDQ KDYH
l = 0, 12 ,1, 23 , 2,... m = −l, −l + 1,..., l
:H ZLOO VHH WKDW WKHUH LV DQ DGGLWLRQDO UHVWULFWLRQ RQ WKH SRVVLEOH YDOXHV RI l LQ
WKH SUHVHQW FDVH EXW WKHVH DUH WKH SRVVLEOH YDOXHV IRU WKH TXDQWXP QXPEHUV ,Q
WHUPV RI WKH TXDQWXP QXPEHUV ZH KDYH WKH HLJHQYDOXH UHODWLRQV
Lˆ2Yl m = L2Yl m = = 2l ( l +1) Yl m
L̂zYl m = =mYl m
1RZ WKH IXQFWLRQV Ylm  WKDW VDWLVI\ WKHVH UHODWLRQV IRU ULJLG URWDWLRQV DUH
FDOOHG 6SKHULFDO +DUPRQLFV ,W LV SRVVLEOH WR GHULYH WKH VSKHULFDO KDUPRQLFV E\
VROYLQJ WKH ' GLIIHUHQWLDO HTXDWLRQ DERYH 0F4XDUULH JRHV WKURXJK D IDLUO\
FRPSOHWH GHULYDWLRQ DQG ZH RXWOLQH WKDW VROXWLRQ LQ WKH DSSHQGL[ WR WKHVH QRWHV
EHORZ  7KH UHVXOW LV WKDW

Prepared By Kevin Boudreaux


96 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Spherical Harmonics page 2

Yl m (θ , φ ) = Alm Pl ( cos θ ) eimφ


m

Pl ( x ) LV DQ DVVRFLDWHG /HJHQGUH
m
ZKHUH Alm LV D QRUPDOL]DWLRQ FRQVWDQW DQG
3RO\QRPLDO 7KH ILUVW IHZ $VVRFLDWHG /HJHQGUH 3RO\QRPLDOV DUH
P00 ( cos θ ) = 1 P10 ( cos θ ) = cos θ
P11 ( cos θ ) = sin θ P20 ( cos θ ) = 1
2 ( 3cos θ −1)
2

P21 ( cos θ ) = 3cos θ sin θ P22 ( cos θ ) = 3sin 2 θ


7KHUH DUH D QXPEHU RI LPSRUWDQW IHDWXUHV RI WKH 6SKHULFDO +DUPRQLFV ZH FDQ
UHFRJQL]H VLPSO\ E\ LQVSHFWLQJ WKHVH VROXWLRQV
• 7KH ZDYHIXQFWLRQV IDFWRUL]H LQWR D SURGXFW RI D IXQFWLRQ RI θ DQG D
IXQFWLRQ RI φ
Yl m (θ , φ ) ∝ f (θ ) g (φ )
7KLV UHVXOW LV YHU\ UHPLQLVFHQW RI WKH UHVXOW ZH IRXQG IRU VHSDUDEOH
+DPLOWRQLDQV ZKLFK LV VRPHZKDW VXUSULVLQJ EHFDXVH WKH +DPLOWRQLDQ
FHUWDLQO\ GLG QRW DSSHDU DW ILUVW VLJKW WR EH VHSDUDEOH LQWR D +DPLOWRQLDQ
IRU θ DQG D +DPLOWRQLDQ IRU φ
−= 2 ª 1 ∂ § ∂ · 1 ∂2 º
Ĥ = « sin θ ∂θ ¨ sin θ ¸ + 2»
≠ Ĥ θ + Ĥ φ
2I ¬ © ∂θ ¹ sin 2
θ ∂φ ¼
+RZHYHU SK\VLFDOO\ LW PDNHV VRPH VHQVH IRU WKH PRWLRQ DORQJ θ DQG φ WR
VHSDUDWH ZH KDYHQ·W DSSOLHG DQ\ SRWHQWLDO WKDW OLQNV WKHP WRJHWKHU VR
SDUWLFOHV VKRXOG EH IUHH WR PRYH DORQJ θ DQG φ LQGHSHQGHQWO\ MXVW DV
SDUWLFOHV LQ D VHSDUDEOH SRWHQWLDO LQ x DQG y FDQ PRYH LQGHSHQGHQWO\ DORQJ
WKRVH D[HV 7KH θ  φ FURVV WHUPV DERYH UHIOHFW WKH FXUYDWXUH RI WKH '
VXUIDFH WKH SDUWLFOHV DUH PRYLQJ RQ
• ,W LV HDV\ WR YHULI\ WKDW WKHVH IXQFWLRQV DUH HLJHQVWDWHV RI L̂z 
∂ m ∂
L̂zYl m = −i= Yl = −i= Alm Pl ( cos θ ) eimφ = =mAlm Pl ( cos θ ) eimφ = =mYl m
m m

∂φ ∂φ
2 m
7KH\ DUH DOVR HLJHQIXQFWLRQV RI L̂  DV FDQ EH SURYHQ IRU DQ\ JLYHQ Yl
DIWHU VRPH DOJHEUD E\ FRPSXWLQJ Lˆ Y DQG YHULI\LQJ WKDW WKH UHVXOW LV
2 m
l

MXVW = l ( l +1) Yl 
2 m

• :H FDQ QRZ VHH ZK\ KDOILQWHJHU YDOXHV RI l DUH QRW DOORZHG KHUH 5HFDOO
WKDW φ LV WKH DQJOH LQ WKH [\ SODQH DQG LW YDULHV IURP  WR π :KDW
VKRXOG KDSSHQ WR Yl
m
(θ , φ ) ZKHQ φ → φ + 2π " 2I FRXUVH WKH YDOXH RI WKH
ZDYHIXQFWLRQ VKRXOG QRW FKDQJH EHFDXVH E\ LQFUHPHQWLQJ φ E\ π ZH·YH

Prepared By Kevin Boudreaux


97 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Spherical Harmonics page 3

MXVW PRYHG WKH SDUWLFOH DURXQG LQ D IXOO FLUFOH 7KXV IRU WKH ZDYH
IXQFWLRQ WR EH VLQJOHYDOXHG ZH PXVW KDYH
Yl m (θ , φ ) = Yl m (θ , φ + 2π )
Ÿ Alm Pl m ( cos θ ) eimφ = Alm Pl m ( cos θ ) eim(φ +2π )
Ÿ eimφ = eim(φ +2π )
Ÿ eimφ = eimφ eim 2π
Ÿ 1 = eim 2π
Ÿ m = an integer
WKXV m PXVW EH DQ LQWHJHU QR PDWWHU ZKDW YDOXH RI l ZH FKRRVH +RZHYHU
VLQFH WKH PLQLPDO YDOXH IRU m LV l l PXVW DOVR EH DQ LQWHJHU 7KLV
FRQWLQXLW\ DUJXPHQW LV WKH UHDVRQ ZK\ KDOI LQWHJHU YDOXHV RI l DUH QRW
DOORZHG IRU ULJLG URWDWLRQV
• 1RWH WKDW WKHUH DUH D IHZ LQWHUHVWLQJ DOJHEUDLF IHDWXUHV RI WKH VSKHULFDO
KDUPRQLFV  WKH φ SDUW RI WKH ZDYHIXQFWLRQ GRHV QRW GHSHQG RQ l  7KH
lWK RUGHU /HJHQGUH SRO\QRPLDO DOZD\V LQYROYHV VXPV RI SURGXFWV RI VLQHV
DQG FRVLQHV VXFK WKDW WKH VXP RI WKH VLQH DQG FRVLQH SRZHUV LV OHVV WKDQ
RU HTXDO WR l  IRU m ≠ 0 WKH VSKHULFDO +DUPRQLFV DUH FRPSOH[ DQG
Ylm * = Yl−m  7KXV RQH FDQ REWDLQ WZR UHDO IXQFWLRQV IURP HDFK ±m SDLU YLD
Rlm = 1
2 (Y
l
m
+ Yl−m ) I lm = 1
i 2 (Y
l
m
− Yl−m )
7KHVH IHDWXUHV DUH KHOSIXO LQ WU\LQJ WR LGHQWLI\ ZKHQ JLYHQ DQ DUELWUDU\
IXQFWLRQ RI WKH DQJOHV ZKLFK VSKHULFDO KDUPRQLFV PLJKW FRQWULEXWH WR WKDW
IXQFWLRQ
• 7\SLFDOO\ WKH VSKHULFDO +DUPRQLFV DUH DVVRFLDWHG ZLWK OHWWHUV DV \RX KDYH
VHHQ LQ \RXU SUHYLRXV FKHPLVWU\ FRXUVHV 7KXV l  LV ¶V· l  LV ¶S· l  LV ¶G·
«
• ,Q WKH DEVHQFH RI D SRWHQWLDO DV LV WKH FDVH IRU ULJLG URWDWLRQV WKH
VSKHULFDO +DUPRQLFV DUH 2l+1-IROG GHJHQHUDWH
l m Yl m ' s 2l + 1
  Y00 
    Y1−1 , Y10 , Y11 
      Y2−2 , Y2−1 , Y20 , Y21 , Y22 
        Y −3 , Y −2 , Y −1 , Y 0 , Y 1 , Y 2 , Y 3 
3 3 3 3 3 3 3

Prepared By Kevin Boudreaux


98 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Spherical Harmonics page 4

$V GLVFXVVHG SUHYLRXVO\ ZH VKRXOG H[SHFW WKLV GHJHQHUDF\ WR EH EURNHQ LI


ZH DSSO\ D SRWHQWLDO WKDW LV QRW VSKHULFDOO\ V\PPHWULF ,Q WKH SUHVHQFH RI
D SRWHQWLDO ZH H[SHFW WKHVH OHYHOV WR EH VSOLW
7KXV WR VXPPDUL]H IRU WKH VSKHULFDO +DUPRQLFV ZH KDYH

Yl m (θ , φ ) = Alm Pl ( cos θ ) eimφ


m

Lˆ2Yl m = L2Yl m = = 2l ( l + 1) Yl m l = 0,1, 2,3...


Lˆ Y m = =mY m
z l l m = −l , −l + 1,..., l

$33(1',; 62/9,1* )25 7+( 63+(5,&$/ +$5021,&6


:H QHHG WR VROYH WKH GLIIHUHQWLDO HTXDWLRQ
=2 ª 1 ∂ § ∂ · 1 ∂2 º
− « sin θ ∂θ ¨ sin θ ∂θ ¸ + sin 2 θ ∂φ 2 » Y (θ , φ ) = EY (θ , φ )
2I ¬ © ¹ ¼
ˆ (θ , φ ) = EY (θ , φ ) IRU 5LJLG URWDWLRQV 5HDUUDQJLQJ WKH (TXDWLRQ
7KLV LV HY

ª ∂ § ∂ · 2IE 2 º ∂2
«sin θ ∂θ ¨ sin θ ∂θ ¸ + = 2 sin θ » Y (θ , φ ) = − ∂φ 2 Y (θ , φ )
¬ © ¹ ¼

RQO\ θ RQO\ φ
:H·YH VHSDUDWHG WKH YDULDEOHV MXVW DV LQ WKH ' KDUPRQLF RVFLOODWRU

∴ 7U\ Y (θ , φ ) = Θ (θ ) Φ (φ ) DV D VROXWLRQ

2IE
'HILQH β≡ ( note β ∝ E )
=2

ª ∂ § ∂ · 2 º ∂2
«sin θ ∂θ ¨ sin θ ∂θ ¸ + β sin θ » Θ (θ ) Φ (φ ) = − ∂φ 2 Θ (θ ) Φ (φ )
¬ © ¹ ¼

'LYLGLQJ E\ Θ (θ ) Φ (φ ) DQG VLPSOLI\LQJ

Prepared By Kevin Boudreaux


99 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Spherical Harmonics page 5

sin θ ∂ § ∂ · 1 ∂2
θ ¸ ( )
Θ θ + β θ = − Φ (φ )
2
¨ sin sin
Θ (θ ) ∂θ © ∂θ ¹ Φ (φ ) ∂φ 2

RQO\ θ RQO\ φ
6LQFH θ DQG φ DUH LQGHSHQGHQW YDULDEOHV HDFK VLGH RI WKH HTXDWLRQ PXVW EH
HTXDO WR D FRQVWDQW ≡ P

1 ∂2
Ÿ Φ (φ ) = − m 2 ,
Φ (φ ) ∂φ 2

sin θ ∂ § ∂ ·
¨ sin θ ¸ Θ (θ ) + β sin θ = m
2 2
DQG Ÿ ,,
Θ (θ ) ∂θ © ∂θ ¹

)LUVW VROYH IRU Φ (φ ) XVLQJ ,

∂ 2 Φ (φ )
= − m 2 Φ (φ )
∂φ 2

6ROXWLRQV DUH Φ (φ ) = Am e
imφ
and A− m e − imφ

%RXQGDU\ FRQGLWLRQV Ÿ TXDQWL]DWLRQ

Φ (φ + 2π ) = Φ (φ )

Ÿ Am eim(φ +2π ) = Am eimφ and A− m e − im(φ +2π ) = A− m e − imφ

∴ eim( 2π ) = 1 and e − im( 2π ) = 1

7KLV LV RQO\ WUXH LI m = 0, ± 1, ± 2, ± 3,....

P LV WKH ´PDJQHWLFµ TXDQWXP QXPEHU

∴ Φ (φ ) = Am eimφ m = 0, ±1, ± 2, ± 3,....

Prepared By Kevin Boudreaux


100 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Spherical Harmonics page 6


1RUPDOL]DWLRQ ³ Φ ∗ (φ )Φ (φ ) d φ = 1
0
1 imφ
Ÿ Φ (φ ) = e m = 0, ± 1, ± 2, ± 3,...

1RZ OHW·V ORRN DW Θ (θ )  1HHG WR VROYH ,,
sin θ ∂ § ∂ ·
¨ sin θ ¸ Θ (θ ) + β sin θ = m
2 2

Θ (θ ) ∂θ © ∂θ ¹
dx
&KDQJH YDULDEOHV x = cos θ Θ (θ ) = P ( x ) = dθ
− sin θ

6LQFH 0 ≤ θ ≤ π Ÿ −1 ≤ x ≤ +1

$OVR sin 2 θ = 1− cos2 θ = 1− x 2

7KLV HTXDWLRQ WXUQV RXW WR EH /HJHQGUH·V HTXDWLRQ LQ WHUPV RI Θ

sin θ
d

ª dΘ º
«sin θ dθ » +
¬ ¼
(β sin2 θ − m2 ) Θ (θ ) = 0
ZKLFK ZH FDQ UHZULWH

d 2Θ
sin 2 θ
dθ 2 + sin θ cos θ

dθ (
+ β sin 2 θ − m 2 Θ (θ ) = 0 )
/HW x = cosθ DQG Θ(θ ) = P (x )

( )
dΘ dP dx dP 12 dP
= = − sin θ = − 1− x 2
dθ dx dθ dx dx
d 2Θ d ª dΘ º ª dx º d ª
( )
1 2 dP º
= = « − 1− x 2
dθ 2 dθ « dθ » « dθ » dx
¬ ¼ ¬ ¼ ¬ dx »¼

ª º
« 1 2 d 2P »
= − sin θ «
x dP
1 2 dx − 1− x
2
( » )
(
« 1− x 2
¬
) dx 2 »
¼

= −x
dP
dx (
+ 1− x 2
d 2P
dx 2
)
6XEVWLWXWLQJ WKHVH UHVXOWV LQWR /HJHQGUH·V HTXDWLRQ JLYHV

Prepared By Kevin Boudreaux


101 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Spherical Harmonics page 7

( ) d 2P
( ) ( ( ) )
2
2 dP
1− x 2 2 − 2x 1− x dx + β 1− x − m P ( x ) = 0
2 2
dx
'LYLGH E\ (1− x 2 ) WR REWDLQ WKH /HJHQGUH HTXDWLRQ LQ D FRQYHQLHQW IRUP

dP ª m2 º
(
1− x 2 ) d 2P
dx 2
− 2x + «
dx «¬
β − » P(x) = 0
1− x 2 »¼

7KH VROXWLRQV WR WKLV HTXDWLRQ DUH NQRZQ EXW YHU\ PHVV\ 7KH\ DUH FDOOHG WKH
DVVRFLDWHG /HJHQGUH SRO\QRPLDOV Pl
m
(x)  1RWH WKDW WKH\ RQO\ GHSHQG RQ _m_
EHFDXVH WKH HTXDWLRQ GHSHQGV RQ m2
Pl
m
(x )= P (cosθ )
l
m

( ) P (cos θ )= (3cos θ − 1)
1
P00 cos θ = 1 2
0 2

2
P (cos θ )= cos θ
1
0
P (cos θ )= 3cos θ sin θ
2
1

P (cos θ )= sin θ
1
1
P (cos θ )= 3sin θ
2
2 2

HWF

6R Θ (θ )= A P
m
(cosθ )
ª§ 2l + 1· l − m !º 2
Alm = «¨ »
( )
lm l
«© 2 ¸¹ l + m !»
¬ ¼ ( )
ZKHUH Alm LV WKH QRUPDOL]DWLRQ FRQVWDQW
2

( )
π
Alm2 ³ ª Pl cos θ º sin θ dθ = 1
m
Ÿ
0 ¬ ¼

6R QRZ SXWWLQJ LW DOO WRJHWKHU

ψ lm (r0 ,θ , φ )= Yl m (θ , φ )= Θl (θ )Φ m (φ )
m

Yl m
( )
ª§ 2l + 1· l − m !º
θ ,φ = «¨
(
» Pl m cos θ eimφ
) 2

( )
¸
«© 4π ¹ l + m !»
¬ ¼ ( )
7KHVH IXQFWLRQV DUH WKH VSKHULFDO KDUPRQLFV

Prepared By Kevin Boudreaux


102 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Spherical Harmonics page 8

63+(5,&$/ +$5021,&6 6800$5<

( ) () ()
Yl m θ , φ = Θ l θ Φ m φ
m

( )
ª§ 2l + 1· l − m !º 2
Yl θ , φ = «¨
m
( )
» Pl m cos θ eimφ ( )
«© 4π ¹̧ l + m !»
¬ ¼ ( )
l = 0, 1, 2,... m = 0, ± 1, ± 2, ± 3,... ± l

Yl m ·V DUH WKH HLJHQIXQFWLRQV WR Ĥψ = Eψ IRU WKH ULJLG URWRU SUREOHP


1
§ 5 ·2
¸ ( 3cos θ −1)
1
Y00 = Y20 = ¨ 2

( 4π ) © 16π ¹
12

1 1
§ 3 ·2 § 15 · 2
Y10 = ¨ ¸ cos θ Y2±1 = ¨ ¸ sin θ cos θ e ±iφ
© 4π ¹ © 8π ¹
1 1
§ 3 · 2 § 15 · 2
Y11 = ¨ ¸ sin θ eiφ Y2±2 = ¨ ¸ sin θ e
2 ±2iφ

© 8π ¹ © 32π ¹
1
§ 3 ·2
Y1−1 = ¨ ¸ sin θ e −iφ
© 8π ¹

³³ Y (θ ,φ )Y (θ ,φ )sin θ dθ dφ = δ δ mm′
m ′∗
Yl m ·V DUH RUWKRQRUPDO l′ l
m
ll ′

­1 if l = l′ ­1 if m = m′ normalization
.U|QHFNHU GHOWD δ ll′ = ® δ mm′ = ®
¯0 if l ≠ l′ ¯0 if m ≠ m′ orthogonality

1RWH 6ZLWFK l → J FRQYHQWLRQDO IRU PROHFXODU URWDWLRQDO TXDQWXP QXPEHU


HJ l ( l + 1) Ÿ J ( J + 1) J = 0, 1, 2,... 

Prepared By Kevin Boudreaux


103 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture # 20 page 1

SPHERICAL HARMONICS

( )
Yl m  ,  =  l   m 
m
() ()
1

( )
2 l + 1 l  m ! 
Yl  ,  = 
m
( )
 Pl m cos  eim
2

( )


4  l + m ! 
  ( )
l = 0, 1, 2,... m = 0, ± 1, ± 2, ± 3,... ± l

Yl m ’s are the eigenfunctions to Ĥ = E for the rigid rotor problem.


1
5 
1
(3cos   1)
2
Y00 = Y20 = 2

( 4 )
12

16 
1 1
3 2 15  2
Y10 =  cos  Y2±1 =  sin  cos  e± i

4 
8 
1 1
3 2 15  2
Y11 =  sin  ei Y2±2 =  sin 2  e±2i

8 
32 
1
3 2
Y11 =  sin  e i

8 

Yl m ’s are orthonormal:  Y ( , ) Y ( , ) sin  d d =   mm


m m
l l ll

1 if l = l 1 if m = m normalization
Krönecker delta  ll =   mm = 
0 if l  l 0 if m  m orthogonality

Energies: ˆ m = E Y m)
(eigenvalues of HYl lm l

Switch l  J conventional for molecular rotational quantum #

Recall
2IE
 = 2 = l l +1  J J +1

( ) (
J = 0, 1, 2,... )

Prepared By Kevin Boudreaux


104 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture # 20 page 2

2
E
 EJ =
2I
J J +1 ( )
62
J=3 E3 =
I
(
Y30 , Y3±1 , Y3±2 , Y3±3 7x degenerate )

32
J=2 E2 =
I
(
Y20 , Y2±1 , Y2±2 5x degenerate )
2
J=1 E1 =
I
(
Y10 , Y10 2x degenerate )
J=0 E0 = 0 (
Y00 nondegenerate )
Degeneracy of each state (
g J = 2J + 1 )
from m = 0, ± 1, ± 2,..., ± J

Spacing between states  as J 

2 2
E J +1  E J = ( )(
 J +1 J + 2  J J +1  =
2I 
) (
 I J +1 ) ( )
Transitions between rotational states can be observed through
spectroscopy, i.e. through absorption or emission of a photon

+ +
h Absorption
EJ EJ+1
- -

+ +
or Emission h
EJ EJ-1
- -

Prepared By Kevin Boudreaux


105 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture # 20 page 3

Molecules need a permanent dipole for rotational transitions.


Oscillating electric field grabs charges and torques the molecule.

2

Strength of transition I JJ
dx J (ξ ⋅ μ) J
dx

electric field dipole moment


of light of rotor

Leads to selection rule for rotational transitions: J = ±1

Recall angular momentum is quantized (in units of ).


Photon carries one quantum of angular momentum.
Conservation of angular momentum J = ±1
Angular momentum of molecule changes by 1 quantum upon absorption or
emission of a photon.

2
E photon = h
J J +1
photon
J J +1
= Erot = E J +1 EJ =
I
( J + 1) photon
J J +1
=
4
h
2
I
( J + 1)
Define
h
B 2
rotational constant (Hz)
8 I
and
h
B 2
rotational constant (cm-1)
8 cI

J J +1 (
(Hz) = 2B J + 1 ) J J +1
(cm -1 ) = 2B J + 1 ( )

Prepared By Kevin Boudreaux


106 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 page 4


Lecture # 20

J=3 E3 = 12Bhc

E23 = 6Bhc

J=2 E2 = 6Bhc

E12 = 4Bhc
J=1 E1 = 2Bhc
E01 = 2Bhc

This gives rise to a rigid rotor absorption spectrum with evenly spaced lines.
J=0

2B

 01 12  23  34  45  56 

Spacing between transitions is 2B (Hz) or 2B (cm -1 )

( )
 J +1 J +2   J  J +1 = 2B  J + 1 + 1  2B J + 1 = 2B ( )
Use this to get microscopic structure of diatomic molecules directly from
the absorption spectrum!

Get B directly from the separation between lines in the spectrum.

Use its value to determine the bond length r0!

Prepared By Kevin Boudreaux


107 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture # 20 page 5

h m1 m2
2B = I = μ r02 μ=
4 2 cI m1 + m2

1 1
 h 2  h 2
 r0 =  2 -1
 (B in cm ) or r0 =  2  (B in Hz)
 8 cB μ   8 B μ 

Prepared By Kevin Boudreaux


108 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #21 page 1

HYDROGEN ATOM
Schrodinger equation in 3D spherical polar coordinates:

2 1   2 
  
2
 r +
1
2 μ r 2 r  r  r 2 sin   
sin  +
1
( )
  r, , + U r, ,  r, , = E r, ,
  r 2 sin 2   2 
( ) ( ) ( )

Ze2
with Coulomb potential U (r) =
4 0 r
Rewrite as

 2  2  
   r
r  r
()
+ 2 μr 2  U r  E  
r, , + L̂2
r, , = 0 ( ) ( )


function of r only function of , only

r is separable  is separable

Angular momentum: solutions are spherical harmonic wavefunctions

( )
 r, , = R r Yl m  , () ( )
with L̂2Yl m ( , ) =  l ( l + 1)Y ( , )
2
l
m
l = 0,1,2,...

Radial equation for the H atom:

2 d  2 dR r   l l + 1
2
()  ( )
 
2 μr 2 dr 
r +
dr
2 μr 2
+ U r  E R r = 0 () ()


()
Solutions R r are the H atom radial wavefunctions

Simplest case: l = 0 yields solution

32
Z
()
R r = 2 
 a0 
e
 Zr a0
exponential decay away from nucleus

Prepared By Kevin Boudreaux


109 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #21 page 2

with

E =  Z 2 e2 8 0 a0 lowest energy eigenvalue


a0   0 h2 μe2 Bohr radius

General case: solutions are products of (exponential) x (polynomial)

Energy eigenvalues:

Z 2 e2 Z 2 μe4
E= = n = 1,2,3,...
8 0 a0 n2 8 02 h2 n2

Radial eigenfunctions:

12
 
( )
n  l  1 !   2Z
l +3 2

Rnl ()
r = 
 
 na
r le
 Z r na0 2l +1 2 Zr
Ln+ l 
( ) na0

3
2n  n + l !
   0

( )
where L2ln++1l 2Zr na0 are the associated Laguerre functions, the first few of which are:
n=1 l=0 L11 = 1

n=2 l=0 L12 = 2! 2  Zr
a0

l =1 L33 = 3!
 2 2
n=3 l=0 L13 = 3! 3  2Zr + 2Z r 2
a0 9a0


l =1 L34 = 4! 4  2Zr
3a0

l=2 L55 = 5!

Normalization:
( ) ( )
2 
Spherical harmonics  0
d  d sin Yl m* , Yl m , = 1
0

() ()

Radial wavefunctions  0
dr r 2 Rnl* r Rnl r = 1

Prepared By Kevin Boudreaux


110 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #21 page 3

TOTAL HYDROGEN ATOM WAVEFUNCTIONS

( ) () ( )
 nlm r, , = Rnl r Yl m  ,

principle quantum number n = 1,2,3,...


angular momentum quantum number l = 0,1,2,..., n  1
magnetic quantum number m = 0, ±1, ±2,..., ±l

ENERGY depends on n: E = Z 2 e2 8 0 a0 n2

ORBITAL ANGULAR MOMENTUM depends on l: (


L =  l l +1 )
ANGULAR MOMENTUM Z-COMPONENT depends on m: Lz = m

Total H atom wavefunctions are normalized and orthogonal:

( ) ( )
2  
 d  sin  d  r 2 dr  nlm r, ,  n ' l ' m ' r, , =  nn ' ll ' mm '
*
0 0 0

() ( )
since components Rnl r Yl m  , are normalized and orthogonal.

Lowest few total H atom wavefunctions, for n = 1 and 2 (with  = Zr a0 ):

Prepared By Kevin Boudreaux


111 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #21 page 4

3/ 2
1 Z
n=1 l=0 m=0  100 = e  =  1s
  a0

3/ 2
Z
n=2 l=0 m=0  200 =
1
(2   ) e  / 2
=  2s
32  a0

3/ 2
Z
1
l =1 m=0  210 =  e  / 2 cos =  2 p
32  a0
z

3/ 2
Z 1
l =1 m = ±1  21±1 =  e  / 2 sin  e± i
64  a0

or the alternate linear combinations


3/ 2
Z
2p =
1
a  e  / 2 sin  cos =
1
( 21+1
+  211 )
x
32  0
2
3/ 2
Z
2p =
1
 e  / 2 sin  sin =
1
( 21+1
  211 )
y
32  a0
2i

The value of l is denoted by a letter: l = 0,1,2,3...


s,p,d,f orbitals

The value of m is denoted by a letter for l = 1: m = 0, ± 1 linear combinations


pz , px ,p y orbitals
HYDROGEN ATOM ENERGIES

Potential energy of two electrons separated by the Bohr radius:


U = e2 4 0 a0 _ one “atomic unit” (a.u.) of energy.

H atom energies: E = Z 2 e2 8 0 a0 n2 = Z 2 2 n2 a.u.

n En (a.u.)
1 -1/2
2 -1/8
3 -1/18
4 -1/32
5 -1/50

Prepared By Kevin Boudreaux


112 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #21 page 5

H atom energies &transitions E E/hc


n (a.u.) (cm-1)
0 0
4 -1/32 -6,855
3 -1/18-12,187

2 -1/8

1 -1/2 -109,680

H atom emission spectra


Lyman series Balmer Paschen

(A)

(1012 Hz)

DEGENERACIES OF H ATOM ENERGY LEVELS

As n increases, the degeneracy of the level increases.


What is the degeneracy gn of each level as a function of n?
Does this help understand the periodic table?

Prepared By Kevin Boudreaux


113 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #21 page 6

SHAPES AND SYMMETRIES OF THE ORBITALS


S ORBITALS
( ) ( )( )
1 2 r / a0 r / 2a0
 1s =  a03 e  2s = 32 a03 2  r a0 e
l=0 spherically symmetric
n - l -1 = 0 radial nodes n - l -1 = 1
l=0 angular nodes l=0
n -1 = 0 total nodes n -1 = 1

( )
2
Electron probability density given by  r, ,
Probability that a 1s electron lies between r and r + dr of the nucleus:

( ) ( ) ( )
2  1 2r / a0 2

0
d  d sin   1s* r, ,  1s r, , r 2 dr = 4  a03
0
e r dr

P ORBITALS: wavefunctions

Not spherically symmetric: depend on  ,

( ) (r a ) e
1/ 2 r / 2a0
m = 0 case:  210 =  2 p = 32 a03 0
cos
z

 2 p independent of  symmetric about z axis


z

radial nodes n - l -1 = 0 ()
(note difference from 2s: Rnl r depends on l as well as n)
angular nodes l =1
total nodes n -1 = 1

xy nodal plane – zero amplitude at nucleus

( ) (r a ) e
1/ 2 r / 2a0
 2 p = 32 a03 0
sin  cos
m = ±1 case: Linear combinations give
x

= ( 32 a ) ( r a ) e
1/ 2 r / 2a0
2p 3
0 0
sin  sin
y

Equivalent probability distributions

Prepared By Kevin Boudreaux


114 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #21 page 7

H atom radial probability densities

r Rnl ( r )
2 2 a0
()
0.6

0.2
0.5

0.1
2s
0.4

0
1s 0 5 10 15 20 25 30
0.3

0.2 0.2

0.1 0.1
2p

0 0
0 5 10 15 0 5 10 15 20 25 30

0.2

0.1
3s
0
0 5 10 15 20 25 30

0.1
3p
0
0 5 10 15 20 25 30

0.1
3d

0
0 5 10 15 20 25 30

s = r/a0

Prepared By Kevin Boudreaux


115 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #21 page 8

MAGNETIC FIELD EFFECTS

Electron orbital angular momentum (circulating charge) magnetic moment

μ=− e
L
2me

Magnetic field B applied along z axis interacts with μ:


Potential energy

U = −μ•Β = −μ zΒz = eBz Lz


2me

Include in potential part of Hamiltonian operator:

eBz
Ĥ = Ĥ 0 + L̂z
2me

H atom wavefunctions are eigenfunctions of both Ĥ 0 and L̂z operators


eigenfunctions of new Ĥ operator.

Energy eigenvalues are the sums

Z 2 e2 eB
E= 2
+ z m
8 0 a0 n 2me

Energy depends on magnetic quantum number m when a magnetic field is applied.

2p orbitals: m = -1,0,+1 states have different energies


Splitting proportional to applied field Bz.

Prepared By Kevin Boudreaux


116 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Fall 2007 Lecture #21 page 9

Applied
magnetic field
No magnetic
m = +1
field
2p m=0
m = -1

Energy

1s m=0
2p 
1s Emission spectra
One line Three lines

Complex functions  211 and  21+1 are eigenfunctions of L̂z with eigenvalues ±m .

 2 p and  2 p are eigenfunctions of Ĥ 0 but not of L̂z  no longer energy


x y

eigenfunctions once magnetic field is applied.

Prepared By Kevin Boudreaux


117 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 23-Electron Spin page 1

(/(&7521 63,1

([SHULPHQWDO HYLGHQFH IRU HOHFWURQ VSLQ

&RPSWRQ 6FDWWHULQJ   $+ &RPSWRQ VXJJHVWHG WKDW ´WKH HOHFWURQ LV


SUREDEO\ WKH XOWLPDWH PDJQHWLF SDUWLFOHµ

6WHUQ*HUODFK ([SHULPHQW   3DVVHG D EHDP RI VLOYHU DWRPV GV


WKURXJK DQ LQKRPRJHQHRXV PDJQHWLF ILHOG DQG REVHUYHG WKDW WKH\ VSOLW LQWR WZR
EHDPV RI VSDFH TXDQWL]HG FRPSRQHQWV

8KOHQEHFN DQG *RXGVPLW  VKRZHG WKDW WKHVH ZHUH WZR DQJXODU
PRPHQWXP VWDWHV ² WKH HOHFWURQ KDV LQWULQVLF DQJXODU PRPHQWXP ² 63,1
DQJXODU PRPHQWXP

3DXOL ([FOXVLRQ 3ULQFLSOH   QR PRUH WKDQ  HOHFWURQV SHU RUELWDO RU
QR WZR HOHFWURQV ZLWK DOO WKH VDPH TXDQWXP QXPEHUV $GGLWLRQDO TXDQWXP
QXPEHU QRZ FDOOHG PV ZDV SRVWXODWHG

3RVWXODWH  $OO HOHFWURQLF ZDYHIXQFWLRQV PXVW EH


DQWLV\PPHWULF XQGHU WKH H[FKDQJH RI DQ\ WZR HOHFWURQV

7KHRUHWLFDO -XVWLILFDWLRQ
'LUDF  GHYHORSHG UHODWLYLVWLF TXDQWXP WKHRU\ GHULYHG HOHFWURQ
VSLQ DQJXODU PRPHQWXP

2UELWDO $QJXODU 0RPHQWXP


/ = RUELWDO DQJXODU PRPHQWXP
/ = = O (O +  )
O = RUELWDO DQJXODU PRPHQWXP TXDQWXP QXPEHU
O ≤ Q −
/] = P =
P =  ± ±! ±O

Prepared By Kevin Boudreaux


118 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 23-Electron Spin page 2

6SLQ $QJXODU 0RPHQWXP


6 ≡ VSLQ DQJXODU PRPHQWXP
6 = = V (V +  ) = =  
V = VSLQ DQJXODU PRPHQWXP TXDQWXP QXPEHU
V = 
6] = PV =
PV = ±  
'HILQH VSLQ DQJXODU PRPHQWXP RSHUDWRUV DQDORJRXV WR RUELWDO DQJXODU PRPHQWXP
RSHUDWRUV

L2Yl m (θ , φ ) = l ( l + 1) = 2Yl m (θ , φ ) l = 0,1, 2,...n for H atom


LzYl m (θ , φ ) = m=Yl m (θ , φ ) m = 0, ±1, ±2,... ± n for H atom

1
Ŝ 2α = s ( s + 1) = 2α Ŝ 2 β = s ( s + 1) = 2 β s= always
2
1 1
Ŝ zα = ms =α msα = Ŝ z β = ms =β msβ = −
2 2

6SLQ HLJHQIXQFWLRQV α and β DUH QRW IXQFWLRQV RI VSDWLDO FRRUGLQDWHV VR WKH


HTXDWLRQV DUH VRPHZKDW VLPSOHU

α ≡ "spin up" β ≡ "spin down"

6SLQ HLJHQIXQFWLRQV DUH RUWKRQRUPDO

³ α α dσ = ³ β β d σ = 1 σ ≡ spin variable
* *

³ α β dσ = ³ β α d σ = 0
* *

6SLQ YDULDEOH KDV QR FODVVLFDO DQDORJ 1HYHUWKHOHVV WKH DQJXODU PRPHQWXP RI


WKH HOHFWURQ VSLQ OHDGV WR D PDJQHWLF PRPHQW VLPLODU WR RUELWDO DQJXODU
PRPHQWXP

Prepared By Kevin Boudreaux


119 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 23-Electron Spin page 3

Electron orbital magnetic moment Electron spin magnetic moment


e e
μL = − L μs = − gS
2me 2me
e= e=
μL = − l ( l + 1) ≡ − β 0 l ( l + 1) μs = − g s ( s + 1) = − β 0 g s ( s + 1)
2me 2me
e e= e e=
μLz = − Lz = − m = − β 0m μSz = − gS z = − gms = − β 0 gms ≈ ± β 0
2me 2me 2me 2me
g ≡ "electronic g factor" = 2.002322

7RWDO HOHFWURQLF ZDYHIXQFWLRQ KDV ERWK 63$7,$/ DQG 63,1 SDUWV


(DFK SDUW LV QRUPDOL]HG VR WKH WRWDO ZDYHIXQFWLRQ LV QRUPDOL]HG

Ψ ( r, θ , φ , σ ) = ψ ( r, θ , φ ) α (σ ) or ψ ( r, θ , φ ) β (σ )

HJ IRU + DWRP WKH JURXQG VWDWH WRWDO ZDYHIXQFWLRQV LQ DWRPLF XQLWV DUH

12 12
§ Z3 · § Z3 ·
Ψ = ¨ ¸ e − Zrα Ψ = ¨ ¸ e − Zr β
©π ¹ ©π ¹
1 1
100 100−
2 2

ZKLFK DUH RUWKRJRQDO DQG QRUPDOL]HG 1RWH WKH TXDQWXP QXPEHUV DUH QRZ 
QOPP6

Prepared By Kevin Boudreaux


120 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 24 Pauli Spin Matrices Page 1

Pauli Spin Matrices

It is a bit awkward to picture the wavefunctions for electron spin because –


the electron isn’t spinning in normal 3D space, but in some internal dimension
that is “rolled up” inside the electron. We have invented abstract states “α”
and “β” that represent the two possible orientations of the electron spin,
but because there isn’t a classical analog for spin we can’t draw “α” and “β”
wavefunctions. This situation comes up frequently in chemistry. We will
often deal with molecules that are large and involve many atoms, each of
which has a nucleus and many electrons … and it will simply be impossible for
us to picture the wavefunction describing all the particles at once.
Visualizing one­particle has been hard enough! In these situations, it is most
useful to have an abstract way of manipulating operators and wavefunctions
without looking explicitly at what the wavefunction or operator looks like in
real space. The wonderful tool that we use to do this is called Matrix
Mechanics (as opposed to the wave mechanics we have been using so far).
We will use the simple example of spin to illustrate how matrix mechanics
works.

The basic idea is that we can write any electron spin state as a linear
combination of the two states α and β:
ψ ≡ cαα + cβ β
Note that, for now, we are ignoring the spatial part of the electron
wavefunction (e.g. the angular and radial parts of ψ). You might ask how we
can be sure that every state can be written in this fashion. To assure
yourself that this is true, note that for this state, the probability of finding
2 2
the electron with spin “up” (“down”) is cα ( cβ ). If there was a state that
could not be written in this fashion, there would have to be some other spin
state, γ, so that
ψ ≡ cαα + cβ β + cγ γ
2
In this case, however, there would be a probability cγ of observing the
electron with spin γ – which we know experimentally is impossible, as the
electron only has two observable spin states.

The basic idea of matrix mechanics is then to replace the wavefunction


with a vector:

Prepared By Kevin Boudreaux


121 of 227
MIT Open Courseware Compillation
Source: 1

http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 24 Pauli Spin Matrices Page 2

⎛ cα ⎞
ψ ≡ cαα + cβ β
ψ≡⎜ ⎟ →
⎝ cβ ⎠
Note that this is not a vector in physical (x,y,z) space but just a convenient
way to arrange the coefficients that define ψ. In particular, this is a nice
way to put a wavefunction into a computer, as computers are very adept at
dealing with vectors.

Now, our goal is to translate everything that we might want to do with the
wavefunction ψ into something we can do to the vector ψ . By going through
this step­by­step, we arrive at a few rules

Integrals are replaced with dot products. We note that the overlap
between any two wavefunctions can be written as a modified dot product
between the vectors. For example, if φ ≡ dαα + d β β then:

1 0 0 1

∫ φ ψ dσ = dα cα ∫ α α dσ + dα cβ ∫ α β dσ + dβ cα ∫ β α dσ + dβ cβ ∫ β β dσ
* * * * * * * * *

= dα *cα + d β *cβ
⎛ cα ⎞
(
= dα * )

* ⎜ ⎟
⎝ cβ ⎠

≡ φ† iψ
Where, on the last line, we defined the adjoint of a vector:

⎛ x1 ⎞
⎜ x ⎟ ≡ x1 x2 .
⎝ 2⎠
* *
( )
Thus, complex conjugation of the wavefunction is replaced by taking the
adjoint of a vector. Note that we must take the transpose of the vector
and not just its complex conjugate. Otherwise when we took the overlap we
would have (column) x (column) and the number of rows and columns would
not match.

These rules lead us to natural definitions of normalization and orthogonality


of the wavefunctions in terms of the vectors:
⎛ cα ⎞
( )
2

2
ψ *ψ dσ = 1 → ψ†iψ = cα * cβ * ⎜ ⎟ = cα + cβ = 1
⎝ cβ ⎠
and

Prepared By Kevin Boudreaux


122 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 24 Pauli Spin Matrices Page 3

⎛ cα ⎞
∫ φ ψ dσ = 0
*
→ (
φ iψ = dα * )
d β * ⎜ ⎟ = dα *cα + d β *cβ = 0
⎝ cβ ⎠

Finally, we’d like to be able to act operators on our states in matrix


mechanics, so that we can compute average values, solve eigenvalue
equations, etc. To understand how we should represent operators, we note
that in wave mechanics operators turn a wavefunction into another
wavefunction. For example, the momentum operator takes one wavefunction
and returns a new wavefunction that is the derivative of the original one:

ψ ( x ) ⎯⎯

→ (x)
dx
Thus, in order for operators to have the analogous behavior in matrix
mechanics, operators must turn vectors into vectors. As it turns out this is
the most basic property of a matrix: it turns vectors into vectors. Thus we
have our final rule: Operators are represented by matrices.

As an illustration, the Ŝz operator will be represented by a 2x2 matrix in


spin space:
⎛? ?⎞
Ŝz → Sz ≡ ⎜ ⎟
⎝? ?⎠
We have yet to determine what the elements of this matrix are (hence the
question marks) but we can see that this has all the right properties: vectors
are mapped into vectors:
⎛ ? ? ⎞ ⎛ cα ⎞ ⎛ cα ' ⎞
Ŝzψ = ψ ′ → ⎜ ⎟⎜ ⎟ = ⎜ ⎟
⎝ ? ? ⎠ ⎝ c β ⎠ ⎝ cβ ' ⎠
and average values are mapped into numbers
⎛ ? ? ⎞ ⎛ cα ⎞
∫ψ Ŝ ψ dσ
*
z → ( cα *
)
cβ * ⎜ ⎟⎜ ⎟
⎝ ? ? ⎠ ⎝ cβ ⎠

Now, the one challenge is that we have to determine which matrix belongs to
a given operator – that is, we need to fill in the question marks above. As an
exercise, we will show how this is done by working out the matrix
representations of Ŝ 2 , Ŝz , Ŝx and Ŝy .

We’ll start with Ŝ 2 . We know how Ŝ 2 acts on the α and β wavefunctions:


3 3
Ŝ 2α = � 2α Ŝ 2 β = � 2 β
4 4

Prepared By Kevin Boudreaux


123 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 24 Pauli Spin Matrices Page 4

Now represent Ŝ 2 as a matrix with unknown elements.


⎛c d ⎞
S2 = ⎜ ⎟
⎝e f ⎠
In wave mechanics, operating Ŝ 2 on α gives us an eigenvalue back, because α
is and eigenfunction of Ŝ 2 (with eigenvalue 34 � 2 ). Translating this into matrix

mechanics, when we multiply the matrix S 2 times the vector α , we should

get the same eigenvalue back times α :
3 � 3 �
Ŝ 2α = � 2α → S 2α = � 2α
4 4
In this way, instead of talking about eigenfunctions, in matrix mechanics we

talk about eigenvectors and we say that α is an eigenvector of S 2 .

We can use this information to obtain the unknown constants in the matrix
S2 :
3 � 3 � ⎛c d ⎞⎛1⎞ 3 2 ⎛1⎞
Ŝ 2α = � 2α → S 2α = � 2α ⇒ ⎜ ⎟⎜ ⎟ = 4 � ⎜ ⎟
4 4 ⎝e f ⎠⎝0⎠ ⎝0⎠
⎛ c ⎞ ⎛ 3 �2 ⎞
⇒⎜ ⎟=⎜4 ⎟
⎝e⎠ ⎝ 0 ⎠
so
c = 43 � 2 e=0
Operating on β
3 2 � 3 � ⎛c d ⎞⎛0⎞ 3 2 ⎛0⎞
Ŝ 2 β = �β → S 2α = � 2α ⇒ ⎜ ⎟⎜ ⎟ = 4 � ⎜ ⎟
4 4 ⎝e f ⎠⎝1⎠ ⎝1⎠
⎛d⎞ ⎛ 0 ⎞
⇒⎜ ⎟=⎜3 2⎟
⎝ f ⎠ ⎝4� ⎠
Thus:
d=0 f = 34 � 2
So for S 2 we have
3 ⎛1 0⎞
S2 = �2 ⎜
4 ⎝ 0 1 ⎟⎠
We can derive Sz in a similar manner. We know its eigenstates as well:
Ŝzα = �2 α Ŝz β = − �2 β
So that the matrix must satisfy:

Prepared By Kevin Boudreaux


124 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 24 Pauli Spin Matrices Page 5

� ⎛c d ⎞⎛1⎞ � ⎛1⎞
Ŝzα = α → ⎜e f ⎟⎜0⎟ = 2 ⎜0⎟
2 ⎝ ⎠⎝ ⎠ ⎝ ⎠
⎛c⎞ ⎛ ⎞

⇒⎜ ⎟=⎜2⎟
⎝e⎠ ⎝0⎠
which gives us

c= e=0
2
Analogous operations on β give
� ⎛c d ⎞⎛0⎞ ⎛0⎞
Ŝz β = − β → ⎜ ⎟⎜ ⎟ = − 2 ⎜ ⎟

2 ⎝e f ⎠⎝1⎠ ⎝1⎠
⎛d⎞ ⎛ 0 ⎞
⇒⎜ ⎟=⎜ �⎟
⎝ f ⎠ ⎝− 2 ⎠

d=0 f =−
2
Compiling these results, we arrive at the matrix representation of Sz :
� ⎛1 0 ⎞
Sz = ⎜
2 ⎝ 0 −1 ⎟⎠
Now, we need to obtain S x and S y , which turns out to be a bit more tricky.
We remember from our operator derivation of angular momentum that we
can re­write the S x and S y in terms of raising and lowering operators:
1 1
Sx = ( S + + S- ) Sy = (S+ − S- )
2 2i
where we know that
Ŝ+ β = c+α Ŝ+ α = 0 and Ŝ−α = c− β Ŝ− β = 0
where c+ and c­ are constants to be determined. Therefore for the raising
operator we have
⎛ c d ⎞ ⎛ 0 ⎞ ⎛ c+ ⎞ ⎛c d ⎞⎛1⎞ ⎛0⎞
Ŝ+ β = c+α → ⎜ ⎟⎜ ⎟ = ⎜ ⎟ Ŝ+ α = 0 → ⎜e f ⎟⎜0⎟ = ⎜0⎟
⎝ e f ⎠ ⎝ 1 ⎠ ⎝ 0 ⎠ ⎝ ⎠⎝ ⎠ ⎝ ⎠
⇒ d = c+ f = 0 ⇒c=0 e=0
And for the lowering operator
⎛ c d ⎞ ⎛ 1 ⎞ ⎛ 0 ⎞ ⎛c d ⎞⎛0⎞ ⎛0⎞
Ŝ−α = c− β → ⎜ ⎟⎜ ⎟ = ⎜ ⎟ Ŝ− β = 0 → ⎜e f ⎟⎜1⎟ = ⎜0⎟
⎝ e f ⎠ ⎝ 0 ⎠ ⎝ c− ⎠ ⎝ ⎠⎝ ⎠ ⎝ ⎠
⇒ c = 0 e = c − ⇒d=0 f =0
Thus
⎛0 1⎞ ⎛0 0⎞
S + = c+ ⎜ ⎟ S - = c− ⎜ ⎟
⎝0 0⎠ ⎝1 0⎠

Prepared By Kevin Boudreaux


125 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 24 Pauli Spin Matrices Page 6

Therefore we find for S x and S y


1⎛ 0 c+ ⎞ 1⎛ 0 −ic+ ⎞
Sx = 1
(S+ + S− ) = ⎜ Sy = 1
(S+ − S− ) =
2
2 ⎝ c− 0 ⎟⎠ 2i ⎜
2 ⎝ ic− 0 ⎟⎠
Thus, we just need to determine the constants: c+ and c­. We do this by
using two bits of information. First, we note that S x and S y are observable,
so they must be Hermitian. Thus,
*
⎛ 0 c+ ⎞ ⎛ 0 ⎞ ⎛ ⎛ 0 c+ ⎞ ⎛ 1 ⎞ ⎞
∫ α * Ŝ β dσ = ( ∫ β * Ŝ α dσ )
* 1 1
→ (1 0 ) ⎜ 1 2
⎟ = ⎜ ( 0 1) ⎜ 1 2

x x
⎝ 2 c−

0 ⎠⎝1⎠ ⎝ ⎟
⎝ 2 c− 0 ⎟⎠ ⎝⎜ 0 ⎟⎠ ⎠
⇒ c + = c− *
This result reduces our work to finding one constant (c+) in terms of which
we have
1⎛ 0 c ⎞ 1 ⎛ 0 −ic+ ⎞
Sx = ⎜ * + ⎟ Sy = ⎜ *
2 ⎝ c+ 0⎠ 2 ⎝ ic+ 0 ⎟⎠
Note that S x and S y have the interesting property that they are equal to
their adjoints:

1⎛ 0 c+ ⎞ 1 ⎛ 0 c+ ⎞
Sx = ⎜ *

⎟ = ⎜ ⎟ = Sx
2 ⎝ c+ 0 ⎠ 2 ⎝ c+ * 0⎠

−ic+ ⎞ 1 ⎛ 0 −ic+ ⎞
1⎛ 0
Sy = ⎜ *

⎟ = ⎜ ⎟ = Sy
0 ⎠ 2 ⎝ ic+ *
2 ⎝ ic+ 0 ⎠
This property turns out to be true in general: Hermitian operators are
represented by matrices that are equal to their own adjoint. These self­
adjoint matrices are typically called Hermitian matrices for this reason, and
the adjoint operation is sometimes called Hermitian conjugation.

To determine the remaining constant, we use the fact that S 2 = S x 2 + S y 2 + S z 2 .


Plugging in our matrix representations for S x , S y , Sz and S 2 we find:
3� 2 ⎛ 1 0 ⎞ � 2 ⎛ 1 0 ⎞ ⎛ 1 0 ⎞ 1 ⎛ 0 c+ ⎞ ⎛ 0 c+ ⎞ 1 ⎛ 0 −ic+ ⎞ ⎛ 0 −ic+ ⎞
= + ⎜ ⎟⎜ ⎟+ ⎜ ⎟⎜ ⎟
4 ⎜⎝ 0 1 ⎟⎠ 4 ⎝⎜ 0 −1 ⎟⎠ ⎜⎝ 0 −1 ⎟⎠ 4 ⎝ c+ * 0 ⎠ ⎝ c+ * 0 ⎠ 4 ⎝ ic+ * 0 ⎠ ⎝ ic+ * 0 ⎠
2 2
� 2 ⎛ 1 0 ⎞ c + ⎛ 1 0 ⎞ c+ ⎛ 1 0 ⎞
= ⎜ + +
4 ⎝ 0 1 ⎟⎠ 4 ⎝⎜ 0 1 ⎟⎠ 4 ⎝⎜ 0 1 ⎠⎟
2 2

3 � 2 c+
c
⇒ �2 = + + +
4 4 4 4
2
⇒ c+ = � 2
⇒ c+ = �

Prepared By Kevin Boudreaux


126 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 24 Pauli Spin Matrices Page 7

where on the last line, we have made an arbitrary choice of the sign of c+ .
Thus we arrive at the final expressions for S x and S y
� ⎛0 1⎞ � ⎛ 0 −i ⎞
⎜ Sx =
⎟ Sy = ⎜
2 ⎝1 0⎠ 2 ⎝ i 0 ⎟⎠
In summary, then, the matrix representations of our spin operators are:
� ⎛0 1⎞ � ⎛ 0 −i ⎞ � ⎛1 0 ⎞ 3 2 ⎛1 0⎞
Sx = Sy = Sz = S2 = �
2 ⎜⎝ 1 0 ⎟⎠ 2 ⎜⎝ i 0 ⎟⎠ 2 ⎝⎜ 0 −1 ⎟⎠ 4 ⎜⎝ 0 1 ⎟⎠
Given our representations of S x , S y and Sz now we can work out any property
we like by doing some linear algebra. For example, we can work out the
commutator of S x and S y :
� ⎛ 0 1 ⎞ � ⎛ 0 −i ⎞ � ⎛ 0 −i ⎞ � ⎛ 0 1 ⎞
⎡⎣S x ,S y ⎤⎦ = S x S y − S yS x = ⎜ −
2 ⎝ 1 0 ⎟⎠ 2 ⎝⎜ i 0 ⎟⎠ 2 ⎜⎝ i 0 ⎟⎠ 2 ⎝⎜ 1 0 ⎟⎠
� 2 ⎡⎛ 0 1 ⎞ ⎛ 0 −i ⎞ ⎛ 0 −i ⎞ ⎛ 0 1 ⎞ ⎤
= ⎢⎜ −
4 ⎣⎝ 1 0 ⎠ ⎝ i 0 ⎟⎠ ⎜⎝ i 0 ⎟⎠ ⎝⎜ 1 0 ⎟⎥
⎟⎜
⎠⎦
�2 ⎡⎛ i 0 ⎞ ⎛ −i 0 ⎞ ⎤ � 2 ⎛ i 0 ⎞
= ⎢⎜ ⎟−⎜ ⎟⎥ = ⎜ ⎟
4 ⎣⎝ 0 −i ⎠ ⎝ 0 i ⎠ ⎦ 2 ⎝ 0 −i ⎠
⇒ ⎡⎣S x ,S y ⎤⎦ = i�S z
It is comforting to see that the matrices that represent the angular
momentum operators obey the commutation relations for angular momentum!
It is similarly possible to work out the eigenvalues of Ŝ 2 , Ŝz and Ŝx by
computing the eigenvalues of S x , S y and Sz ; average values can be obtained
by taking (row)*x(matrix)x(column) products. We are thus in a position to
compute anything we want for these operators. As it turns out, any operator
in this 2x2 space can be written as a linear combination of S x , S y , Sz and S2 .
So, in some sense, we now have all the information we could possibly want
about spin­½ systems.

In honor of Pauli, it is conventional to define the dimensionless versions of


S x , S y , Sz (i.e. matrices without the leading factor of � / 2 ) as Pauli Spin
matrices:
⎛ 0 1⎞ ⎛ 0 −i ⎞ ⎛1 0⎞
σx ≡ ⎜ ⎟ σy ≡ ⎜ ⎟ σz ≡ ⎜ ⎟
⎝ 1 0⎠ ⎝i 0⎠ ⎝ 0 −1⎠
One of the reasons the Pauli spin matrices are useful (or equivalently the
matrix representations of S x , S y and Sz ) is that they are the simplest

Prepared By Kevin Boudreaux


127 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 24 Pauli Spin Matrices Page 8

possible place to practice our skills at using matrix mechanics. We will see
later on that more complicated systems can also be described by matrix­
vector operations. However, in these cases the matrices will have many
more elements – sometimes even infinite numbers of elements! Therefore, it
is best to get some practice on these simple 2x2 matrices before taking on
the more complicated cases.

Prepared By Kevin Boudreaux


128 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 25 Helium Atom page 1

HELIUM ATOM

Now that we have treated the Hydrogen like atoms in some detail, we now
proceed to discuss the next­simplest system: the Helium atom. In this
situation, we have tow electrons – with coordinates
z r1 and r2 – orbiting a nucleus with charge Z = 2
located at the point R. Now, for the hydrogen atom
we were able to ignore the motion of the nucleus
r2 by transforming to the center of mass. We then
R obtained a Schrödinger equation for a single
y
effective particle – with a reduced mass that was
very close to the electron mass – orbiting the
x origin. It turns out to be fairly difficult to
r1
transform to the center of mass when dealing with
three particles, as is the case for Helium. However, because the nucleus is
much more massive than either of the two
electrons (MNuc ≈ 7000 mel) it is a very good z
approximation to assume that the nucleus sits at
the center of mass of the atom. In this
approximate set of COM coordinates, then, R=0 r2
and the electron coordinates r1 and r2 measure
y
the between each electron and the nucleus.
Further, we feel justified in separating the
motion of the nucleus (which will roughly x
correspond to rigidly translating the COM of the
r1
atom) from the relative d the electrons orbiting
the nucleus within the COM frame. Thus, in what follows, we focus only on the
motion of the electrons and ignore the motion of the nucleus.

We will treat the quantum mechanics of multiple particles (1,2,3…) in much the
same way as we described multiple dimensions. We will invent operators r̂1 , r̂2 ,
r̂3 , … and associated momentum operators p̂1 , p̂ 2 , p̂3 …. The operators for a
given particle (i) will be assumed to commute with all operators associated with
any other particle (j):
[r̂1, p̂ 2 ] = [ p̂2 , r̂3 ] = [r̂2 , r̂3 ] = [ p̂1, p̂3 ] = ... ≡ 0

Prepared By Kevin Boudreaux


129 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 25 Helium Atom page 2

Meanwhile, operators belonging to the same particle will obey the normal
commutation relations. Position and momentum along a given axis do not
commute:
⎡ x̂1, p̂x ⎤ = i� ⎡ ŷ , p̂ ⎤ = i� ⎡ ẑ , p̂ ⎤ = i�
⎣ 1⎦ ⎣ 1 y1 ⎦ ⎣ 1 z1 ⎦
while all components belonging to different axes commute:
xˆ1 yˆ1 = yˆ1 xˆ1 p̂z1 yˆ1 = yˆ1 pˆ z1 pˆ z1 p̂x1 = pˆ x1 pˆ z1 etc.
As you can already see, one of the biggest challenges to treating multiple electrons
is the explosion in the number of variables required!

In terms of these operators, we can quickly write down the Hamiltonian for the
Helium atom:

Kinetic Energy Nucleus­Electron 1


Of Electron 1 Electron­Electron

Attraction
Repulsion

p̂12 p̂22 Ze2 1 Ze2 1 e2 1


Ĥ ≡ + − − +
2me 2me 4πε0 r̂1 4πε0 r̂2 4πε 0 r̂ −r̂
1 2
Kinetic Energy Nucleus­Electron 2
Of Electron 2 Attraction

This Hamiltonian looks very intimidating, mainly because of all the constants (e, me,
ε0, etc.) that appear in the equation. These constants result from our decision to
use SI units (meter, gram, second) to express our lengths, masses and energies.
This approach is quite awkward when the typical mass we’re dealing with is 10­28
grams, the typical distance is 10­10 meters and the typical energy unit is 10­18
Joules. It is therefore much simpler to work everything out in what are called
atomic units. In this system of units we choose our unit of mass to be equal to the
electron mass, me, our unit of charge to be equal to the electron charge, e, and our
unit of angular momentum to be � . Further, we choose to work in electrostatic
units, so that the permittivity of free space (4π ε0) is also 1. The result of these
four choices is twofold. First of all, the Hamiltonian of the Helium atom (and
indeed of any atom or molecule) simplifies greatly because all the constants are
unity and can be omitted in writing the equations:

Prepared By Kevin Boudreaux


130 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 25 Helium Atom page 3

p̂12 p̂22 Z Z
Ĥ ≡ + − − + 1
2 2 r̂1 r̂2 r̂ − r̂
1 2

∇12 ∇22
=− − −Z −Z + 1
2 2 r̂1 r̂2 r̂ − r̂
1 2
This will greatly simplify our algebra in the future and we will typically use atomic

units from this point forward. The second benefit of this choice of units is that

now our units of energy, mass, distance, angular momentum, etc. are of the

appropriate size for dealing with atoms and molecules:

Atomic units and their SI equivalents

Quantity Natural unit SI equivalent

Electron mass m =1 9.11x10 −31 kg

Charge e =1 1.06x10−19 C

Angular momentum � =1 1.05x10−34 J ⋅ s


Permittivity κ 0 = 4πε 0 = 1 1.11x10−10 C2 ⋅ J -1 ⋅ m -1
Length κ 0� 2 me 2 = a0 = 1 (bohr) 5.29x10−11 m
(Bohr radius)
Energy me 4 κ 02 � 2 = e2 κ 0a0 = 1 (hartree) 4.36x10−18 J = 27.2 eV
(twice the ionization energy of H)
Time κ 02�3 me 4 = 1 2.42x10−17 s
(period of an electron in the first Bohr orbit)

Speed e 2 κ 0� = 1 2.19 x106 kg

(speed of an electron in the first Bohr orbit)

Electric potential me3 κ 02 � 2 = e 2 κ 0 a0 = 1 27.21 V

(potential of an electron in the first Bohr orbit)

Magnetic dipole moment e� m = 1 1.85x10−23 J ⋅ T-1

(twice a Bohr magneton)

Thus, so long as we work consistently in atomic units, we will tend to find

energies that are of order unity, distances that are of order unity, momenta

that are of order unity … without having to keep track of so much scientific

notation.

Prepared By Kevin Boudreaux


131 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 25 Helium Atom page 4

Once Schrodinger had solved the Hydrogen atom, it was generally believed that
the solution of the Helium atom would follow not long afterward. However,
Scientists have tried for decades to solve this three body problem without
succeeding. Very, very accurate approximations were developed, but no exact
solutions were found. As it turns out, even with the simplifications described
above it is impossible to determine the eigenstates of the Helium atom. This
situation is common in chemistry, as most of the problems we are interested in
cannot be solved exactly. We therefore must resort to making approximations, as
we will do throughout this course.

Non­Interacting Electron Approiximation


For Helium, the first thing we notice is that the Hamiltonian becomes separable if
we neglect the electron­electron repulsion term:

∇12
Z Z ∇22
∇12 Z ∇22 Z ˆ ˆ
Ĥ ind = + − − = − + − ≡ H1 + H 2
2 2 r̂1 r̂2 2 r̂1 2 r̂2
Electron One Electron Two
Thus, if we neglect the interaction between the electrons, the Hamiltonian reduces
to the sum of two hydrogenic atom Hamiltonians (with Z=2). Based on our
experience with separable Hamiltonians in the past (Hamiltonians Add �
Wavefunctions Multiply � Energies Add), we can immediately write down the
correct form for the eigenfunctions and eigenvalues of this Hamiltonian:
Ψ n l m s ;n l m s (r1,σ 1; r2 ,σ 2 ) = ψ n l m s (r1,σ 1 )ψ n l m s (r2 ,σ 2 )
11 1 1 2 2 2 2 11 1 1 22 2 2
2 2
Z Z
En1n2 = En1 + En2 = − 2

2n1 2n2 2
where, in the second line, we have made use of atomic units ( � =me=e=4π ε0=1) to
simplify our energy expression. Thus, by neglecting the electron repulsion (an
approximation) we move from a problem that is impossible to solve to one that we
can already solve easily.

However, the outstanding question is: how good is this approximation? The easiest
way to test this is to look at the ground state. According to the Pauli exclusion
principle, we can’t put two electrons in the same state. However, as we learned in
freshman chemistry, we can put one electron in 1s with spin up and the other in 1s
with spin down:
Ψ1sα ;1sβ (r1,σ 1; r2 ,σ 2 ) = ψ100α (r1,σ 1 )ψ100 β (r2 ,σ 2 )

Prepared By Kevin Boudreaux


132 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 25 Helium Atom page 5

This wavefunction has an energy


Z2 Z2
E11 = − − = − Z 2 =-4 a.u. = −108.8 eV
2 2
How good is this result? Well, we can determine the ground state energy of
Helium by removing one electron to create He+ and then removing the second to
create He2+. As it turns out, the first electron takes 24.2 eV to remove and the
second takes 54.4 eV to remove, which means the correct ground state energy is
78.8 eV. So our non­interacting electron picture is off by 30 eV, which is a lot of
energy. To give you an idea of how much energy that is, you should note that a
typical covalent chemical bond is worth about 5 eV of energy. So totally neglecting
the electron interaction is not a very good approximation.

Independent Electron Approximation


So how can we go about improving this approximation? Well, first we note that the
product wavefunction described above is not antisymmetric. To test antisymmetry,
all we have to do is recall that, since the electrons are identical, the labels “1” and
“2” are arbitrary. By swapping these labels we can’t possibly change the outcome
of any measurement. Upon interchanging the labels “1” and “2”, an antisymmetric
wavefunction will give the same wavefunction back times a minus sign. However,
our proposed wavefunction does not do this:
Interchange
ψ1sα ( r1,σ 1 )ψ 1sβ ( r2 ,σ 2 ) ⎯⎯⎯⎯→ψ 1sα ( r2 ,σ 2 )ψ 1sβ ( r1,σ 1 )
1 and 2

≠ −ψ 1sα (r1,σ )ψ1sβ (r2 ,σ )


1 2

This is a problem, because we said that all electron wavefunctions should be


antisymmetric under exchange. We can fix this problem by taking the “­“
combination of the wavefunction we proposed and its exchange partner:
Ψ1sα ;1sβ ( r1,σ 1 ; r2 , σ 2 ) ≡ 1
2
(ψ1sα (r1,σ )ψ1sβ (r2 ,σ ) −ψ1sα (r2 ,σ )ψ1sβ (r1,σ ))
1 2 2 1

where the leading factor of 1/√2 ensures that the new wavefunction is normalized.
We check that this is antisymmetric:
Ψ1sα ;1sβ ( r1 , σ 1 ; r2 , σ 2 ) ≡
1
2
( )
ψ 1sα ( r1 , σ 1 )ψ 1sβ ( r2 , σ 2 ) − ψ 1sα ( r2 , σ 2 )ψ 1sβ ( r1 , σ 1 )

⎯⎯⎯
1↔2

1
2
( )
ψ 1sα ( r2 , σ 2 )ψ 1sβ ( r1 , σ 1 ) − ψ 1sα ( r1 , σ 1 )ψ 1sβ ( r2 , σ 2 )

=−
1
2
( )
ψ 1sα ( r1 , σ 1 )ψ 1sβ ( r2 , σ 2 ) − ψ 1sα ( r2 , σ 2 )ψ 1sβ ( r1 , σ 1 )

= −Ψ1sα ;1sβ ( r1 , σ 1 ; r2 , σ 2 )

Prepared By Kevin Boudreaux


133 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 25 Helium Atom page 6

Does this new wavefunction give a better energy? As it turns out, this change by
itself does nothing for the energy prediction. The new wavefunction is a linear
combination of two degenerate eigenstates of the independent electron
Hamiltonian. As we learned before, any sum of degenerate eigenstates is also an
eigenstate with the same eigenvalue. So, while quantum mechanics says we have to
make the wavefunction antisymmetric, antisymmetry by itself does not affect our
energy for Helium.

The simplest way for us to improve our estimate of the helium ground state energy
is to consider not the eigenvalues of our approximate Hamiltonian with our
approximate eigenfunctions, but instead look at the average energy of our
approximate function with the exact Hamiltonian. That is to say, a better
approximation to the energy can be got from
Ĥ = ∫ Ψ*1sα ;1sβ Ĥ Ψ 1sα ;1sβ dτ1dτ 2
where dτ1 = dr1 dσ1 and similarly for dτ2. We refer to this picture as an
independent electron approximation. Within the wavefunction the electrons
behave as if they do not interact, because we have retained the separable form.
However, in computing the energy, we fold these interactions back in in an
approximate way by computing the average energy including the interaction.

We can simplify the average energy pretty quickly:


⎛ 1 ⎞
∫Ψ ⎜ Ĥ1 + Ĥ2 + ⎟Ψ dτ dτ
*
1sα ;1s β ⎜ r1 − r2 ⎟ 1sα ;1sβ 1 2
⎝ ⎠
⎛ 1 ⎞
⇒ ∫ Ψ 1sα ;1sβ
* ⎜ −2 + −2 + ⎟Ψ dτ dτ
⎜ r1 − r2 ⎟ 1sα ;1sβ 1 2
⎝ ⎠
1
⇒ −4 ∫ Ψ*1sα ;1sβ Ψ 1sα ;1sβ dτ1dτ 2 + ∫ Ψ*1sα ;1sβ Ψ dτ dτ
r1 − r2 1sα ;1sβ 1 2
2 1
Ψ 1sα ;1sβ
⇒ −4 + ∫ dτ1dτ 2
r1 − r2
We thus have for the average energy:

2
Ψ 1sα ;1sβ
Ĥ = −4 + ∫ dτ1dτ 2
r1 − r2

Prepared By Kevin Boudreaux


134 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 25 Helium Atom page 7

The first term is simply the non­interacting electron energy from above. The
second term is the average value of the electron­electron repulsion. Because the
integrand is strictly positive (as one would expect for electron repulsion) this new
term will only raise the average energy, which is a good thing, since our non­
interacting picture gave an energy that was 30 eV too low! We can further expand
the repulsion term based on the antisymmetric form of the wavefunction. First,
we note that we can factorize the antisymmetric wavefunction into a space part
times a spin part:
1

(ψ 1sα (r1 , σ1 )ψ 1sβ ( r2 , σ 2 ) − ψ 1sα ( r2 )ψ 1sβ ( r1 , σ1 ))



1
2

(ψ 1s ( r1 ) α (σ1 )ψ 1s ( r2 ) β (σ 2 ) − ψ 1s ( r2 ) α (σ 2 )ψ 1s ( r1 ) β (σ1 ) )
⇒ ψ 1s ( r1 )ψ 1s ( r2
)
1
2

(α (σ1 ) β (σ 2 ) − β (σ1 ) α (σ 2 ) )

Ψ space (r1 , r2 ) Ψ spin (σ 1 , σ 2 )

With these definitions, it is easy to verify that the space and spin wavefunctions
are individually normalized. Note, in the absence of a magnetic field, you will
always be able to write the eigenfunctions of the Hamiltonian in this form because
H is separable into a space and spin part
Ĥ = Ĥspace + Ĥ spin
With the spin part being (trivially) zero. As a result, eigenfunctions of H will always
be products of a space part and a spin part as above. With this space/spin
separation in hand, we can simplify the repulsion term:
2

Ψ space ( r1 , r2 ) Ψ spin (σ1 , σ 2 )


2
2 2
Ψ Ψ space Ψ spin
∫ dτ1dτ 2 = ∫ dr1dr2 dσ 1dσ 2 = ∫ dr1dr2 dσ 1dσ 2
1sα ;1 s β

r1 − r2
r1 − r2 r1 − r2
Ψ space ( r1 , r2 )
2

(σ 1 , σ 2 )
2

= ∫
Ψ spin dσ 1dσ 2 ∫ r1 − r2

dr1dr2

1
Ψ space ( r1 , r2 )
2

= ∫ r1 − r2

dr1dr2

The evaluation of this 6 dimensional integral is very tedious (cf. McWeeny


problems 8­39 and 8­40) but the result is that
2

Ψ space 5Z 5
∫ r1 − r2
dr1dr2 =
8
=
a.u. = +34 eV
4

Prepared By Kevin Boudreaux


135 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry 25 Helium Atom page 8

Adding this average repulsion term to our non­interacting energy gives a ground
state estimate of ­108.8 eV +34 eV= ­74.8 eV, which is only 4 eV off of the correct
energy. The energy is still not very close, but at least we are making progress.

As we have seen already, the independent electron picture is not all that accurate
for describing atoms. However, chemists are very pragmatic and realize that the
ease of solving non­interacting problems is extremely valuable and as we will soon
see, it gives us a picture (molecular orbital theory) that allows us to describe a
wide range of chemistry. Therefore, chemists are extremely reluctant to abandon
an independent particle picture. Instead, a great deal of work has gone into making
more accurate models based on independent particles – either by making more
sophisticated corrections like the one above or by coming up with a different non­
interacting Hamiltonian that gives us a better independent particle model. We will
spend the next several lectures discussing the qualitative features of electronic
structure and bonding that come out of this picture.

Prepared By Kevin Boudreaux


136 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #26 page 1

TWO ELECTRONS: EXCITED STATES

In the last lecture, we learned that the independent particle model gives a reasonable
description of the ground state energy of the Helium atom. Before moving on to talk
about many­electron atoms, it is important to point out that we can describe many
more properties of the system using the same type of approximation. By using the
same independent particle prescription we can come up with wavefunctions for
excited states and determine their energies, their dependence on electron spin, etc.
by examining the wavefunctions themselves. That is to say, there is much we can
determine from simply looking at Ψ without doing any significant computation.

We will use the excited state 1s2s configuration of Helium as an example. For the
ground state we had:
Ψ space (r1,r2 ) × Ψspin (σ1,σ 2 )

1s ⇒ ψ1s (r1 )ψ1s (r2 ) ⎜ ( 1) ( 2 ) ( 1 ) ( 2 ) ⎟⎠


1 ⎛α σ β σ − β σ α σ ⎞
2⎝
In constructing excited states it is useful to extend the stick diagrams we have used
before to describe electronic configurations. Then there are four different
configurations we can come up with:
Ψ space (r1,r2 ) × Ψspin (σ1,σ 2 )
2s
? ?

1s

2s
? ?
1s

2s
? ?
1s

2s
? ?
1s

Prepared By Kevin Boudreaux


137 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #26 page 2

Where the question marks indicate that we need to determine the space and spin
wavefunctions that correspond to these stick diagrams. It is fairly easy to come up
with a reasonable guess for each configuration. For example, in the first case we
might write down a wavefunction like:
ψ1sα (r1σ1 )ψ 2sα (r2σ 2 ) =ψ1s (r1 )ψ 2s (r2 ) α (σ1 ) α (σ 2 )
However, we immediately note that this wavefunction is not antisymmetric. We can
perform the same trick as before to make an antisymmetric wavefunction out this:

⇒ 1 ⎛ψ
(r1 )ψ 2s (r2 )α (σ1 )α (σ 2 ) −ψ1s (r2 )ψ 2s (r1 )α (σ 2 )α (σ1 ) ⎟⎞⎠
2 ⎜⎝ 1s

⇒ 12 ⎛⎜ψ1s (r1 )ψ 2s (r2 ) −ψ1s (r2 )ψ 2s (r1 ) ⎟⎞ α (σ1 ) α (σ 2 )


⎝ ⎠

Ψ space Ψ spin
Applying the same principle to the 1s↑2s↓ configuration gives us a bit of trouble:
⇒ 1 ⎛ψ
2 ⎜⎝ 1s (r1 )ψ 2s (r2 )α (σ1 ) β (σ 2 ) −ψ1s (r2 )ψ 2s (r1 )α (σ 2 ) β (σ1 ) ⎟⎠⎞ ≠ ΨspaceΨspin
Hence, the pure ↑↓ configuration can’t be separated in terms of a space part and a
spin part. We find a similar result for 1s↓2s↑:
⇒ 1 ⎛ψ
2 ⎜⎝ 1s (r1 )ψ 2s (r2 ) β (σ1 )α (σ 2 ) −ψ1s (r2 )ψ 2s (r1 ) β (σ 2 )α (σ1 ) ⎟⎞⎠ ≠ ΨspaceΨspin
Since we know the wavefunction should separate, we have a problem. The solution
comes in realizing that for an open shell configuration like this one, the 1s↑2s↓ and
1s↓2s↑ states are degenerate eigenstates and so we can make any linear combinations
of them we like and we’ll still obtain an eigenstate. If we make the “+” and “­“
combinations of 1s↑2s↓ and 1s↓2s↑ we obtain:

( 1s (r1 )ψ 2s (r2 ) α (σ1 ) β (σ 2 ) −ψ1s (r2 )ψ 2s (r1 ) α (σ 2 ) β (σ1 )) ±


⇒ ψ

(ψ1s (r1 )ψ 2s (r2 ) β (σ1 ) α (σ 2 ) − ψ1s (r2 )ψ 2s (r1 ) β (σ 2 ) α (σ1 ))


⇒ (ψ (r )ψ (r ) � ψ (r )ψ (r )) (α (σ ) β (σ ) ± β (σ ) α (σ ) )
1s 1 2s 2 1s 2 2s 1 1 2 2 1

Ψ space Ψ spin

which separates nicely. Performing similar manipulations for the ↓↓ configuration and
taking care to make sure that all our spatial and spin wavefunctions are individually
normalized allows us to complete the table we set out for the 1s2s excited states:

Prepared By Kevin Boudreaux


138 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #26 page 3

Ψ space (r1,r2 ) × Ψspin (σ1,σ 2 )


2s
⇒ 1
2 (ψ 1s (r1 )ψ 2s (r2 ) −ψ1s (r2 )ψ 2s (r1 )) × α (σ1 ) α (σ 2 )
1s

2s

1s
1
(
ψ ( r )ψ ( r ) + ψ1s ( r2 )ψ 2s ( r1 ) ×
2 1s 1 2s 2
) 1
2
(
α (σ1 ) β (σ 2 ) − β (σ1 ) α (σ 2 ) )
and
2s 1
(
ψ ( r )ψ ( r ) −ψ1s ( r2 )ψ 2s ( r1 ) ×
2 1s 1 2s 2
) 1
2
(
α (σ1 ) β (σ 2 ) + β (σ1 ) α (σ 2 ) )
1s

2s
⇒ 1
2 (ψ1s (r1 )ψ 2s (r2 ) −ψ1s (r2 )ψ 2s (r1 )) × β (σ1 ) β (σ 2 )
1s
We notice several things about these wavefunctions:
• While the overall wavefunction is always antisymmetric by construction, the
spatial part can be either antisymmetric (cases 1, 3 and 4) or symmetric (case
2). This effect is compensated for in the spin part, which can also be
antisymmetric (case 2) or symmetric (cases 1,3 and 4). The resulting
wavefunction always has a symmetric part times an antisymmetric part,
resulting in an antisymmetric wavefunction.
• The spin part of Case 2 is exactly the same as the spin part of the ground state
of the helium atom. Thus, just as we thought of the electrons in the ground
state as being “paired”, we say the electrons in Case 2 are paired.
• The spatial parts of three of the states above (cases 1,3 and 4) are the same.
Case 2 has a different spatial part. Because the Hamiltonian only depends on
spatial variables and not spin, we immediately conclude that 1,3 and 4 will be
degenerate states – even when we take into account the electron­electron
interaction. State 2, however, will generally have a different energy once we
account for interactions. In common spectroscopic parlance the three
degenerate states are called a triplet and the unique state is called a singlet.
Further, because these states arise from degenerate spin states, they are
called singlet and triplet spin states.

Prepared By Kevin Boudreaux


139 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #26 page 4

Energies of Singlet and Triplet States


As we showed above, we expect the singlet and triplet states to have different
energies once electron repulsion is taken into account. Which one will be lower? To
decide this, we note that the triplet spatial wavefunction is zero when the two
electrons are at the same position:
r1 = r2 ⇒ ΨT =
1
(
ψ ( r )ψ ( r ) −ψ1s ( r1 )ψ 2s ( r1 ) = 0
2 1s 1 2s 1
)
whereas the singlet wavefunction is non­zero:
r1 = r2 ⇒ Ψ S =
1
( )
ψ ( r )ψ ( r ) + ψ1s ( r1 )ψ 2s ( r1 ) = 2 ψ 1s ( r1 )ψ 2s ( r1 ) ≠ 0
2 1s 1 2s 1
Because the electrons repel each other more when they are close to one another, we
therefore expect the singlet to have more electron­electron repulsion and a higher
energy. This rule turns out to hold quite generally and is called Hund’s rule: for
degenerate non­interacting states, the configuration with highest spin multiplicity
lies lowest in energy. Hund actually has three rules (of which this is the first)
concerning the ordering of degenerate non­interacting states. The others apply only
to atoms and will not be discussed here, but see McQuarrie Section 9.11­9.12 for
more on this topic.

So we expect the triplet to be lower. How much lower? To answer this question, we
have to compute the average energies of the singlet and triplet wavefunctions. Recall
that the spin part never matters for the energy:

∫ Ψ Ĥ Ψdrdσ = ∫ Ψ ∫ ∫
Ψ spin* Ĥ Ψ space Ψ spin drdσ = Ψ spin* Ψ spin dσ Ψ space* Ĥ Ψ space dr
* *
space

= ∫Ψ
1
space Ĥ Ψ space dr
*

The influence of the spin wavefunction is only indirect: if the spin part is
antisymmetric (e.g. singlet) then the spatial part must be symmetric and vice versa.
To simplify our algebra, it is convenient to create the obvious shorthand notation:
( )
ΨT = 1 ψ 1s ( r1 )ψ 2s ( r2 ) −ψ 1s ( r2 )ψ 2s ( r1 ) ≡ 1 (1s2s − 2s1s )
2 2

ΨS = 1 (ψ 1s ( r1 )ψ 2s ( r2 ) +ψ 1s ( r2 )ψ 2s ( r1 ) ) ≡ 1
(1s2s + 2s1s )
2 2
where we just need to remember that the first function in a product will be the one
that has electron “1” while the second will have electron “2”. Proceeding then:

Prepared By Kevin Boudreaux


140 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #26 page 5

1 ⎛ ⎞
ES /T = ∫ Ψ S /T * Ĥ Ψ S /T dr1dr2 = ( ) ˆ + Hˆ + 1 ⎟ (1s 2s ± 2s1s ) dr1dr2
2 ∫ 1s 2s ± 2 s1s ⎜H

1 2
r1 − r2 ⎟
⎝ ⎠
1 ⎛ ⎞
( ) ⎜ −2 − 1 + 1 ⎟ (1s2s ± 2s1s ) dr1dr2
2∫
= 1s 2 s ± 2 s1s
⎜ 2 r1 − r2 ⎟
⎝ ⎠
5 1 1
= − + ∫ (1s 2s ± 2s1s ) (1s2s ± 2s1s ) dr1dr2
2 2 r1 − r2
On the second line, we have used the fact that both ΨS and ΨT are eigenstates of the
independent particle Hamiltonian (by construction) and on the third line, we have
taken the independent particle energy outside the integral because ΨS and ΨT are
normalized. Thus, we see that the average energy takes on the familiar form of
(noninteracting energy)+(interactions). The interaction term can be simplified
further:
1
2∫
(1s2s ± 2s1s ) r −1 r (1s2s ± 2s1s ) dr1dr2
1 2
1 1 1 1 1
2∫
⇒ 1s2s 1s 2s −1s2s 2s1s − 2s1s 1s2s + 2s1s 2s1s dr1dr2
r1 − r2 r1 − r2 r1 − r2 r1 − r2
We note that the first and last terms are the same if we just interchange the dummy
integration variables:
1 1
∫ 1s (r1 ) 2s (r2 ) r1 − r2 1s (r1 ) 2s ( r2 ) dr1dr2 ⎯⎯⎯→∫ 1s ( r2 ) 2s (r1 ) r2 − r1 1s (r2 ) 2s (r1 ) dr2 dr1
1↔2

1
= ∫ 2s ( r1 )1s ( r2 ) 2s ( r1 )1s ( r2 ) dr1dr2 ≡ J12
r1 − r2
Meanwhile the second and third terms are also the same:
1 1
∫ 1s (r1 ) 2s (r2 ) r1 − r2 2s (r1 ) 1s (r2 ) dr1dr2 ⎯⎯⎯→∫ 1s ( r2 ) 2s (r1 ) r2 − r1 2s (r2 ) 1s ( r1 ) dr2dr1
1↔2

1
= ∫ 2s ( r1 ) 1s ( r2 ) 1s ( r2 ) 2s ( r1 ) dr1dr2 ≡ K12
r1 − r2
These integrals are called Coulomb(J) and exchange(K) integrals, respectively. Both
are positive numbers (because they arise from electron repulsion) and it can be
rigorously proved that J>K always (i.e. no matter what functional form the 1s and 2s
wavefunctions have). Thus, in terms of J and K the energies of the singlet and triplet
states become:
5
ES /T = − + J12 ± K12
2
Thus we see that, as expected, the singlet state is higher in energy than the
triplet. In fact, we can even give a numerical estimate for the splitting by evaluating
K12. Plugging in the forms of the 1s and 2s orbitals of helium and doing the integrals,
we obtain K12=32/729=1.2 eV and a splitting of 2K12=2.4 eV. The latter is quite a bit
larger than the experimental singlet­triplet splitting in helium, which comes out to

Prepared By Kevin Boudreaux


141 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #26 page 6

only .8 eV. Once again, we see the independent particle model gives us a qualitatively
correct picture (i.e. the sign of the splitting is correct and of the right order of
magnitude) but we fail to obtain quantitative results. We therefore arrive at the
following qualitative picture of the 1s2s excited state of Helium:

It is interesting to note that the exchange interaction results from the fact that the
electrons are indistinguishable. Notice that, if we had not antisymmetrized our
wavefunctions, the spatial part would have just been a direct product 1s2s instead of
the symmetric/antisymmetric 1s2s±2s1s combinations we obtained for the singlet and
triplet. In the former case, the electrons are being treated as distinguishable (e.g.
electron “1” is always 1s while electron “2” is always 2s) and the exchange term
disappears from the interaction:
1 1 1
2 ∫ (1s 2s ± 2s1s )
r1 − r2
(1s 2s ± 2s1s ) dr1dr2 ⎯⎯⎯⎯⎯
distinguishable
⎯→ ∫ 1s2s
r1 − r2
1s 2sdr1dr2 = J12

Clearly exchange – which arises from the cross terms on the left – is absent on the
right. Thus, the K integrals only arise when we have terms in the wavefunction where
two electrons have exchanged places. Hence the name “exchange.” It is important to
note that, next to the Pauli exclusion principle, this is the biggest impact that
antisymmetry has on chemistry.

There was a lot of interest in class about how we reconcile the fact that, in other
chemistry courses you’ve been taught that there are only two spin states for a pair of
electrons: ↑↓ and ↑↑. The former represented the singlet state and the latter the
triplet state. You referred to the singlet electrons as being “paired” and the triplets
as being “unpaired.” However, how do these strange spin states we’ve derived connect
with the “paired” and “unpaired” ideas? To answer this question, we first of all we
should note that neither of the antiparallel states we’ve derived is strictly ↑↓.
Instead, they look like ↑↓±↓↑, with ↑↓-↓↑ being the singlet and ↑↓+↓↑ being part of the
triplet. The idea that the singlet state is ↑↓ is a white lie that we tell in order to
simplify our arguments: as long as the subtle difference between ↑↓ and ↑↓-↓↑ isn’t
important, we can get away with explaining much (though not all) chemistry by treating
the singlet state as ↑↓.

Prepared By Kevin Boudreaux


142 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #26 page 7

In the more precise picture we’ve derived here, it the spin part of the wavefunction
determines whether the electrons are paired or not. An electron pair has the
characteristic spin part αβ−βα. That is to say, paired electrons form a singlet. Spin
parts that look like αα, αβ+βα, or ββ are unpaired triplet configurations. As we have
seen above, pairing two electrons raises the energy through the exchange integral.
In some situations, this is called the “pairing energy.” The counterintuitive thing that
we have to re­learn is that αβ+βα does not describe an electron pair. In every way it
behaves like αα or ββ : the energies are the same and (as you will show on the
2
homework) the eigenvalues of Ŝtotal are the same. This idea really does not fit into the
simple qualitative picture of triplet states being ↑↑, but it is nonetheless true.

The fact that there are three elements of the triplet state is not a coincidence. As
you will show, the eigenvalues of Sˆtotal
2
for the triplet states are all 2� 2 , which is
consistent with a total spin of S=1, because the eigenvalues of Sˆ 2 would then be total

� S ( S + 1) = 2� . This picture is also


2 2

consistent with the idea that, if we add z


two spins with s=½ parallel to one
M S = +1
another we should get a total S of S=
½+½=1. Given this picture, we note that
the three triplet states would then
y
correspond to the three possible z­ MS = 0
projections of spin. That is to say the x
three triplet states should have M S = −1
MS=+1,0 and­1, respectively. This gives
us at least some qualitative picture of what the αβ+βα state means and why it
corresponds to unpaired electrons. In the αβ+βα state the spins are oriented parallel
to each other, but they are both oriented perpendicular to the z axis, so that on
average you will always find one spin­up and one spin­down along z. This is a very
simple example of the addition of angular momentum, a topic which is covered in
much greater depth in McQuarrie.

Prepared By Kevin Boudreaux


143 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #27 page 1

MANY ELECTRON ATOMS

Thus far, we have learned that the independent particle model (IPM) gives a
qualitatively correct picture of the eigenstates of the helium atom. What about
atoms with more than two electrons, such as lithium or carbon? As it turns out, the
IPM is capable of giving a realistic picture of atomic structure in essentially an
analogous fashion to the helium case. To begin with, we set up our coordinates so that
the nucleus is at the origin and the N electrons are at positions r1, r2, r3, …rN. In terms
of these variables, we can quickly write down the many-electron Hamiltonian (in atomic
units):

1 N 2 Z N N N
1
Hˆ = − ∑∇ j −∑ + ∑∑
2 j=1 j=1 r̂ j i=1 j >i r̂i − r̂ j

Kinetic Energy Electron-Nuclear Electron-Electron


Attraction Repulsion
Thus, the Hamiltonian has the same three sources of energy as in the two electron
case, but the sheer number of electrons makes the algebra more complicated. As
before, we note that we can make the Hamiltonian separable if we neglect the
electron-electron repulsion:

1 N N Z N ⎛ Z ⎟
⎞ N
Ĥ =− 2
∑ ∇ − ∑ ⎜
= ∑ − ∇ − 1 2 ≡ ∑ ĥ
j = 1 ⎝⎜
NI 2 j = 1 j j = 1 r̂ 2 j r̂ ⎟ i
j j ⎠ j =1
where each of the independent Hamiltonians ĥi describes a single electron in the field
of a nucleus of charge +Z. Based on our experience with separable Hamiltonians, we
can immediately write down the eigenstates of this Hamiltonian as products with
energies given as sums of the independent electron energies:
Ψ = ψ k (1)ψ k ( 2 )ψ k ( 3) ...ψ k (N ) E = Ek + Ek + Ek +... + Ek
1 2 3 N 1 2 3 N

Where (1) is a shorthand for (r1,σ1) and ki ≡ {ni ,li , mi , si } specifies all the quantum
numbers for a given hydrogen atom eigenstate. Of course, there is a problem with
these eigenstates: they are not antisymmetric. For the Helium atom, we fixed this by
making an explicitly antisymmetric combination of two degenerate product states:
ψ1sα (1) ψ1sα ( 2 )
1
(ψ (1)ψ1sβ ( 2) −ψ1sα ( 2)ψ1sβ (1)) =
2 1sα
1

1sβ (1) ψ1sβ ( 2)

Prepared By Kevin Boudreaux


144 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #27 page 2

where on the right we have noted that this antisymmetric product can also be written
as a determinant of a 2x2 matrix. As it turns out, it is straightforward to extend this
idea to generate an N particle antisymetric state by computing an NxN determinant
called a Slater Determinant:
ψ k (1) ψ k (1) ψ k (1) � ψ k (1)
1 2 3 N

ψ k ( 2) ψ k ( 2) ψ k ( 2) � ψ k ( 2)
1 2 3 N

Ψ (1,2,..., N ) =
1
N ! k1 ( )
ψ 3 ψ k ( 3) ψ k ( 3) � ψ k ( 3)
2 3 N

� � � � �
ψk (N ) ψk (N ) ψk (N ) � ψk (N )
1 2 3 N

As you can imagine, the algebra required to compute integrals involving Slater
determinants is extremely difficult. It is therefore most important that you realize
several things about these states so that you can avoid unnecessary algebra:
• A Slater determinant corresponds to a single stick diagram. This is easy to
see by example:
2px 1sα (1) 1s β (1) 2sα (1) 2 pxα (1)
1sα ( 2 ) 1s β ( 2 ) 2sα ( 2 ) 2 pxα ( 2 )
2s ⇒ Ψ (1, 2, 3, 4 ) =
1sα ( 3 ) 1s β ( 3 ) 2sα ( 3 ) 2 pxα ( 3 )
1s 1sα ( 4 ) 1s β ( 4 ) 2sα ( 4 ) 2 pxα ( 4 )
It should be clear that we can extend this idea to associate a determinant with
an arbitrary stick diagram. Further, recall that for the excited states of
helium we had a problem writing certain stick diagrams as a (space)x(spin)
product and had to make linear combinations of certain states to force things
to separate. Because of the direct correspondence of stick diagrams and
Slater determinants, the same pitfall arises here: Slater determinants
sometimes may not be representable as a space)x(spin) product, in which
case a linear combination of Slater determinants must be used instead.
This generally only happens for systems with unpaired electrons, like the 1s↑2s↓
configuration of helium or the …2px↑2py↓ configuration of carbon.
• A Slater determinant is anitsymmetric upon exchange of any two electrons.
We recall that if we take a matrix and interchange two its rows, the
determinant changes sign. Thus, interchanging 1 and 2 above, for example:

Prepared By Kevin Boudreaux


145 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #27 page 3

ψ k ( 2 ) ψ k ( 2 ) ψ k ( 2 ) � ψ k ( 2 )
1 2 3 N

ψ k (1) ψ k (1) ψ k (1) � ψ kN (1)


1 2 3

Ψ ( 2,1,..., N ) =
1
N ! k1 ( )
ψ 3 ψ k ( 3) ψ k ( 3) � ψ k ( 3)
2 3 N

� � � � �
ψ k ( N ) ψ k ( N ) ψ k ( N ) � ψ k ( N )
1 2 3 N

ψ k (1) ψ k (1) ψ k (1) � ψ k (1)


1 2 3 N

ψ k ( 2 ) ψ k ( 2 ) ψ k ( 2 ) � ψ k ( 2 )

Interchange
1 2 3 N

ψ 3 ψ k ( 3) ψ k ( 3) � ψ k ( 3) = −Ψ (1,2,..., N )
1

N ! k1 ( )
Rows 1&2 → −
⎯⎯⎯⎯⎯⎯
2 3 N

� � � � �

ψ k ( N ) ψ k ( N ) ψ k ( N ) � ψ k ( N )

1 2 3 N

A similar argument applies to any pair of indices, and so the Slater determinant
antisymmetric under any i↔j interchange.
• The determinant is zero if the same orbital appears twice. We recall that
if we take a matrix and interchange two of its columns, the determinant also
changes sign. Assuming k1=k3 above:
ψ k (1) ψ k (1) ψ k (1) � ψ k (1)
1 2 1 N

ψ k ( 2 ) ψ k ( 2 ) ψ k ( 2 ) � ψ k ( 2 )
1 2 1 N

Ψ (1,2,..., N ) =
1
N ! k1 ( )
ψ 3 ψ k ( 3) ψ k (3) � ψ k (3)
2 1 N

� � � � �
ψ k ( N ) ψ k ( N ) ψ k ( N ) � ψ k ( N )
1 2 1 N

ψ k (1) ψ k (1) ψ k (1) � ψ k (1)


1 2 1 N

Interchange

ψ k ( 2) ψ k ( 2) ψ k ( 2) � ψ k ( 2)
1 2 1 N

ψ 3 ψ k ( 3) ψ k ( 3) � ψ k ( 3) = −Ψ (1,2,..., N )
1

N ! k 1 ( )
Columns
⎯⎯⎯⎯⎯⎯⎯ 1&3→ −
2 1 N

� � � � �
ψk (N ) ψk (N ) ψk (N ) � ψk (N )
1 2 1 N

The only way a number can be equal to its opposite is if it is zero. Thus, in
enforcing antisymmetry the determinant also enforces the Pauli exclusion
principle. Thus, Slater determinants immediately lead us to the aufbau

Prepared By Kevin Boudreaux


146 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #27 page 4

principle we learned in Freshman chemistry: because we cannot place two


electrons in orbitals with the same quantum numbers, we are forced to fill up
orbitals with successively higher energies as we add more electrons: 1s↑, 1s↓,
2s↑, 2s↓, …. This filling up is dictated by the fact that the wavefunction is
antisymmetric, which, in turn, results from the fact that the electron is spin-½.
As it turns out, all half-integer spin particles (Fermions) have antisymmetric
wavefunctions and all integer spin particles (Bosons) have symmetric
wavefunctions. Imagine how different life would be if electrons were Bosons
instead of Fermions!
• Ψ is normalized if the orbitals ψ ki are normalized and two determinants are
orthogonal if they differ in any single orbital. These two facts are relatively
tedious to prove, but are useful in practice.
• The non­interacting energy of a Slater determinant is the energy of the
orbitals that make it up. Just as was the case for helium, our antisymmetric
wavefunction is a linear combination of degenerate non-interacting eigenstates:
Ψ (1,2,..., N ) =
1
(ψ k (1)ψ k ( 2 )ψ k ( 3) �ψ k ( N ) − ψ k ( 2 )ψ k (1)ψ k ( 3) �ψ k ( N )
N! 1 2 3 N 1 2 3 N

+ ψ k ( 3)ψ k (1)ψ k ( 2 ) �ψ k ( N ) − ψ k (1)ψ k ( 3)ψ k ( 2 ) �ψ k ( N ) ...)


1 2 3 N 1 2 3 N

Thus the determinant itself is an eigenstate of the non-interacting Hamiltonian


with the same eigenvalue as each state in the sum:
EΨ = Ek + Ek + Ek +... + Ek
1 2 3 N

We have only completed half of the independent particle picture at this point. We
have the noninteracting energy; what remains is the computation of the average
energy. To compute this, we would need to do a 2N-dimensional integral involving a
Slater Determinant on the left, the Hamiltonian in the middle (including all interaction
terms) and a Slater Determinant on the right. The book-keeping is quite tedious, but
can be worked out in the most general case. The result is that the energy breaks
down into terms that we already recognize:
Noninteracting Average
Energy Repulsion

N N
Ĥ = ∑ Ei + ∑ J�ij − K� ij
i=1 i< j

J�ij ≡ ∫∫ψ k* (1)ψ k* ( 2) ( ) k j ( ) 1 2 1 2


1 ψ 1 ψ 2 dr dr dσ dσ
i j r1 − r2 ki

K� ij ≡ ∫∫ψ k* (1)ψ k* (2) ( ) kj ( ) 1 2 1 2


1 ψ 2 ψ 1 dr dr dσ dσ
i j r1 − r2 ki

Prepared By Kevin Boudreaux


147 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #27 page 5

This energy expression has a nice intuitive feel to it: we get contributions from the
interaction of each electron orbital with the nucleus (first term) as well as from the
mutual repulsion of each pair of orbitals (second term). The repulsion has the
characteristic “Coulomb minus Exchange” form because of the antisymmetry of the
Slater determinant, which leads to a minus sign every time we exchange 1 and 2. For
example, when we expand the determinant in terms of product functions, we will have
a term like:
X ≡ ψ k1 (1)ψ k2 ( 2 )ψ k3 ( ? ) �ψ k N ( ? )
we will have a corresponding term with 1↔2 with a minus sign:
Y ≡ ψ k1 ( 2 )ψ k2 (1)ψ k3 ( ? ) �ψ k N ( ? )
When we compute the average repulsion we will have:
1 1
≈ ( X − Y ... ) ( X − Y ... ) dr1 dr2 ...
∫∫
r12 r12
The +XX and +YY terms will give us Coulomb integrals and the cross terms -XY and
-YX will give us exchange integrals. We note that, as defined, the Coulomb and
Exchange terms are both positive, so that the exchange integral always reduces the
repulsion energy for a Slater determinant. These arguments do not constitute a proof
of the energy expression above – they are merely intended to give you a general
feeling that this expression is plausible. Deriving the average energy actually takes
quite a bit of time and careful bookkeeping of different permutations (e.g. 1↔2↔4
versus 4↔2↔3 …). If you want to delve deeper into these kinds of things, we
recommend you take 5.73, which covers much of the material here in greater depth.

In any case, most of the time we need not worry about the details of how the average
energy expression is derived for a general determinant. Usually what we want to do is
use the formula to compute something interesting rather than re-derive it. Toward
this end we note that the Coulomb and exchange integrals above involve integration
over spin variables, which we can be done trivially. Basically, the integration over σ1
(σ2) will give unity if the left and right functions for electron 1 (electron 2) have the
same spin, and zero otherwise. For the Coulomb integral, the left and right functions
are the same, so the spin integration always gives 1 and we can get rid of the spin
integration:
1
J = ∫∫ ψ * (1)ψ * ( 2 ) ψ (1)ψ ( 2 ) dr dr
ij i j r −r i j 1 2
1 2

Prepared By Kevin Boudreaux


148 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #27 page 6

where here we use the shorthand i ≡ {ni , mi , li } to represent all the spatial quantum
numbers for the ith orbital, and similarly for the jth one. For the exchange integrals,
the left and right functions are different, and so the exchange integral is only
nonzero if the ith orbital and the jth orbital have the same spin part:
⎧ * 1
⎪ ∫∫ ψ i (1)ψ j ( 2 ) ψ ( 2 )ψ (1) dr dr if s = s
*
− i j 1 2 i j
K =⎨ r
1 2
r
ij

⎩ 0 Otherwise

In practice (for example on a Problem Set) it is useful to begin with the general
expression for the independent particle energy and then integrate out the spin part
to get a final expression that only involves the spatial integrals J and K rather than
the space-spin integrals J� and K� . For example, say we are interested in the
1s↑1s↓2s↑ configuration of lithium. The energy is given by:
N N
Hˆ = ∑ E + ∑ J� − K�
i ij ij
i =1 i< j J1s1s 0 J1s2s K1s2s J1s2s 0
= E + E + E + J� − K� + J� − K� + J� − K�
1s 1s 2s 1sα ;1s β 1sα ;1s β 1sα ; 2 sα 1sα ; 2 sα 1s β ; 2 sα 1s β ; 2 sα
= 2E + E + J + 2J −K
1s 2s 1s1s 1s 2 s 1s 2 s
Thus, in going from space-spin integrations to space integrations, we acquire a few
factors of 2, but the formula still looks similar to the expressions we’ve seen before.

The independent particle energy expression is extremely powerful: it allows us to


make rough predictions of things like atomic ionization energies and excitation
energies once we have the Coulomb and exchange integrals. The independent particle
energy also gives us a rigorous explanation of another freshman chemistry effect:
shielding. As we know, while 2s and 2p are degenerate for the hydrogen atom they
are not degenerate as far as the order of filling up orbitals. 2s comes first and 2p
comes second. As you know, physically this arises because the 2s orbital sees an
effective nuclear charge that is bigger than 2p because it is shielded less by the 1s
orbital. It turns out that the IPM has a simple means of explaining this result. If we
look at the 1s↑1s↓2px↑ configuration of lithium and go through exactly the same
manipulations as we did above for 1s↑1s↓2s↑ we get:
Ĥ = 2E + E +J + 2J −K
1s 2 px 1s1s 1s 2 p x 1s 2 p x
taking the difference between 1s↑1s↓2px↑ and 1s↑1s↓2s↑ gives:

Prepared By Kevin Boudreaux


149 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #27 page 7

ΔE = E1s2 2 p − E1s2 2s = 2E + E +J + 2J −K −
1s 2 px 1s1s 1s2 px 1s2 px

(2E1s + E2s + J1s1s + 2J1s2s − K1s2s )


= 2J − 2J +K −K ≠0
1s2 px 1s2s 1s2s 1s2 px
Thus, while 1s↑1s↓2px↑ and 1s↑1s↓2s↑ give the same energies if the electrons do not
interact, they give different energies once we include the average interaction.
Further, it is clear from the above expression that the relevant interactions are
between the 1s and 2p orbitals on the one hand and the 1s and 2s orbitals on the
other. Using our physical argument of shielding, we assert that the 1s and 2s orbitals
will repel each other less than the 1s and 2p orbitals, leading to a net stabilization of
1s↑1s↓2s↑ relative to 1s↑1s↓2px↑. We explore this idea in more detail on the problem
set.

Prepared By Kevin Boudreaux


150 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


1

MOLECULAR ORBITAL THEORY­ PART I


At this point, we have nearly completed our crash­course introduction to
quantum mechanics and we’re finally ready to deal with molecules. Hooray!
To begin with, we are going to treat what is absolutely the simplest
molecule we can imagine: H+2 . This simple molecule will allow us to work
out the basic ideas of what will become molecular orbital (MO) theory.

We set up our coordinate r


system as shown at right, e­
with the electron positioned rB rA
at r, and the two nuclei
positioned at points RA and
RB, at a distance R from one HA HB
another. The Hamiltonian is RA R RB
easy to write down: H2+ Coordinates
∇2 ∇2 1 1 1
Ĥ = − 1 ∇r2 − A − B − − +
2 2M A 2MB R̂ A − r̂ ˆ − r̂
R R̂ A − R̂ B
B

Electron HA HB e­­HA e­­HB HA­HB


Kinetic Kinetic Kinetic Attraction Attraction Repulsion
Energy Energy Energy
Now, just as was the case for atoms, we would like a picture where we can
separate the electronic motion from the nuclear motion. For helium, we
did this by noting that the nucleus was much heavier than the electrons
and so we could approximate the center of mass coordinates of the
system by placing the nucleus at the origin. For molecules, we will make a
similar approximation, called the Born­Oppenheimer approximation.
Here, we note again that the nuclei are much heavier than the electrons.
As a result, they will move much more slowly than the light electrons.
Thus, from the point of view of the electrons, the nuclei are almost
sitting still and so the moving electrons see a static field that arises
from fixed nuclei. A useful analogy here is that of gnats flying around on
the back of an elephant. The elephant may be moving, but from the gnats’
point of view, the elephant is always more or less sitting still. The
electrons are like the gnats and the nucleus is like the elephant.

Prepared By Kevin Boudreaux


151 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


2

The result is that, if we are interested in the electrons, we can to a good


approximation fix the nuclear positions – RA and RB – and just look at the
motion of the electrons in a molecule. This is the B­Oppenheimer
approximation, which is sometimes also called the clamped­nucleus
approximation, for obvious reasons. Once the nuclei are clamped, we can
make two simplifications of our Hamiltonian. First, we can neglect the
kinetic energies of the nuclei because they are not moving. Second, because
the nuclei are fixed, we can replace the operators R̂ A and R̂ B with the
numbers RA and RB. Thus, our Hamiltonian reduces to
∇2
(
Hˆ R A ,R B = − r −
el 2 ) 1

1
+
1
R A − r̂ R B − r̂ R A − R B
where the last term is now just a number – the electrostatic repulsion
between two protons at a fixed separation. The second and third terms
depends only on the position of the electron, r, and not its momentum, so
we immediately identify those as a potential and write:
∇2
( ) RA , RB
Hˆ el R A ,R B = − r +Veff
2
( r̂ ) +
1
RA − RB
This Hamiltonian now only contains operators for the electrons (hence the
subscript “el”), but the eigenvalues of this Hamiltonian depend on the
distance, R, between the two nuclei. For example, the figure below shows
the difference between the effective potentials the electron “feels” when
the nuclei are close together versus far apart:
R Small R Large

Veff(r)

Likewise, because the electron feels a different potential at each bond

distance R, the wavefunction will also depend on R. In the same limits as

above, we will have:


ψel(r)

R Small R Large

Prepared By Kevin Boudreaux


152 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


3

Finally, because the electron eigenfunction, ψel, depends on R then the


eigenenergy of the electron, Eel(R), will also depend on the bond length.
Mechanically, then, what we have to do is solve for the electronic
eigenstates, ψel, and their associated eigenvalues, Eel(R), at many different
fixed values of R. The way that these eigenvalues change with R will tell us
about how the energy of the molecule changes as we stretch or shrink the
bond. This is the central idea of the Born­Oppenheimer approximation, and
it is really very fundamental to how chemists think about molecules. We
think about classical point­like nuclei clamped at various different positions,
with the quantum mechanical electrons whizzing about and gluing the nuclei
together. When the nuclei move, the energy of the system changes because
the energies of the electronic orbitals change as well. There are certain
situations where this approximation breaks down, but for the most part the
Born­Oppenheimer picture provides an extremely useful and accurate way to
think about chemistry.

How are we going to solve for these eigenstates? It should be clear that
looking for exact solutions is going to lead to problems in general. Even for
H2+ the solutions are extremely complicated and for anything more complex
than H2+exact solutions are impossible. So we have to resort to
approximations again. The first thing we note is that if we look closely at
our intuitive picture of the H2+ eigenstates above, we recognize that these
molecular eigenstates look very much like the sum of the 1s atomic orbitals
for the two hydrogen atoms. That is, we note that to a good approximation
we should be able to write:
ψ el ( r ) ≈ c11s A ( r ) + c2 1sB ( r )
where c1 and c2 are constants. In the common jargon, the function on the
left is called a molecular orbital (MO), whereas the functions on the right
are called atomic orbitals (AOs). If we write our MOs as sums of AOs, we
are using what is called the linear combination of atomic orbitals (LCAO)
approach. The challenge, in general, is to determine the “best” choices for c1
and c2 – for H2+ it looks like the best choice for the ground state will be
c1=c2. But how can we be sure this is really the best we can do? And what
about if we want something other than the ground state? Or if we want to
describe a more complicated molecule like HeH+2?

Prepared By Kevin Boudreaux


153 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


4

THE VARIATIONAL PRINCIPLE


In order to make further progress, we will use the Variational Principle
to predict a better estimate of the ground state energy. This method is
very general and its use in physical chemistry is widespread. Assume you
have a Hamiltonian (such as the Helium atom) but you don’t know the
ground state energy E0 and or ground state eigenfunction φ0.
Ĥφ0 = E0φ0 ⇒ Ĥ = ∫ φ0*Ĥ φ0dτ = ∫ φ0*E0φ0dτ = E0
Now, assume we have a guess, ψ , at the ground state wavefunction, which we
will call the trial wavefunction. Compute the average value of the energy for
the trial wavefunction:
∫ψ *Ĥψ dτ = ψ * Hˆψ dτ (if ψ normalized)
Eavg = ∫
∫ψ *ψ dτ
the Variational Theorem tells us that Eavg ≥ E0 for any choice of the trial
function ψ! This make physical sense, because the ground state energy is,
by definition, the lowest possible energy, so it would be nonsense for the
average energy to be lower.

SIDEBAR: PROOF OF VARIATIONAL THEOREM


Expand ψ (assumed normalized) as linear combination of the unknown
eigenstates, φn , of the Hamiltonian:
ψ = ∑ anφn
n

Note that in practice you will not know these eigenstates. The important
point is that no matter what function you choose you can always expand it in
terms of the infinite set of orthonormal eigenstates of Ĥ .

∫ψ *ψ dτ = ∑ an am ∫ φn *φm dτ = ∑ an amδ mn = ∑ an
2
* *
=1
n ,m n ,m n

Eavg = ∫ψ *Ĥψ dτ = ∑ an*am ∫ φn *Ĥ φm dτ = ∑ an*am ∫ φn *Emφm dτ


n ,m n ,m

= ∑ an*am Emδ mn = ∑ an En
2

n ,m n
Now, subtracting the ground state energy from the average

Prepared By Kevin Boudreaux


154 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


5

E0 = 1• E0 = ∑ an E0
2

⇒ Eavg − E0 = ∑ an En − ∑ an E0 = ∑ an ( En − E0 ) ≥ 0 since En ≥ E0
2 2 2

n n n
Where, in the last line we have noted that the sum of terms that are non­
negative must also be non­negative. It is important to note that the
equals sign is only obtained if an=0 for all states that have En>E0. In this
situation, ψ actually is the ground state of the system (or at least one
ground state, if the ground state is degenerate).

The variational method does two things for us. First, it gives us a means of
comparing two different wavefunctions and telling which one is closer to the
ground state – the one that has a lower average energy is the better
approximation. Second, if we include parameters in our wavefunction
variation gives us a means of optimizing the parameters in the following way.
Assume that ψ depends on some parameters c – such as is the case for our
LCAO wavefunction above. We’ll put the parameters in brackets ­ψ[c] – in
order to differentiate them from things like positions that are inside of
parenthesis ­ψ(r).Then the average energy will depend on these parameters:
ψ [ c ] Ĥψ [c ] dτ
(c) = ∫
*

Eavg
∫ψ [c] *ψ [c] dτ
Note that, using the variational principle, the best parameters will be the
ones that minimize the energy. Thus, the best parameters will satisfy
∂Eave ( c ) ∂ ∫ ψ [ c ] Hψ [c ] dτ
*
ˆ
= =0
∂ci ∂ci ∫ ψ [c ]*ψ [c ] dτ
Thus, we can solve for the optimal parameters without knowing anything
about the exact eigenstates!

Let us apply this in the particular case of our LCAO­MO treatment of H2+.
Our trial wavefunction is:
ψ el [ c ] = c11s A + c2 1sB

Prepared By Kevin Boudreaux


155 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


6

where c=(c1 c2). We want to determine the best values of c1 and c2 and
the variational theorem tells us we need to minimize the average energy
to find the right values. First, we compute the average energy. The
numerator gives:
∫ψ el Ĥ elψ el dτ = ∫ ( c11sA + c21sB ) Ĥ ( c11sA + c21sB ) dτ
* *

= c1*c1 ∫ 1s A Ĥ el 1s A dτ + c1* c2 ∫ 1s A Ĥ el 1sB dτ + c2* c1 ∫ 1sB Ĥ el 1s A dτ + c2* c2 ∫ 1sB Ĥ el 1sB dτ

≡H11 ≡H12 ≡H21 ≡H22


≡ c1* H11c1 + c1* H12 c2 + c2* H 21c1 + c2* H 22 c2
while the normalization integral gives:
∫ψ ψ el dτ = ∫ ( c11s A + c2 1sB ) ( c11s A + c2 1sB ) dτ
* *
el

= c1*c1 ∫ 1s A1s A dτ + c1*c2 ∫ 1s A1sB dτ + c2* c1 ∫ 1sB 1s A dτ + c2* c2 ∫ 1sB 1sB dτ

≡S11 ≡S12 ≡S21 ≡S22


≡ c1* S11c1 + c1* S12 c2 + c2* S21c1 + c2* S22 c2

So that the average energy takes the form:


c * H c + c * H c + c2* H 21c1 + c2* H 22 c2
Eavg ( c ) = 1 * 11 1 1 * 12 2
c1 S11c1 + c1 S12 c2 + c2* S21c1 + c2* S22 c2
We note that there are some simplifications we could have made to this
formula: for example, since our 1s functions are normalized S11=S22=1. By
not making these simplifications, our final expressions will be a little more
general and that will help us use them in more situations.

Now, we want to minimize this average energy with respect to c1 and c2.
Taking the derivative with respect to c1 and setting it equal to zero [Note:
when dealing with complex numbers and taking derivatives one must treat
variables and their complex conjugates as independent variables. Thus d/dc1
has no effect on c1*]:
∂E avg c1* H11 + c2* H 21
=0= *
∂c1 c1 S11c1 + c1* S12 c2 + c2* S21c1 + c2* S22 c2
c1* H11c1 + c1* H12 c2 + c2* H 21c1 + c2* H 22 c 2
− (c*
S + c2* S21 )
(c S c + c S c + c2 S21c1 + c2 S22 c2 )
2 1 11
* * * *
1 11 1 1 12 2

Prepared By Kevin Boudreaux


156 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


7

c1* H11c1 + c1* H12 c2 + c2* H 21c1 + c2* H 22 c2 *


⇒ 0 = ( c1* H11 + c2* H 21 ) − ( c1 S11 + c2* S21 )
c1 S11c1 + c1 S12 c2 + c2 S21c1 + c2 S22 c2
* * * *

⇒ 0 = ( c1* H11 + c2* H 21 ) − Eavg ( c1* S11 + c2* S21 )


Applying the same procedure to c2:
∂Eavg c1* H12 + c2* H 22

=0= *
∂c2 c1 S11c1 + c1* S12 c2 + c2* S21c1 + c2*
S22 c2
c1* H11c1 + c1* H12 c2 + c2* H 21c1 + c2* H 22 c2
− (c*
S + c2* S22 )
(c S c + c S c + c2 S21c1 + c2 S22 c2 )
2 1 12
* * * *
1 11 1 1 12 2

c1* H11c1 + c1* H12 c2 + c2* H 21c1 + c2* H 22 c2 *


⇒ 0 = ( c H12 + c2 H 22 ) −
* *
( c1 S12 + c2* S22 )
c1 S11c1 + c1 S12 c2 + c2 S21c1 + c2 S22 c2
1 * * * *

⇒ 0 = ( c1* H12 + c2* H 22 ) − Eavg ( c1* S12 + c2* S22 )


We notice that the expressions above look strikingly like matrix­vector
operations. We can make this explicit if we define the Hamiltonian matrix:
⎛ 1s A Ĥ el 1sB dτ ⎞
H≡⎜
⎛ H11 H12 ⎞ ⎜ 1s A Ĥ el 1s A dτ ∫ ∫⎟
⎟≡⎜ ⎟
H
⎝ 21 H 22 ⎠
⎝ ∫
⎜ 1sB Ĥ el 1s A dτ ∫
1sB Ĥ el 1sB dτ ⎟

and the Overlap matrix:
⎛ 1s A 1sB dτ ⎞
S≡⎜
⎛ S11 S12 ⎞ ⎜ 1s A 1s A dτ ∫ ⎟ ∫
⎟≡⎜ ⎟
⎝ S21 S22 ⎠ ⎜ 1sB 1s A dτ
⎝ ∫
1sB 1sB dτ ⎟
⎠ ∫
Then the best values of c1 and c2 satisfy the matrix eigenvalue equation:
⎛H H12 ⎞ ⎛S S12 ⎞
(c1
*
)
c2* ⎜ 11
⎝ H 21

H 22 ⎠
(
= Eavg c1* )
c2* ⎜ 11
⎝ S21 S22 ⎟⎠

Which means that:


∂Eavg
= 0 ⇔ c† iH = Eavg c† iS Eq. 1
∂c
This equation doesn’t look so familiar yet, so we need to massage it a bit.
First, it turns out that if we had taken the derivatives with respect to c1*
and c2* instead of c1 and c2, we would have gotten a slightly different
equation:
⎛ H11 H12 ⎞ ⎛ c1 ⎞ ⎛S S12 ⎞ ⎛ c1 ⎞
⎜H ⎟ ⎜ ⎟ = Eavg ⎜ 11 ⎟⎜ ⎟
⎝ 21 H 22 ⎠ ⎝ c2 ⎠ ⎝ S21 S22 ⎠ ⎝ c2 ⎠
or

Prepared By Kevin Boudreaux


157 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


8

∂Eavg
= 0 ⇔ H ic = Eavg Sic Eq. 2
∂c *
Taking the derivatives with respect to c1* and c2* is mathematically
equivalent to taking the derivatives with respect to c1 and c2 [because we
can’t change c1 without changing its complex conjugate, and vice versa].
Thus, the two matrix equations (Eqs. 1&2) above are precisely equivalent, and
the second version is a little more familiar. We can make it even more
familiar if we think about the limit where 1sA and 1sB are orthogonal (e.g.
when the atoms are very, very far apart). Then we would have for the
Overlap matrix:
⎛ 1s 1s dτ 1s A 1sB dτ ⎞ ⎛ 1 0 ⎞
S≡ ⎜ ∫
A A
⎟=∫ =1
⎜⎜ ⎟⎟ ⎝⎜ 0 1 ⎠⎟
⎝ ∫
1sB 1s A dτ 1sB 1sB dτ
⎠ ∫
Thus, in an orthonormal basis the overlap matrix is just the identity matrix
and we can write Eq. 2 as:
∂Eavg
= 0 ⇔ H ic = Eavg c
∂c *
Now this equation is in a form where we certainly recognize it: this is an
eigenvalue equation. Because of its close relationship with the standard
eigenvalue equation, Eq. 2 is usually called a Generalized Eigenvalue
Equation.

In any case, we see the quite powerful result that the Variational theorem
allows us to turn operator algebra into matrix algebra. Looking for the
lowest energy LCAO state is equivalent to looking for the lowest eigenvalue
of the Hamiltonian matrix H. Further, looking for the best c1 and c2 is
equivalent to finding the lowest eigenvector of H.

Let’s go ahead and apply what we’ve learned to obtain the MO coefficients c1
and c2 for H2+. At this point we make use of several simplifications. The
off­diagonal matrix elements of H are identical because the Hamiltonian is
Hermitian and the orbitals are real:

(∫ 1s ) = ∫ 1s Hˆ 1s dτ ≡ V
*
∫ 1s A Hˆ el 1sB dτ = Hˆ el 1s A dτ
*
B B el A 12

Meanwhile, the diagonal elements are also equal, but for a different reason.
The diagonal elements are the average energies of the states 1sA and 1sB. If
these energies were different, it would imply that having the electron on one

Prepared By Kevin Boudreaux


158 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


9

side of H2+ was more energetically favorable than having it on the other,
which would be very puzzling. So we conclude

∫ 1s A Ĥ el 1s A dτ ∫
= 1sB Ĥ el 1sB dτ ≡ ε
Finally, we remember that 1sA and 1sB are normalized, so that

∫ 1s 1s
A A dτ ∫
= 1sB 1sB dτ = 1
and because the 1s orbitals are real, the off­diagonal elements of S are also
the same:

∫ ∫
S12 = 1s A 1sB dτ = 1sB 1s A dτ = S21 .
Incorporating all these simplifications gives us the generalized eigenvalue
equation:
⎛ ε V12 ⎞ ⎛ c1 ⎞ ⎛ 1 S12 ⎞ ⎛ c1 ⎞
⎜V ⎟ ⎜ ⎟ = Eavg ⎜ ⎟⎜ ⎟
⎝ 12 ε ⎠ ⎝ c2 ⎠ ⎝ S21 1 ⎠ ⎝ c2 ⎠
All your favorite mathematical programs (Matlab, Mathematica, Maple,
MathCad…) are capable of solving for the generalized eigenvalues and
eigenvectors, and for more complicated cases we suggest you use them.
However, this case is simple enough that we can solve it by guess­and test.
Based on our physical intuition above, we guess that the correct eigenvector
will have c1=c2. Plugging this in, we find:
⎛ ε V12 ⎞ ⎛ c1 ⎞ ⎛ 1 S12 ⎞ ⎛ c1 ⎞
⎜ ⎟ ⎜ ⎟ = Eavg ⎜ ⎟⎜ ⎟
⎝ V12 ε ⎠ ⎝ c1 ⎠ ⎝ S21 1 ⎠ ⎝ c1 ⎠
⎛ (ε + V12 ) c1 ⎞ ⎛ (1 + S12 ) c1 ⎞
⇒⎜ ⎟ = Eavg ⎜ ⎟
⎝ (ε + V12 ) c1 ⎠ ⎝ (1 + S12 ) c1 ⎠
ε + V12
⇒ Eavg = ≡ Eσ
1 + S12
Thus, our guess is correct and one of the eigenvectors of this matrix has
c1=c2. This eigenvector is the σ­bonding state of H2+, and we can write down
the associated orbital as:
ψ elσ = c11s A + c2 1sB = c11s A + c11sB ∝ 1s A + 1sB
where in the last expression we have noted that c1 is just a normalization
constant. In freshman chemistry, we taught you that the σ­bonding orbital
existed, and this is where it comes from.

We can also get the σ∗­antibonding orbital from the variational procedure.
Since the matrix is a 2x2 it has two unique eigenvalues: the lowest one
(which we just found above) is bonding and the other is antibonding. We can

Prepared By Kevin Boudreaux


159 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


10

again guess the form of the antibonding eigenvector, since we know it has
the characteristic shape +/­, so that we guess the solution is c1=­c2:
⎛ ε V12 ⎞ ⎛ c1 ⎞ ⎛ 1 S12 ⎞ ⎛ c1 ⎞
⎜ ⎟⎜ ⎟ = Eavg ⎜ ⎟⎜ ⎟
⎝ V12 ε ⎠ ⎝ − c1 ⎠ ⎝ S21 1 ⎠ ⎝ − c1 ⎠
⎛ (ε − V12 ) c1 ⎞ ⎛ (1 − S12 ) c1 ⎞
⇒⎜ ⎟ = Eavg ⎜ ⎟
⎝ − (ε − V12 ) c1 ⎠ ⎝ − (1 − S12 ) c1 ⎠
ε − V12
⇒ Eavg = = Eσ *

1 − S
12
so, indeed the other eigenvector
has c1=­c2. The corresponding
antibonding orbital is given by: ψσ(r)
ψ elσ * = c11s A + c2 1sB = c11s A − c11sB ∝ 1s A − 1sB

where we note again that c1 is

just a normalization constant.

Given these forms for the

bonding and antibonding orbitals, ψσ(r)

we can draw a simple picture for

the H2+ MOs (see right).

We can incorporate the energies obtained above into a simple MO diagram of

H2+:

ε − V12
Eσ∗=
1 − S12

E1sA=ε E1sB=ε

ε + V12
Eσ=
1 + S12
On the left and right, we draw the energies of the atomic orbitals (1sA and

1sB) that make up our molecular orbitals (σ and σ*) in the center. We note

Prepared By Kevin Boudreaux


160 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #28


11

that when the atoms come together the energy of the bonding and
antibonding orbitals are shifted by different amounts:
ε − V12 ε − V12 ε (1 − S12 ) ε S12 − V12
Eσ * − E1s = −ε = − =
1 − S12 1 − S12 1 − S12 1 − S12
ε + V12 ε (1 + S12 ) ε + V12 ε S12 − V12
E1s − Eσ * = ε − = − =
1 + S12 1 + S12 1 + S12 1 + S12
Now, S12 is the overlap between two 1s orbitals. Since these orbitals are
never negative, S12 must be a positive number. Thus, the first denominator
is greater than the second, from which we conclude
ε S − V12 ε S12 − V12
Eσ * − E1s = 12 > = E1s − Eσ *
1 − S12 1 + S12
Thus, the antibonding orbital is destabilized more than the bonding orbital is
stabilized. This conclusion need not hold for all diatomic molecules, but it is
a good rule of thumb. This effect is called overlap repulsion. Note that in
the special case where the overlap between the orbitals is negligible, S12
goes to zero and the two orbitals are shifted by equal amounts. However,
when is S12 nonzero there are two effects that shift the energies: the
physical interaction between the atoms and the fact that the 1sA and 1sB
orbitals are not orthogonal. When we diagonalize H, we account for both
these effects, and the orthogonality constraint pushes the orbitals upwards
in energy.

Prepared By Kevin Boudreaux


161 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30


1

MOLECULAR ORBITAL THEORY­ PART II

For the simple case of the one­electron bond in H2+ we have seen that using
the LCAO principle together with the variational principle led to a recipe for
computing some approximate orbitals for a system that would be very
difficult to solve analytically. To generalize this to the more interesting
case of many electrons, we take our direction from our experience with the
independent particle model (IPM) applied to atoms and we build up
antisymmetrized wavefunctions out of the molecular orbitals. This is the
basic idea behind molecular orbital theory – there are many variations on the
central theme, but the same steps are always applied. Rather than go step­
by­step and deal with H2 and then Li2 and then LiH … we will instead begin by
stating the general rules for applying MO theory to any system and then
proceed to show some illustrations of how this works out in practice.

1) Define a basis of atomic orbitals


For H2+ our atomic orbital basis was simple: we used the 1s functions
from both hydrogen atoms and wrote our molecular orbitals as linear
combinations of our basis functions:
ψ = c11s A + c2 1sB
Note that the AO basis determines the dimension of our MO vector and
also determines the quality of our result – if we had chosen the 3p
orbitals instead of the 1s orbitals, our results for H2+ would have been
very wrong!

For more complicated systems, we will require a more extensive AO basis.


For example, in O2 we might want to include all the 2s and 2p orbitals on
both oxygens, in which case our MOs would take the form
ψ = c1 2s A + c2 2 pxA + c3 2 p yA + c4 2 pzA + c5 2sB + c6 2 pxB + c7 2 p yB + c8 2 pzB
Meanwhile, for methane we might want to include the 1s functions on all
four hydrogens and the 2s and 2p functions on carbon:
ψ = c11s1 + c2 1s2 + c3 1s3 + c4 1s4 + c5 2s + c6 2 px + c7 2 p y + c8 2 pz
In the general case, we will write:
N
ψ = ∑ ciφiAO
i=1

and represent our MOs by column vectors:

Prepared By Kevin Boudreaux


162 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30


2

⎛ c1 ⎞
⎜ ⎟
� c
ψ =⎜ 2 ⎟
⎜ ... ⎟
⎜ ⎟
⎝ cN ⎠
We note that for the sake of accuracy it is never a bad idea to include
more AO functions than you might think necessary – more AO functions
will always lead to more accurate results. The price is that the more
accurate computations also tend to be more complicated and time
consuming. To illustrate, note that we could have chosen to write the H2+
MOs as linear combinations of four functions – the 1s and 2s states on
each atom:
ψ = c11s A + c2 1sB + c3 2s A + c4 2sB
Now, when we use the variational principle to get the coefficients of the
lowest MO , c0, we are guaranteed that there is no set of coefficients
that will give us a lower energy. This is the foundation of the variational
method. Note that one possible set of coefficients is c3=c4=0, in which
case our 4­function expansion reduces to the 2­function expansion above.
Thus, the variationally optimal 4­function MO will always have an energy
less than or equal to the optimal 2­function MO. As a result, the
expansion with four functions allows the approximate MO to get closer to
the ground state energy. This makes sense, as the four AO expansion
has more flexibility than the constrained two AO expansion used
previously. The reason we didn’t use the four function expansion from
the beginning is that all the algebra is twice as difficult when we use four
functions as two: the vectors are twice as long, the matrices are twice as
big…. At least for a first try, it is generally good to start with the
smallest conceivable set of AOs for performing a calculation. If higher
accuracy is required, a longer expansion can be tried.
2) Compute the relevant matrix representations
For H2+ we had to compute two matrices – the Hamiltonian and the
overlap, which were both 2­by­2 by virtue of the two AO basis functions:
⎛ 1s A Ĥ el 1sB dτ ⎞
H≡⎜

⎛ H11 H12 ⎞ ⎜ 1s A Ĥ el 1s A dτ ∫ ⎟
⎟≡⎜ ⎟
⎝ ∫
⎝ H 21 H 22 ⎠ ⎜ 1sB Ĥ el 1s A dτ ∫
1sB Ĥ el 1sB dτ ⎟

⎛ 1s A 1sB dτ ⎞
S≡⎜

⎛ S11 S12 ⎞ ⎜ 1s A 1s A dτ ∫ ⎟
⎟≡⎜ ⎟
S
⎝ 21 S22 ⎠
⎝ ∫
⎜ 1sB 1s A dτ 1sB 1sB dτ ⎟∫ ⎠

Prepared By Kevin Boudreaux


163 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30


3

In the general case, the Hamiltonian and overlap become N­by­N

matrices of the form:

⎛ φ1AO Ĥ φ1AO φ1AO Ĥ φNAO ⎞


⎛ H11 H12 ... H1N ⎞ ⎜ ∫ φ1AO Ĥ φ2AO∫ ⎟ ∫
⎜ ⎟ ⎜ AO ˆ AO AO ˆ AO ⎟
H 22 ... H 2 N ⎟ ⎜ φ2 H φ1 ∫ φ2 H φ2 ∫ φ2 H φN ⎟ ∫
AO ˆ AO
H
H ≡ ⎜ 21 ≡
⎜ ... ... ... ... ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
⎝ H N1 H N 2 ... H NN ⎠ ⎜ φ AO Ĥ φ AO AO ⎟
∫ φN Ĥ φ2 ∫ φN Ĥ φN ∫
AO AO AO
⎝ N 1 ⎠
⎛ φ φ φ1 φN ⎞
∫ φ1 φ2 ∫ ∫
AO AO AO AO AO AO
⎛ S11 S12 ... S1N ⎞ ⎜ 1 1 ⎟
⎜ ⎟ ⎜ AO AO AO AO ⎟
S22 ... S2 N ⎟ ⎜ φ2 φ1 ∫ φ2 φ2 ∫φ2 φN ⎟ ∫
AO AO
S
S ≡ ⎜ 21 ≡
⎜ ... ... ... ... ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
⎝ S N1 S N 2 ... S NN ⎠ ⎜ φ AOφ AO AO AO ⎟
∫ φN φ2 ∫φ N φN ∫
AO AO
⎝ N 1 ⎠
This step is where much of the hard work is done in most MO
calculations. Not only can the integrals between the different AO
functions be very tricky to work out, there are a lot of them to be
computed – N2 of them, to be exact! This hard work is best done in an
automated fashion by a computer, and in practice we will usually give you 
explicit values for the matrix elements for this step. However, it is
important for you to realize what the matrix elements mean. The
diagonal elements of H represent the average energies of putting
electrons in each AO and the off­diagonal terms tell us how strongly
coupled one AO is to another. The diagonal elements of S are
normalization integrals and the off­diagonal terms tell us how much
spatial overlap there is between the different AOs
3) Solve the generalized eigenvalue problem
For every MO problem, the central step is determining the MOs, which
always involves solving the generalized eigenvalue problem:
H icα = Eα Sicα
The eigenvalues from this equation are the MO energies. The
eigenvectors are the coefficients of the molecular orbitals, written as
sums of AOs:
N

ψ α ( r ) = ∑ ciα φiAO ( r
)
i =1

In general, we will obtain N molecular orbitals out of N atomic orbitals.


This step is precisely the same as what we did for H2+, just generalized
to the N­orbital case. We note that for anything larger than a 2­by­2, it

Prepared By Kevin Boudreaux


164 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30


4

is usually best to ask a computer to solve the generalized eigenvalue


problem for you.
4) Occupy the orbitals according to a stick diagram
At this point, we must depart from the H2+ model and begin to account
for the fact that we have multiple electrons. To do so, we follow the
prescription of the independent particle model and build a Slater
determinant out of our orbitals. However, whereas for atoms we built
the determinant out of atomic orbitals, for molecules we will build the
determinant out of molecular orbitals:
ψ 1↑ (1) ψ 1↓ (1) ... ψ N ↓ (1)
ψ 1↑ ( 2 ) ψ 1↓ ( 2 ) ... ψ N ↓ ( 2 )
Ψ≡
... ... ... ...
ψ 1↑ ( N ) ψ 1↓ ( N ) ... ψ N ↓ ( N )
As was the case for atoms, it is much easier to reason in terms of stick
diagrams, rather than write out all of the orbitals in determinant form.
So, for example, we would associate a stick diagram like this
ψ2

ψ1
with a determinant:

ψ 1↑ (1) ψ 1↓ (1) ψ 2↑ (1) ψ 2↓ (1)


ψ 1↑ ( 2 ) ψ 1↓ ( 2 ) ψ 2↑ ( 2 ) ψ 2↓ ( 2 )
Ψ (1, 2, 3, 4 ) ≡
ψ 1↑ ( 3) ψ 1↓ ( 3) ψ 2↑ ( 3) ψ 2↓ ( 3)
ψ 1↑ ( 4 ) ψ 1↓ ( 4 ) ψ 2↑ ( 4 ) ψ 2↓ ( 4 )
But all the information we would need is contained in the stick diagram
and, of course, the MOs.

5) Compute the energy


There are a variety of ways to compute the energy once the MOs have
been obtained. The simplest is to use the non­interacting particle picture
we used for atoms. Here, the energy of N electrons is just given by the
sum of the energies of the N orbitals that are occupied:

Prepared By Kevin Boudreaux


165 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30


5

N
E = ∑ Ei
i=1

A more accurate way is to use the independent particle model to add


an average electron­electron repulsion to the energy:
N N
E = Ĥ = ∑ Ei +∑ J�ij − K� ij
i=1 i< j
Where now the Coulomb and exchange integrals use molecular orbitals
rather than atomic orbitals:
J�ij ≡ ∫∫ψ i* (1)ψ *j ( 2) ( ) j( ) 1 2 1 2
1 ψ 1 ψ 2 dr dr dσ dσ
r1 − r2 i

K� ij ≡ ∫∫ψ i* (1)ψ *j ( 2) ( ) j( ) 1 2 1 2
1 ψ 2 ψ 1 dr dr dσ dσ
r1 − r2 i
In fact, as we will see later on, there are even more elaborate ways to
obtain the energy from an MO calculation. When we work things out
by hand, the non­interacting picture is easiest and we will usually work
in that approximation when dealing with MOs.

Diatomic molecules
As a first application of MO theory, it is useful to consider first­row
diatomic molecules (B2, C2, N2,O2, CO,CN, NO, etc.), which actually map
rather nicely on to an MO picture. We’ll go step­by step for the generic
“AB” diatomic to show how this fits into the MO theory framework.

1) Define a basis of atomic orbitals. To begin with, one would consider a


set consisting of 10 atomic orbitals – 5 on A and 5 on B:
ψ = c11s A + c2 1sB + c3 2s A + c4 2 sB + c5 2 pzA + c6 2 pzB + c7 2 p yA + c8 2 p yB + c9 2 pxA + c10 2 pxB

However, for all the diatomics above, the 1s orbitals on both atoms will
be doubly occupied. Since we will primarily be interested in comparing the
MO descriptions of different diatomics the eternally occupied 1s orbital
will have no qualitative effect on our comparisons. It is therefore
customary to remove the 1s orbitals from the expansion:
ψ = c1 2s A + c2 2sB + c3 2 pzA + c4 2 pzB + c5 2 p yA + c6 2 p yB + c7 2 p xA + c8 2 pxB
The latter approximation is referred to as the valence electron or
frozen core approximation. The advantage is that it reduces the length
of our vectors from 10 to 8.

Prepared By Kevin Boudreaux


166 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30


6

2) Compute the Matrix Representations. Here, we have the rather


daunting task of computing two 8­by­8 matrices. As mentioned above, we
won’t be concerned in this class about filling in precise values for matrix
elements here. However, we will be very interested in obtaining the
proper shape of the matrix by determining which matrix elements are
zero and which are not. The Hamiltonian takes the shape:
⎛ ⎞
⎜ ∫ s ˆ
A Hs A ∫s ˆ
A HsB ∫s ˆ
A Hp zA ∫s ˆ
A Hp zB ∫s ˆ
A Hp yA ∫s ˆ
A Hp yB ∫s ˆ
A Hp xA ∫s ˆ
A Hp xB

⎜ ⎟
⎜ ∫s ˆ
B Hs A ∫s ˆ
B Hs B ∫s ˆ
B Hp zA ∫s ˆ
B Hp zB ∫s ˆ
B Hp yA ∫s ˆ
B Hp yB ∫s ˆ
B Hp xA ∫s ˆ
B Hp xB ⎟
⎜ ⎟
⎜∫p

ˆ
zA Hs A ∫p ˆ
zA HsB ∫p ˆ
zA Hp zA ∫p ˆ
zA Hp zB ∫p ˆ
zA Hp yA ∫p ˆ
zA Hp yB ∫p ˆ
zA Hp xA ∫p ˆ
zA Hp xB ⎟

⎜∫p
H ≡ ⎜
ˆ
zB Hs A ∫p ˆ
zB HsB ∫p ˆ
zB Hp zA ∫p ˆ
zB Hp zB ∫p ˆ
zB Hp yA ∫p ˆ
zB Hp yB ∫p ˆ
zB Hp xA ∫p
zB Hp xB ⎟
ˆ

⎜ ⎟
⎜∫p ˆ
yA Hs A ∫p ˆ
yA HsB ∫p ˆ
yA Hp zA ∫p ˆ
yA Hp zB ∫p ˆ
yA Hp yA ∫p ˆ
yA Hp yB ∫p ˆ
yA Hp xA ∫p ˆ
yA Hp xB

⎜ ⎟
⎜∫ p ˆ
yB Hs A ∫p ˆ
yB HsB ∫p ˆ
yB Hp zA ∫p ˆ
yB Hp zB ∫p ˆ
yB Hp yA ∫p ˆ
yB Hp yB ∫p ˆ
yB Hp xA ∫p ˆ
yB Hp xB ⎟
⎜ ⎟
⎜∫p

ˆ
xA Hs A ∫p ˆ
xA HsB ∫p ˆ
xA Hp zA ∫p ˆ
xA Hp zB ∫p ˆ
xA Hp yA ∫p ˆ
xA Hp yB ∫p ˆ
xA Hp xA ∫p ˆ
xA Hp xB ⎟


⎝∫p ˆ
xB Hs A ∫p ˆ
xB HsB ∫p ˆ
xB Hp zA ∫p ˆ
xB Hp zB ∫p ˆ
xB Hp yA ∫p ˆ
xB Hp yB ∫p ˆ
xB Hp xA ∫p
xB
ˆ
Hp xB ⎟

We assume, for simplicity, that the AB­bond

lies along the z­axis. Then it is relatively easy y

to see that the molecule is symmetric upon

reflection around the x and y axes. As a

result, the Hamiltonian for AB is also z

A B
symmetric (even) with respect to reflection

about x and y. Similarly, the s and p orbitals ­x

all have definite reflection symmetries:

X Reflection Y Reflection
Hamiltonian + +
s + +
pz + +
py + ­
px ­ +

Further, we note that if we perform an integral, if the integrand is odd

with respect to reflection about either x or y the integrand will be zero.

Prepared By Kevin Boudreaux


167 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30 7

This knocks out a bunch of integrals for us. For example, looking at

reflection about x:

s Hpˆ

= ( + )( + )( −) = − ⇒ 0
A/ B xA / B

∫p zA / B
ˆ
Hp xA / B = ( + )( + )( − ) = − ⇒ 0

∫p yA / B
ˆ
Hp xA / B = ( + )( + )( − ) = − ⇒ 0

∫p xA / B
ˆ
Hp xA / B = ( − )( + )( − ) = + ⇒ ≠0
We have analogous expressions for reflection about y zero several more

integrals. The result is that the Hamiltonian simplifies to

⎛ s A Hs ⎞
⎜ ∫ ˆ
A ∫s ˆ
A HsB ∫s ˆ
A Hp zA ∫s ˆ
A HpzB 0 0 0 0

⎜ ⎟
∫ ˆ
⎜ sB Hs A ∫s ˆ
B HsB ∫s ˆ
B HpzA ∫s ˆ
B Hp zB 0 0 0 0 ⎟
⎜ ⎟


⎜ pzA Hsˆ
A ∫p ˆ
zA HsB ∫p ˆ
zA Hp zA ∫p ˆ
zA Hp zB 0 0 0 0 ⎟

H≡⎜

⎜ pzB Hsˆ
A ∫p ˆ
zB HsB ∫p ˆ
zB Hp zA ∫pzB Hp zB
ˆ 0 0 0 0 ⎟

⎜ ⎟

0 0 0 0 ∫p ˆ
yA Hp yA ∫p ˆ
yA Hp yB 0 0

⎜ ⎟
⎜ 0 0 0 0 ∫p ˆ
yB Hp yA ∫p ˆ
yB Hp yB 0 0 ⎟
⎜ ⎟


0 0 0 0 0 0 ∫p ˆ
xA Hp xA ∫ ˆ
pxA Hp xB ⎟



0 0 0 0 0 0 ∫p ˆ
xB Hp xA ∫ ˆ
pxB Hp xB ⎟

which we write:
⎛ H11 H12 H13 H14 0 0 0 0 ⎞
⎜H H 22 H 23 H 24 0 0 0 0 ⎟⎟
⎜ 21
⎜ H 31 H 32 H 33 H 34 0 0 0 0 ⎟
⎜ ⎟
H H 42 H 43 H 44 0 0 0 0 ⎟
H ≡ ⎜ 41
⎜ 0 0 0 0 H 55 H 56 0 0 ⎟
⎜ ⎟
⎜ 0 0 0 0 H 65 H 66 0 0 ⎟
⎜ 0 0 0 0 0 0 H 77 H 78 ⎟
⎜⎜ ⎟
⎝ 0 0 0 0 0 0 H 87 H 88 ⎟⎠
It is easy to show that the overlap matrix has the same overall shape

Prepared By Kevin Boudreaux


168 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30


8

⎛ S11 S12 S13 S14 0 0 0 0 ⎞


⎜S S22 S23 S24 0 0 0 0 ⎟⎟
⎜ 21
⎜ S31 S32 S33 S34 0 0 0 0 ⎟
⎜ ⎟
S S42 S43 S44 0 0 0 0 ⎟
S ≡ ⎜ 41
⎜ 0 0 0 0 S55 S56 0 0 ⎟
⎜ ⎟
⎜ 0 0 0 0 S65 S66 0 0 ⎟
⎜ 0 0 0 0 0 0 S77 S78 ⎟

⎜ 0 ⎟
⎝ 0 0 0 0 0 S87 S88 ⎟⎠
There are some additional symmetries in these matrices but the
reflection symmetry properties are the most important.
3) Solve the generalized eigenvalue problem. This part would be
impossible if we hadn’t simplified our matrices above. However, with the
simplifications, it is clear that our matrices are block diagonal. For
example:

sA sB pzA pzB pyA pyB pxA pxB

⎛ H11 H12 H13 H14 0 0 0 0 ⎞ sA


⎜ H H 22 H 23 H 24 0 0 0 0 ⎟⎟ sB
⎜ 21
⎜ H 31 H 32 H 33 H 34 0 0 0 0 ⎟ pzA
⎜ ⎟ pzB
H H 42 H 43 H 44 0 0 0 0 ⎟
H ≡ ⎜ 41
⎜ 0 0 0 0 H 55 H 56 0 0 ⎟ pyA
⎜ ⎟ pyB
⎜ 0 0 0 0 H 65 H 66 0 0 ⎟
⎜ 0 0 0 0 0 0 H 77 H 78 ⎟ pxA
⎜⎜ ⎟
⎝ 0 0 0 0 0 0 H 87 H 88 ⎟⎠ pxB

And similarly for the overlap matrix. The nice thing about block diagonal
matrices is you can reduce a large eigenvalue to several smaller ones. In
this case, our matrices break down into a 4­by­4 block (sA, sB, pzA, pzB) a
2­by­2 block (pyA,pyB) and another 2­by­2 block (pxA,pxB). All the rest of
the matrix is zero. As a result, we can decompose the above 8­by­8 into
three separate eigenvalue problems:
A) The first eigenvalue problem to be solved is a 4­by­4:

Prepared By Kevin Boudreaux


169 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30


9

α α
⎛ H11 H12 H14 ⎞ ⎛ c1 ⎞
H13 ⎛ S11 S12 S13 S14 ⎞ ⎛ c1 ⎞
⎜H ⎟ ⎜ α⎟ ⎜ ⎜ ⎟
⎜ 21 H 22 H 23 H 24 ⎟ ⎜ c2 ⎟ α ⎜ S21 S 22 S23 S24 ⎟⎟ ⎜ c2α ⎟
⎜ ⎟ = E
⎜ H 31 H 32 H 33 H 34 ⎟ c3α ⎜ S31 S32 S33 S34 ⎟ ⎜ c3α ⎟
⎜ ⎟⎜ ⎟ ⎜ ⎟⎜ ⎟
⎝ H 41 H 42 H 43 H 44 ⎠ ⎜ c α ⎟ ⎝ S41 S42 S43 S44 ⎠ ⎜ c α ⎟
⎝ 4 ⎠ ⎝ 4 ⎠
which will give us four molecular orbitals that can be written as linear
combinations of the first four AOs (sA, sB, pzA, pzB)
ψ α = c1α 2s A + c2α 2sB + c3α 2 pzA + c4α 2 pzB
Because these orbitals are symmetric with respect to reflection about
both x and y, they will look something like the H2+ bonding and
antibonding orbitals, and so they are referred to as σ­orbitals. For
example, we can make the +/­ combinations of the 2s orbitals to
obtain one bonding orbital and one antibonding:
σ1­orbital

+ A B

A B

­
σ1∗­orbital

A B

we can make the similar linear combinations of the 2pz orbitals to


obtain:
σ2∗­orbital
A B
+

A B
­ σ2­orbital

A B

where we label the upper orbital σ* because of the nodes between the
nuclei, whereas the σ orbital has no nodes between the nuclei. Note

Prepared By Kevin Boudreaux


170 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30

10

that these +/­ combinations are just to illustrate what the orbitals
will look like; in order to get the actual molecular orbitals we would
need to diagonalize the 4­by­4 and get the eigenvectors. However, if
we do that for a molecule like N2 we actually get orbitals that look 
strikingly similar to the ones above:

σ2*

σ2

σ1*

σ1

B) The second eigenvalue problem to be solved is a 2­by­2:


α ⎛ α⎞
⎛ H55 H 56 ⎞ ⎛ c5 ⎞ α ⎛ S55 S56 ⎞ c5
⎜H ⎟⎜ α ⎟ = E ⎜S ⎟⎜ α ⎟
⎝ 65 H66 ⎠ ⎜⎝ c6 ⎟⎠ ⎝ 65 S66 ⎠ ⎜⎝ c6 ⎟⎠
which will give us two molecular orbitals that can be written as linear
combinations of the next two AOs (pyA,pyB):
ψ α = c5α 2 p yA + c6α 2 p yB
These orbitals get “­“ signs upon reflection about y, so we designate
them πy orbitals. We can again make the +/­ combinations to get an
idea what these orbitals look like:

Prepared By Kevin Boudreaux


171 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30

11

πy­orbital

A B

+
πy*­orbital
A B
­

A B

Here again, the use of +/­ combinations is only for illustration


purposes. Unless we have a homonuclear diatomic, the coefficients
will not be ±1. However, for a molecule like N2, the orbitals look
strikingly similar again:

πy

πy*

C) The last eigenvalue problem is also 2­by­2:


α ⎛ α⎞
⎛ H77 H 78 ⎞ ⎛ c7 ⎞ α ⎛ S77 S78 ⎞ c7
⎜H ⎜
⎟ α ⎟ = E ⎜S ⎟⎜ α ⎟
⎝ 87 H88 ⎠ ⎜⎝ c8 ⎟⎠ ⎝ 87 S88 ⎠ ⎜⎝ c8 ⎟⎠
which will give us two molecular orbitals that can be written as linear
combinations of the last two AOs (pxA,pxB):
ψ α = c5α 2 p xA + c6α 2 p xB
These orbitals get “­“ signs upon reflection about x, so we designate
them πx orbitals. The qualitative picture of the πx orbitals is the same

Prepared By Kevin Boudreaux


172 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30

12

as for the πy orbitals above, expect that the πx orbitals come out of
the page.
4) Occupy the orbitals based on a stick diagram. The most important
thing here is to know the energetic ordering of the orbitals. This would
come out of actually evaluating the non­zero matrix elements in matrices
above and then solving the generalized eigenvalue problem, which is
tedious to do by hand. As a general rule however, there are only two
commonly found MO diagrams for diatomics:
σ2∗ σ2∗

πx*,πy* πx*,πy*

σ2
πx,πy
πx,πy
σ2

σ1∗ Versus σ1∗

σ1 σ1
Hence, the only question is whether the second σ­bonding orbital is above
or below the π­bonding orbitals. In practice, the σ­orbital (which has
significant pz character) is stabilized as you move from left to right along
the periodic table, with the σ­orbital being less stable for atoms to the
left of and including nitrogen and more stable for atoms to the right of
N. Once we have the orbital energy diagram in hand, we can assign the
electrons based on stick diagrams. For example, for CO we have 10
valence electrons and we predict a stick diagram for the ground state
like the one at left below .Meanwhile for NO, which has 11 valence
electrons, we have the stick diagram shown on the right.

Prepared By Kevin Boudreaux


173 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30

13

σ2∗ σ2∗

πx*,πy* πx*,πy*

σ2
πx,πy
πx,πy
σ2

σ1∗ σ1∗

σ1 σ1
CO MO diagram NO MO diagram
We note one important feature we get directly out of the stick 
diagrams: the highest occupied molecular orbital (HOMO) and the
lowest unoccupied molecular orbital (LUMO). For example, in CO the
HOMO is a π­bonding orbital, whereas the LUMO is a π* antibonding
orbital. Meanwhile, for NO the HOMO and LUMO are both π*
orbitals. These orbitals determine reactivity in a crude fashion, as
when electrons are taken out of the molecule, they are removed from
the HOMO, and when electrons are added, they are added to the
LUMO.
5) Compute the energy. Here we can say very little about diatomics,
because we don’t even know the orbital energies exactly, making it
difficult to predict the energies of the whole molecule. If we knew
the orbital energies, the total energy for CO, for example, would be:
ECO=2Eσ1+2Eσ1∗+2Eπx+2Eπy+2Eσ2
As we don’t know these orbital energies, we cannot evaluate the
accuracy of this independent electron model for diatomics. However,
the bond order is a useful descriptor that correlates very well with
the MO energy. For a diatomic, the bond order is simply:
((# Bonding Electrons)­(# Antibonding Electrons))/2
The factor of two reflects the requirement of two electrons for
forming a bond. A higher bond order implies a stronger bond and a
lower bond order a weaker bond. Thus, MO theory predicts CO will
have a stronger bond (bond order 3) than NO (bond order 2.5), which

Prepared By Kevin Boudreaux


174 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #30

14

is experimentally verifiable: the bond energy in NO is 6.5 eV, while


the bond energy in CO is 11.1 eV.

There are a number of other successful predictions of MO theory for


diatomics: O2 is correctly predicted to be a spin triplet ground state, CO is
correctly predicted to be slightly more stable than N2, the highest occupied
molecular orbital for C2 is predicted to be degenerate …. Overall, given its
basis on the independent particle model, MO theory predicts a surprisingly
large array of chemical features correctly.

Prepared By Kevin Boudreaux


175 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #31


1

HÜCKEL MOLECULAR ORBITAL THEORY

In general, the vast majority polyatomic molecules can be thought of as


consisting of a collection of two­electron bonds between pairs of atoms. So
the qualitative picture of σ and π­bonding and antibonding orbitals that we
developed for a diatomic like CO can be carried over give a qualitative
starting point for describing the C=O bond in acetone, for example. One
place where this qualitative picture is extremely useful is in dealing with
conjugated systems – that is, molecules that contain a series of alternating
double/single bonds in their Lewis structure like 1,3,5­hexatriene:

Now, you may have been taught in previous courses that because there are

other resonance structures you can draw for this molecule, such as:

that it is better to think of the molecule as having a series of bonds of


order 1 ½ rather than 2/1/2/1/… MO theory actually predicts this
behavior, and this prediction is one of the great successes of MO
theory as a descriptor of chemistry. In this lecture, we show how even a
very simple MO approximation describes conjugated systems.

Conjugated molecules of tend to be planar, so that we can place all the atoms
in the x­y plane. Thus, the molecule will have reflection symmetry about the
z­axis:

Now, for diatomics, we had reflection symmetry about x and y and this gave
rise to πx and πy orbitals that were odd with respect to reflection and σ
orbitals that were even. In the same way, for planar conjugated systems the
orbitals will separate into σ orbitals that are even with respect to reflection

Prepared By Kevin Boudreaux


176 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #31


2

and πz orbitals that are odd with respect to reflection about z. These πz
orbitals will be linear combinations of the pz orbitals on each carbon atom:

In trying to understand the chemistry of these compounds, it makes sense


to focus our attention on these πz orbitals and ignore the σ orbitals. The πz
orbitals turn out to be the highest occupied orbitals, with the σ orbitals
being more strongly bound. Thus, the forming and breaking of bonds – as
implied by our resonance structures – will be easier if we talk about making
and breaking π bonds rather than σ. Thus, at a basic level, we can ignore the
existence of the σ­orbitals and deal only with the π­orbitals in a qualitative
MO theory of conjugated systems. This is the basic approximation of
Hückel theory, which can be outlined in the standard 5 steps of MO theory:

1) Define a basis of atomic orbitals. Here, since we are only interested


in the πz orbitals, we will be able to write out MOs as linear
combinations of the pz orbitals. If we assume there are N carbon
atoms, each contributes a pz orbital and we can write the µth MOs as:
N
π µ = ∑ ciµ pzi
i=1

2) Compute the relevant matrix representations. Hückel makes some


radical approximations at this step that make the algebra much
simpler without changing the qualitative answer. We have to compute
two matrices, H and S which will involve integrals between pz orbitals
on different carbon atoms:
H ij = ∫ p zi Hˆ pzj dτ Sij = ∫ pzi p zj dτ
The first approximation we make is that the pz orbitals are
orthonormal. This means that:
⎧1 i = j
Sij = ⎨
⎩0 i ≠ j

Prepared By Kevin Boudreaux


177 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #31


3

Equivalently, this means S is the identity matrix, which reduces our


generalized eigenvalue problem to a normal eigenvalue problem
Hicα = Eα Sic µ ⇒ Hi c µ = E µ c µ
The second approximation we make is to assume that any Hamiltonian
integrals vanish if they involve atoms i,j that are not nearest
neighbors. This makes some sense, because when the pz orbitals are
far apart they will have very little spatial overlap, leading to an
integrand that is nearly zero everywhere. We note also that the
diagonal (i=j) terms must all be the same because they involve the
average energy of an electron in a carbon pz orbital:
H ii = ∫ p zi Hˆ p zi dτ ≡ α
Because it describes the energy of an electron on a single carbon, α is
often called the on­site energy. Meanwhile, for any two nearest
neighbors, the matrix element will also be assumed to be constant:
H ij = ∫ pzi Ĥ pzj dτ ≡β i,j neigbors
This last approximation is good as long as the C­C bond lengths in the
molecule are all nearly equal. If there is significant bond length
alternation (e.g. single/double/single…) then this approximation can be
relaxed to allow β to depend on the C­C bond distance. As we will see,
β allows us to describe the electron delocalization that comes from
multiple resonance structures and hence it is often called a resonance
integral. There is some debate about what the “right” values for the
α, β parameters are, but one good choice is α=­11.2 eV and β=­.7 eV.
3) Solve the generalized eigenvalue problem. Here, we almost always
need to use a computer. But because the matrices are so simple, we
can usually find the eigenvalues and eigenvectors very quickly.
4) Occupy the orbitals according to a stick diagram. At this stage, we
note that from our N pz orbitals we will obtain N π orbitals. Further,
each carbon atom has one free valence electron to contribute, for a
total of N electrons that will need to be accounted for (assuming the
molecule is neutral). Accounting for spin, then, there will be N/2
occupied molecular orbitals and N/2 unoccupied ones. For the ground
state, we of course occupy the lowest energy orbitals.
5) Compute the energy. Being a very approximate form of MO theory,
Hückel uses the non­interacting electron energy expression:

Prepared By Kevin Boudreaux


178 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #31


4

N
Etot = ∑ Ei
i =1

where Ei are the MO eigenvalues determined in the third step.

To illustrate how we apply Hückel in practice, let’s work out the energy of
benzene as an example.

1
6 2

5 3
4
1) Each of the MOs is a linear combination of 6 pz orbitals
⎛ c1µ ⎞
⎜ µ⎟
⎜ c2 ⎟
6 ⎜ cµ ⎟
ψ µ = ∑ ciµ pzi → c µ = ⎜ 3µ ⎟
i=1 ⎜ c4 ⎟
⎜ cµ ⎟
⎜⎜ 5µ ⎟⎟
⎝ c6 ⎠
2) It is relatively easy to work out the Hamiltonian. It is a 6­by­6 matrix.
The first rule implies that every diagonal element is α:
⎛α ⎞
⎜ ⎟
⎜ α ⎟
⎜ α ⎟
H=⎜ ⎟
⎜ α ⎟
⎜ α ⎟
⎜⎜ ⎟
⎝ α ⎟⎠
The only other non­zero terms will be between neighbors: 1­2, 2­3, 3­4, 4­5,
5­6 and 6­1. All these elements are equal to β:
⎛α β β⎞
⎜ ⎟
⎜β α β ⎟
⎜ β α β ⎟
H=⎜ ⎟
⎜ β α β ⎟
⎜ β α β⎟

⎜ β ⎟
⎝ β α ⎟⎠
All the rest of the elements involve non­nearest neighbors and so are zero:

Prepared By Kevin Boudreaux


179 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #31


5

⎛α β 0 0 0 β⎞

⎜β α β 0 0 0 ⎟⎟
⎜0 β α β 0 0⎟
H=⎜ ⎟
⎜0 0 β α β 0⎟
⎜0 0 0 β α β⎟
⎜⎜ ⎟
⎝β 0 0 0 β α ⎟⎠
3) Finding the eigenvalues of H is easy with a computer. We find 4 distinct
energies:
E6=α−2β

E4=E5=α−β

E2=E3=α+β

E1=α+2β
The lowest and highest energies are non­degenerate. The second/third and
fourth/fifth energies are degenerate with one another. With a little more
work we can get the eigenvectors. They are:
⎛ +1⎞ ⎛ +1 ⎞ ⎛ +1⎞ ⎛ +1⎞ ⎛ +1 ⎞ ⎛ +1⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ −1⎟ ⎜ −2 ⎟ ⎜0⎟ ⎜0⎟ ⎜ +2 ⎟ ⎜ +1⎟
1 ⎜ +1⎟ 5 1 ⎜ +1 ⎟ 4 1 ⎜ −1⎟ 3 1 ⎜ −1⎟ 2 1 ⎜ +1 ⎟ 1 1 ⎜ +1⎟
c6 = ⎜ ⎟ c = ⎜ ⎟ c = ⎜ ⎟ c = ⎜ ⎟ c = ⎜ ⎟ c = ⎜ ⎟
6 ⎜ −1⎟ 12 ⎜ +1 ⎟ 4 ⎜ +1⎟ 4 ⎜ −1⎟ 12 ⎜ −1 ⎟ 6 ⎜ +1⎟
⎜ +1⎟ ⎜ −2 ⎟ ⎜0⎟ ⎜0⎟ ⎜ −2 ⎟ ⎜ +1⎟
⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟
⎝ −1⎠ ⎝ +1 ⎠ ⎝ −1⎠ ⎝ +1⎠ ⎝ −1 ⎠ ⎝ +1⎠

The pictures at the bottom illustrate the MOs by denting positive (negative)
lobes by circles whose size corresponds to the weight of that particular pz
orbital in the MO. The resulting phase pattern is very reminiscent of a
particle on a ring, where we saw that the ground state had no nodes, the
first and second excited states were degenerate (sine and cosine) and had
one node, the third and fourth were degenerate with two nodes. The one

Prepared By Kevin Boudreaux


180 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #31


6

difference is that, in benzene the fifth excited state is the only one with

three nodes, and it is non­degenerate.

4) There are 6 π electrons in benzene, so we doubly occupy the first 3 MOs:

E6=α−2β

E4=E5=α−β

E2=E3=α+β

E1=α+2β

5) The Hückel energy of benzene is then:


E = 2 E1 + 2 E2 + 2 E3 = 6α + 8β

Now, we get to the interesting part. What does this tell us about the
bonding in benzene? Well, first we note that benzene is somewhat more
stable than a typical system with three double bonds would be. If we do
Hückel theory for ethylene, we find that a single ethylene double bond has
an energy
EC =C = 2α + 2β
Thus, if benzene simply had three double bonds, we would expect it to have a
total energy of
E = 3EC =C = 6α + 6 β
which is off by 2β. We recall that β is negative, so that the π­electrons in
benzene are more stable than a collection of three double bonds. We call
this aromatic stabilization, and Hückel theory predicts a similar stabilization
of other cyclic conjugated systems with 4N+2 electrons. This energetic
stabilization explains in part why benzene is so unreactive as compared to
other unsaturated hydrocarbons.

We can go one step further in our analysis and look at the bond order. In
Hückel theory the bond order can be defined as:
occ
Oij ≡ ∑ ciµ c µj
µ =1

This definition incorporates the idea that, if molecular orbital µ has a bond
between the ith and jth carbons, then the coefficients of the MO on those
carbons should both have the same sign (e.g. we have pzi + pzj). If the orbital

Prepared By Kevin Boudreaux


181 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #31


7

is antibonding between i and j, the coefficients should have opposite


signs(e.g. we have pzi ­ pzj). The summand above reflects this because
ciµ c µj > 0 if ciµ , c µj have same sign
ciµ c µj < 0 if ciµ , c µj have opposite sign
Thus the formula gives a positive contribution for bonding orbitals and a
negative contribution for antibonding. The summation over the occupied
orbitals just sums up the bonding or antibonding contributions from all the
occupied MOs for the particular ij­pair of carbons to get the total bond
order. Note that, in this summation, a doubly occupied orbital will appear
twice. Applying this formula to the 1­2 bond in benzene, we find that:
O12 ≡ 2c1µ =1c2µ =1 + 2c1µ =2 c2µ =2 + 2c1µ =3 c2
µ =3

⎛ +1 ⎞ ⎛ + 1 ⎞ ⎛ +1 ⎞ ⎛ +2 ⎞ ⎛ +1 ⎞ ⎛ 0 ⎞
= 2⎜ ⎟×⎜ ⎟ + 2⎜ ⎟×⎜ ⎟ + 2⎜ ⎟×⎜ ⎟
⎝ 6⎠ ⎝ 6⎠ ⎝ 12 ⎠ ⎝ 12 ⎠ ⎝ 4⎠ ⎝ 4⎠
1 2 2
=2 +2 =
6 12 3
Thus, the C1 and C2 formally appear to share 2/3 of a π­bond [Recall that we
are omitting the σ­orbitals, so the total bond order would be 1 2/3 including
the σ bonds]. We can repeat the same procedure for each C­C bond in
benzene and we will find the same result: there are 6 equivalent π­bonds,
each of order 2/3. This gives us great confidence in drawing the Lewis
structure we all learned in freshman chemistry:

You might have expected this to give a bond order of 1/2 for each C­C π­
bond rather than 2/3. The extra 1/6 of a bond per carbon comes directly
from the aromatic stabilization: because the molecule is more stable than
three isolated π­bonds by 2β, this effectively adds another π­bond to the
system, which gets distributed equally among all six carbons, resulting in an
increased bond order. This effect can be confirmed experimentally, as
benzene has slightly shorter C­C bonds than non­aromatic conjugated
systems, indicating a higher bond order between the carbons.

Just as we can use simple MO theory to describe resonance structures and


aromatic stabilization, we can also use it to describe crystal field and ligand
field states in transition metal compounds and the sp, sp2 and sp3 hybrid

Prepared By Kevin Boudreaux


182 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #31


8

orbitals that arise in directional bonding. These results not only mean MO
theory is a useful tool – in practice these discoveries have led to MO theory
becoming part of the way chemists think about molecules.

Prepared By Kevin Boudreaux


183 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32


1

MODERN ELECTRONIC STRUCTURE THEORY


At this point, we have more or less exhausted the list of electronic
structure problems we can solve by hand. If we were limited to solving
problems manually, there would be a lot of chemistry we wouldn’t be able to
explain! Fortunately, the advent of fast personal computers allows chemists
to routinely use more accurate models of molecular electronic structure.
These types of calculations typically play a significant role in interpreting
experimental results: calculations can be used to assign spectra, evaluate
reaction mechanisms and predict structures of molecules. In this way
computation is complementary to experiment: when the two agree we have
confidence that our interpretation is correct.

The basic idea of electronic structure theory is that, within the Born
Oppenheimer approximation, we can fix the M nuclei in our molecule at some
positions RI. Then, we are left with the Hamiltonian for the electrons
moving in the effective field created by the nuclei:
N N M N
ZI 1
Ĥ ≡ − 1
2 ∑ ∇ -∑∑
i=1
2
i
i=1 I =1
+∑
r̂i − R I i< j r̂i − r̂j Eq. 1

Where the first term is the kinetic energy of all N electrons, the second
term is the attraction between the electrons and nuclei and the third is the
pairwise repulsion between all the electrons. The central aim of electronic
structure theory is to find all the eigenfunctions of this Hamiltonian. As
we have seen, the eigenvalues we
Reaction get will depend on our choice of
Unstable
Barrier the positions of the nuclei –
intermediate
Eel(R1,R2,R3,…RM). As was the
case with diatomics, these
Eel(R1,R2)
energies will tell us how stable
the molecule is with a given
R2 Equilibrium configuration of the nuclei {RI} –
Conformation if Eel is very low, the molecule will
R1 be very stable, while if Eel is high,
the molecule will be unstable in
that configuration. The energy Eel(R1,R2,R3,…RM) is called the potential
energy surface, and it contains a wealth of information, as illustrated in the
picture at above. We can determine the equilibrium configuration of the

Prepared By Kevin Boudreaux


184 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32


2

molecule by looking for the minimum energy point on the potential energy
surface. We can find metastable intermediate states by looking for local
minima – i.e. minima that are not the lowest possible energy states, but which
are separated from all other minima by energy barriers. In both of these
cases, we are interested in points where ∇Eel = 0 . Further, the potential
surface can tell us about the activation energies between different minima
and the pathways that are required to get from the “reactant” state to the
“product” state.

Solving the electronic Schrödinger also gives us the electronic


wavefunctions Ψel(r1,r2,r3,…rN), which allow us to compute all kinds of
electronic properties – average positions, momenta, uncertainties, etc – as
we have already seen for atoms.

We note that while the Hamiltonian above will have many, many eigenstates,
in most cases we will only be interested in the lowest eigenstate – the
electronic ground state. The basic reason for this is that in stable
molecules, the lowest excited states are usually several eV above the ground
state and therefore not very important in chemical reactions where the
available energy is usually only tenths of an eV. In cases where multiple
electronic states are important, the
Hamiltonian above will give us
separate potential surfaces E1el, E2el, σ* potential
E3el … and separate wavefunctions surface
1 2 3
Ψ el, Ψ el, Ψ el. The different
potential surfaces will tell us about
the favored conformations of the
molecules in the different electronic σ potential
states. We have already seen this Erxn surface
for H2+. When we solved for the
electronic states, we got two
eigenstates: σ and σ*. If we put the electron in the σ orbital, the molecule
was bound and had a potential surface like the lower surface at right.
However, if we put the electron in the σ∗ orbital the molecule was not bound
and we got the upper surface.

Prepared By Kevin Boudreaux


185 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32


3

So, at the very least our task is clear cut: solve for the eigenstates of Eq. 1.
Unfortunately, this task is also impossible in practice, even on a computer.
Thus, over the years chemists have developed a vast array of sophisticated
techniques that allow us to at least approximate these solutions to within a
tolerable degree of accuracy. Learning all the details of these
approximations would require a course unto itself: the derivations of the
individual approximations are exceedingly complex, and the sheer number of
different approximations that can be made is quite impressive. These
detailed derivations teach us a lot about what molecules and properties we
should expect our approximations to work for and how we should think about
improving our calculations in cases where the theory fails. However, the
thing that has really brought computational chemistry into the mainstream is
the fact that one does not have to understand every nuance of a method
in order to know how to use it successfully. It suffices to have a simple,
qualitative understanding of how each method works and when it can be
applied. Then, coupling that knowledge with a little technical proficiency at
using commercial chemistry software packages allows us to run fairly
sophisticated calculations on our desktop computer. The next two lectures
are intended to prepare us to run these types of calculations.

First, we note that nearly all the popular approximations are still based on
MO theory – MO theory on steroids in some cases, but MO theory
nonetheless. Thus, there are still 5 steps in the calculation
1) Choose an Atomic Orbital Basis
2) Build the Relevant Matrices
3) Solve the Eigenvalue Problem
4) Occupy the orbitals based on a stick diagram
5) Compute the energy
In a typical calculation, the computer automatically handles steps 2­4
automatically – we don’t have to tell it anything at all. It is sometimes
helpful to know that the computer is doing these things (e.g. The calculation
crashed my computer. What was it doing when it crashed? Oh, it was trying
to solve the eigenvalue problem.) but we essentially never have to do them
ourselves. This leaves two steps (1 and 5) that require some input from us
for the calculation to run properly

Choosing an Atomic Orbital Basis

Prepared By Kevin Boudreaux


186 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32


4

The first point here is that for real electronic structure calculations, you 
typically use a basis set that is much larger than your intuition might tell
you. For example, for H2+ we guessed that we would get a decent result if
we wrote:
ψ = c11sA + c2 1sB
A basis of this type would be called a minimal basis set, because it is the
smallest one that is even close to right. In a more accurate calculation, you 
might use a basis that looks more like:
ψ = c11s A + c2 1sB + c3 2s A + c4 2sB + c5 2 pxA + c6 2 pxB + c7 2 p yA
+ c8 2 p yA + c9 2 pzA + c10 2 pzB + c11 3s A + c12 3sB
The reason we use such extended basis sets arises from a point that was
discussed earlier concerning MO theory. Because our results are variational,
a bigger basis always gets us a lower energy, and therefore a longer AO
expansion always gets us closer to the ground state energy. In the worst
case, the unimportant basis functions will just have coefficients that are
very near zero. While such extended basis sets would be a significant
problem if we were working things out by hand, computers have no problem
dealing with expansions involving even 10,000 basis functions.

The second important point about the atomic orbitals we use is that they are
not hydrogenic orbitals. The reason for this is that the two­electron
integrals involving hydrogenic orbitals cannot all be worked out analytically,
making it very difficult to complete Step 2. Instead of hydrogenic orbitals –
which decay like e−r – we will use Gaussian orbitals that decay like e−α r .
2

Gaussians do not look very much like


hydrogenic orbitals – they don’t have a cusp
at r=0 and they decay much too fast at
large distances. About the only good thing
about them is that they have a mximum at
r=0 and decay. These differences between
Gaussians and hydogenic orbitals are not a
problem, though, because we use extended
basis sets as emphasized above. Basically,
given enough Gaussians, you can expand
anything you like – including a hydrogenic
orbital, as shown in the picture at right. So

Prepared By Kevin Boudreaux


187 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32


5

while using Gaussians may mean we have to use a few extra AOs, if we use
enough of them we should be able to get the same answer.

So we plan to use relatively large Gaussian basis sets for our calculations.
How exactly do we choose those basis sets? Thankfully, a significant amount
of trial­and­error research has distilled the choices down to a few key basis
set features. Depending on the problem at hand and the accuracy desired
we only need to consider three aspects of the AO basis.

Single, Double, Triple, Quadruple Zeta Basis Sets


As we have already discussed for MO theory of diatomics, the smallest basis
we can think of for describing bonding would include all the valence orbitals
of each atom involved. Thus, for H we had 1 s­function, for C there were 2
s­functions and one set of p’s. Similarly, for sulfur we would have needed 3
s­functions and 2 p’s …. A basis of this size is called a minimal or single zeta
basis. The term “single zeta” refers to the fact that we have only a single
set of the valence functions (Note: single valence might seem like a more
appropriate name, but history made a different choice). The most important
way to expand the basis is to include more than a single set of valence
functions. Thus, in a double zeta (DZ) basis set, one would include 2 s­
functions for H, 3 s­ and 2 p­functions for C and 4 s­ and 3 p­functions for
S. Qualitatively, we think of these basis functions as coming from increasing
the n quantum number: the first s function on each atom is something like 1s,
the second something like 2s, the third like 3s …. Of course, since we are
using Gaussians, they’re not exactly 1s, 2s, 3s … but that’s the basic idea.
Going one step further, a triple zeta (TZ) basis would have: H=3s, C=4s3p,
S=5s4p. For Quadruple zeta (QZ): H=4s, C=5s4p, S=6s5p and so on for 5Z,
6Z, 7Z. Thus, one has:
H,He Li­Ne Na­Ar Names
Minimal 1s 2s1p 3s2p STO­3G
DZ 2s 3s2p 4s3p 3­21G,6­31G, D95V
TZ 3s 4s3p 5s4p 6­311G,TZV

Unfortunately, the commonly used names for basis sets follow somewhat
uneven conventions. The basic problem is that many different folks develop
basis sets and each group has their own naming conventions. At the end of
the table above, we’ve listed a few names of commonly used SZ,DZ and TZ
basis sets. There aren’t any commonly used QZ basis sets, because once

Prepared By Kevin Boudreaux


188 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32


6

your basis is that large, it is best to start including polarization functions


(see below).

Polarization Basis Functions


Note that no matter how high you go in the DZ, TZ, QZ hierarchy, you will
never, for example, get a p­function on hydrogen or a d­function on carbon.
These functions tend to be important for describing polarization of the
electrons; at a qualitative level, the p­functions aren’t as flexible in their
angular parts and it’s hard to get them to “point” in as many directions as d­
functions. Thus, particularly when dealing with directional bonding in
molecules, it can be important to include some of these higher angular
momentum functions in your AO basis. In this situation the basis set is said
to contain some “polarization” functions. The general nomenclature of
polarization functions is to add the letter “P” to a basis set with a single set
of polarization functions, and “2P” to a basis with two sets. Thus, a DZP
basis would have: 2s1p on hydrogen, 3s2p1d on carbon and 4s3p1d on sulfur.
A TZP basis set would have 3s1p on hydrogen, 4s3p1d on carbon and 5s4p1d
on sulfur.

H,He Li­Ne Na­Ar Names


DZP 2s1p 3s2p1d 4s3p1d 6­31G(d,p), D95V
TZP 3s1p 4s3p1d 5s4p1d 6­311G(d,p),TZVP

We note that in practice it is possible to mix­and­match different numbers


of polarization functions with different levels of zeta basis sets. The
nomenclature here is to put (xxx,yyy) after the name of the basis set. “xxx”
specifies the number and type of polarization functions to be placed on
Hydrogen atoms and “yyy” specifies the number and type of polarization
functions to be placed on non­hydrogen atoms. Thus, we would have, for
example:

H,He Li­Ne Na­Ar


6­311G(2df,p) 3s1p 4s3p2d1f 5s4p2d1f

Diffuse Functions
Occasionally, and particularly when dealing with anions, the SZ/DZ/TZ/…
hierarchy converges very slowly. For anions, this is because the extra
electron is only very weakly bound, and therefore spends a lot of time far

Prepared By Kevin Boudreaux


189 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32


7

from the nucleus. It is therefore best to include a few basis functions that
decay very slowly to describe this extra electron. Functions of this type
are called “diffuse” functions. They are still Gaussians ( e−α r ), but the value
2

of α is very, very small causing the atomic orbital to decay slowly. Similar to
the situation for polarization functions, diffuse functions can be added in a
mix­and­match way to standard basis sets. Here, the notation “+” or “aug­“
is added to a basis set to show that diffuse functions have been added.
Thus, we have basis sets like 3­21++G, 6­31+G(d,p), aug­TZP.

Aside: Transition Metals


Those of you interested in inorganic chemistry will note that no transition
metals appear in the tables above. This is not because there aren’t basis
sets for transition metals – it is just more complicated to compare different
transition metal basis sets. First, we note that many of the basis sets above
are defined for transition metals. Thus, for example, a 6­31G(d,p) basis on
iron is 5s4p2d1f while a TZV basis for iron is 6s5p3d. The reason we didn’t
include this above is that the idea of “valence” for a transition metal is a
subject of debate: is the valence and s­ and d­ function? An s a p and a d?
Hence, depending on who put the basis set together, there will be some
variation in the number of functions. However, one still expects the same
ordering in terms of quality: TZ will be better than DZ, DZP will be better
than a minimal basis, etc. Thus, you can freely use the above basis sets for
all the elements between K and Kr without significant modification.
Extending the above table for specific basis sets gives:

K­Ca Sc­Zn Ga­Kr


3­21G 5s4p 5s4p2d 5s4p1d
6­31G(d,p) 5s4p1d 5s4p2d1f N/A
6­311G(d,p) 8s7p2d N/A 8s7p3d
TZV 6s3p 6s3p2d 6s5p2d

Things also become more complicated when dealing with second row
transition metals. Here, relativistic effects become important, because the
Schrödinger equation predicts that the 1s electrons in these atoms are
actually moving at a fair fraction of the speed of light. Under these
circumstances, the Schrödinger equation is not strictly correct and we need

Prepared By Kevin Boudreaux


190 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32


8

to start considering corrections for relativistic effects. The most efficient


way to incorporate the relativity is to use an effective core potential (ECP).
An effective core potential removes the core electrons from the problem
and replaces them with an effective potential that the valence electrons
feel. This potential reflects the combined interaction with the nucleus and
the (relativistic) core electrons. Thus, for an ECP we specify both how many
core electrons we want to neglect and how many basis functions we want to
use to describe the valence electrons. For example, one popular double zeta
ECP is the LANL2DZ basis. As an example, for ruthenium LANL2DZ
replaces the 28 core electrons (1s22s22p63s23p63d10=Argon) with an
effective potential and uses a 3s3p2d basis to describe the valence orbitals.
Thus, for the second transition series we have (using
[CoreSize]/ValenceBasisSize as our shorthand):
Y­Cd Hf­Hg
LANL2DZ [Argon]/3s3p2d N/A
SDD [Argon]/8s7p6d [Kr4d 4f14]/8s7p6d
10

As one progresses further up the periodic table, fewer and fewer basis sets
are available, simply because less is known about their chemistry.

This is just a very brief overview of what basis sets are available and how
good each one is expected to be. The general idea of using basis sets is to
use larger and larger basis sets until the quantity we are computing stops
changing. This is based on the idea that we are really using the AO
expansion to approximate the exact solution of the Schrödinger equation. If
we had an infinite basis, then we would get the exact answer, but with a
large and finite basis we should be able to get close enough. Note, however,
that the calculation will typically take more time in a larger basis than in a
smaller one. Thus, we really want to make the basis just big enough to get
the answer right, but no larger.

Computing the Energy


For simple MO theory, we used the non­interacting (NI) electron model for
the energy:

Prepared By Kevin Boudreaux


191 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32


9

N N
ENI = ∑ Ei = ∑ ∫ ψ i (1) Ĥψ i (1) dτ
i=1 i=1

Where, on the right hand side we have noted that we can write the NI
energy as a sum of integrals involving the orbitals. We already know from
looking at atoms that this isn’t going to be good enough to get us really
accurate answers; the electron­electron interaction is just too important.
In real calculations, one must choose a method for computing the energy
from among several choices.

The Hartree Fock (HF) Approximation


The Hartree­Fock method uses the IPM energy expression we’ve already
encountered:
N
EIPM = ∑ Ei + ∑ J�ij − K� ij
i=1
N
Ei = ∑ ∫ψ i (1) Hˆψ i (1) dτ
i=1

J�ij ≡ ∫∫ψ i* (1)ψ *j ( 2) ( ) j( ) 1 2 1 2


1 ψ 1 ψ 2 dr dr dσ dσ
r1 − r2 i

K� ij ≡ ∫∫ψ i* (1)ψ *j ( 2) ( ) j( ) 1 2 1 2
1 ψ 2 ψ 1 dr dr dσ dσ
r1 − r2 i
Since the energy contains the average repulsion, we expect our results will
be more accurate. However, there is an ambiguity in this expression. The
IPM energy above is correct for a determinant constructed out of any set of
orbitals {ψ i } and the energy will be different depending on the orbitals we
choose. For example, we could have chosen a different set of orbitals, {ψ 'i } ,
and gotten a different energy:
N
E ' NI = ∑ E 'i + ∑ J� 'ij − K� 'ij
i =1

How do we choose the best set of orbitals then? Hartree­Fock uses the
variational principle to determine the optimal orbitals. That is, in HF we find
the set of orbitals that minimize the independent particle energy. These
orbitals will be different from the non­interacting orbitals because they will
take into account the average electron­electron repulsion terms in the
Hamiltonian. Thus, effects like shielding that we have discussed
qualitatively will be incorporated into the shapes of the orbitals. This will
tend to lead to slightly more spread out orbitals and will also occasionally

Prepared By Kevin Boudreaux


192 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32

10

change the ordering of different orbitals (e.g. σ might shift below π once
interactions are included).

Now, the molecular orbitals (and hence the energy) are determined by their
coefficients. Finding the best orbitals is thus equivalent to finding the best
coefficients. Mathematically, then, we want to find the orbitals that make
the derivative of the IPM energy zero:
∂EIPM ∂ N
∂ciα
= ∑ Ei + ∑ J�ij − K� ij = 0
∂ciα i=1
In order to satisfy this condition, one typically resorts to an iterative
procedure, where steps 2­5 of our MO procedure are performed repeatedly:
1) Choose an AO Basis

1’) Guess an IPM Hamiltonian Heff

2) Build Heff, S matrices

Choose 3) Solve the Eigenvalue Problem


Better
Heff 4) Occupy Lowest Orbitals
dE
5) Compute E,
dc
dE
= 0? Done
No dc Yes
Here, HF makes use of the fact that defining an IPM Hamiltonian, Heff,
completely determines the molecular orbital coefficients, c. Thus, the most
convenient way to change the orbitals is actually to change the Hamiltonian
that generates the orbitals. The calculation converges when we find the
molecular orbitals that give us the lowest possible energy, because then
dE
= 0 . These iterations are called self­consistent field (SCF) iterations
dc
and the effective Hamiltonian Heff is often called the Fock operator, in honor
of one of the developers of the Hartree­Fock approximation.

Generally Hartree Fock is not very accurate, but it is quite fast. On a


decent computer, you can run a Hartree Fock calculation on several hundred
atoms quite easily, and the results are at least reasonable.

Prepared By Kevin Boudreaux


193 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32

11

Density Functional Theory (DFT)


Here, we still use a Slater determinant to describe the electrons. Hence,
the things we want to optimize are still the MO coefficients cα. However, we
use a different prescription for the energy – one that is entirely based on
the electron density. For a single determinant, the electron density, ρ(r) is
just the probability of finding an electron at the point r. In terms of the
occupied orbitals, the electron density for a Slater Determinant is:
N
ρ (r ) = ∑ ψ α (r )
2
Eq. 2
α =1

This has a nice interpretation: ψ i ( r ) is the probability of finding an


2

electron in orbital i at a point r. So the formula above tells us that for a


determinant the probability of finding an electron at a point r is just the
sum of the probabilities of finding it in one of the orbitals at that point.

There is a deep theorem (the Hohenberg­Kohn Theorem) that states:

There exists a functional Ev[ρ] such that, given the ground state
density, ρ0, Ev[ρ0]=E0 where E0 is the exact ground state energy.
Further, for any density, ρ’, that is not the ground state density,
Ev[ρ’]>E0.

This result is rather remarkable. While solving the Schrödinger Equation


required a very complicated 3N dimensional wavefunction Ψel(R1, R2,…RN),
this theorem tells us we only need to know the density ­ which is a 3D
function! – and we can get the exact ground state energy. Further, if we
don’t know the density, the second part of this theorem gives us a simple
way to find it: just look for the density that minimizes the functional Ev.

The unfortunate point is that we don’t know the form of the functional Ev.
We can prove it exists, but we can’t construct it. However, from a
pragmatic point of view, we do have very good approximations to Ev, and the
basic idea is to choose an approximate (but perhaps very, very good) form
for Ev and then minimize the energy as a function of the density. That is, we
dEv
look for the point where = 0 . Based on Eq. 2 above, we see that ρ just

Prepared By Kevin Boudreaux


194 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32

12

depends on the MOs and hence on the MO coefficients, so once again we are
dEv
looking for the set of MO coefficients such that = 0 . Given the
dc
similarity between DFT and HF, it is not surprising that DFT is also solved
by self consistent field iterations. In fact, in a standard electronic
structure code, DFT and HF are performed in exactly the same manner (see
flow chart above). The only change is the way one computes the energy and
dE
.
dc

Now, as alluded to above, there exist good approximations (note the plural)
to Ev. Just as was the case with approximate AO basis sets, these
approximate energy expressions have strange abbreviations. We won’t go
into the fine differences between different DFT energy expressions here.
I’ll simply note that roughly, the quality of the different functionals is
expected to follow:
LSDA < PBE ≈ BLYP < PBE0 ≈ B3LYP
Thus, LSDA is typically the worst DFT approximation and B3LYP is typically
among the best. I should mention that this is just a rule of thumb; unlike
the case for basis sets where we were approaching a well­defined limit, here
we are trying various uncontrolled approximations to an unknown functional.
Experience shows us that B3LYP is usually the best, but this need not always
be the case.

Finally, we note that the speed of a DFT calculation is about the same as
Hartree Fock – both involve self­consistent field iterations to determine the
best set of orbitals, and so both take about the same amount of computer
time. However, for the same amount of effort, you can get quite accurate
results. As a rule of thumb, with B3LYP you can get energies correct to
within 3 kcal/mol and distances correct to within .01 Å.

Post­Hartree Fock Calculations
Here, the idea is to employ wavefunctions that are more flexible than a
Slater determinant. This can be done by adding up various combinations of
Slater determinants, by adding terms that explicitly correlate pairs of
electrons (e.g. functions that depend on r1 and r2 simulataneously) and a
variety of other creative techniques. These approaches are all aimed at

Prepared By Kevin Boudreaux


195 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32

13

incorporating the correlation between electrons – i.e. the fact that


electrons tend to spend more time far apart from one another as opposed to
close together. This correlation reduces the average repulsion employed in
HF and brings us closer to the true ground state energy. In each case, one
usually does a Hartree­Fock calculation first and then includes the
correlation afterward, leading to the heading of “post­Hartree Fock”
methods. Once again, there are a number of acronyms, and we merely assert
that the quality of the results goes approximately like:
HF < CASSCF < MP2 < CCSD < MP4 < CCSD(T)
Here, the ordering is rigorous: we have something solid that we are tyring to
approximate and going from left to right we are making better and better
approximations. On the scale above, DFT typically gives results of about
MP2 quality.

As a general rule, post­HF calculations are much, much more expensive than
HF or DFT and also require bigger basis sets: whereas HF might converge
with a DZ or TZ basis, a post­HF calculation might require QZ or even 5Z.
Hence, they should only be attempted for relatively small molecules where
high accuracy is required

Combining what we have learned, then, the approximations we can make fit
nicely into a two­dimensional parameter space:

Exact
Method
Answer

CCSD(T)
CCSD

MP2
Feasible
DFT Calculations
CASSCF
HF Basis

STO­3G 6­31G(d,p) 6­311G+(2df,p)


3­21G TZVP

Prepared By Kevin Boudreaux


196 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #32

14

On the one axis, we have the basis sets we can choose from. On the other,
we have the different methods for approximating the energy. The get close
to the exact answer, we need to employ a large basis set and an accurate
energy method. Unfortunately, both increasing the size of the basis and
improving the method tend to slow our calculations down. Given that we don’t
want to wait years and years to find out the result of a calculation, modern
computers therefore limit how accurate our answers can be (as illustrated
with the red line above). As we become experts at what is and is not
feasible with current computing power, we become better able to get good
accuracy for a variety of problems.

Prepared By Kevin Boudreaux


197 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


1

SPECTROSCOPY: PROBING MOLECULES WITH LIGHT


In practice, even for systems that are very complex and poorly
characterized, we would like to be able to probe molecules and find out as
much about the system as we can so that we can understand reactivity,
structure, bonding, etc. One of the most powerful tools for interrogating
molecules is spectroscopy. Here, we tickle the system with electromagnetic
radiation (i.e. light) and see how the molecules respond. The motivation for
this is that different molecules respond to light in different ways. Thus, if
we are creative in the ways that we probe the system with light, we can hope
to find a unique spectral fingerprint that will differentiate one molecule
from all other possibilities. Thus, in order to understand how spectroscopy
works, we need to answer the question: how do electromagnetic waves
interact with matter?

The Dipole Approximation


An electromagnetic wave of wavelength λ, produces an electric field, E(r,t),
and a magnetic field, B(r,t), of the form:
E(r,t)=E0 cos(k·r – ωt) B(r,t)=B0 cos(k·r – ωt)
Where ω=2πν is the angular frequency of the wave, the wavevector k has a
magnitude 2π/λ and k (the direction the wave propagates) is perpendicular to
E0 and B0. Further, the electric and magnetic fields are related:
E0· B0=0 |E0|=c|B0|
Thus, the electric and magnetic
http://www.monos.leidenuniv.nl
fields are orthogonal and the
magnetic field is a factor of c (the
speed of light, which is 1/137 in
atomic units) smaller than the
electric field. Thus we obtain a
picture like the one at right, where
the electric and magnetic fields
oscillate transverse to the
direction of propagation.

Now, in chemistry we typically deal with the part of the spectrum from
ultraviolet (λ≈100 nm) to radio waves (λ≈10 m)1. Meanwhile, a typical molecule

1
There are a few examples of spectroscopic measurements in the X­Ray region. In these
cases, the wavelength can be very small and the dipole approximation is not valid.

Prepared By Kevin Boudreaux


198 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


2

is about 1 nm in size. Let us assume that the molecule is sitting at the origin.
Then, in the 1 nm3 volume occupied by the molecule we have:
k·r ≈ |k| |r| ≈ 2p/(100 nm) 1 nm = .06
Where we have assumed UV radiation (longer wavelengths would lead to even
smaller values for k·r). Thus, k·r is a small number and we can expand the
electric and magnetic fields in a power series in k·r:
E(r,t)≈E0 [cos(k·0-ωt)+O(k·r)]≈E0 cos(ωt)

B(r,t)≈B0 [cos(k·0-ωt)+O(k·r)]≈B0 cos(ωt)

Where we are neglecting terms of order at most a few percent. Thus, in


most chemical situations, we can think of light as applying two time
dependent fields: an oscillating, uniform electric field (top) and a
uniform, oscillating magnetic field (bottom). This approximation is called
the Dipole approximation – specifically when applied to the electric
(magnetic) field it is called the electric (magnetic) dipole approximation. If
we were to retain the next term in the expansion, we would have what is
called the quadrupole approximation. The only time it is advisable to go to
higher orders in the expansion is if the dipole contribution is exactly zero as
happens, for example, due to symmetry in some cases. In this situation, even
though the quadrupole contributions may be small, they are certainly large
compared to zero and would need to be computed.

The Interaction Hamiltonian


How do these oscillating electric and magnetic fields couple to the molecule?
Well, for a system interacting with a uniform electric field E(t) the
interaction energy is
Hˆ E ( t ) = −µˆ iE ( t ) = −e rˆ i E ( t )
where µ is the electric dipole moment of the system. Thus, uniform electric
fields interact with molecular dipole moments.

Similarly, the magnetic field couples to the magnetic dipole moment, m.


Magnetic moments arise from circulating currents and are therefore
proportional to angular momentum – larger angular momentum means higher
circulating currents and larger magnetic moments. If we assume that all the
angular momentum in our system comes from the intrinsic spin angular
momentum, I=(Ix , Iy ,Iz), then the magnetic moment is strictly proportional to
I. For example, for a particle with charge q and mass m then
q gˆ
Hˆ B ( t ) = −m
ˆ iB ( t ) = − Ii B ( t )
2m

Prepared By Kevin Boudreaux


199 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34 3

where g is a phenomenological factor (creatively called the “g­factor”) that


takes into account the internal structure of the particle containing the
intrinsic spin – for an electron ge=2.0023, while for a proton gp=5.5857.

So now suppose that we have a molecule we are interested in, and it has a
Hamiltonian, Ĥ 0 , when the field is off. Then, when the field is on, the
Hamiltonian will be
Ĥ ( t ) = Ĥ 0 + Ĥ E ( t ) + Ĥ B ( t )
Actually, in most cases, the simultaneous effects of electric and magnetic
fields are not important and we will consider one or the other:
Ĥ ( t ) ≡ Ĥ 0 + Ĥ1 ( t ) Ĥ1 ( t ) ≡ Ĥ E ( t ) or Ĥ B ( t ) .
Thus, in the presence of light, a molecule feels a time­dependent
Hamiltonian. This situation is quite different with what we have discussed
so far. Previously, our Hamiltonian has been time independent and our job
has simply reduced to finding the eigenstates of Ĥ . Now, we have a
Hamiltonian that varies with time, meaning that the energy eigenvalues and
eigenstates of Ĥ also change with time. What can we say that is meaningful
about a system that is constantly changing?

Time Dependent Eigenstates


As it turns out, the best way to think about this problem is to think about
the eigenstates of Ĥ 0 . When the field is off, each of these eigenstates
evolves by just getting a phase factor:
Ĥ 0φn ( t ) = Enφn ( t ) ⇒ φn ( t ) = e− iE t / �φn ( 0 )
n

Thus, things like the probability density do not change because multiplying
by the complex conjugate wipes out the phase factor:
φn ( t ) = {e−iE t / �φn ( 0 )} * e−iE t / �φn ( 0 ) =eiE t / �φn ( 0 ) * e−iE t / �φn ( 0 ) = φn ( 0 )
2 2
n n n n

Thus, when considering measurable quantities (which always involve complex


conjugates) the eigenstates of the Hamiltonian appear not to change with
time. However, when the field is on the eigenstates will change with
time. In particular, we will be interested in the rate at which the field
induces transitions between an initial eigenstate φi and a final state φf.

Prepared By Kevin Boudreaux


200 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


4

To work out these rates, we first work out the time dependence of some
arbitrary state, ψ(t). We can expand ψ(t) as a linear combination of the
eigenstates:
ψ ( t ) = ∑ cn ( t ) φn ( t )
n

where cn(t) are the coefficients to be determined. Next, we plug this into
the TDSE:
i�ψ� ( t ) = Ĥψ ( t )

⇒ i� ∑ cn ( t ) φn ( t ) = Ĥ ∑n cn ( t ) φn ( t )
∂t n
( )
⇒ i� ∑ c�n ( t ) φn ( t ) + cn ( t ) φ�n ( t ) = ∑ cn ( t ) Ĥ 0 + Ĥ1 ( t ) φn ( t )
n n

⇒ i� ∑ c�n ( t ) φn ( t ) −
n
iEn

( )
cn ( t ) φn ( t ) = ∑ cn ( t ) En + Ĥ1 ( t ) φn ( t )
n

n n n
(
⇒ i� ∑ c�n ( t ) φn ( t ) + ∑ En cn ( t ) φn ( t ) = ∑ cn ( t ) En + Ĥ1 ( t ) φn ( t ) )
⇒ i� ∑ c�n ( t ) φn ( t ) = ∑ cn ( t ) Ĥ1
( t ) φn ( t )
n n

Next, we multiply both sides by the final state we are interested in (φf*) and
then integrate over all space. On the left hand side, we get:
i� ∫ φ f * ( t ) ∑ c�n ( t )φn ( t ) dτ = i�∑ c�n ( t ) ∫ φ f * ( t ) φn ( t ) dτ = i�c� f ( t )
n n

δnf
Meanwhile, on the right we get:
∫ φ f ( t ) ∑ cn ( t )Hˆ 1 ( t ) φn ( t ) dτ = ∑ cn ( t )∫ φ f ( t ) Ĥ1 ( t ) φn ( t ) dτ
* *

n n

Combining terms gives:

⇒ i�c� f ( t ) = ∑ ∫ φ f * ( t ) Ĥ1 ( t ) φn ( t ) dτ cn ( t ) Eq. 1


n

Up to this point, we haven’t used the form of H1 at all. We note that we can
re­write the light­matter interaction as:
Ĥ1 ( t ) = V̂ cos (ω t )
qgˆ
where, for electric fields V̂ ≡ −erˆ i E 0 and for magnetic fields Vˆ ≡ − Ii B .
2m
In either case, we can re­write the cosine in terms of complex exponentials:

Prepared By Kevin Boudreaux


201 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


5

Ĥ1 ( t ) = V̂ 1
2 (e iωt
+ e− iωt )
Plugging this into Eq.1 above gives:

( )
i�c� f ( t ) = ∑ ∫ φ f * ( t ) 12 V̂ eiωt + e−iωt φn ( t ) dτ cn ( t )
n

= ∑ ∫ φf * (0) e
iE f t /� 1
2 ( )
V̂ eiωt + e −iωt e−iEnt /�φn ( 0 ) dτ cn ( t )
n

= ∑ ∫ φ f * ( 0 ) 12 V̂ φn ( 0 ) dτ
n
(e −i ( En −E f −�ω ) t / �
+e
−i ( En −E f +�ω ) t / �
)c n (t )

=∑
n
1
2 V fn (e −i ( En −E f −�ω ) t / �
+e
−i ( En −E f +�ω ) t / �
)c n (t )

Tickling the Molecule With Light


To this point we haven’t made any approximations to the time evolution. We
now make some assumptions that allow us to focus on one particular i→f
transition. We make two physical assumptions:
1) The molecule starts in a particular eigenstate, φi, at t=0. This sets
the initial conditions for our coefficients: only the coefficient of
state i can be non­zero initially:
cn ( 0 ) = 0 if n ≠ i ci ( 0 ) = 1
It is easy to verify that this choice gives the desired initial state:
ψ ( 0 ) = ∑ cn ( 0 ) φn ( 0 ) = 0 + 0 + ...1iφi ( 0 ) + 0.... = φi ( 0 )
n

2) The interaction only has a small effect on the dynamics. This is


certainly an approximation, and it will not always be true. We can
certainly guarantee its validity in one limit: if we reduce the
intensity of our light source sufficiently, we will reduce the
strength of the electric and magnetic fields to the point where the
influence of the light is small. As we turn up the intensity, there
may be additional effects that will come into play, and we will come
back to this possibility later on. However, if we take this
assumption at face value, we can assume on the right hand side

Prepared By Kevin Boudreaux


202 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


6

that the coefficient, cn, of a state other than φi will be much


smaller than ci for all times:
cn ( t ) � ci ( t ) if n ≠ i ci ( t ) ≈ 1
Where, in the second equality we have noted that if all the other
coefficients are tiny, ci must be approximately 1 if we want our
state to stay normalized.

These two assumptions lead to an equation for the coefficients of the form:
i�c� f
( t ) = ∑ 12 V fn e
n

( − i( En −E f −�ω ) t / �
+e
−i ( En −E f +�ω ) t / �
)c n (t )
⇒ i�c� f ( t ) = 12 V fi e ( −i ( Ei −E f −�ω ) t / �
+e
−i ( Ei −E f +�ω ) t / �
) c (t )
i

= 12 V fi e ( −i ( Ei −E f −�ω ) t / �
+e
−i ( Ei −E f +�ω ) t / �
)
Now we can integrate this new equation to obtain c f ( t ) :

( ) dt
T
i�c f ( T ) = V fi ∫ e
−i ( Ei −E f −�ω ) t / � −i ( Ei −E f +�ω ) t / �
1
2 +e
0

(e ) dt
T
V fi
⇒ c f (T ) =
−i ( Ei −E f −�ω ) t / � −i ( Ei −E f +�ω ) t / �

2i� ∫
+e Eq. 2
0

Now, this formula for c f (T ) is only approximate, because of assumption 2).


If we wanted to improve our result for, we could plug our approximate final
expression (Eq. 2) back in on the RHS of Eq. 1 and then integrate the
equation again. This would lead to a better approximate solution for c f ( t ) .
Most importantly, while our approximate solution is linear the interaction
matrix element, V fi , after plugging the result back in, we would get terms
that were quadratic in V fi . By assumption 2) above, these quadratic terms
will be much smaller than the linear ones we have retained above and so we
feel safe in neglecting them. For these reasons, assumption 2) is known as
the linear response approximation.

We now make the final rearrangement: we recall that we are interested in


the probability of finding the system in the state f. This is given by c f (T ) :
2

Prepared By Kevin Boudreaux


203 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


7

2 2

∫ (e ) dt
T
V fi
Pf ( T ) = c f ( T ) =
2 −i ( Ei −E f −�ω ) t / � −i ( Ei −E f +�ω ) t / �
+e
4� 2 0

Fermi’s Golden Rule


Now, usually our experiments take a long time from the point of view of
electromagnetic waves. In a single second a light wave will oscillate billions
of times. Thus, our observations are likely to correspond to the long­time
limit of the above expression:
2 2

∫ (e ) dt
T
V fi −i ( Ei −E f −�ω ) t / � −i ( Ei −E f +�ω ) t / �
Pf = lim +e
4� 2 T →∞
0

and in fact, we are usually not interested in probabilities, but rates, which
are probabilities per unit time:
2 2

∫ (e ) dt
T
V fi 1 −i ( Ei −E f −�ω ) t / � −i ( Ei −E f +�ω ) t / �
W fi = lim
2 T →∞
+e
4� T 0

This integral looks very difficult. However, it is easy to work out with
pictures because it is almost always zero. Note that both the real and
imaginary parts of the integrand oscillate. Thus, we will be computing the
integral of something that looks like:

Thus, as long as the integrand oscillates, the positive regions will cancel out
the negative ones and the integral will be zero. There only two situations
where the integrand is not oscillatory: Ei − E f − �ω = 0 (in which case the
first term is unity) and Ei − E f + �ω = 0 (in which case the second term is
unity). We can therefore write
2
V fi
W fi ∝ ⎡δ ( Ei − E f − �ω ) + δ ( Ei − E f + �ω ) ⎤
4� ⎣ 2 ⎦
where δ(x) is a function that is defined to be non­zero only when x=0. This
result is called Fermi’s golden rule. It gives us a way of predicting the rate
of any i→f transition in any molecule induced by an electromagnetic field of

Prepared By Kevin Boudreaux


204 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


8

arbitrary frequency coming from any direction. This formula – as well as


generalizations that relax the electric dipole and linear response
approximations – is probably the single most important relationship in terms
of how chemists think about spectroscopy, and so we will dwell a bit on the
interpretation of the various terms.

On the one hand, the probability of an i→f transition is proportional to


2 2
V fi = ∫ φ f *V̂ φi dτ
Thus, if the matrix element of the interaction operator V̂ between the
initial and final states is zero, then the transition never happens. This is
called a selection rule, and a transition that does not occur because of a
selection rule is said to be forbidden. For example, in the case of the
electric field,
2 2 2 2
V fi = ∫ φ f *µ̂
µ iE 0φi dτ = E 0 i ∫ φ f *µ̂
µφi dτ = E 0 iµ fi
Thus, for molecules interacting with electric fields, the transition i→f is
forbidden unless the matrix element of the dipole operator between i&f is
nonzero. Meanwhile, in the case of a magnetic field,
2 2
2 2 qg qg
V fi = ∫ φ f m i B0φi dτ
*
= B0 i ∫ φ f * Îφi dτ = B0 iIˆ fi
2m 2m
Thus, a magnetic field can only induce an i→f transition if the matrix
element of one of the spin angular momentum operators is non­zero between
the initial and final states. Selection rules of this type are extremely
important in determining which transitions will and will not appear in our
spectra. E Ef i

The second thing we note about Fermi’s


Golden Rule is that it enforces energy
conservation. We note that the energy �ω �ω
carried by a photon is �ω . The δ­function
portion is only non­zero if E f − Ei = �ω
Ei Ef
(second term) or Ei − E f = �ω (first
term). Thus, the transition only occurs E f − Ei = �ω E f − Ei = −�ω
if the energy difference between the
two states exactly matches the energy
of the photon we are sending in. This is depicted in the picture at right.

Prepared By Kevin Boudreaux


205 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


9

The way these terms are interpreted are as follows: in the first case, the
light increases the energy in the system by exactly one photon worth of
energy. Here, we think of a photon being absorbed by the molecule and
exciting the system. In the second case, the light reduces the energy of
the system by exactly one photon worth of energy. Here, we think of the
molecule emitting a photon and relaxing to a lower energy state. The fact
that photon emission by a molecule can be induced by light is called
stimulated emission, and is the principle on which lasers are built: basically,
when you shine light on an excited molecule, you get more photons out than
you put in.

In order to make much more progress with spectroscopy, we have to


consider some specific choices of the molecular Hamiltonian, Ĥ 0 , which we
do in the next several lectures. Depending on the system at hand the energy
conservation and selection rules give different spectral signatures that
ultimately allow us to interpret the spectra of real molecules and to
characterize their physical properties.

Prepared By Kevin Boudreaux


206 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #35


1

VIBRATIONAL SPECTROSCOPY
As we’ve emphasized many times in this course, within the Born­
Oppenheimer approximation,
the nuclei move on a potential Harmonic
energy surface (PES) Approximation
determined by the electrons. R
For example, the potential R0 A + B separated atoms
felt by the nuclei in a V(R)
diatomic molecule is shown in
cartoon form at right. At low
energies, the molecule will sit
near the bottom of this
potential energy surface. In
this case, no matter what the equilibrium bond length
detailed structure of the potential is, locally the nuclei will “feel” a nearly
harmonic potential. Generally, the motion of the nuclei along the PES is
called vibrational motion, and clearly at low energies a good model for the
nuclear motion is a Harmonic oscillator.

Simple Example: Vibrational Spectroscopy of a Diatomic


If we just have a diatomic molecule, there is only one degree of freedom
(the bond length), and so it is reasonable to model diatomic vibrations using a
1D harmonic oscillator:
P̂ 2 1 ˆ 2 P̂ 2 1
Ĥ = + ko R = + mωo 2 Rˆ 2
2µ 2 2µ 2
where ko is a force constant that measures how stiff the bond is and can be
approximately related to the second derivative to the true (anharmonic) PES
near equilibrium:
∂ 2V
ko ≈ 2
∂R R
0

Applying Fermi’s Golden Rule, we find that when we irradiate the molecule,
the probability of a transition between the ith and fth Harmonic oscialltor
states is:
2
V fi
W fi ∝ ⎡δ ( Ei − E f − �ω ) + δ ( Ei − E f + �ω )⎤
4� 2 ⎣ ⎦

Prepared By Kevin Boudreaux


207 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


2

where ω is the frequency of the light (not to be confused with the


frequency of the oscillator, ωo). Because the vibrational eigenstates involve
spatial degrees of freedom and not spin, we immediately recognize that it is
the electric field (and not magnetic) that is important here. Thus, we can
write the transition matrix element as:
2 2 2 2
V fi = ∫ φ f *µ̂
µ iE 0φi dτ = E 0 i ∫ φ f *µ̂
µφi dτ = E 0 i ∫ φ f * eR̂φi dτ
Now, we define the component of the electric field, ER, that is along the
bond axis which gives
2 2 2
V fi = E R ∫ φ f * eR̂φi dτ = e2 E R ∫ φ f R̂φi dτ
2 *

So the rate of transitions is proportional to the square of the strength of


the electric field (first two terms) as well as the square of the transition
dipole matrix element (third term). Now, because of what we know about
the Harmonic oscillator eigenfunctions, we can simplify this. First, we re­
write the position operator, R, in terms of raising and lowering operators:
2
� �e 2 2
( aˆ + + aˆ − ) φi dτ = ∫ φ f ( aˆ + + aˆ − ) φi dτ
2
∫ φf
2 2
V fi = e ER 2 *
ER *

2 µω 2 µω
�e 2
( (i + 1) δ + i δ f ,i −1 )
2 2
⇒ V fi = ER f ,i +1 i + 1 φi +1 iφi−1
2 µω
where above it should be clarified that in this expression “i" never refers to
√­1 – it always refers to the initial quantum number of the system. Thus, we
immediately see that a transition will be forbidden unless the initial and
final states differ by one quantum of excitation. Further, we see that
the transitions become more probable for more highly excited states. That
is, Vfi gets bigger as i gets bigger.

Combining the explicit expression for the transition matrix element with
Fermi’s Golden Rule again gives:
e2
W fi ∝
8�µω
( ) (
ER ( i + 1) δ f ,i +1 + i δ f ,i −1 ⎡⎣δ Ei − E f − �ω + δ Ei − E f + �ω ⎤⎦
2
) ( )
⇒ W fi ∝ E R
2
{(i + 1) δ f ,i +1
⎡⎣δ ( Ei − Ei+1 − �ω ) + δ ( Ei − Ei+1 + �ω )⎤⎦

+ i δ f ,i −1 ⎡⎣δ ( Ei − Ei−1 − �ω ) + δ ( Ei − Ei−1 + �ω )⎤⎦ }

Prepared By Kevin Boudreaux


208 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


3

0
⇒ W fi ∝ ER
2
{(i + 1) δ f ,i+1 ⎡⎣δ ( −�ωo − �ω ) + δ ( −�ωo + �ω )⎤⎦
+ i δ f ,i −1 ⎡⎣δ ( �ωo − �ω ) + δ ( �ωo + �ω ) ⎤⎦ }
0

⇒ W fi ∝ ER
2
{(i + 1) δ f ,i +1 + i δ f ,i −1 }δ ( �ωo − �ω )
Thus, we see that a harmonic oscillator will only absorb or emit photons
of frequency �ωo , where ωo is the frequency of the oscillator. Thus, if
we look at the absorption spectrum, for example, we will see absorption at
only one frequency:
ωo Light Frequency (ω)

Absorption
Intensity

Molecular force constants are typically on the order of an eV per Å, which


leads to vibrational frequencies that are typically
between 500­3500 cm­1 and places these absorption
features in the infrared. As a result, this form of
spectroscopy is traditionally called IR spectroscopy. We
associate the spectrum above as arising from all the
n→n+1 transitions in the Harmonic oscillator (see left).
Of course, most of the time the molecule will start in its
ground state, so that the major contribution comes from the 0→1 transition.
However, the other transitions occur at the same frequency and also
contribute to the absorption.

This is the classic paradigm for IR vibrational spectroscopy: each diatomic


molecule absorbs radiation only at one frequency that is characteristic of
the curvature of the PES near its minimum. Thus, in a collection of
different molecules one expects to be able differentiate one from the other
by looking for the frequency appropriate to each one. In particular there is

Prepared By Kevin Boudreaux


209 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


4

a nice correlation between the “strength” of the bond and the frequency at
which it will absorb.

Of course, we are not always or even usually interested in diatomics, and


even diatomics are not perfect Harmonic oscillators. Thus, there are a
number of reasons why IR absorption spectra do not really look like the
classic Harmonic oscillator spectrum shown above, but more like:

Heterogeneity
The primary reason the real spectrum above looks different than the model
is because the real spectrum was taken in solution. The model is correct for
a single diatomic, or for many, many copies of identical diatomic molecules.
However, in solution, ever molecule is just slightly different, because every
molecule has a slightly different arrangement of solvent molecules around it.
These solvent molecules subtly change the PES, slightly shifting the
vibrational frequency of each molecule and also modifying the transition
dipole a bit. Thus, while a single hydrogen fluoride molecule might have a
spectrum like the model above, a solution with many HF molecules would look
something like:

Prepared By Kevin Boudreaux


210 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


5

ωo Light Frequency (ω)

Absorption
Intensity

Going over to the situation where there are 1023 HF molecules and
recognizing that our spectra will tend to add the intensity of lines that are
closer together than our spectrometer can differentiate, we anticipate that
for a diatomic molecule in solution, the vibrational spectrum should look 
something like:
ωo Light Frequency (ω)

Absorption
Intensity

The resulting feature in the spectrum is usually called a lineshape. It


primarily reflects the distribution of different environments surrounding
your oscillators. Thus, by analyzing the lineshape of a well­known type of
vibration (such as a C=O stretch) one can get an idea about the environments
those CO groups live in: How polar are the surroundings? Are they near
electron withdrawing groups? What conformations give rise to the
spectrum? Finally, we should note that vibrational spectra recorded in the
gas phase have very narrow linewidths, qualitatively resembling our model
above.

Anharmonicity

Prepared By Kevin Boudreaux


211 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


6

Another reason real spectra differ from our model is that assuming the PES
is harmonic is only a model. If we want high accuracy, we need to account
for anharmonic terms in the potential:
1 1 1
V ( R ) = mωo 2 ( R − R0 ) + α ( R − R0 ) + β ( R − R0 ) + ...
2 3 4

2 6 24
One can investigate the quantitative effects of the anharmonic terms on the
spectrum by performing variational calculations. However, at a basic level
there are two ways that anharmonic terms impact vibrational spectra:
1) The energy differences between adjacent states are no longer
constant. Clearly, the eigenvalues of an anharmonic Hamiltonian
will not be equally spaced – this was a special feature of the
Harmonic system. Thus, for a real system we should expect the
0→1 transition to have a slightly different frequency than 1→2,
which in turn will be different that 2→3 …. Generally, the higher
transitions have lower (i.e. redshifted) energies because of the
shape of the molecular PES – rather than tend toward infinity at
large distances as the harmonic potential does, a molecular PES
tends toward a constant dissociation limit. Thus, the higher
eigenstates are lower in energy than they would be for the
corresponding harmonic potential. Taking this information, we
would then expect a single anharmonic oscillator to have a
spectrum something like:
ωo Light Frequency (ω)

2→ 3
Absorption
1→ 2
Intensity

0→ 1

where we note that while the rate of, say, 1→2 is about twice that
of 0→1 (because the transition dipole is twice as big) the intensity
of 0→1 is greater because the intensity is (Probability of i being
occupied)x(Rate of i→f) and at room temperature the system

Prepared By Kevin Boudreaux


212 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


7

spends most of its time in the ground vibrational state. The 1→2,
2→3 … lines in the spectrum are called hot bands.
2) Anharmonicity relaxes the Δn=±1 selection rule. Note that the
rules we arrived at were based on the fact that â +φi ∝ φi+1 . This is
only true for the Harmonic oscillator states. For anharmonic
eigenstates â +φi ∝ φi+1 + ε1φi+2 + ..... . Thus transitions with Δn=±2,±3…
will no longer be forbidden for anharmonic oscillators. Rather, in
the presence of a bit of anharmonicity, they will be weakly allowed.
Combining this observation with point 1) above results in a more
complete picture for what the IR spectrum of an anharmonic
oscillator should look like:
ωo 2ωo Light Frequency (ω)

2→ 4
2→ 3
Absorption 1→ 3
1→ 2
Intensity
0→ 2

0→ 1

The peaks at around 2ωo are called overtones. Meanwhile, those at


around ωo are called fundamentals.

Polyatomic Molecules
The final difference between the model above and a general IR spectrum is
that in chemistry, we are not always dealing with diatomic molecules. For a
polyatomic molecule, we can still think of the potential as a Harmonic
potential, but it has to be many­dimensional – it has to depend on several
variables R1, R2, R3, …. The most general Harmonic potential we can come up
with is then of the form:
V ( R1 , R2 , R3 ,...) = 12 k11 R12 + 12 k12 R1 R2 + 12 k13 R1 R3 + .... + 12 k21 R2 R1 + 12 k22 R2 2 + 12 k23 R2 R3 + ....
+ 12 k31 R3 R1 + 12 k32 R3 R2 + 12 k33 R32 + ...
where it is important to notice the cross terms involving, say R1 and R2,
which couple the different vibrations. At first sight, it seems like we

Prepared By Kevin Boudreaux


213 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


8

can’t solve this Hamiltonian; the only many­dimensional Harmonic potential


we would know how to solve would be one that is separable:
V� ( R1 , R2 , R3 ,...) = 12 k11 R12 + 12 k22 R2 2 + 12 k33 R32 + ...
If the Harmonic potential were of this form, we would be able to write down
the eigenstates as products of the 1D eigenstates and get the energies as
sums of the 1D eigenenergies. As it turns out, by changing coordinates we
can turn a quadratic system with off­diagonal cross terms (like the first
potential) into one with no cross terms (like the second). These new
coordinates, in terms of which the Hamiltonian separates, are called normal
modes and they allow us to reduce a polyatomic molecule to a collection of
independent 1D oscillators.

First, we note that V can be re­written concisely in matrix notation [Note: it


may be useful to consult McQuarrie’s supplement on Matrix Eigenvalue
problems if the following seems unfamiliar.]:
V ( R1 , R2 , R3 ,...) = 12 R T iK iR
⎛ R1 ⎞ ⎛ k11 k12 k13 ... ⎞
⎜ ⎟ ⎜ ⎟
R k k22 k23 ... ⎟
R ≡ ⎜ 2 ⎟ K ≡ ⎜ 21
⎜ R3 ⎟ ⎜ k31 k23 k33 ... ⎟
⎜ ⎟ ⎜ ⎟
⎝ � ⎠ ⎝ � � � �⎠
Now, the Hamiltonian is of the form:
P̂ 2 1
Ĥ = ∑ i + R̂T iK iR
ˆ
i 2 µ i 2
It is convenient to first transform to mass-weighted coordinates:
P̂ kij
p̂i ≡ i x̂i ≡ µi R̂i k�ij ≡
µi µi µ j
in terms of which we can write the Hamiltonian:
p̂i 2 1 T �
Ĥ = ∑ + xˆ iK ixˆ
i 2 2
As is clear from the kinetic energy above, in these coordinates, every
degree of freedom has the same reduced mass.

Now we perform the normal mode transformation. We want to write:

Prepared By Kevin Boudreaux


214 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


9

⎛ k11' 0 0 ... ⎞
⎜ ⎟
0 k22' 0 ... ⎟
� ixˆ = yˆ T iK ′iy
x̂T iK K′ = ⎜
⎜ 0 0 k33' ... ⎟
⎜⎜ ⎟
⎝ ... ... ... ... ⎟⎠
Further, we will assume there is a matrix U that transforms from x to y
y = Uix ⇒ y T = xT i U T
where in the second equality, we recall the general rule that the transpose
of a product is the product of the transposes, but in the opposite order.
Combining these two equations:
� ixˆ = yˆ T iK ′iy = xˆ T iUT iK ′iUixˆ
x̂T iK
⇒K � = UT iK ′iU
The last equation is a common problem encountered in linear algebra: the
� ) and place it in diagonal form (the right
quest to take a given matrix ( K
hand side). For a symmetric matrix like K � the solution to this problem is
well known: the diagonal entries of K ′ are the eigenvalues of K � and the
columns of the transformation matrix U are the eigenvectors of K � . The
transformed variables y are called the normal modes. These modes are
linear combinations of the local degrees of freedom R1, R2, R3, … that we
started out with. Thus, while the initial motions might correspond clearly to
local stretching of one bond or bending of an angle, the normal modes will
generally be complicated mixtures of different molecular motions. We can
visualize this in the simple case of two degrees of freedom. The local modes
R1, R2 can be thought of as the two orthogonal axes in a plane. Meanwhile,
the normal modes y1, y2 are also orthogonal axes, but rotated from the
original set:
R2 R2
y2 Normal modes
Local modes

y1

R1 R1
The local modes have interactions between each other: local stretches are
coupled to local bends, etc. As a result, the Hamiltonian is not separable in
terms of the local modes R1, R2. However, by design the Hamiltonian is
separable when written in terms of the normal modes:

Prepared By Kevin Boudreaux


215 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


10

p̂i 2 1
Ĥ = ∑ + kii′ ŷi 2
i 2 2

The eigenvalues, kii , tell us the “stiffness” of the PES in the particular
direction yi. Note that the Hamiltonian above is exactly the same
Hamiltonian as the one we started with. The coupling terms have simply
been rotated away by changing the coordinates. As discussed before, we can
immediately interpret the spectra of this Hamiltonian in terms of a sum of
many independent oscillators. Thus, for a polyatomic molecule within the
harmonic approximation we expect to see lines at each of the normal mode
frequencies:
ω1 ω2 ω3 ω4 ω5 Light Frequency (ω)

Absorption
Intensity

Where we have noted that the different oscillators will typically also have
different transition dipoles. (For obvious reasons, in vibrational
spectroscopy the square of the transition dipole is often called the
oscillator strength) We can, of course, combine this polyatomic picture with
the anharmonicity effects above to get a more general picture that looks
like:
2ω1 2ω2 ω5-ω1 ω2+ω3
ω1 ω2 ω3 ω4 ω5 Light Frequency (ω)

Absorption
Intensity

Prepared By Kevin Boudreaux


216 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #34


11

where we predict the existence of various hotbands and overtones for each
of the normal mode oscillators in the molecule. Note that while the
overtones always involve multiple quanta, the quanta need not come from the
same normal mode – hence we expect not only overtones at 2ω1 , but also a
combination bands at ω2 + ω3 and ω5 − ω1 . The picture above is qualitatively
correct for the IR spectrum of a single molecule. In solution, heterogeneity
leads to a smearing out and broadening of the peaks, leading to the complex
IR fingerprints we are used to.

As should be clear from the above discussion, IR spectra contain a wealth of


information about the molecule: the stiffness of each normal mode, the
degree of anhormonic effects, the character of the local environment felt
by the oscillators …. Of course, in order to extract this information, one
must be able to assign the spectrum – i.e. one must be able to distinguish
hotbands from overtones and associate the various normal modes (at least
qualitatively) with physical motions of the molecule. This task can be
extremely challenging – and computation must be used as a guide in many
cases – but when it is accomplished, one typically has a very sensitive
fingerprint of molecule under consideration.

Prepared By Kevin Boudreaux


217 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #36 Page


1

NUCLEAR MAGNETIC RESONANCE


Just as IR spectroscopy is the simplest example of transitions being induced by light’s
oscillating electric field, so NMR is the simplest example of transitions induced by the
oscillating magnetic field. Because the strength of matter­magnetic field interactions are
typically two orders of magnitude smaller than the corresponding electric field interactions,
NMR is a much more delicate probe of molecular structure and properties. The NMR spin
Hamiltonians and wavefunctions are particularly simple, and permit us to demonstrate several
fundamental principles (about raising and lowering operators, energy levels, transition
probabilities, etc.) with a minimal amount of algebra. The principles and procedures are
applicable to other areas of spectroscopy ­­ electronic, vibrational, rotational, etc. – but for
these cases the algebra is more extensive.

Nuclear Spins in a Static Magnetic Field


For a single isolated spin in a static magnetic field, the contribution to the energy is:
ˆ iB 0 = −γ Iˆ iB 0
Ĥ 0 = − m
where γ is called the gyromagnetic ratio. If we choose our z axis to point in the direction of
the magnetic field then:
Ĥ 0 = − mˆ z B0 = −γ Iˆz B0
If we assume the nuclear spin is ½ (As it is for a proton) then we can easily work out the
energy levels of this Hamiltonian:
E± = ± 12 γ �B0 ≡ ± 12 �ω0
where ω0 = γB0 is called the nuclear Larmor frequency (rad/sec). Now, nuclei are never
isolated in chemistry – they are always surrounded by electrons. As we learned for the
hydrogen atom, the electrons near the nucleus shield the outer electrons from the bare
electric field produced by the nucleus. Similarly, the electrons shield the nucleus from the
bare electric field we apply in the laboratory. More specifically, the electron circulation
produces a field, B’ opposed to B0 and of
magnitude equal to σ B0 where σ is a constant. nucleus with
bare
Thus, the effective field, B, at the nucleus is nucleus electrons
B0 B0(1-σ)
B = (1 − σ )B0
Note that σ is different for each chemically
different nuclear spin – this is the famous
Energy

chemical shift – and permits resolution of lines in hω 0 hω 0 (1 − σ )

NMR spectra corresponding to chemically


different sites. The Hamiltonian is modified
accordingly
Hˆ 0 = −mˆ z B0 (1 − σ ) = −γ Iˆz B0 (1 − σ )
Zero Field High Field
Thus, instead of “seeing” a magnetic field of
magnitude B0, a proton in a molecule will see a

Prepared By Kevin Boudreaux


218 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #36 Page 2

magnetic field of magnitude (1-σ)B0 and the associated Hamiltonian and spin state energies
will become:
E± = ± 12 γ �B0 (1 − σ ) ≡ ± 12 �ω0 (1 − σ )
This is illustrated in the figure above. Note the sign of the Hamiltonian is chosen so that the
α state (spin parallel to B0) is lower in energy than the β state ( spin antiparallel to B0).

Now, in the simplest NMR experiment, we probe this system with an oscillating magnetic field
perpendicular to the static field. By convention, we take this field to be along the x axis:
Ĥ1 ( t ) = − m
ˆ iB1 ( t ) = −γ Iˆ iB1 ( t ) = −γ Iˆx Bx cos (ωt )
We use Fermi’s Golden Rule to describe the spectrum of the spin in the oscillating field. The
selection rule is:
2 2 2
V fi = ∫ φ f * miB1φi dτ = γ Bx i ∫ φ f * Iˆxφi dτ
Now, we recall that Iˆx can be written in terms of the raising and lowering operators for
angular momentum:
(
Iˆx ∝ Iˆ+ + Iˆ− )
So that:

( )
2 2
V fi ∝ γ Bx i ∫ φ f * Iˆ+ + Iˆ− φi dτ
We immediately see that the integral is non­zero only if the initial and final spin states
differ by ±1 quantum of angular momentum (i.e. ΔM = ±1 ) , because the operator must either
raise or lower the eigenvalue of Iˆz . Thus, there are two possible transitions: α→β and β→α.
Futher, the energy conservation rule tells us that these transitions will only occur when the
photon energy matches the energy gap between the two states. As a result, we can
immediately draw the spectrum of a single shielded spin:

Intensity

(1−σ)ωo Frequency (ω)

This is perhaps not all that shocking: there is only one transition here, and so we might have
guessed that the spectrum would involve the frequency of that transition. However, we note
two generalizations of this result. First, we note that if we had chosen to apply the
oscillating field parallel to the static field, we would not have generated any transitions; we

Prepared By Kevin Boudreaux


219 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #36 Page 3

only changed the spin state because we could decompose the x­oscillating field into raising
and lowering operators. If the field was z­oscillating, then we would have had
2 2
V fi ∝ γ Bx i ∫ φ f * Iˆzφi dτ
which is only non­zero for the trivial α→α and β→β transitions. Second, we note that if the
spin was bigger than ½ (e.g. a spin­3/2 nucleus) then our selection rule above would be
precisely the same. Thus, we would have allowed transitions − 23 ↔ − 12 , − 12 ↔ + 12 and
+ 12 ↔ + 32 and all of these transitions would occur at the same frequency. Thus, spin­3/2
transitions like − 23 ↔ + 12 or − 32 ↔ + 32 are strictly forbidden.

Now, as noted above, depending on their environment, different protons will be shielded
differently, resulting in a spectrum that will look qualitatively like:

Intensity

(1−σ1)ωo (1−σ2)ωo (1−σ3)ωo (1−σ4)ωo Frequency (ω)

We note that the transition moment above is independent of the chemical environment: it
does not depend on shielding or any other property of the molecule. Thus, the area under an
NMR peak is strictly proportional to the number of spins that have transitions at that
frequency. This stands in contrast to IR spectroscopy, where the intensity of each oscillator
depended on the character of the oscillator, the initial state ….

Two Spins – J Couplings


Now, we are not usually interested in two isolated spins. For two uncoupled spins with
different chemical shifts ( σ1≠ σ2 ) in an external field we obtain a Hamiltonian of the form:
Ĥ 0 = −γ Iˆ1 z B0 (1 − σ 1 ) − γ Iˆ2 z B0 (1 − σ 2 )
Because this Hamiltonian is separable, we can immediately work out the energies:
ωο
E↓↓ = E1↓ + E2↓ = + [(1 − σ 1 ) + (1 − σ 2 )]
2
ωο
E↓↑ = E1↓ + E2↑ = + [(1 − σ 1 ) − (1 − σ 2 )]
2

ωο
E↑↓ = E1↑ + E2↓ = − [(1 − σ 1 ) − (1 − σ 2 )]

ωο
E↑↑ = E1↑ + E2↑ = − [(1 − σ 1 ) + (1 − σ 2 )]

Prepared By Kevin Boudreaux


220 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #36 Page 4

where we have assumed for simplicity that σ1< σ2 so that E↓↑ > E↑↓ . Now, the selection rule is
the of the same form as for a single spin, but Ix decomposes into a sum of Ix for spin 1 and an
Ix for spin 2:

( )
2 2 2
V fi = γ Bx i ∫ φ f * Iˆxφi dτ = γ Bx i ∫ φ f * Iˆ1x + Iˆ2 x φi dτ

( )
2
∝ ∫ φ f * Iˆ1+ + Iˆ1− + Iˆ2+ + Iˆ2− φi dτ
The remaining integral is only nonzero if ΔM 1 = ±1 or ΔM 2 = ±1 , because the operators must
raise or lower the spin state of either spin 1 or spin 2 (but not both). If we wanted to change
both spins, we would need an operator like Î1+ Î 2− , which would allow us to raise 1 while also
lowering 2. Since we do not have any of these cross terms, we conclude only one or the other
spin can flip in an allowed transition – any two­spin transitions are forbidden.

Combining these results for two uncoupled spins, we obtain the picture at left. We note that
the ↑↑↔↓↓ and ↓↑↔↑↓ transitions are forbidden, since they require flipping both spins
simultaneously. For the allowed transitions, we can easily work out
E↓↓
the energies:
ΔM2=±1 ω ω
ΔM1=±1 E↓↓ − E↑↓ = + ο [(1 − σ 1 ) + (1 − σ 2 )] + ο [(1 − σ 1 ) − (1 − σ 2 )] = ωο (1 − σ 1 )
2 2
E↓↑ ωο ωο
E↑↓
E↓↑ − E↑↑ = + [(1 − σ 1 ) − (1 − σ 2 )] + [(1 − σ 1 ) + (1 − σ 2 )] = ωο (1 − σ 1 )
2 2
ΔM1=±1 ωο ωο
ΔM2=±1 E↓↓ − E↓↑ = + [(1 − σ 1 ) + (1 − σ 2 )] − [(1 − σ 1 ) − (1 − σ 2 )] = ωο (1 − σ 2 )
2 2
ωο ωο
E↑↑ E↑↓ − E↑↑ = − [(1 − σ 1 ) − (1 − σ 2 )] +
[(1 − σ 1 ) + (1 − σ 2 )] = ωο (1 − σ 2 )
2 2
So we have only two transition energies, corresponding to each of the isolated transitions,
just as predicted above:

Intensity

(1−σ2)ωo (1−σ1)ωo Frequency (ω)

Where we note that there are actually two degenerate transitions contributing to each line.

Prepared By Kevin Boudreaux


221 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #36 Page


5
We now permit the two spins to be coupled to one another in a simple way. We include a J­
J
coupling of the spins: Î1z Î 2z , where the factors of � are included so that J has units of
�2
energy. Thus, the Hamiltonian assumes the form
J
Ĥ 0 = −γ Iˆ1z B0 (1 − σ 1 ) − γ Iˆ2 z B0 (1 − σ 2 ) + 2 Iˆ1z Iˆ2 z

Now, we can work out the eigenvalues of this new Hamiltonian quite easily because we know
the eigenvalues of Î1z and Î 2z . For example, for the ↓↓ state:
⎛ J ˆ ˆ ⎞
Ĥ 0φ↓↓ = ⎜ −γ Iˆ1z B0 (1 − σ 1 ) − γ Iˆ2 z B0 (1 − σ 2 ) + I I
2 1z 2z ⎟ φ↓↓
⎝ � ⎠
⎛ J ⎛ −� ⎞ ⎛ −� ⎞ ⎞
=⎜ E1↓ + E 2↓ + ⎜ ⎟⎜ ⎟ φ
⎝ � 2 ⎝ 2 ⎠ ⎝ 2 ⎠ ⎟⎠ ↓↓

⎛ J⎞
= ⎜ E1↓ +E2↓ + ⎟ φ↓↓

⎝ 4 ⎠

Similar algebra for the other states gives:


⎛ J ⎛ � ⎞ ⎛ −� ⎞ ⎞ ⎛ J⎞
Ĥ 0φ↑↓ = ⎜ E1↑ + E2↓ + 2 ⎜ ⎟ ⎜ ⎟ ⎟ φ↑↓ = ⎜ E1↑ +E1↓ − ⎟ φ↑↓
⎝ � ⎝ 2 ⎠ ⎝ 2 ⎠⎠ ⎝ 4⎠
⎛ J ⎛ −� ⎞ ⎛ � ⎞ ⎞ ⎛ J⎞
Ĥ 0φ↓↑ = ⎜ E1↓ + E2↑ + 2 ⎜ ⎟ ⎜ ⎟ ⎟ φ↓↑ = ⎜ E1↓ +E2↑ − ⎟ φ↓↑
⎝ � ⎝ 2 ⎠⎝ 2 ⎠⎠ ⎝ 4⎠
⎛ J ⎛ � ⎞⎛ � ⎞⎞ ⎛ J⎞
Ĥ 0φ↑↑ = ⎜ E1↑ + E2↑ + 2 ⎜ ⎟ ⎜ ⎟ ⎟ φ↑↑ = ⎜ E1↑ +E1↓ + ⎟ φ↑↑
⎝ � ⎝ 2 ⎠⎝ 2 ⎠⎠ ⎝ 4⎠
Thus, in the presence of the coupling, our energy diagram changes:
E↓↓ E↓↓
ΔM2=±1 ΔM2=±1
ΔM1=±1 ΔM1=±1
E↓↑ E↓↑
E↑↓ E↑↓
ΔM1=±1 ΔM2=±1 ΔM1=±1 ΔM2=±1

E↑↑ E↑↑
Uncoupled J Coupling

Where we have noted that states where the spins are parallel shift upward in energy and
those where the spins are antiparallel shift down, and we have exaggerated the magnitude of
the shift for visual effect. Note that the selection rules do not change, because the states
have not changed – only the energies are different with the coupling on. The energies of the
allowed transitions are:
ω ⎡ J⎤ ω ⎡ J⎤ J
E↓↓ − E↑↓ = + ο ⎢(1 − σ 1 ) + (1 − σ 2 ) + ⎥ − ο ⎢ −(1 − σ 1 ) + (1 − σ 2 ) − ⎥ = ωο (1 − σ 1 ) +
2 ⎣ 4⎦ 2 ⎣ 4⎦ 2

Prepared By Kevin Boudreaux


222 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #36 Page


6
ωο ⎡ J ⎤ ωο ⎡ J⎤ J
E↓↑ − E↑↑ = + (1 − σ ) − (1 − σ ) − − − (1 − σ ) − (1 − σ ) + = ω (1 − σ ) −
2 ⎢⎣ 4 ⎥⎦ 2 ⎢⎣ 4 ⎥⎦
1 2 1 2 ο 1
2
ω ⎡ J⎤ ω ⎡ J⎤ J
E↓↓ − E↓↑ = + ο ⎢(1 − σ 1 ) + (1 − σ 2 ) + ⎥ − ο ⎢(1 − σ 1 ) − (1 − σ 2 ) − ⎥ = ωο (1 − σ 2 ) +
2 ⎣ 4⎦ 2 ⎣ 4⎦ 2
ω ⎡ J⎤ ω ⎡ J⎤ J
E↑↓ − E↑↑ = ο ⎢ −(1 − σ 1 ) + (1 − σ 2 ) − ⎥ − ο ⎢ −(1 − σ 1 ) − (1 − σ 2 ) + ⎥ = ωο (1 − σ 2 ) −
2 ⎣ 4⎦ 2 ⎣ 4⎦ 2
Thus, whereas we had two doubly degenerate transitions in the absence of coupling, in the
presence of coupling we have four distinct transitions:

Intensity

J J

(1−σ2)ωo (1−σ1)ωo Frequency (ω)

where here we have noted the physical fact that J is typically much smaller than the
difference in chemical shielding σ between distinct protons. Thus, we see that the splitting
of NMR peaks is mediated by the coupling between the nuclear spins. This coupling is
typically mediated via the electrons – nucleus 1 pushes on the electrons, which are delocalized
and in turn push on nucleus 2. While one can routinely compute these couplings via DFT or
HF, it is much more common to use empirical rules to determine which protons will be coupled
and how large we expect the coupling to be. We should note that the magnitude of the J­
splitting is independent of the magnetic field strength. Meanwhile, the Larmor frequency
increases with increasing B0. Thus, in a strong enough magnet, the peaks with shielding near σ1
will be very far from those with shielding σ2.

Spin Dynamics and Pulsed NMR


One of the extremely appealing aspects of NMR is we can exactly work out virtually any
property we’re interested in knowing. In particular, we can get a picture of the dynamics of
the spin in an external magnetic field. This gives us a qualitative picture of what we are doing
when we take an NMR spectrum and also serves as the basis for modern pulsed NMR
experiments. Consider an arbitrary initial state written as a linear combination of the two
spin states:
ψ ( t ) = cα ( t )ψ α + cβ ( t )ψ β
where we have noted that the time dependence of the state comes through the time
dependence of the coefficients. We can write this in matrix mechanics:

Prepared By Kevin Boudreaux


223 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #36 Page


7
⎛ cα ( t ) ⎞

ψ ( t ) = ⎜⎜ ⎟⎟
⎝ cβ ( t ) ⎠
We can also write the time­dependent Schrödinger equation in Matrix mechanics:
⎛ c� ( t ) ⎞ ⎛ Hαα Hαβ ⎞ ⎛ cα ( t ) ⎞
i�ψ� ( t ) = Hˆ ψ ( t ) ⇒ i� ⎜ α =
⎜ c�β ( t ) ⎟⎟ ⎜⎜ H ⎟⎟ ⎜⎜ ⎟⎟
⎝ ⎠ ⎝ βα H ββ ⎠ ⎝ cβ ( t ) ⎠
Now, for a spin in a static field, we know the Hamiltonian
⎛ − �ω� 0 ⎞
⎜ 2 0 ⎟
Ĥ = −ω0 (1 − σ ) Î z ≡ −ω� 0 Î z ⇒ H = −ω� 0 I z = ⎜ ⎟
⎜ 0 �ω� 0 ⎟
⎜ ⎟
⎝ 2 ⎠
Thus, the TDSE becomes:
⎛ − �ω� 0 ⎞
⎛ c (t ) ⎞ ⎛ cα ( t ) ⎞
� ⎜ 0 ⎟
2
i� ⎜ α =
⎜ c�β ( t ) ⎟⎟ ⎜⎜ ⎟ ⎜⎜
−�ω� 0 ⎟ ⎝ cβ ( t ) ⎟⎠

⎝ ⎠ ⎜ 0 ⎟
⎝ 2 ⎠
Which reduces to two independent differential equations for the coefficients:
− �ω� 0 + �ω� 0
i�c�α ( t ) = cα ( t ) i�c�β ( t ) = cβ ( t )
2 2
These equations can easily be integrated to yield:
+iω� 0 t −iω� 0 t

cα ( t ) = e 2
cα ( 0 ) cβ ( t ) = e 2
cβ ( 0 )
where we will assume for simplicity that the initial values, cα ( 0 ) , cβ ( 0 ) are real. Thus, the
magnitude of each coefficient is constant with time; we only acquire a phase factor for each
coefficient. However, these coefficients completely describe the time evolution of an
arbitrary spin state in the static magnetic field.

Now that we have solved for the coefficients of the time dependent wavefunction, let’s look
at some interesting properties of the system. First, let’s compute the z­component of the
spin:
⎛� ⎞
⎜ 2 0 ⎟ ⎛ cα ( t ) ⎞ �
Î z ( t ) = ( cα ( t ) * cβ ( t ) *) ⎜

⎟⎟ = ⎡⎢ cα ( t ) − cβ ( t ) ⎤⎥ = ⎡⎢ cα ( 0 ) − cβ ( 0 ) ⎤⎥
2 2 2 2
⎟ ⎜⎜
⎜ 0 − � ⎟ ⎝ cβ ( t ) ⎠ 2 ⎣ ⎦ 2⎣ ⎦
⎜ ⎟
⎝ 2⎠
Thus, the z­component of the spin does not change with time! This is perhaps a bit
surprising. We continue to compute the x and y components:
⎛ �⎞
⎜ 0 2 ⎟ ⎛ cα ( t ) ⎞ �
Î x ( t ) = ( cα ( t ) * cβ ( t ) *) ⎜ ⎟ ⎜⎜ ⎟⎟ = ⎡c ⎣ α ( t ) * cβ ( t ) + cβ ( t ) * cα ( t )⎤⎦
⎜ � 0 ⎟ ⎝ cβ ( t ) ⎠ 2
⎜ ⎟
⎝2 ⎠

= cα ( 0 ) cβ ( 0 ) ⎡⎣ e−iω0 t + e+iω0 t ⎦⎤ = �cα ( 0 ) cβ ( 0 ) cos ω� 0 t
� �

Prepared By Kevin Boudreaux


224 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #36 Page


8
⎛ −i� ⎞
Î y ( t ) = ( cα ( t ) * cβ ( t ) *) ⎜
⎜0 2 ⎟ ⎛ cα ( t ) ⎞ = −i� ⎡c t * c t − c t * c t ⎤
⎟ ⎜⎜ ⎟ α ( ) β ( ) β ( ) α ( )⎦
⎜ i� ⎟ ⎝ cβ ( t ) ⎟⎠ 2 ⎣
⎜ 0 ⎟
⎝2 ⎠
−i�
= cα ( 0 ) cβ ( 0 ) ⎣⎡ e −iω0t − e+iω0 t ⎤⎦ = −�cα ( 0 ) cβ ( 0 ) sin ω� 0 t
� �

2
Thus, the x and y components oscillate with time at the shielded Larmor frequency ω� 0 . It is
convenient to define a magnetization vector that contains these three expectation values:
( )

M ( t ) ≡ Iˆx ( t ) Î y ( t ) Î z ( t )

It is fairly easy to see that the magnetization is precessing about the magnetic field: the
projection onto the magnetic field axis is constant, while the perpendicular motion is tracing
out a circular path. This is precisely the behavior one would expect from a classical magnetic
moment in a magnetic field. In this case, the magnetic field would exert a torque on the
magnetic moment according to:

dM ( t ) � �
= M ( t ) × γ B eff
dt
where we note that the magnetic moment feels the shielded magnetic field Beff. This gives us
three differential equations for the components of the magnetization, called Bloch
Equations:
dM x ( t )
= γ ⎡⎣ M y ( t ) Bzeff − M z ( t ) Byeff ⎤⎦ = γ M y ( t ) Bzeff
dt
dM y ( t )
= −γ ⎡⎣ M x ( t ) Bzeff − M z ( t ) Bxeff ⎤⎦ = −γ M x ( t ) Bzeff
dt
dM z ( t )
= γ ⎡⎣ M x ( t ) Byeff − M y ( t ) Bxeff ⎦⎤ = 0
dt
Where we have noted that only the z­component of the magnetic field is non­zero. Further,
it is easy to see by substitution that our quantum mechanical predictions for Î x (t ) and

Î y (t ) satisfy the equations above for M x ( t ) and M y ( t ) , respectively (try it and see). Thus,
we find that the quantum evolution of the average spin exactly follows the classical
equations of motion! We find comfort in this conclusion, because it is usually much easier to
think in terms of classical properties whenever possible, giving us a very nice semiclassical
way of interpreting NMR.

This rather surprising result turns out to be true for a single spin evolving in an arbitrary
time depenedent magnetic field Beff(t). To prove this, we have to use Ehrenfest’s theorem,
which states that for an arbitrary operator O, the time dependent average value of O
satisfies
d i
Ô ( t ) = ⎡ Hˆ , Oˆ ⎤ ( t )
dt � ⎣ ⎦

Prepared By Kevin Boudreaux


225 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #36 Page


9
We proved a shortened version of this on one of the Problem Sets. Applying Ehrenfests
theorem to the three operators Î x , Î y , Î z in an arbitrary magnetic field Beff(t) gives equations

of motion that are exactly the same as the classical equations for M ( t ) . Thus, one can prove
quite rigorously that the classical picture is exactly right for describing spin dynamics in a
magnetic field.

What does this gain us? Well, with this result in hand it is relatively easy to derive the
correct differential equations for our favorite time dependent magnetic field:

B eff ( t ) = − Bzeff − Bx cos (ω t )
This is the magnetic field we apply in an NMR experiment and being able to visualize the
dynamics will help us understand how the experiment works. It is relatively straightforward
to work out the associated Bloch Equations for this magnetic field. They are:
dM x ( t )
= γ M y ( t ) Bzeff
dt
dM y ( t )
= −γ ⎡⎣ M x ( t ) Bzeff + M z ( t ) Bx cos (ω t )⎤⎦
dt
dM z ( t )
= −γ M y ( t ) Bx cos (ωt )
dt
These equations can actually be solved analytically to obtain the magnetization as a function
of time. From these equations we obtain the picture below:

ω � ω0 ω = ω0 ω � ω0
Here, we are plotting the magnetization as a function of time for various choices of the
frequency of the oscillating magnetic field component. If our field oscillates too quickly
(first case) then the magnetization just sees the average field and noting interesting
happens – we just get precession about the average field. If the oscillating field is too slow,
the magnetization oscillates around the instantaneous field and we get a sort of hula­hoop
motion of the magnetization. However, if we hit the frequency just right (middle) we can get
the magnetization to invert – to go from “up” to “down”. Thus, we see that the absorption
condition in NMR is associated with flipping the magnetization of the system.

Now, we note that at resonance, with the field on continuously, the spin will actually flip from
“up” to “down” and back to “up” and back to “down”… as a function of time. It is this

Prepared By Kevin Boudreaux


226 of 227
MIT Open Courseware Compillation
Source:
http://ocw.mit.edu/courses/chemistry/5-61-physical-chemistry-fall-2007/lecture-notes/

5.61 Physical Chemistry Lecture #36 Page 10

oscillation that shows up in our NMR spectrum. However, it is possible to turn the
oscillating field on and off as a function of time. Thus, for example, if we kept the field on
for exactly π/ω0 then the system would only have time to flip one time – all the up spins would
be converted to down and vice versa. Such a pulsed magnetic field is called an inversion
pulse, for obvious reasons. Meanwhile, if we kept the field on for exactly π/2ω0 we could
drive all the magnetization into the x­y plane. This is called a π/2 pulse. Further, we note
these pulses only work if we are on resonance with a particular proton’s Larmor frequency;
from the above figure it is clear that if we are off resonance, we can’t get the spins to flip.
Thus, one can imagine fairly complex sequences of inversion pulses and π/2 pulses applied at
various frequencies being used to isolate different couplings within a complicated molecule
(like a protein). Thus, it should not be surprising that cutting edge NMR experiments are all
time­resolved in order to extract the maximum information from the molecule.

Prepared By Kevin Boudreaux


227 of 227

Das könnte Ihnen auch gefallen