You are on page 1of 771

CHEM2350

Foundations of Physical Chemistry

Quantum Mechanics

Marcelo P. de Miranda
m.miranda@leeds.ac.uk
room 2.92a

Slide 1
Remember this?

Slide 2 · Introduction
Course structure

Foundations of physical chemistry:


• quantum mechanics
• chemical thermodynamics
• chemical kinetics

Quantum mechanics:
• 22 lectures (weeks 1–8, 1/3 discussion)
• 2 tutorials (weeks 5 and 9)
• 3 online assessments (weeks 3, 7 and 11)

Slide 3 · Introduction
Assessments

Online:
• go hand in hand with “questions you should ask”
• worth 15% of final mark
• for each assessment, three attempts allowed:
I one attempt: assessment mark is the one obtained in this attempt
I two attempts: assessment mark is average of two marks
I three attempts: assessment mark is average of best two marks

Tutorials:
• worth 15% of final mark
• it’s your job to contact the tutor

Slide 4 · Introduction
Reading and online material

Textbooks:
• Engel & Reid’s “Physical Chemistry”
• McQuarrie & Simon’s “Physical Chemistry”
• Atkins’ “Quanta — A Handbook of Concepts”
• Atkins’ “Physical Chemistry”
• Atkins’ “Elements of Physical Chemistry”

Online:
• commented versions of all my slides / handouts
• maple worksheets used or mentioned in lectures
• other maple worksheets with derivation of results
• additional reading suggestions

Slide 5 · Introduction
While this doesn’t happen...

...some bad news:


• this is a hard course
• this is a hard year

Slide 6 · Introduction
While this doesn’t happen...

...some bad news:


• this is a hard course
• this is a hard year

First-year material needed:


• QM part of chem1210
• mathematical concepts
(complex numbers,
differentials, integrals)
• maple

Slide 6 · Introduction
Quantum mechanics: what and why

What is quantum mechanics?


• theory of how objects behave in presence of other objects
• microscopic point of view
• bottom-up approach

Why quantum mechanics?


• explains existence, structure, states
and properties of atoms, molecules, etc.
• explains otherwise unexplained experimental observations
• predicts experimental observations

Slide 7 · Introduction
Some things QM predicts and explains

Atoms and molecules have different states...


• what states are those?
• what properties do they have?
Eg, what’s the energy of a given state?
• what keeps atoms together in a molecule?
In other words, what is a chemical bond?

...and their states can change.


• how long will atoms and molecules stay in a given state?
• how reactive is a given state?
• how are chemical bonds formed or broken?

Slide 8 · Introduction
QM and the foundations of physical chemistry

Thermodynamics... ...kinetics... ...and dynamics

Slide 9 · Introduction
QM and the foundations of physical chemistry

Thermodynamics... ...kinetics... ...and dynamics


What equilibrium
state?

Slide 9 · Introduction
QM and the foundations of physical chemistry

Thermodynamics... ...kinetics... ...and dynamics


How fast towards
equilibrium?

Slide 9 · Introduction
QM and the foundations of physical chemistry

Thermodynamics... ...kinetics... ...and dynamics


Processes leading to
equilibrium?

Slide 9 · Introduction
QM and the foundations of physical chemistry

Thermodynamics... ...kinetics... ...and dynamics


What equilibrium How fast towards Processes leading to
state? equilibrium? equilibrium?

The bottom line: chemistry is a molecular science.


• macroscopic behaviour explained in terms of molecular properties
(CO2 a greenhouse gas, H2 a good fuel, etc.)
• quantum devices and techniques used extensively
(lasers, spectroscopy, computational chemistry, etc.)
• technology development (materials and drug design,
molecular biology, nanotechnology, quantum computers, etc.)

Slide 9 · Introduction
Quantum v classical mechanics

Quantum mechanics Classical mechanics


• fundamental theory: • phenomenological theory:
what objects really are, useful model of reality that
how they really behave actually misrepresents it
• valid at every scale • valid in restricted domain
• global • local
• computationally hard • computationally easy
• partially deterministic • fully deterministic
• less intuitive • more intuitive

When is QM necessary? Things to consider:


wave-particle duality, quantisation, uncertainty, tunneling.

Slide 10 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty
10. Superposition of wavefunctions

Slide 11 · Introduction
The things we’ll see

QM principles and concepts Models and approximations


1. Complex numbers 11. Translation:
2. Questions and answers molecule in a 3D box
3. Wavefunctions 12. Vibration:
harmonic oscillator
4. Observables and operators
13. Rotation: rigid rotor
5. Expectation values
and eigenvalues 14. Electronic structure:
hydrogen atom
6. TI and TD Schrödinger equations
15. Spin
7. Phase and wave-particle duality
8. Overlap and probability
9. Uncertainty (all of this in 22 lectures!)
10. Superposition of wavefunctions

Slide 11 · Introduction
While this doesn’t happen...

...some bad news:


• this is a hard course
• this is a hard year

Slide 12 · Introduction
Part I

The basic ingredients


of quantum mechanics:
ideas, principles,
vocabulary and concepts

Slide 13 · QM principles and concepts


Part I — QM principles and concepts

1 Complex numbers

2 Questions and answers

3 Wavefunctions

Complex numbers: 4 Observables and operators

5 Expectation values and eigenvalues


numbers with 6 TI and TD Schrödinger equations

magnitude and direction 7 Phase and wave-particle duality

8 Overlap and probability

9 Uncertainty

10 Superposition of wavefunctions

Slide 14 · QM principles and concepts · Complex numbers


Why are some numbers complex?

From the
Oxford English Dictionary
complex, a. consisting of
various parts united or
connected together; formed
by combination of
different elements;
composite, compound.

Note: no mention of
“complicated” or “difficult.”
A complex number is a number
with more than one part.

Slide 15 · QM principles and concepts · Complex numbers


Why are some numbers complex?

From the
Why more than one part?
Oxford English Dictionary
complex, a. consisting of Never, ever, ever forget this:
various parts united or a complex number is a number
connected together; formed that has a magnitude and also
by combination of a direction.
different elements;
Numbers that consist of only
composite, compound.
one part (natural, real, etc.)
have magnitude but no
Note: no mention of
direction.
“complicated” or “difficult.”
Complex numbers, however,
A complex number is a number
have magnitude and direction.
with more than one part.

Slide 15 · QM principles and concepts · Complex numbers


The imaginary unit (i)
One of the components of a complex number is “i ”, the imaginary unit.
It has the following very important property:

p
i= −1

We deal with i in the same way we would deal with any other number.
Apart from being the square root of −1, there’s nothing special about it.

Here are some consequences of that property:

i 2 = −1;
(−i )2 = (−1)2 × i 2 = 1 × (−1) = −1;
i × (−i ) = −(i × i ) = −(i 2 ) = −(−1) = 1.
√i π/2
(And here is a teaser: can you prove that i =e ?)
Slide 16 · QM principles and concepts · Complex numbers
The parts of a complex number

A complex number consists of two parts. Depending on how you look


at it, the two parts are:
• the real part and the imaginary part
• the modulus and the phase

These parts are all related to each other. Each of them, including the
imaginary part, is a real number; the modulus is nonnegative.

We can understand all that by looking at the graphical representation


of complex numbers.

For instance, we’ll see that the modulus and the phase are the
numbers that specify the magnitude and the direction of a complex
number.

Slide 17 · QM principles and concepts · Complex numbers


Cartesian and polar forms of a complex number

Cartesian representation Polar representation


The two parts of a z are: The two parts of a z are:
• the real part, Re(z) • the modulus, |z|
• the imaginary part, Im(z) • the phase, arg(z)
If we represent the value of the If we represent the value of the
real part by “x” and the value of modulus by “r” and the value of
the imaginary part by “y”, then the phase by “ϕ”, then

z = x + i y. z = rei ϕ .

This is the Cartesian form of z. This is the polar form of z.

Slide 18 · QM principles and concepts · Complex numbers


Cartesian and polar forms of a complex number

Cartesian representation Polar representation


The two parts of a z are: The two parts of a z are:
• the real part, Re(z) • the modulus, |z|
• the imaginary part, Im(z) • the phase, arg(z)
If we represent the value of the If we represent the value of the
real part by “x” and the value of modulus by “r” and the value of
the imaginary part by “y”, then the phase by “ϕ”, then

z = x + i y. z = rei ϕ .

This is the Cartesian form of z. This is the polar form of z.

Slide 18 · QM principles and concepts · Complex numbers


Cartesian and polar forms of a complex number

Cartesian representation Polar representation


The two parts of a z are: The two parts of a z are:
• the real part, Re(z) • the modulus, |z|
• the imaginary part, Im(z) • the phase, arg(z)
If we represent the value of the If we represent the value of the
real part by “x” and the value of modulus by “r” and the value of
the imaginary part by “y”, then the phase by “ϕ”, then

z = x + i y. z = rei ϕ .

This is the Cartesian form of z. This is the polar form of z.

Slide 18 · QM principles and concepts · Complex numbers


Cartesian and polar forms of a complex number

Cartesian representation Polar representation


The two parts of a z are: The two parts of a z are:
• the real part, Re(z) • the modulus, |z|
• the imaginary part, Im(z) • the phase, arg(z)
If we represent the value of the If we represent the value of the
real part by “x” and the value of modulus by “r” and the value of
the imaginary part by “y”, then the phase by “ϕ”, then

z = x + i y. z = rei ϕ .

This is the Cartesian form of z. This is the polar form of z.


Note that x, y, r and ϕ are all real numbers and that r ≥ 0.
e = 2.71828 . . . is the number we know from the exponential and
logarithm functions: we have exp(1) = ln(1) = e = 2.71828 . . .

Slide 18 · QM principles and concepts · Complex numbers


Graphical representation of complex numbers (1)

We will look at the graphical representations of two complex numbers,


z1 and z2 , chosen as follows:


z1 : modulus r1 = 2, phase ϕ1 = π/3 = 60
z2 : modulus r2 = 3/2, phase ϕ2 = 7π/6 = 210◦

Among other things we will see that this implies that the real and
imaginary parts of z1 and z2 are given by


z1 : real part x1 = 1, imaginary part y1 = 3

z2 : real part x2 = −3 3/4, imaginary part y2 = −3/4

Slide 19 · QM principles and concepts · Complex numbers


Graphical representation of complex numbers (2)

The complex number z1 The complex number z2


Im(z) Im(z)
z1
3 3
2 2

1 1
1 1
2 2

Re(z) Re(z)
− 32 −1 − 1 1 1 3 − 32 −1 1 1 3
2 2 2 2 2
− 12 − 12
z2
−1 −1

− 32 − 32

Never, ever, ever forget this: a complex number is a


number that has a magnitude and also a direction.

Slide 20 · QM principles and concepts · Complex numbers


Graphical representation of complex numbers (3)

The complex number z1 The complex number z2


Im(z) Im(z)
z1
3 3
2 2

r1 1
ϕ2
ϕ1
Re(z) Re(z)
− 32 −1 − 1 1 1 3 − 32 −1 1 1 3
2 2 2 2 2
− 12
z2 r1
−1 −1

− 32 − 32

The modulus gives the magnitude,


whereas the phase gives the direction.

Slide 21 · QM principles and concepts · Complex numbers


Graphical representation of complex numbers (4)

The complex number z1 The complex number z2


Im(z) Im(z)
y1 z1
3
2

1 1
1 1
2 2
x2
Re(z) Re(z)
− 32 −1 − 1 1
x1 3 − 32 −1 1 1 3
2 2 2 2 2
− 12 − 12
z2 y2
−1 −1

− 32 − 32

The horizontal component gives the real part,


whereas the vertical component gives the imaginary part.

Slide 22 · QM principles and concepts · Complex numbers


Some important relationships (1)

Note that x, y, r and ϕ are related through a right-angle triangle:

• x = r cos ϕ
• y = r sin ϕ
• r 2 = x2 + y 2, r =
p
x2 + y 2
r |y|
• ϕ̃ = arctan(y/x) = tan−1 (y/x)
• ϕ = ϕ̃ if x > 0,
ϕ = ϕ̃ + π if x<0
ϕ̃
|x|

Slide 23 · QM principles and concepts · Complex numbers


Some important relationships (2)

• we have two expressions for the complex number z:



z = x + i y and z = re
• together these expressions imply x + i y = rei ϕ
• the expressions for x and y in the previous slide imply
x + i y = (r cos ϕ) + i (r sin ϕ) = r(cos ϕ + i sin ϕ)
• the two equations imply ei ϕ = cos ϕ + i sin ϕ
• Using cos(−ϕ) = cos ϕ and sin(−ϕ) = − sin ϕ we get

±i ϕ
e = cos ϕ ± i sin ϕ,

a very important formula known as “Euler’s identity”.

Slide 24 · QM principles and concepts · Complex numbers


Some important relationships (2)

• we have two expressions for the complex number z:



z = x + i y and z = re
• together these expressions imply x + i y = rei ϕ
• the expressions for x and y in the previous slide imply
x + i y = (r cos ϕ) + i (r sin ϕ) = r(cos ϕ + i sin ϕ)
• the two equations imply ei ϕ = cos ϕ + i sin ϕ
• Using cos(−ϕ) = cos ϕ and sin(−ϕ) = − sin ϕ we get

±i ϕ
e = cos ϕ ± i sin ϕ,

a very important formula known as “Euler’s identity”.

Slide 24 · QM principles and concepts · Complex numbers


Some important relationships (2)

• we have two expressions for the complex number z:



z = x + i y and z = re
• together these expressions imply x + i y = rei ϕ
• the expressions for x and y in the previous slide imply
x + i y = (r cos ϕ) + i (r sin ϕ) = r(cos ϕ + i sin ϕ)
• the two equations imply ei ϕ = cos ϕ + i sin ϕ
• Using cos(−ϕ) = cos ϕ and sin(−ϕ) = − sin ϕ we get

±i ϕ
e = cos ϕ ± i sin ϕ,

a very important formula known as “Euler’s identity”.

Slide 24 · QM principles and concepts · Complex numbers


Some important relationships (2)

• we have two expressions for the complex number z:



z = x + i y and z = re
• together these expressions imply x + i y = rei ϕ
• the expressions for x and y in the previous slide imply
x + i y = (r cos ϕ) + i (r sin ϕ) = r(cos ϕ + i sin ϕ)
• the two equations imply ei ϕ = cos ϕ + i sin ϕ
• Using cos(−ϕ) = cos ϕ and sin(−ϕ) = − sin ϕ we get

±i ϕ
e = cos ϕ ± i sin ϕ,

a very important formula known as “Euler’s identity”.

Slide 24 · QM principles and concepts · Complex numbers


Some important relationships (2)

• we have two expressions for the complex number z:



z = x + i y and z = re
• together these expressions imply x + i y = rei ϕ
• the expressions for x and y in the previous slide imply
x + i y = (r cos ϕ) + i (r sin ϕ) = r(cos ϕ + i sin ϕ)
• the two equations imply ei ϕ = cos ϕ + i sin ϕ
• Using cos(−ϕ) = cos ϕ and sin(−ϕ) = − sin ϕ we get

±i ϕ
e = cos ϕ ± i sin ϕ,

a very important formula known as “Euler’s identity”.

Slide 24 · QM principles and concepts · Complex numbers


Some important relationships (2)

• we have two expressions for the complex number z:



z = x + i y and z = re
• together these expressions imply x + i y = rei ϕ
• the expressions for x and y in the previous slide imply
x + i y = (r cos ϕ) + i (r sin ϕ) = r(cos ϕ + i sin ϕ)
• the two equations imply ei ϕ = cos ϕ + i sin ϕ
• Using cos(−ϕ) = cos ϕ and sin(−ϕ) = − sin ϕ we get

±i ϕ
e = cos ϕ ± i sin ϕ,

a very important formula known as “Euler’s identity”.

Slide 24 · QM principles and concepts · Complex numbers


Complex conjugation (1)


Conjugation of the complex number z is represented by z , and is
done by replacing i by −i . Some examples:
∗
(1 + i )∗ = 1 − i , 2 exp(i π/4) = 2 exp(−i π/4)


√ √ ∗ √ √ "r #∗ r
2−i 3 2+i 3
q q
2 −i 3 2 i 3
= , e 2 = e 2
2 2 3 3

cos(1 − i x) ∗ cos(1 + i x)
 
∗ −i y
iy
= (−i x) =

(i x) ,
sin(2i x) − 4i sin(−2i x) + 4i

Slide 25 · QM principles and concepts · Complex numbers


Complex conjugation (1)


Conjugation of the complex number z is represented by z , and is
done by replacing i by −i . Some examples:
∗
(1 + i )∗ = 1 − i , 2 exp(i π/4) = 2 exp(−i π/4)


√ √ ∗ √ √ "r #∗ r
2−i 3 2+i 3
q q
2 −i 3 2 i 3
= , e 2 = e 2
2 2 3 3

cos(1 − i x) ∗ cos(1 + i x)
 
∗ −i y
iy
= (−i x) =

(i x) ,
sin(2i x) − 4i sin(−2i x) + 4i

Slide 25 · QM principles and concepts · Complex numbers


Complex conjugation (1)


Conjugation of the complex number z is represented by z , and is
done by replacing i by −i . Some examples:
∗
(1 + i )∗ = 1 − i , 2 exp(i π/4) = 2 exp(−i π/4)


√ √ ∗ √ √ "r #∗ r
2−i 3 2+i 3
q q
2 −i 3 2 i 3
= , e 2 = e 2
2 2 3 3

cos(1 − i x) ∗ cos(1 + i x)
 
∗ −i y
iy
= (−i x) =

(i x) ,
sin(2i x) − 4i sin(−2i x) + 4i

Slide 25 · QM principles and concepts · Complex numbers


Complex conjugation (1)


Conjugation of the complex number z is represented by z , and is
done by replacing i by −i . Some examples:
∗
(1 + i )∗ = 1 − i , 2 exp(i π/4) = 2 exp(−i π/4)


√ √ ∗ √ √ "r #∗ r
2−i 3 2+i 3
q q
2 −i 3 2 i 3
= , e 2 = e 2
2 2 3 3

cos(1 − i x) ∗ cos(1 + i x)
 
∗ −i y
iy
= (−i x) =

(i x) ,
sin(2i x) − 4i sin(−2i x) + 4i

Slide 25 · QM principles and concepts · Complex numbers


Complex conjugation (1)


Conjugation of the complex number z is represented by z , and is
done by replacing i by −i . Some examples:
∗
(1 + i )∗ = 1 − i , 2 exp(i π/4) = 2 exp(−i π/4)


√ √ ∗ √ √ "r #∗ r
2−i 3 2+i 3
q q
2 −i 3 2 i 3
= , e 2 = e 2
2 2 3 3

cos(1 − i x) ∗ cos(1 + i x)
 
∗ −i y
iy
= (−i x) =

(i x) ,
sin(2i x) − 4i sin(−2i x) + 4i

Slide 25 · QM principles and concepts · Complex numbers


Complex conjugation (1)


Conjugation of the complex number z is represented by z , and is
done by replacing i by −i . Some examples:
∗
(1 + i )∗ = 1 − i , 2 exp(i π/4) = 2 exp(−i π/4)


√ √ ∗ √ √ "r #∗ r
2−i 3 2+i 3
q q
2 −i 3 2 i 3
= , e 2 = e 2
2 2 3 3

cos(1 − i x) ∗ cos(1 + i x)
 
∗ −i y
iy
= (−i x) =

(i x) ,
sin(2i x) − 4i sin(−2i x) + 4i

Slide 25 · QM principles and concepts · Complex numbers


Complex conjugation (1)


Conjugation of the complex number z is represented by z , and is
done by replacing i by −i . Some examples:
∗
(1 + i )∗ = 1 − i , 2 exp(i π/4) = 2 exp(−i π/4)


√ √ ∗ √ √ "r #∗ r
2−i 3 2+i 3
q q
2 −i 3 2 i 3
= , e 2 = e 2
2 2 3 3

cos(1 − i x) ∗ cos(1 + i x)
 
∗ −i y
iy
= (−i x) =

(i x) ,
sin(2i x) − 4i sin(−2i x) + 4i

It doesn’t matter how complicated a complex expression is. Its


complex conjugate is always obtained through replacement of i by −i .

Slide 25 · QM principles and concepts · Complex numbers


Complex conjugation (2)

Multiplication of a complex number by its conjugate results in a real


number — the square of the modulus of the complex number:

z × z ∗ = z ∗ × z = |z|2

Slide 26 · QM principles and concepts · Complex numbers


Complex conjugation (2)

Multiplication of a complex number by its conjugate results in a real


number — the square of the modulus of the complex number:

z × z ∗ = z ∗ × z = |z|2

Here’s a way of seeing it:


• every complex expression can be written in polar form: z = rei ϕ
• the complex conjugate of z = rei ϕ is z ∗ = re−i ϕ
• their product is z ∗ z = re−i ϕ × rei ϕ = r 2 ei ϕ−i ϕ = r 2 e0 = r 2 = |z|2

Slide 26 · QM principles and concepts · Complex numbers


Complex conjugation (2)

Multiplication of a complex number by its conjugate results in a real


number — the square of the modulus of the complex number:

z × z ∗ = z ∗ × z = |z|2

Here’s a way of seeing it:


• every complex expression can be written in polar form: z = rei ϕ
• the complex conjugate of z = rei ϕ is z ∗ = re−i ϕ
• their product is z ∗ z = re−i ϕ × rei ϕ = r 2 ei ϕ−i ϕ = r 2 e0 = r 2 = |z|2

Slide 26 · QM principles and concepts · Complex numbers


Complex conjugation (2)

Multiplication of a complex number by its conjugate results in a real


number — the square of the modulus of the complex number:

z × z ∗ = z ∗ × z = |z|2

Here’s a way of seeing it:


• every complex expression can be written in polar form: z = rei ϕ
• the complex conjugate of z = rei ϕ is z ∗ = re−i ϕ
• their product is z ∗ z = re−i ϕ × rei ϕ = r 2 ei ϕ−i ϕ = r 2 e0 = r 2 = |z|2

Slide 26 · QM principles and concepts · Complex numbers


Complex conjugation (2)

Multiplication of a complex number by its conjugate results in a real


number — the square of the modulus of the complex number:

z × z ∗ = z ∗ × z = |z|2

Here’s a way of seeing it:


• every complex expression can be written in polar form: z = rei ϕ
• the complex conjugate of z = rei ϕ is z ∗ = re−i ϕ
• their product is z ∗ z = re−i ϕ × rei ϕ = r 2 ei ϕ−i ϕ = r 2 e0 = r 2 = |z|2

Slide 26 · QM principles and concepts · Complex numbers


Complex conjugation (2)

Multiplication of a complex number by its conjugate results in a real


number — the square of the modulus of the complex number:

z × z ∗ = z ∗ × z = |z|2

Here’s a way of seeing it:


• every complex expression can be written in polar form: z = rei ϕ
• the complex conjugate of z = rei ϕ is z ∗ = re−i ϕ
• their product is z ∗ z = re−i ϕ × rei ϕ = r 2 ei ϕ−i ϕ = r 2 e0 = r 2 = |z|2

Slide 26 · QM principles and concepts · Complex numbers


Complex conjugation (2)

Multiplication of a complex number by its conjugate results in a real


number — the square of the modulus of the complex number:

z × z ∗ = z ∗ × z = |z|2

Here’s a way of seeing it:


• every complex expression can be written in polar form: z = rei ϕ
• the complex conjugate of z = rei ϕ is z ∗ = re−i ϕ
• their product is z ∗ z = re−i ϕ × rei ϕ = r 2 ei ϕ−i ϕ = r 2 e0 = r 2 = |z|2

Slide 26 · QM principles and concepts · Complex numbers


Complex conjugation (2)

Multiplication of a complex number by its conjugate results in a real


number — the square of the modulus of the complex number:

z × z ∗ = z ∗ × z = |z|2

Here’s a way of seeing it:


• every complex expression can be written in polar form: z = rei ϕ
• the complex conjugate of z = rei ϕ is z ∗ = re−i ϕ
• their product is z ∗ z = re−i ϕ × rei ϕ = r 2 ei ϕ−i ϕ = r 2 e0 = r 2 = |z|2

Slide 26 · QM principles and concepts · Complex numbers


Complex conjugation (2)

Multiplication of a complex number by its conjugate results in a real


number — the square of the modulus of the complex number:

z × z ∗ = z ∗ × z = |z|2

Here’s a way of seeing it:


• every complex expression can be written in polar form: z = rei ϕ
• the complex conjugate of z = rei ϕ is z ∗ = re−i ϕ
• their product is z ∗ z = re−i ϕ × rei ϕ = r 2 ei ϕ−i ϕ = r 2 e0 = r 2 = |z|2

∗ ∗
Note also that z = z.

Slide 26 · QM principles and concepts · Complex numbers


The importance of complex numbers for QM

2 2
Like x + 5 = 0 or x − 2x + 2 = 0, many equations cannot be solved
with real numbers only. Their solutions are complex.

This is the case with many of the equations of quantum mechanics:


their solutions are complex.

To solve the equations of QM — to understand nature, and chemistry


in particular — we must use complex numbers.

Slide 27 · QM principles and concepts · Complex numbers


Some things you should remember...

• complex numbers are of the form z = rei ϕ = x + i y


• real numbers are complex numbers with Im(z) = 0
and either arg(z) = 0 or arg(z) = π
• imaginary numbers are complex numbers with Re(z) = 0
and either arg(z) = π/2 or arg(z) = −π/2
• the imaginary part of a complex number is a real number
• complex conjugation is done by replacing i by −i
• multiplication of a complex number by its conjugate
gives the square of the modulus of the complex number
• many of the equations of QM have complex solutions
• imaginary numbers are called imaginary because (like negative
and irrational numbers before them) they were victims of prejudice

Slide 28 · QM principles and concepts · Complex numbers


Some things you should remember...

• complex numbers are of the form z = rei ϕ = x + i y


• real numbers are complex numbers with Im(z) = 0
and either arg(z) = 0 or arg(z) = π
• imaginary numbers are complex numbers with Re(z) = 0
and either arg(z) = π/2 or arg(z) = −π/2
• the imaginary part of a complex number is a real number
• complex conjugation is done by replacing i by −i
• multiplication of a complex number by its conjugate
gives the square of the modulus of the complex number
• many of the equations of QM have complex solutions
• imaginary numbers are called imaginary because (like negative
and irrational numbers before them) they were victims of prejudice

Slide 28 · QM principles and concepts · Complex numbers


Some things you should remember...

• complex numbers are of the form z = rei ϕ = x + i y


• real numbers are complex numbers with Im(z) = 0
and either arg(z) = 0 or arg(z) = π
• imaginary numbers are complex numbers with Re(z) = 0
and either arg(z) = π/2 or arg(z) = −π/2
• the imaginary part of a complex number is a real number
• complex conjugation is done by replacing i by −i
• multiplication of a complex number by its conjugate
gives the square of the modulus of the complex number
• many of the equations of QM have complex solutions
• imaginary numbers are called imaginary because (like negative
and irrational numbers before them) they were victims of prejudice

Slide 28 · QM principles and concepts · Complex numbers


Some things you should remember...

• complex numbers are of the form z = rei ϕ = x + i y


• real numbers are complex numbers with Im(z) = 0
and either arg(z) = 0 or arg(z) = π
• imaginary numbers are complex numbers with Re(z) = 0
and either arg(z) = π/2 or arg(z) = −π/2
• the imaginary part of a complex number is a real number
• complex conjugation is done by replacing i by −i
• multiplication of a complex number by its conjugate
gives the square of the modulus of the complex number
• many of the equations of QM have complex solutions
• imaginary numbers are called imaginary because (like negative
and irrational numbers before them) they were victims of prejudice

Slide 28 · QM principles and concepts · Complex numbers


Some things you should remember...

• complex numbers are of the form z = rei ϕ = x + i y


• real numbers are complex numbers with Im(z) = 0
and either arg(z) = 0 or arg(z) = π
• imaginary numbers are complex numbers with Re(z) = 0
and either arg(z) = π/2 or arg(z) = −π/2
• the imaginary part of a complex number is a real number
• complex conjugation is done by replacing i by −i
• multiplication of a complex number by its conjugate
gives the square of the modulus of the complex number
• many of the equations of QM have complex solutions
• imaginary numbers are called imaginary because (like negative
and irrational numbers before them) they were victims of prejudice

Slide 28 · QM principles and concepts · Complex numbers


Some things you should remember...

• complex numbers are of the form z = rei ϕ = x + i y


• real numbers are complex numbers with Im(z) = 0
and either arg(z) = 0 or arg(z) = π
• imaginary numbers are complex numbers with Re(z) = 0
and either arg(z) = π/2 or arg(z) = −π/2
• the imaginary part of a complex number is a real number
• complex conjugation is done by replacing i by −i
• multiplication of a complex number by its conjugate
gives the square of the modulus of the complex number
• many of the equations of QM have complex solutions
• imaginary numbers are called imaginary because (like negative
and irrational numbers before them) they were victims of prejudice

Slide 28 · QM principles and concepts · Complex numbers


Some things you should remember...

• complex numbers are of the form z = rei ϕ = x + i y


• real numbers are complex numbers with Im(z) = 0
and either arg(z) = 0 or arg(z) = π
• imaginary numbers are complex numbers with Re(z) = 0
and either arg(z) = π/2 or arg(z) = −π/2
• the imaginary part of a complex number is a real number
• complex conjugation is done by replacing i by −i
• multiplication of a complex number by its conjugate
gives the square of the modulus of the complex number
• many of the equations of QM have complex solutions
• imaginary numbers are called imaginary because (like negative
and irrational numbers before them) they were victims of prejudice

Slide 28 · QM principles and concepts · Complex numbers


Some things you should remember...

• complex numbers are of the form z = rei ϕ = x + i y


• real numbers are complex numbers with Im(z) = 0
and either arg(z) = 0 or arg(z) = π
• imaginary numbers are complex numbers with Re(z) = 0
and either arg(z) = π/2 or arg(z) = −π/2
• the imaginary part of a complex number is a real number
• complex conjugation is done by replacing i by −i
• multiplication of a complex number by its conjugate
gives the square of the modulus of the complex number
• many of the equations of QM have complex solutions
• imaginary numbers are called imaginary because (like negative
and irrational numbers before them) they were victims of prejudice

Slide 28 · QM principles and concepts · Complex numbers


...and something you should forget

Eleventeen, thirty-twelve, etc, are not imaginary numbers.

Slide 29 · QM principles and concepts · Complex numbers


Part I — QM principles and concepts

1 Complex numbers

2 Questions and answers

Questions and answers: 3 Wavefunctions

4 Observables and operators


stuff we know, 5 Expectation values and eigenvalues
stuff we want to know, 6 TI and TD Schrödinger equations

who to ask, how to ask 7 Phase and wave-particle duality

and how to interpret the answers 8 Overlap and probability

9 Uncertainty

10 Superposition of wavefunctions

Slide 30 · QM principles and concepts · Questions and answers


Stuff we know, stuff we want to know

We know how things acquire energy. For the purposes of this course,
they do it in one of two ways:
• Motion leads to kinetic energy. The faster things move,
the larger their kinetic energy.
• Interaction leads to potential energy. The more strongly things
interact, the larger (in modulus) their potential energy.

We want to know what complex things (say, molecules) are like:


• What states can we find them in?
• What are the physical and chemical properties
associated with each of the possible states?
• How can states change or be changed?

Slide 31 · QM principles and concepts · Questions and answers


Stuff we know, stuff we want to know

We know how things acquire energy. For the purposes of this course,
they do it in one of two ways:
• Motion leads to kinetic energy. The faster things move,
the larger their kinetic energy.
• Interaction leads to potential energy. The more strongly things
interact, the larger (in modulus) their potential energy.

We want to know what complex things (say, molecules) are like:


• What states can we find them in?
• What are the physical and chemical properties
associated with each of the possible states?
• How can states change or be changed?

Slide 31 · QM principles and concepts · Questions and answers


Who to ask, how to ask

According to QM, the state and properties of every physical system


are described by its wavefunction, usually represented by Ψ or ψ.
In a sense, Ψ is like an oracle: it can tell us all there is to know
about the state and properties of the system it refers to.
If we want to know about the state and properties of a system, all we
have to do is ask the wavefunction.

In order to extract information from the wavefunction, we need


mathematical operations. We need to operate on it with mathematical
operators, usually represented by roman letters with a circumflex
accent: things like x̂, Ô, Ŝ or ̂.
Operators involve relatively simple mathematical operations
(multiplication, differentiation, etc) or combinations of them.

Slide 32 · QM principles and concepts · Questions and answers


Who to ask, how to ask

According to QM, the state and properties of every physical system


are described by its wavefunction, usually represented by Ψ or ψ.
In a sense, Ψ is like an oracle: it can tell us all there is to know
about the state and properties of the system it refers to.
If we want to know about the state and properties of a system, all we
have to do is ask the wavefunction.

In order to extract information from the wavefunction, we need


mathematical operations. We need to operate on it with mathematical
operators, usually represented by roman letters with a circumflex
accent: things like x̂, Ô, Ŝ or ̂.
Operators involve relatively simple mathematical operations
(multiplication, differentiation, etc) or combinations of them.

Slide 32 · QM principles and concepts · Questions and answers


Who to ask, how to ask

According to QM, the state and properties of every physical system


are described by its wavefunction, usually represented by Ψ or ψ.
In a sense, Ψ is like an oracle: it can tell us all there is to know
about the state and properties of the system it refers to.
If we want to know about the state and properties of a system, all we
have to do is ask the wavefunction.

In order to extract information from the wavefunction, we need


mathematical operations. We need to operate on it with mathematical
operators, usually represented by roman letters with a circumflex
accent: things like x̂, Ô, Ŝ or ̂.
Operators involve relatively simple mathematical operations
(multiplication, differentiation, etc) or combinations of them.

Slide 32 · QM principles and concepts · Questions and answers


How to interpret the answers
True to its “oracle” nature, the wavefunction often answers our
questions in a less-than-straightforward way:
• Sometimes we find that the answer we want does not exist.
• If it exists, the answer may come in the form of an
expectation value, and be less precise than we would wish.
• Sometimes the answer comes in the form of an eigenvalue.
This is more like a clear, unambiguous answer.

In order to understand the information extracted from wavefunctions,


it is very important that we understand these things well:
• What are eigenvalues and expectation values?
• When and why will our questions be answered by either of those?
• When will those answers be limited by uncertainty?
• When will it be impossible for us to obtain answers?

Slide 33 · QM principles and concepts · Questions and answers


How to interpret the answers
True to its “oracle” nature, the wavefunction often answers our
questions in a less-than-straightforward way:
• Sometimes we find that the answer we want does not exist.
• If it exists, the answer may come in the form of an
expectation value, and be less precise than we would wish.
• Sometimes the answer comes in the form of an eigenvalue.
This is more like a clear, unambiguous answer.

In order to understand the information extracted from wavefunctions,


it is very important that we understand these things well:
• What are eigenvalues and expectation values?
• When and why will our questions be answered by either of those?
• When will those answers be limited by uncertainty?
• When will it be impossible for us to obtain answers?

Slide 33 · QM principles and concepts · Questions and answers


Summary: questions and answers

Our general strategy is this:


1. We know how things acquire energy.
2. We want to know what complex things are like, and in particular
what are their states and properties.
3. We can obtain answers to our question by asking the
wavefunction.
4. The way to ask the wavefunction is to operate on it with
mathematical operators.
5. We must know how to interpret the answers we get.

Slide 34 · QM principles and concepts · Questions and answers


Summary: questions and answers

Our general strategy is this:


1. We know how things acquire energy.
2. We want to know what complex things are like, and in particular
what are their states and properties.
3. We can obtain answers to our question by asking the
wavefunction.
4. The way to ask the wavefunction is to operate on it with
mathematical operators.
5. We must know how to interpret the answers we get.

Slide 34 · QM principles and concepts · Questions and answers


Summary: questions and answers

Our general strategy is this:


1. We know how things acquire energy.
2. We want to know what complex things are like, and in particular
what are their states and properties.
3. We can obtain answers to our question by asking the
wavefunction.
4. The way to ask the wavefunction is to operate on it with
mathematical operators.
5. We must know how to interpret the answers we get.

Slide 34 · QM principles and concepts · Questions and answers


Summary: questions and answers

Our general strategy is this:


1. We know how things acquire energy.
2. We want to know what complex things are like, and in particular
what are their states and properties.
3. We can obtain answers to our question by asking the
wavefunction.
4. The way to ask the wavefunction is to operate on it with
mathematical operators.
5. We must know how to interpret the answers we get.

Slide 34 · QM principles and concepts · Questions and answers


Summary: questions and answers

Our general strategy is this:


1. We know how things acquire energy.
2. We want to know what complex things are like, and in particular
what are their states and properties.
3. We can obtain answers to our question by asking the
wavefunction.
4. The way to ask the wavefunction is to operate on it with
mathematical operators.
5. We must know how to interpret the answers we get.

Slide 34 · QM principles and concepts · Questions and answers


Summary: questions and answers

Our general strategy is this:


1. We know how things acquire energy.
2. We want to know what complex things are like, and in particular
what are their states and properties.
3. We can obtain answers to our question by asking the
wavefunction.
4. The way to ask the wavefunction is to operate on it with
mathematical operators.
5. We must know how to interpret the answers we get.

But there’s a problem: we don’t have the wavefunction.

Slide 34 · QM principles and concepts · Questions and answers


Part I — QM principles and concepts

1 Complex numbers

2 Questions and answers

3 Wavefunctions

Wavefunctions: 4 Observables and operators

5 Expectation values and eigenvalues


what they are 6 TI and TD Schrödinger equations

and what they mean 7 Phase and wave-particle duality

8 Overlap and probability

9 Uncertainty

10 Superposition of wavefunctions

Slide 35 · QM principles and concepts · Wavefunctions


Wavefunctions

According to QM, the states and properties of every physical system


are described by this system wavefunctions. Note:
• each particular system has a set of wavefunctions.
• each wavefunction corresponds to a particular state of the system.
• each wavefunction depends on parameters and variables.

Example of wavefunction:
r
2  nπ 
ψn (x) = sin x , n = 1, 2, 3, 4, 5, . . .
L L
This is the wavefunction of a particle of mass m
trapped in a 1-D box stretching from x = 0 to x = L.

Slide 36 · QM principles and concepts · Wavefunctions


Parameters

• specify either what system is being considered,


or else what quantum state is being considered.
• mass and length are examples of system parameters.
• quantum numbers are quantum state parameters.
• for a given system and quantum state,
the values of the parameters cannot vary.
• changing parameter values is equivalent
to changing the system or its quantum state.

Slide 37 · QM principles and concepts · Wavefunctions


Parameters

• specify either what system is being considered,


or else what quantum state is being considered.
• mass and length are examples of system parameters.
• quantum numbers are quantum state parameters.
• for a given system and quantum state,
the values of the parameters cannot vary.
• changing parameter values is equivalent
to changing the system or its quantum state.

Slide 37 · QM principles and concepts · Wavefunctions


Parameters

• specify either what system is being considered,


or else what quantum state is being considered.
• mass and length are examples of system parameters.
• quantum numbers are quantum state parameters.
• for a given system and quantum state,
the values of the parameters cannot vary.
• changing parameter values is equivalent
to changing the system or its quantum state.

Slide 37 · QM principles and concepts · Wavefunctions


Parameters

• specify either what system is being considered,


or else what quantum state is being considered.
• mass and length are examples of system parameters.
• quantum numbers are quantum state parameters.
• for a given system and quantum state,
the values of the parameters cannot vary.
• changing parameter values is equivalent
to changing the system or its quantum state.

Slide 37 · QM principles and concepts · Wavefunctions


Parameters

• specify either what system is being considered,


or else what quantum state is being considered.
• mass and length are examples of system parameters.
• quantum numbers are quantum state parameters.
• for a given system and quantum state,
the values of the parameters cannot vary.
• changing parameter values is equivalent
to changing the system or its quantum state.

Slide 37 · QM principles and concepts · Wavefunctions


Parameters

• specify either what system is being considered,


or else what quantum state is being considered.
• mass and length are examples of system parameters.
• quantum numbers are quantum state parameters.
• for a given system and quantum state,
the values of the parameters cannot vary.
• changing parameter values is equivalent
to changing the system or its quantum state.

The wavefunction shown on slide 36 depends on two parameters:


• L, the length of the box (identifies the system).
• n, the quantum number (specifies the quantum state).

Note that this wavefunction does not depend on the particle mass.

Slide 37 · QM principles and concepts · Wavefunctions


Variables

• specify neither the system nor its quantum state.


• coordinates and time are examples of variables.
• coordinates (x, y, etc.) specify the positions
of the various components (particles) of the system.
• for a particular system and quantum state,
the values of the variables can and do vary.
• a change in the value of a variable does not imply
a change of the system or its quantum state.

Slide 38 · QM principles and concepts · Wavefunctions


Variables

• specify neither the system nor its quantum state.


• coordinates and time are examples of variables.
• coordinates (x, y, etc.) specify the positions
of the various components (particles) of the system.
• for a particular system and quantum state,
the values of the variables can and do vary.
• a change in the value of a variable does not imply
a change of the system or its quantum state.

Slide 38 · QM principles and concepts · Wavefunctions


Variables

• specify neither the system nor its quantum state.


• coordinates and time are examples of variables.
• coordinates (x, y, etc.) specify the positions
of the various components (particles) of the system.
• for a particular system and quantum state,
the values of the variables can and do vary.
• a change in the value of a variable does not imply
a change of the system or its quantum state.

Slide 38 · QM principles and concepts · Wavefunctions


Variables

• specify neither the system nor its quantum state.


• coordinates and time are examples of variables.
• coordinates (x, y, etc.) specify the positions
of the various components (particles) of the system.
• for a particular system and quantum state,
the values of the variables can and do vary.
• a change in the value of a variable does not imply
a change of the system or its quantum state.

Slide 38 · QM principles and concepts · Wavefunctions


Variables

• specify neither the system nor its quantum state.


• coordinates and time are examples of variables.
• coordinates (x, y, etc.) specify the positions
of the various components (particles) of the system.
• for a particular system and quantum state,
the values of the variables can and do vary.
• a change in the value of a variable does not imply
a change of the system or its quantum state.

Slide 38 · QM principles and concepts · Wavefunctions


Variables

• specify neither the system nor its quantum state.


• coordinates and time are examples of variables.
• coordinates (x, y, etc.) specify the positions
of the various components (particles) of the system.
• for a particular system and quantum state,
the values of the variables can and do vary.
• a change in the value of a variable does not imply
a change of the system or its quantum state.

The wavefunction shown on slide 36 depends on only one variable:


• x, the particle coordinate (position).

Note that this wavefunction does not depend on time.

Slide 38 · QM principles and concepts · Wavefunctions


Born interpretation of the wavefunction

Never, ever forget this: the squared modulus of the wavefunction ψ(x)
1
gives the probability density of finding the particle at the position x.

Mathematically this is written as

ρ(x) = |ψ(x)|2

Note:
• we take the squared modulus, |ψ|2 (rather than ψ 2 )
• this gives the probability density (rather than the probability)

1
The reason why this result is known as an “interpretation” is historical.
Slide 39 · QM principles and concepts · Wavefunctions
Where in the 1-D box is the electron? (1)

1
)

r
2  nπx 
−1/2

0.5
• ψ(x) = sin
ψ4 (x) (nm

0
L L
• wavefunction has units
−0.5
• wavefunction
−1 can be negative

0 0.25 0.5 0.75 1


x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 40 · QM principles and concepts · Wavefunctions


Where in the 1-D box is the electron? (2)

2 h  nπx i2
1.5 • |ψ(x)|2 = sin
|ψ4 (x)| (nm )

L L
−1

• squared modulus of
1
wavefunction has units
2

• squared modulus of
0.5
wavefunction cannot be
negative
0
0 0.25 0.5 0.75 1
x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 41 · QM principles and concepts · Wavefunctions


Where in the 1-D box is the electron? (3)

• |ψ(0)|2 = 0 (because of
1.5
|ψ4 (x)| (nm )
−1

boundary conditions)
1
2

0.5

0 ·
0 0.25 0.5 0.75 1
x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 42 · QM principles and concepts · Wavefunctions


Where in the 1-D box is the electron? (3)

2 ·

1.5
|ψ4 (x)| (nm )
−1

1
• |ψ(0.125 nm)|2 = 2.00 nm−1
2

0.5

0
0 0.25 0.5 0.75 1
x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 42 · QM principles and concepts · Wavefunctions


Where in the 1-D box is the electron? (3)

2
·
1.5
|ψ4 (x)| (nm )
−1

1
2

• |ψ(0.4 nm)|2 = 1.81 nm−1


0.5

0
0 0.25 0.5 0.75 1
x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 42 · QM principles and concepts · Wavefunctions


Where in the 1-D box is the electron? (3)

1.5
|ψ4 (x)| (nm )
−1

1
2

·
0.5
• |ψ(0.7 nm)|2 = 0.69 nm−1

0
0 0.25 0.5 0.75 1
x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 42 · QM principles and concepts · Wavefunctions


Where in the 1-D box is the electron? (3)

1.5
|ψ4 (x)| (nm )
−1

1
2

0.5

• |ψ(0.75 nm)|2 = 0 (it’s a node)


0
0 0.25 0.5
·
0.75 1
x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 42 · QM principles and concepts · Wavefunctions


Probability v probability density
In practice, what we often want is not the probability density
of finding a particle exactly at a particular position.
Instead, what we often want is the probability
of finding the particle in a region of space.
For this we must add the contributions (the probability densities)
of all points within that region.

Slide 43 · QM principles and concepts · Wavefunctions


Probability v probability density
In practice, what we often want is not the probability density
of finding a particle exactly at a particular position.
Instead, what we often want is the probability
of finding the particle in a region of space.
For this we must add the contributions (the probability densities)
of all points within that region.
The mathematical technique for adding the values of a function
whose variables change continuously is integration. The integral
of a function is a continuous summation of the function values.
In the case of the probability density of slide 42, the 1-D integral
ZB ZB
PA,B = ρ(x) dx = |ψ(x)|2 dx.
A A

gives the probability of finding the particle between points A and B.


Slide 43 · QM principles and concepts · Wavefunctions
One-dimensional integration

This integral of a 1-D function also gives the “net” area under the curve
representing the function we are integrating. An example:

2 • integral (“area”) positive where


f (x) > 0, negative where f (x) < 0
1

f (x) 0

−1

−2
−1 − 2 − 1 0 1 2 1
3 3 3 3
x

Slide 44 · QM principles and concepts · Wavefunctions


One-dimensional integration

This integral of a 1-D function also gives the “net” area under the curve
representing the function we are integrating. An example:

2 • integral (“area”) positive where


f (x) > 0, negative where f (x) < 0
1 • these integrals are zero:
R1/3 R1
−1
f (x) dx = −1/3
f (x) dx
f (x) 0 R2/3
= f (x) dx = 0
−2/3
−1

−2
−1 − 2 − 1 0 1 2 1
3 3 3 3
x

Slide 44 · QM principles and concepts · Wavefunctions


One-dimensional integration

This integral of a 1-D function also gives the “net” area under the curve
representing the function we are integrating. An example:

2 • integral (“area”) positive where


f (x) > 0, negative where f (x) < 0
1 • these integrals are zero:
R1/3 R1
−1
f (x) dx = −1/3
f (x) dx
f (x) 0 R2/3
= f (x) dx = 0
−2/3
−1

−2
−1 − 2 − 1 0 1 2 1
3 3 3 3
x

Slide 44 · QM principles and concepts · Wavefunctions


One-dimensional integration

This integral of a 1-D function also gives the “net” area under the curve
representing the function we are integrating. An example:

2 • integral (“area”) positive where


f (x) > 0, negative where f (x) < 0
1 • these integrals are zero:
R1/3 R1
−1
f (x) dx = −1/3
f (x) dx
f (x) 0 R2/3
= f (x) dx = 0
−2/3
−1

−2
−1 − 2 − 1 0 1 2 1
3 3 3 3
x

Slide 44 · QM principles and concepts · Wavefunctions


One-dimensional integration

This integral of a 1-D function also gives the “net” area under the curve
representing the function we are integrating. An example:

2 • integral (“area”) positive where


f (x) > 0, negative where f (x) < 0
1 • these integrals are zero:
R1/3 R1
−1
f (x) dx = −1/3
f (x) dx
f (x) 0 R2/3
= f (x) dx = 0
−2/3
−1 • these integrals are not:
R−1/3 R1
−1
f (x) dx > 0, 1/3
f (x) dx > 0
−2 R1/3 R1
−1 − 2 − 1 0 1 2 1
3 3 3 3 f (x) dx < 0, −1 f (x) dx >0
x −1/3

Slide 44 · QM principles and concepts · Wavefunctions


One-dimensional integration

This integral of a 1-D function also gives the “net” area under the curve
representing the function we are integrating. An example:

2 • integral (“area”) positive where


f (x) > 0, negative where f (x) < 0
1 • these integrals are zero:
R1/3 R1
−1
f (x) dx = −1/3
f (x) dx
f (x) 0 R2/3
= f (x) dx = 0
−2/3
−1 • these integrals are not:
R−1/3 R1
−1
f (x) dx > 0, 1/3
f (x) dx > 0
−2 R1/3 R1
−1 − 2 − 1 0 1 2 1
3 3 3 3 f (x) dx < 0, −1 f (x) dx >0
x −1/3

Slide 44 · QM principles and concepts · Wavefunctions


One-dimensional integration

This integral of a 1-D function also gives the “net” area under the curve
representing the function we are integrating. An example:

2 • integral (“area”) positive where


f (x) > 0, negative where f (x) < 0
1 • these integrals are zero:
R1/3 R1
−1
f (x) dx = −1/3
f (x) dx
f (x) 0 R2/3
= f (x) dx = 0
−2/3
−1 • these integrals are not:
R−1/3 R1
−1
f (x) dx > 0, 1/3
f (x) dx > 0
−2 R1/3 R1
−1 − 2 − 1 0 1 2 1
3 3 3 3 f (x) dx < 0, −1 f (x) dx >0
x −1/3

Slide 44 · QM principles and concepts · Wavefunctions


One-dimensional integration

This integral of a 1-D function also gives the “net” area under the curve
representing the function we are integrating. An example:

2 • integral (“area”) positive where


f (x) > 0, negative where f (x) < 0
1 • these integrals are zero:
R1/3 R1
−1
f (x) dx = −1/3
f (x) dx
f (x) 0 R2/3
= f (x) dx = 0
−2/3
−1 • these integrals are not:
R−1/3 R1
−1
f (x) dx > 0, 1/3
f (x) dx > 0
−2 R1/3 R1
−1 − 2 − 1 0 1 2 1
3 3 3 3 f (x) dx < 0, −1 f (x) dx >0
x −1/3

Slide 44 · QM principles and concepts · Wavefunctions


Where in the 1-D box is the electron? (4)

2
Z1
1.5 • P0,1 = |ψ(x)|2 dx = 1 = 100%
|ψ4 (x)| (nm )
−1

0
• probability of finding electron
1
inside the box is 100%.
2

0.5
(this is a consequence of
0
normalization to unity)
0 0.25 0.5 0.75 1
x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 45 · QM principles and concepts · Wavefunctions


Where in the 1-D box is the electron? (5)

1.5
• P0.250,0.500 = 1/4 = 25%
|ψ4 (x)| (nm )
−1

• P0,0.125 = 1/8 = 12.5%


1 • P0.6,0.9 = 0.347 = 34.7%
2

0.5
(probability is given by the
area under the curve)
0
0 0.25 0.5 0.75 1
x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 46 · QM principles and concepts · Wavefunctions


Where in the 1-D box is the electron? (5)

1.5
• P0.250,0.500 = 1/4 = 25%
|ψ4 (x)| (nm )
−1

• P0,0.125 = 1/8 = 12.5%


1 • P0.6,0.9 = 0.347 = 34.7%
2

0.5
(probability is given by the
area under the curve)
0
0 0.25 0.5 0.75 1
x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 46 · QM principles and concepts · Wavefunctions


Where in the 1-D box is the electron? (5)

1.5
• P0.250,0.500 = 1/4 = 25%
|ψ4 (x)| (nm )
−1

• P0,0.125 = 1/8 = 12.5%


1 • P0.6,0.9 = 0.347 = 34.7%
2

0.5
(probability is given by the
area under the curve)
0
0 0.25 0.5 0.75 1
x (nm)

(box of length 1 nm, electron in n = 4 state;


see maple worksheet where is it.mws)

Slide 46 · QM principles and concepts · Wavefunctions


Summary: wavefunctions

• Wavefunctions are functions that have as variables


the coordinates of the system components.
• Time-dependent functions have an extra variable: time.
• Wavefunctions in general have complex values.
Some particular wavefunctions have real values.
• Each particular system has its own wavefunctions,
and there is one wavefunction for each system state.
• The squared modulus of the wavefunction ψ(x) gives the
probability density of finding the particle at the position x.
• The integral of the probability density ρ(x) over a region of space
gives the probability of finding the particle in the specified region.
• Probability and probability density are not the same thing.

Slide 47 · QM principles and concepts · Wavefunctions


Summary: wavefunctions

• Wavefunctions are functions that have as variables


the coordinates of the system components.
• Time-dependent functions have an extra variable: time.
• Wavefunctions in general have complex values.
Some particular wavefunctions have real values.
• Each particular system has its own wavefunctions,
and there is one wavefunction for each system state.
• The squared modulus of the wavefunction ψ(x) gives the
probability density of finding the particle at the position x.
• The integral of the probability density ρ(x) over a region of space
gives the probability of finding the particle in the specified region.
• Probability and probability density are not the same thing.

Slide 47 · QM principles and concepts · Wavefunctions


Summary: wavefunctions

• Wavefunctions are functions that have as variables


the coordinates of the system components.
• Time-dependent functions have an extra variable: time.
• Wavefunctions in general have complex values.
Some particular wavefunctions have real values.
• Each particular system has its own wavefunctions,
and there is one wavefunction for each system state.
• The squared modulus of the wavefunction ψ(x) gives the
probability density of finding the particle at the position x.
• The integral of the probability density ρ(x) over a region of space
gives the probability of finding the particle in the specified region.
• Probability and probability density are not the same thing.

Slide 47 · QM principles and concepts · Wavefunctions


Summary: wavefunctions

• Wavefunctions are functions that have as variables


the coordinates of the system components.
• Time-dependent functions have an extra variable: time.
• Wavefunctions in general have complex values.
Some particular wavefunctions have real values.
• Each particular system has its own wavefunctions,
and there is one wavefunction for each system state.
• The squared modulus of the wavefunction ψ(x) gives the
probability density of finding the particle at the position x.
• The integral of the probability density ρ(x) over a region of space
gives the probability of finding the particle in the specified region.
• Probability and probability density are not the same thing.

Slide 47 · QM principles and concepts · Wavefunctions


Summary: wavefunctions

• Wavefunctions are functions that have as variables


the coordinates of the system components.
• Time-dependent functions have an extra variable: time.
• Wavefunctions in general have complex values.
Some particular wavefunctions have real values.
• Each particular system has its own wavefunctions,
and there is one wavefunction for each system state.
• The squared modulus of the wavefunction ψ(x) gives the
probability density of finding the particle at the position x.
• The integral of the probability density ρ(x) over a region of space
gives the probability of finding the particle in the specified region.
• Probability and probability density are not the same thing.

Slide 47 · QM principles and concepts · Wavefunctions


Summary: wavefunctions

• Wavefunctions are functions that have as variables


the coordinates of the system components.
• Time-dependent functions have an extra variable: time.
• Wavefunctions in general have complex values.
Some particular wavefunctions have real values.
• Each particular system has its own wavefunctions,
and there is one wavefunction for each system state.
• The squared modulus of the wavefunction ψ(x) gives the
probability density of finding the particle at the position x.
• The integral of the probability density ρ(x) over a region of space
gives the probability of finding the particle in the specified region.
• Probability and probability density are not the same thing.

Slide 47 · QM principles and concepts · Wavefunctions


Summary: wavefunctions

• Wavefunctions are functions that have as variables


the coordinates of the system components.
• Time-dependent functions have an extra variable: time.
• Wavefunctions in general have complex values.
Some particular wavefunctions have real values.
• Each particular system has its own wavefunctions,
and there is one wavefunction for each system state.
• The squared modulus of the wavefunction ψ(x) gives the
probability density of finding the particle at the position x.
• The integral of the probability density ρ(x) over a region of space
gives the probability of finding the particle in the specified region.
• Probability and probability density are not the same thing.

Slide 47 · QM principles and concepts · Wavefunctions


Summary: wavefunctions

• Wavefunctions are functions that have as variables


the coordinates of the system components.
• Time-dependent functions have an extra variable: time.
• Wavefunctions in general have complex values.
Some particular wavefunctions have real values.
• Each particular system has its own wavefunctions,
and there is one wavefunction for each system state.
• The squared modulus of the wavefunction ψ(x) gives the
probability density of finding the particle at the position x.
• The integral of the probability density ρ(x) over a region of space
gives the probability of finding the particle in the specified region.
• Probability and probability density are not the same thing.

Now, how can we extract information from a given wavefunction?

Slide 47 · QM principles and concepts · Wavefunctions


Part I — QM principles and concepts

1 Complex numbers

2 Questions and answers

3 Wavefunctions

Extracting information 4 Observables and operators

from wavefunctions. 5 Expectation values and eigenvalues

6 TI and TD Schrödinger equations

a) Observables and operators 7 Phase and wave-particle duality

8 Overlap and probability

9 Uncertainty

10 Superposition of wavefunctions

Slide 48 · QM principles and concepts · Observables and operators


Observables

“Observable” is the QM jargon for all things we can observe


(measure in an experiment). If a given quantity can be measured,
then this quantity is said to be an observable.
Examples: position, momentum, energy.
According to QM, in order to find the values of observables
we have to ask our oracle — the wavefunction.
The language we use to ask is not plain language. It’s
mathematics. In order to extract information from the
wavefunction, we need mathematical operations.
We need to operate on the wavefunction with mathematical operators.

Slide 49 · QM principles and concepts · Observables and operators


Observables

“Observable” is the QM jargon for all things we can observe


(measure in an experiment). If a given quantity can be measured,
then this quantity is said to be an observable.
Examples: position, momentum, energy.
According to QM, in order to find the values of observables
we have to ask our oracle — the wavefunction.
The language we use to ask is not plain language. It’s
mathematics. In order to extract information from the
wavefunction, we need mathematical operations.
We need to operate on the wavefunction with mathematical operators.

Slide 49 · QM principles and concepts · Observables and operators


Observables

“Observable” is the QM jargon for all things we can observe


(measure in an experiment). If a given quantity can be measured,
then this quantity is said to be an observable.
Examples: position, momentum, energy.
According to QM, in order to find the values of observables
we have to ask our oracle — the wavefunction.
The language we use to ask is not plain language. It’s
mathematics. In order to extract information from the
wavefunction, we need mathematical operations.
We need to operate on the wavefunction with mathematical operators.

Slide 49 · QM principles and concepts · Observables and operators


Observables

“Observable” is the QM jargon for all things we can observe


(measure in an experiment). If a given quantity can be measured,
then this quantity is said to be an observable.
Examples: position, momentum, energy.
According to QM, in order to find the values of observables
we have to ask our oracle — the wavefunction.
The language we use to ask is not plain language. It’s
mathematics. In order to extract information from the
wavefunction, we need mathematical operations.
We need to operate on the wavefunction with mathematical operators.

Slide 49 · QM principles and concepts · Observables and operators


Observables

“Observable” is the QM jargon for all things we can observe


(measure in an experiment). If a given quantity can be measured,
then this quantity is said to be an observable.
Examples: position, momentum, energy.
According to QM, in order to find the values of observables
we have to ask our oracle — the wavefunction.
The language we use to ask is not plain language. It’s
mathematics. In order to extract information from the
wavefunction, we need mathematical operations.
We need to operate on the wavefunction with mathematical operators.

Slide 49 · QM principles and concepts · Observables and operators


Operators

An operator is a mathematical symbol (an instruction)


that tells you to do something to whatever follows it.
That “something” will be a mathematical operation of one kind
or another. Normally addition, multiplication, differentiation
or a combination of these.
Operators are usually represented by Roman letters with a circumflex
accent (a “hat”): things like x̂, Ô or p̂z .
In this course we’ll apply operators to functions, and to
wavefunctions in particular. We may use use maple for that.

Slide 50 · QM principles and concepts · Observables and operators


Operators

An operator is a mathematical symbol (an instruction)


that tells you to do something to whatever follows it.
That “something” will be a mathematical operation of one kind
or another. Normally addition, multiplication, differentiation
or a combination of these.
Operators are usually represented by Roman letters with a circumflex
accent (a “hat”): things like x̂, Ô or p̂z .
In this course we’ll apply operators to functions, and to
wavefunctions in particular. We may use use maple for that.

Slide 50 · QM principles and concepts · Observables and operators


Operators

An operator is a mathematical symbol (an instruction)


that tells you to do something to whatever follows it.
That “something” will be a mathematical operation of one kind
or another. Normally addition, multiplication, differentiation
or a combination of these.
Operators are usually represented by Roman letters with a circumflex
accent (a “hat”): things like x̂, Ô or p̂z .
In this course we’ll apply operators to functions, and to
wavefunctions in particular. We may use use maple for that.

Slide 50 · QM principles and concepts · Observables and operators


Operators

An operator is a mathematical symbol (an instruction)


that tells you to do something to whatever follows it.
That “something” will be a mathematical operation of one kind
or another. Normally addition, multiplication, differentiation
or a combination of these.
Operators are usually represented by Roman letters with a circumflex
accent (a “hat”): things like x̂, Ô or p̂z .
In this course we’ll apply operators to functions, and to
wavefunctions in particular. We may use use maple for that.

Slide 50 · QM principles and concepts · Observables and operators


Observables and operators

Observables and operators are directly related.


Each and every observable is associated with an operator
having some particular mathematical properties.
In order to find the result of a given experiment, you can either do the
experiment, or else you can get the answer from the wavefunction.
In the latter case you’ll have to:
1. Find the expression for operator associated with the observable to
be measured in your experiment.
2. Act on the system wavefunction with this operator.
The result of your experiment will depend on what you get after
applying your operator to the system’s wavefunction.

Slide 51 · QM principles and concepts · Observables and operators


Observables and operators

Observables and operators are directly related.


Each and every observable is associated with an operator
having some particular mathematical properties.
In order to find the result of a given experiment, you can either do the
experiment, or else you can get the answer from the wavefunction.
In the latter case you’ll have to:
1. Find the expression for operator associated with the observable to
be measured in your experiment.
2. Act on the system wavefunction with this operator.
The result of your experiment will depend on what you get after
applying your operator to the system’s wavefunction.

Slide 51 · QM principles and concepts · Observables and operators


Position and momentum operators

Two of the most important operators in QM are the position and the
momentum operators:

position operator: x̂ = x

momentum operator: p̂x = −i ħ
∂x
(and similarly for ŷ, p̂y , ẑ, p̂z ).

Besides being associated with important observables,


these operators help us find expressions for other operators.

Slide 52 · QM principles and concepts · Observables and operators


Finding expressions for operators (1)

It’s quite easy to find classical (ie, Newtonian) expressions for


observables in terms of positions and momenta.
Finding the corresponding quantum operators is quite easy too.
All you need to do is to replace the classical positions
and momenta by the corresponding quantum operators.
Let us consider the kinetic energy (K ). The classical formula for the
kinetic energy of an object of mass m moving in 3-D space is
2
p2x py p2z
K = Kx + Ky + Kz = + + .
2m 2m 2m

Using the expressions p̂2x = (p̂x )(p̂x ) and p̂x = −i ħ∂/∂x we get...

Slide 53 · QM principles and concepts · Observables and operators


Finding expressions for operators (1)

It’s quite easy to find classical (ie, Newtonian) expressions for


observables in terms of positions and momenta.
Finding the corresponding quantum operators is quite easy too.
All you need to do is to replace the classical positions
and momenta by the corresponding quantum operators.
Let us consider the kinetic energy (K ). The classical formula for the
kinetic energy of an object of mass m moving in 3-D space is
2
p2x py p2z
K = Kx + Ky + Kz = + + .
2m 2m 2m

Using the expressions p̂2x = (p̂x )(p̂x ) and p̂x = −i ħ∂/∂x we get...

Slide 53 · QM principles and concepts · Observables and operators


Finding expressions for operators (2)

...we get
 ∂  ∂ 
p̂2x = −i ħ −i ħ
∂x ∂x

Slide 54 · QM principles and concepts · Observables and operators


Finding expressions for operators (2)

...we get
 ∂  ∂   ∂  ∂ 
p̂2x = −i ħ −i ħ = (−i )2 ħ2
∂x ∂x ∂x ∂x

Slide 54 · QM principles and concepts · Observables and operators


Finding expressions for operators (2)

...we get
2
 ∂  ∂   ∂  ∂  ∂
p̂2x = −i ħ −i ħ = (−i )2 ħ2 = − ħ2 ,
∂x ∂x ∂x ∂x ∂x 2

Slide 54 · QM principles and concepts · Observables and operators


Finding expressions for operators (2)

...we get
2
 ∂  ∂   ∂  ∂  ∂
p̂2x = −i ħ −i ħ = (−i )2 ħ2 = − ħ2 ,
∂x ∂x ∂x ∂x ∂x 2

and similarly for p̂2y and p̂2z .

Slide 54 · QM principles and concepts · Observables and operators


Finding expressions for operators (2)

...we get
2
 ∂  ∂   ∂  ∂  ∂
p̂2x = −i ħ −i ħ = (−i )2 ħ2 = − ħ2 ,
∂x ∂x ∂x ∂x ∂x 2

and similarly for p̂2y and p̂2z .

Now, back to the classical formula on the previous slide. Replacing the
classical variables (K , px , py and pz ) by the corresponding operators
(K̂ , p̂x , p̂y and p̂z ) we finally get

1
K̂ = (p̂2 + p̂2y + p̂2z )
2m x

Slide 54 · QM principles and concepts · Observables and operators


Finding expressions for operators (2)

...we get
2
 ∂  ∂   ∂  ∂  ∂
p̂2x = −i ħ −i ħ = (−i )2 ħ2 = − ħ2 ,
∂x ∂x ∂x ∂x ∂x 2

and similarly for p̂2y and p̂2z .

Now, back to the classical formula on the previous slide. Replacing the
classical variables (K , px , py and pz ) by the corresponding operators
(K̂ , p̂x , p̂y and p̂z ) we finally get
2 2 2 2
1 ħ  ∂ ∂ ∂ 
K̂ = (p̂2x + p̂2y + p̂2z ) = − + + .
2m 2m ∂x 2 ∂y 2 ∂z 2

Slide 54 · QM principles and concepts · Observables and operators


Finding expressions for operators (2)

...we get
2
 ∂  ∂   ∂  ∂  ∂
p̂2x = −i ħ −i ħ = (−i )2 ħ2 = − ħ2 ,
∂x ∂x ∂x ∂x ∂x 2

and similarly for p̂2y and p̂2z .

Now, back to the classical formula on the previous slide. Replacing the
classical variables (K , px , py and pz ) by the corresponding operators
(K̂ , p̂x , p̂y and p̂z ) we finally get
2 2 2 2
1 ħ  ∂ ∂ ∂ 
K̂ = (p̂2x + p̂2y + p̂2z ) = − + + .
2m 2m ∂x 2 ∂y 2 ∂z 2

The nice thing is, this procedure works for all observables.

Slide 54 · QM principles and concepts · Observables and operators


Applying operators

Let us apply operators to the wavefunction of a particle in a 1-D box.


Application of the position operator (x̂ = x, “multiply by x”)

x̂ψn (x) = xψn (x)

Slide 55 · QM principles and concepts · Observables and operators


Applying operators

Let us apply operators to the wavefunction of a particle in a 1-D box.


Application of the position operator (x̂ = x, “multiply by x”)
"r #
2  nπx 
x̂ψn (x) = xψn (x) = x sin
L L

Slide 55 · QM principles and concepts · Observables and operators


Applying operators

Let us apply operators to the wavefunction of a particle in a 1-D box.


Application of the position operator (x̂ = x, “multiply by x”)
"r # r
2  nπx  2  nπx 
x̂ψn (x) = xψn (x) = x sin = x sin .
L L L L

Slide 55 · QM principles and concepts · Observables and operators


Applying operators

Let us apply operators to the wavefunction of a particle in a 1-D box.


Application of the position operator (x̂ = x, “multiply by x”)
"r # r
2  nπx  2  nπx 
x̂ψn (x) = xψn (x) = x sin = x sin .
L L L L

Application of D̂x = d /dx (“differentiate with regard to x”)

d ψn (x)
D̂x ψn (x) =
dx

Slide 55 · QM principles and concepts · Observables and operators


Applying operators

Let us apply operators to the wavefunction of a particle in a 1-D box.


Application of the position operator (x̂ = x, “multiply by x”)
"r # r
2  nπx  2  nπx 
x̂ψn (x) = xψn (x) = x sin = x sin .
L L L L

Application of D̂x = d /dx (“differentiate with regard to x”)

d ψn (x) d (2/L)1/2 sin(nπx/L)


D̂x ψn (x) = =
dx dx

Slide 55 · QM principles and concepts · Observables and operators


Applying operators

Let us apply operators to the wavefunction of a particle in a 1-D box.


Application of the position operator (x̂ = x, “multiply by x”)
"r # r
2  nπx  2  nπx 
x̂ψn (x) = xψn (x) = x sin = x sin .
L L L L

Application of D̂x = d /dx (“differentiate with regard to x”)

d ψn (x) d (2/L)1/2 sin(nπx/L)


r
2 d sin(nπx/L)
D̂x ψn (x) = = =
dx dx L dx

Slide 55 · QM principles and concepts · Observables and operators


Applying operators

Let us apply operators to the wavefunction of a particle in a 1-D box.


Application of the position operator (x̂ = x, “multiply by x”)
"r # r
2  nπx  2  nπx 
x̂ψn (x) = xψn (x) = x sin = x sin .
L L L L

Application of D̂x = d /dx (“differentiate with regard to x”)

d ψn (x) d (2/L)1/2 sin(nπx/L)


r
2 d sin(nπx/L)
D̂x ψn (x) = = =
dx dx L dx
r 
2 nπ  nπx 
= cos
L L L

Slide 55 · QM principles and concepts · Observables and operators


Applying operators

Let us apply operators to the wavefunction of a particle in a 1-D box.


Application of the position operator (x̂ = x, “multiply by x”)
"r # r
2  nπx  2  nπx 
x̂ψn (x) = xψn (x) = x sin = x sin .
L L L L

Application of D̂x = d /dx (“differentiate with regard to x”)

d ψn (x) d (2/L)1/2 sin(nπx/L)


r
2 d sin(nπx/L)
D̂x ψn (x) = = =
dx dx L dx
1/2
r  
2 nπ  nπx  2 nπ  nπx 
= cos = cos .
L L L L3/2 L

Slide 55 · QM principles and concepts · Observables and operators


Applying operators:
numerical and graphical examples

At this point, you should take a look at the maple worksheet


Applying operators.mws. You’ll find the following:
• examples of how you can use maple
to define wavefunctions and operators
• examples of how you can use maple
to apply operators to wavefunctions;
• graphical examples of how the application of operators
changes or does not change wavefunctions

Slide 56 · QM principles and concepts · Observables and operators


Kinetic and potential energy operators

For our purposes, things acquire energy in one of two ways:


• through motion, they acquire kinetic energy.
The faster they move, the larger is K , their kinetic energy.
• through interaction with other things, they acquire potential energy.
2
The stronger the interaction, the larger is V , the potential energy.
K and V are both observable. The corresponding operators are the
kinetic energy operator, K̂ , and the potential energy operator, V̂ .

2
The potential energy can be positive or negative. People often use the term
“large” meaning “large and positive” or “large and negative.”
Slide 57 · QM principles and concepts · Observables and operators
Kinetic and potential energy operators

For our purposes, things acquire energy in one of two ways:


• through motion, they acquire kinetic energy.
The faster they move, the larger is K , their kinetic energy.
• through interaction with other things, they acquire potential energy.
2
The stronger the interaction, the larger is V , the potential energy.
K and V are both observable. The corresponding operators are the
kinetic energy operator, K̂ , and the potential energy operator, V̂ .

We have already obtained an expression for the kinetic energy


operator of a particle moving in 3-D space — see slides 53–54.

As for the potential energy operator...

2
The potential energy can be positive or negative. People often use the term
“large” meaning “large and positive” or “large and negative.”
Slide 57 · QM principles and concepts · Observables and operators
Potential energy operator

The expression for the potential energy operator changes


from system to system. There is no single formula for it.
Reason: the way things interact depends on what things
we are considering.
For instance, the way electrons interact with atomic nuclei
differs from the way neutral molecules interact with each other.

Slide 58 · QM principles and concepts · Observables and operators


Potential energy operator

The expression for the potential energy operator changes


from system to system. There is no single formula for it.
Reason: the way things interact depends on what things
we are considering.
For instance, the way electrons interact with atomic nuclei
differs from the way neutral molecules interact with each other.
But we know this: because V only depends on the positions
of the system components, V̂ is a function of those positions.
We can therefore write the potential operator in the form V̂ = V (x, y, z).
The action of V̂ on the wavefunction ψ is simply this:
multiply ψ by the function V (x, y, z).

Slide 58 · QM principles and concepts · Observables and operators


Total energy operator: the Hamiltonian
As mentioned, things can acquire energy through motion
(kinetic energy, K ) or interaction (potential energy, V ).

The total energy of a system, E , is therefore given by


the sum of the potential and kinetic energies: E = K + V .

The total energy is an observable. The corresponding operator


(the total energy operator) is so important that it has a special name:
it’s called the Hamiltonian.

The Hamiltonian is given by the sum of the kinetic


and potential energy operators.

Hamiltonian operator: Ĥ = K̂ + V̂

Slide 59 · QM principles and concepts · Observables and operators


Summary: observables and operators

• A quantity is “observable” if it can be measured in an experiment.


• An “operator” is a mathematical symbol that tells you to carry out
a set of mathematical operations on whatever follows it.
• Every observable is associated with an operator.
• The formula for a quantum operator is obtained by taking its
classical counterpart and replacing the classical positions and
momenta by the corresponding quantum operators.
• Important operators we’ve seen: position (x̂), momentum (p̂),
kinetic energy (K̂ ), potential energy (V̂ ), Hamiltonian (Ĥ).
• The operator associated with the total energy is the Hamiltonian.
• To extract information about a system from its wavefunction, we
need to operate on the wavefunction with mathematical operators.

Slide 60 · QM principles and concepts · Observables and operators


Summary: observables and operators

• A quantity is “observable” if it can be measured in an experiment.


• An “operator” is a mathematical symbol that tells you to carry out
a set of mathematical operations on whatever follows it.
• Every observable is associated with an operator.
• The formula for a quantum operator is obtained by taking its
classical counterpart and replacing the classical positions and
momenta by the corresponding quantum operators.
• Important operators we’ve seen: position (x̂), momentum (p̂),
kinetic energy (K̂ ), potential energy (V̂ ), Hamiltonian (Ĥ).
• The operator associated with the total energy is the Hamiltonian.
• To extract information about a system from its wavefunction, we
need to operate on the wavefunction with mathematical operators.

Slide 60 · QM principles and concepts · Observables and operators


Summary: observables and operators

• A quantity is “observable” if it can be measured in an experiment.


• An “operator” is a mathematical symbol that tells you to carry out
a set of mathematical operations on whatever follows it.
• Every observable is associated with an operator.
• The formula for a quantum operator is obtained by taking its
classical counterpart and replacing the classical positions and
momenta by the corresponding quantum operators.
• Important operators we’ve seen: position (x̂), momentum (p̂),
kinetic energy (K̂ ), potential energy (V̂ ), Hamiltonian (Ĥ).
• The operator associated with the total energy is the Hamiltonian.
• To extract information about a system from its wavefunction, we
need to operate on the wavefunction with mathematical operators.

Slide 60 · QM principles and concepts · Observables and operators


Summary: observables and operators

• A quantity is “observable” if it can be measured in an experiment.


• An “operator” is a mathematical symbol that tells you to carry out
a set of mathematical operations on whatever follows it.
• Every observable is associated with an operator.
• The formula for a quantum operator is obtained by taking its
classical counterpart and replacing the classical positions and
momenta by the corresponding quantum operators.
• Important operators we’ve seen: position (x̂), momentum (p̂),
kinetic energy (K̂ ), potential energy (V̂ ), Hamiltonian (Ĥ).
• The operator associated with the total energy is the Hamiltonian.
• To extract information about a system from its wavefunction, we
need to operate on the wavefunction with mathematical operators.

Slide 60 · QM principles and concepts · Observables and operators


Summary: observables and operators

• A quantity is “observable” if it can be measured in an experiment.


• An “operator” is a mathematical symbol that tells you to carry out
a set of mathematical operations on whatever follows it.
• Every observable is associated with an operator.
• The formula for a quantum operator is obtained by taking its
classical counterpart and replacing the classical positions and
momenta by the corresponding quantum operators.
• Important operators we’ve seen: position (x̂), momentum (p̂),
kinetic energy (K̂ ), potential energy (V̂ ), Hamiltonian (Ĥ).
• The operator associated with the total energy is the Hamiltonian.
• To extract information about a system from its wavefunction, we
need to operate on the wavefunction with mathematical operators.

Slide 60 · QM principles and concepts · Observables and operators


Summary: observables and operators

• A quantity is “observable” if it can be measured in an experiment.


• An “operator” is a mathematical symbol that tells you to carry out
a set of mathematical operations on whatever follows it.
• Every observable is associated with an operator.
• The formula for a quantum operator is obtained by taking its
classical counterpart and replacing the classical positions and
momenta by the corresponding quantum operators.
• Important operators we’ve seen: position (x̂), momentum (p̂),
kinetic energy (K̂ ), potential energy (V̂ ), Hamiltonian (Ĥ).
• The operator associated with the total energy is the Hamiltonian.
• To extract information about a system from its wavefunction, we
need to operate on the wavefunction with mathematical operators.

Slide 60 · QM principles and concepts · Observables and operators


Summary: observables and operators

• A quantity is “observable” if it can be measured in an experiment.


• An “operator” is a mathematical symbol that tells you to carry out
a set of mathematical operations on whatever follows it.
• Every observable is associated with an operator.
• The formula for a quantum operator is obtained by taking its
classical counterpart and replacing the classical positions and
momenta by the corresponding quantum operators.
• Important operators we’ve seen: position (x̂), momentum (p̂),
kinetic energy (K̂ ), potential energy (V̂ ), Hamiltonian (Ĥ).
• The operator associated with the total energy is the Hamiltonian.
• To extract information about a system from its wavefunction, we
need to operate on the wavefunction with mathematical operators.

Slide 60 · QM principles and concepts · Observables and operators


Summary: observables and operators

• A quantity is “observable” if it can be measured in an experiment.


• An “operator” is a mathematical symbol that tells you to carry out
a set of mathematical operations on whatever follows it.
• Every observable is associated with an operator.
• The formula for a quantum operator is obtained by taking its
classical counterpart and replacing the classical positions and
momenta by the corresponding quantum operators.
• Important operators we’ve seen: position (x̂), momentum (p̂),
kinetic energy (K̂ ), potential energy (V̂ ), Hamiltonian (Ĥ).
• The operator associated with the total energy is the Hamiltonian.
• To extract information about a system from its wavefunction, we
need to operate on the wavefunction with mathematical operators.
Now, how exactly does this “information extraction” work?

Slide 60 · QM principles and concepts · Observables and operators


Part I — QM principles and concepts

1 Complex numbers

2 Questions and answers

3 Wavefunctions
Extracting information 4 Observables and operators
from wavefunctions. 5 Expectation values and eigenvalues

6 TI and TD Schrödinger equations


b) Expectation values
7 Phase and wave-particle duality
and eigenvalues
8 Overlap and probability

9 Uncertainty

10 Superposition of wavefunctions

Slide 61 · QM principles and concepts · Expectation values and eigenvalues


The outcome of identical experiments
If a perfect experiment is repeated in exactly the same way
and under exactly the same conditions,
then the experimental result does not change, right?

This is not because of experimental limitations (we are talking about


perfect experiments). This happens because of restrictions imposed
by nature itself.
Experiments in general do not have a single, “right” result.
In other words: in general, it’s impossible to predict the exact result
of a single, individual experiment.
We’ll see when and why this general rule is broken and exact
prediction of the result of a single experiment becomes possible.

Slide 62 · QM principles and concepts · Expectation values and eigenvalues


The outcome of identical experiments
If a perfect experiment is repeated in exactly the same way
and under exactly the same conditions,
then the experimental result does not change, right?
Wrong. It’s only in some special cases that identical experiments
lead to identical results. In general they don’t.
This is not because of experimental limitations (we are talking about
perfect experiments). This happens because of restrictions imposed
by nature itself.
Experiments in general do not have a single, “right” result.
In other words: in general, it’s impossible to predict the exact result
of a single, individual experiment.
We’ll see when and why this general rule is broken and exact
prediction of the result of a single experiment becomes possible.

Slide 62 · QM principles and concepts · Expectation values and eigenvalues


The outcome of identical experiments
If a perfect experiment is repeated in exactly the same way
and under exactly the same conditions,
then the experimental result does not change, right?
Wrong. It’s only in some special cases that identical experiments
lead to identical results. In general they don’t.
This is not because of experimental limitations (we are talking about
perfect experiments). This happens because of restrictions imposed
by nature itself.
Experiments in general do not have a single, “right” result.
In other words: in general, it’s impossible to predict the exact result
of a single, individual experiment.
We’ll see when and why this general rule is broken and exact
prediction of the result of a single experiment becomes possible.

Slide 62 · QM principles and concepts · Expectation values and eigenvalues


The outcome of identical experiments
If a perfect experiment is repeated in exactly the same way
and under exactly the same conditions,
then the experimental result does not change, right?
Wrong. It’s only in some special cases that identical experiments
lead to identical results. In general they don’t.
This is not because of experimental limitations (we are talking about
perfect experiments). This happens because of restrictions imposed
by nature itself.
Experiments in general do not have a single, “right” result.
In other words: in general, it’s impossible to predict the exact result
of a single, individual experiment.
We’ll see when and why this general rule is broken and exact
prediction of the result of a single experiment becomes possible.

Slide 62 · QM principles and concepts · Expectation values and eigenvalues


The truth about experimental results
The truth about experimental results and their prediction:
• In general, the exact result of an individual experiment cannot be
predicted. Identical experiments may lead to different results.
• We can, however, predict what experimental results are possible
and the probabilities of obtaining each of the possible results.
• We can also predict the average result
of a large number of identical experiments.

Slide 63 · QM principles and concepts · Expectation values and eigenvalues


The truth about experimental results
The truth about experimental results and their prediction:
• In general, the exact result of an individual experiment cannot be
predicted. Identical experiments may lead to different results.
• We can, however, predict what experimental results are possible
and the probabilities of obtaining each of the possible results.
• We can also predict the average result
of a large number of identical experiments.

Slide 63 · QM principles and concepts · Expectation values and eigenvalues


The truth about experimental results
The truth about experimental results and their prediction:
• In general, the exact result of an individual experiment cannot be
predicted. Identical experiments may lead to different results.
• We can, however, predict what experimental results are possible
and the probabilities of obtaining each of the possible results.
• We can also predict the average result
of a large number of identical experiments.

Slide 63 · QM principles and concepts · Expectation values and eigenvalues


The truth about experimental results
The truth about experimental results and their prediction:
• In general, the exact result of an individual experiment cannot be
predicted. Identical experiments may lead to different results.
• We can, however, predict what experimental results are possible
and the probabilities of obtaining each of the possible results.
• We can also predict the average result
of a large number of identical experiments.
Note:
• The term “in general” is important — there are exceptions.
• While result of single experiment in general isn’t exactly
predictable, it isn’t totally unpredictable either.
• While result of single experiment in general isn’t exactly
predictable, average of many experiments is.

Slide 63 · QM principles and concepts · Expectation values and eigenvalues


The truth about experimental results
The truth about experimental results and their prediction:
• In general, the exact result of an individual experiment cannot be
predicted. Identical experiments may lead to different results.
• We can, however, predict what experimental results are possible
and the probabilities of obtaining each of the possible results.
• We can also predict the average result
of a large number of identical experiments.
Note:
• The term “in general” is important — there are exceptions.
• While result of single experiment in general isn’t exactly
predictable, it isn’t totally unpredictable either.
• While result of single experiment in general isn’t exactly
predictable, average of many experiments is.

Slide 63 · QM principles and concepts · Expectation values and eigenvalues


The truth about experimental results
The truth about experimental results and their prediction:
• In general, the exact result of an individual experiment cannot be
predicted. Identical experiments may lead to different results.
• We can, however, predict what experimental results are possible
and the probabilities of obtaining each of the possible results.
• We can also predict the average result
of a large number of identical experiments.
Note:
• The term “in general” is important — there are exceptions.
• While result of single experiment in general isn’t exactly
predictable, it isn’t totally unpredictable either.
• While result of single experiment in general isn’t exactly
predictable, average of many experiments is.

Slide 63 · QM principles and concepts · Expectation values and eigenvalues


Expectation values

The average result of many identical measurements of the observable


A associated with the operator  is given by hAi, the expectation value
of that observable.

The expectation value is related to the wavefunction Ψ and to the


operator  by an integral “over all space” — i.e., an integral covering all
values of all coordinates upon which the wavefunction depends:
Z
expectation value of A: hAi = [Ψ(ω)]∗ Â Ψ(ω) d ω,
all
space

where ω collectively represents all the coordinates of all the particles


in the system. In general, the integral is multidimensional.

Slide 64 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 1 (an analogy)

A B C D E F
(“system state”) (white) (light grey) (grey) (dark grey) (black)

• The “system” is a disk.


• A represents the “state” of the system.
• The “observable” is colour.
• In this state, the system has
two characteristic colours: black and white.
• The two characteristic colours are present in equal amounts.

Slide 65 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 1 (an analogy)

A B C D E F
(“system state”) (white) (light grey) (grey) (dark grey) (black)

• The “system” is a disk.


• A represents the “state” of the system.
• The “observable” is colour.
• In this state, the system has
two characteristic colours: black and white.
• The two characteristic colours are present in equal amounts.

Slide 65 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 1 (an analogy)

A B C D E F
(“system state”) (white) (light grey) (grey) (dark grey) (black)

• The “system” is a disk.


• A represents the “state” of the system.
• The “observable” is colour.
• In this state, the system has
two characteristic colours: black and white.
• The two characteristic colours are present in equal amounts.

Slide 65 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 1 (an analogy)

A B C D E F
(“system state”) (white) (light grey) (grey) (dark grey) (black)

• The “system” is a disk.


• A represents the “state” of the system.
• The “observable” is colour.
• In this state, the system has
two characteristic colours: black and white.
• The two characteristic colours are present in equal amounts.

Slide 65 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 1 (an analogy)

A B C D E F
(“system state”) (white) (light grey) (grey) (dark grey) (black)

• The “system” is a disk.


• A represents the “state” of the system.
• The “observable” is colour.
• In this state, the system has
two characteristic colours: black and white.
• The two characteristic colours are present in equal amounts.

Slide 65 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 1 (an analogy, cont.)

A B C D E F
(“system state”) (white) (light grey) (grey) (dark grey) (black)

• If we repeatedly place the system in state A and then measure its


colour, we’ll find that it’s 50% black and 50% white.
• The average colour is grey = 0.5 black + 0.5 white.
• The expectation value of the colour observable is hcolouri = grey.
• On average, the state of the system is as shown in D.
• On average, the state of the system is not as shown in B, C, E or F.

Slide 66 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 1 (an analogy, cont.)

A B C D E F
(“system state”) (white) (light grey) (grey) (dark grey) (black)

• If we repeatedly place the system in state A and then measure its


colour, we’ll find that it’s 50% black and 50% white.
• The average colour is grey = 0.5 black + 0.5 white.
• The expectation value of the colour observable is hcolouri = grey.
• On average, the state of the system is as shown in D.
• On average, the state of the system is not as shown in B, C, E or F.

Slide 66 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 1 (an analogy, cont.)

A B C D E F
(“system state”) (white) (light grey) (grey) (dark grey) (black)

• If we repeatedly place the system in state A and then measure its


colour, we’ll find that it’s 50% black and 50% white.
• The average colour is grey = 0.5 black + 0.5 white.
• The expectation value of the colour observable is hcolouri = grey.
• On average, the state of the system is as shown in D.
• On average, the state of the system is not as shown in B, C, E or F.

Slide 66 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 1 (an analogy, cont.)

A B C D E F
(“system state”) (white) (light grey) (grey) (dark grey) (black)

• If we repeatedly place the system in state A and then measure its


colour, we’ll find that it’s 50% black and 50% white.
• The average colour is grey = 0.5 black + 0.5 white.
• The expectation value of the colour observable is hcolouri = grey.
• On average, the state of the system is as shown in D.
• On average, the state of the system is not as shown in B, C, E or F.

Slide 66 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 1 (an analogy, cont.)

A
(“system state”)
×B
(white)
× C
(light grey)
D
(grey)
× E
(dark grey)
× F
(black)

• If we repeatedly place the system in state A and then measure its


colour, we’ll find that it’s 50% black and 50% white.
• The average colour is grey = 0.5 black + 0.5 white.
• The expectation value of the colour observable is hcolouri = grey.
• On average, the state of the system is as shown in D.
• On average, the state of the system is not as shown in B, C, E or F.

Slide 66 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 2

We have an electron in a 1-D box stretching from x = 0 to x = 10 nm.


We want to calculate the average position of the electron.
• The observable is the position, x.
• The corresponding operator is the position operator,
x̂ = x (“multiply by x”).
• The average position of the electron is given
by the expectation value of the position:
Z 10 nm
∗
hxi =

ψn (x) xψn (x) dx.
0

• We can use the maple worksheet expect-1Dbox.mws to


calculate hxi for different values of the quantum number n.
• We have hxi = 5 nm regardless of the value of n. How come?

Slide 67 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 2

We have an electron in a 1-D box stretching from x = 0 to x = 10 nm.


We want to calculate the average position of the electron.
• The observable is the position, x.
• The corresponding operator is the position operator,
x̂ = x (“multiply by x”).
• The average position of the electron is given
by the expectation value of the position:
Z 10 nm
∗
hxi =

ψn (x) xψn (x) dx.
0

• We can use the maple worksheet expect-1Dbox.mws to


calculate hxi for different values of the quantum number n.
• We have hxi = 5 nm regardless of the value of n. How come?

Slide 67 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 2

We have an electron in a 1-D box stretching from x = 0 to x = 10 nm.


We want to calculate the average position of the electron.
• The observable is the position, x.
• The corresponding operator is the position operator,
x̂ = x (“multiply by x”).
• The average position of the electron is given
by the expectation value of the position:
Z 10 nm
∗
hxi =

ψn (x) xψn (x) dx.
0

• We can use the maple worksheet expect-1Dbox.mws to


calculate hxi for different values of the quantum number n.
• We have hxi = 5 nm regardless of the value of n. How come?

Slide 67 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 2

We have an electron in a 1-D box stretching from x = 0 to x = 10 nm.


We want to calculate the average position of the electron.
• The observable is the position, x.
• The corresponding operator is the position operator,
x̂ = x (“multiply by x”).
• The average position of the electron is given
by the expectation value of the position:
Z 10 nm
∗
hxi =

ψn (x) xψn (x) dx.
0

• We can use the maple worksheet expect-1Dbox.mws to


calculate hxi for different values of the quantum number n.
• We have hxi = 5 nm regardless of the value of n. How come?

Slide 67 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 2

We have an electron in a 1-D box stretching from x = 0 to x = 10 nm.


We want to calculate the average position of the electron.
• The observable is the position, x.
• The corresponding operator is the position operator,
x̂ = x (“multiply by x”).
• The average position of the electron is given
by the expectation value of the position:
Z 10 nm
∗
hxi =

ψn (x) xψn (x) dx.
0

• We can use the maple worksheet expect-1Dbox.mws to


calculate hxi for different values of the quantum number n.
• We have hxi = 5 nm regardless of the value of n. How come?

Slide 67 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: example 2

We have an electron in a 1-D box stretching from x = 0 to x = 10 nm.


We want to calculate the average position of the electron.
• The observable is the position, x.
• The corresponding operator is the position operator,
x̂ = x (“multiply by x”).
• The average position of the electron is given
by the expectation value of the position:
Z 10 nm
∗
hxi =

ψn (x) xψn (x) dx.
0

• We can use the maple worksheet expect-1Dbox.mws to


calculate hxi for different values of the quantum number n.
• We have hxi = 5 nm regardless of the value of n. How come?

Slide 67 · QM principles and concepts · Expectation values and eigenvalues


Expectation values: more examples and exercises

The maple worksheet expect-1Dbox.mws includes calculations of


other expectation values for the system we have just considered:
2 2
hx i, hpx i, hpx i and hKx i.

• The value of hx 2 i does depend on the system state,


even though the value of hxi does not. How come?
• The values that maple obtains for hpx i are wrong.
What should the correct results be?
• Maple obtains numerical values for hp2x i and hKx i. Can you write
general mathematical expressions for hp2x i and hKx i, valid for
whatever values of the system and state parameters (L, m and n)?
• Maple does not calculate expectation values of the total energy.
Can you obtain these yourself?

Slide 68 · QM principles and concepts · Expectation values and eigenvalues


Possible experimental results

Slide 63 contained the following statement:

We can predict what experimental results are possible


and the probabilities of obtaining each of the possible results.

These predictions require understanding of two very important


mathematical concepts: eigenvalues and eigenfunctions.

Slide 69 · QM principles and concepts · Expectation values and eigenvalues


Possible experimental results

Slide 63 contained the following statement:

We can predict what experimental results are possible


and the probabilities of obtaining each of the possible results.

These predictions require understanding of two very important


mathematical concepts: eigenvalues and eigenfunctions.

 
Earlier in this course I have emphasised the importance of the vocabulary.
 
 
 Words prefixed with “eigen” are very important in QM. 
 
 
You must understand the meaning of the prefix “eigen”
 

and also the meaning of all the words prefixed with it.
 

 
In past years, this was probably the part of the course that students
 
found most difficult to understand. Can we change that?

Slide 69 · QM principles and concepts · Expectation values and eigenvalues


Eigenvalues and eigenfunctions (1)
In general, application of an operator to a function leads some other,
quite different function. Examples:

d
 = and f (x) = sin(2x) ⇒ Âf (x) = 2 cos(2x)
dx

Slide 70 · QM principles and concepts · Expectation values and eigenvalues


Eigenvalues and eigenfunctions (1)
In general, application of an operator to a function leads some other,
quite different function. Examples:

d
 = and f (x) = sin(2x) ⇒ Âf (x) = 2 cos(2x)
dx
d2
B̂ = and g(x) = 2x 3 ⇒ B̂g(x) = 12x
dx 2

Slide 70 · QM principles and concepts · Expectation values and eigenvalues


Eigenvalues and eigenfunctions (1)

In some special cases, however, application of an operator to a


function does not lead to a very different function. Instead, it leads to
the same function multiplied by a real number. Examples:

d 3 3 3x 3
 = and f (x) = e 4 x ⇒ Âf (x) = e 4 = f (x)
dx 4 4

Slide 70 · QM principles and concepts · Expectation values and eigenvalues


Eigenvalues and eigenfunctions (1)

In some special cases, however, application of an operator to a


function does not lead to a very different function. Instead, it leads to
the same function multiplied by a real number. Examples:

d 3 3 3x 3
 = and f (x) = e 4 x ⇒ Âf (x) = e 4 = f (x)
dx 4 4
2
d
B̂ = and g(x) = cos x ⇒ B̂g(x) = − cos x = −g(x)
dx 2

Slide 70 · QM principles and concepts · Expectation values and eigenvalues


Eigenvalues and eigenfunctions (2)

If this happens — if the result of applying the operator to a function is a


nonzero real number times the same function — we have an

eigenvalue equation: Âf (x) = af (x)

If things are like that, then f (x) is an eigenfunction of the operator Â,
and the corresponding eigenvalue is a.

The prefix “eigen” is German. It means “characteristic.”

Eigenvalues and eigenfunctions are numbers and functions that are


characteristic of the operator involved in the eigenvalue equation.

Slide 71 · QM principles and concepts · Expectation values and eigenvalues


Eigenvalues and eigenfunctions (3)

Back to the examples on slide 70:


d
• sin(2x) = 2 cos(2x) the function sin(2x) is not an
dx
eigenfunction of the operator d /dx.
d2 3
• 2x 3 = 12x the function 2x is not an eigenfunction
dx 2 2 2
of the operator d /dx .
d 3x 3 3x
• e4 = e4 the function exp(3x/4) is an eigenfunction
dx 4
of the operator d /dx.
The corresponding eigenvalue is 3/4.
d2
• cos x = − cos x the function cos x) is an eigenfunction
dx 2 2
of the operator d /dx .
2

The corresponding eigenvalue is −1.

Slide 72 · QM principles and concepts · Expectation values and eigenvalues


Eigenvalues and eigenfunctions (3)

Back to the examples on slide 70:


d
• sin(2x) = 2 cos(2x) the function sin(2x) is not an
dx
eigenfunction of the operator d /dx.
d2 3
• 2x 3 = 12x the function 2x is not an eigenfunction
dx 2 2 2
of the operator d /dx .
d 3x 3 3x
• e4 = e4 the function exp(3x/4) is an eigenfunction
dx 4
of the operator d /dx.
The corresponding eigenvalue is 3/4.
d2
• cos x = − cos x the function cos x) is an eigenfunction
dx 2 2
of the operator d /dx .
2

The corresponding eigenvalue is −1.

Slide 72 · QM principles and concepts · Expectation values and eigenvalues


Eigenvalues and eigenfunctions (3)

Back to the examples on slide 70:


d
• sin(2x) = 2 cos(2x) the function sin(2x) is not an
dx
eigenfunction of the operator d /dx.
d2 3
• 2x 3 = 12x the function 2x is not an eigenfunction
dx 2 2 2
of the operator d /dx .
d 3x 3 3x
• e4 = e4 the function exp(3x/4) is an eigenfunction
dx 4
of the operator d /dx.
The corresponding eigenvalue is 3/4.
d2
• cos x = − cos x the function cos x) is an eigenfunction
dx 2 2
of the operator d /dx .
2

The corresponding eigenvalue is −1.

Slide 72 · QM principles and concepts · Expectation values and eigenvalues


Eigenvalues and eigenfunctions (3)

Back to the examples on slide 70:


d
• sin(2x) = 2 cos(2x) the function sin(2x) is not an
dx
eigenfunction of the operator d /dx.
d2 3
• 2x 3 = 12x the function 2x is not an eigenfunction
dx 2 2 2
of the operator d /dx .
d 3x 3 3x
• e4 = e4 the function exp(3x/4) is an eigenfunction
dx 4
of the operator d /dx.
The corresponding eigenvalue is 3/4.
d2
• cos x = − cos x the function cos x) is an eigenfunction
dx 2 2
of the operator d /dx .
2

The corresponding eigenvalue is −1.

Slide 72 · QM principles and concepts · Expectation values and eigenvalues


Eigenthis and eigenthat (1)

The same thing again, only with more detail in the notation.

If application of an operator  to the function fpars (vars) results in that


same function multiplied by a nonzero real number apars , then:
• fpars (vars) is an eigenfunction of the operator Â.
• apars is the eigenvalue of the operator Â
corresponding to the eigenfunction fpars (vars).
• the function fpars (vars) satisfies the eigenvalue equation
Âfpars (vars) = apars fpars (vars)
Note:
• apars must be real and nonzero.
• the numerical value of apars can depend on the parameters of
fpars (vars), but it cannot depend on the variables of fpars (vars)

Slide 73 · QM principles and concepts · Expectation values and eigenvalues


Eigenthis and eigenthat (2)

If all of these things are true...


• Ψn ,n ,... (x1 , x2 , x3 , . . .) is the wavefunction of a system in the state
1 2
associated with the quantum numbers n1 , n2 , etc
• Ô is the operator associated with the observable O
• Ψn ,n ,... (x1 , x2 , x3 , . . .) is an eigenfunction of the operator Ô
1 2

...then we say that the system state identified by the quantum numbers
n1 , n2 , etc is an eigenstate of the observable O.

Note the correspondence between wavefunction and system state


and also the correspondence between observable and operator.

If the wavefunction is an eigenfunction of the operator,


then the state is an eigenstate of the observable.

Slide 74 · QM principles and concepts · Expectation values and eigenvalues


The outcome of a single experiment (1)
Let’s say you want to make an experiment in which
you’ll measure the value of the observable property A.
Q: Can the result of a single experiment be predicted?

Slide 75 · QM principles and concepts · Expectation values and eigenvalues


The outcome of a single experiment (1)
Let’s say you want to make an experiment in which
you’ll measure the value of the observable property A.
Q: Can the result of a single experiment be predicted?
A: It depends on whether the system you’ll be dealing with
is or is not in an eigenstate of the observable A.

Slide 75 · QM principles and concepts · Expectation values and eigenvalues


The outcome of a single experiment (1)
Let’s say you want to make an experiment in which
you’ll measure the value of the observable property A.
Q: Can the result of a single experiment be predicted?
A: It depends on whether the system you’ll be dealing with
is or is not in an eigenstate of the observable A.

System state is an System state is not an


eigenstate of the observable eigenstate of the observable
the result will be the the result will be one of the
eigenvalue associated with eigenvalues associated
that observable and with that observable and
eigenstate. some eigenstate.

Slide 75 · QM principles and concepts · Expectation values and eigenvalues


The outcome of a single experiment (1)
Let’s say you want to make an experiment in which
you’ll measure the value of the observable property A.
Q: Can the result of a single experiment be predicted?
A: It depends on whether the system you’ll be dealing with
is or is not in an eigenstate of the observable A.

System state is an System state is not an


eigenstate of the observable eigenstate of the observable
the result will be the the result will be one of the
eigenvalue associated with eigenvalues associated
that observable and with that observable and
eigenstate. some eigenstate.

Slide 75 · QM principles and concepts · Expectation values and eigenvalues


The outcome of a single experiment (1)
Let’s say you want to make an experiment in which
you’ll measure the value of the observable property A.
Q: Can the result of a single experiment be predicted?
A: It depends on whether the system you’ll be dealing with
is or is not in an eigenstate of the observable A.

System state is an System state is not an


eigenstate of the observable eigenstate of the observable
the result will be the the result will be one of the
eigenvalue associated with eigenvalues associated
that observable and with that observable and
eigenstate. some eigenstate.

Slide 75 · QM principles and concepts · Expectation values and eigenvalues


The outcome of a single experiment (1)
Let’s say you want to make an experiment in which
you’ll measure the value of the observable property A.
Q: Can the result of a single experiment be predicted?
A: It depends on whether the system you’ll be dealing with
is or is not in an eigenstate of the observable A.

System state is an System state is not an


eigenstate of the observable eigenstate of the observable
the result will be the the result will be one of the
eigenvalue associated with eigenvalues associated
that observable and with that observable and
eigenstate. some eigenstate.

(later on we’ll see how to determine the probability


of obtaining each of the possible eigenvalues)

Slide 75 · QM principles and concepts · Expectation values and eigenvalues


The outcome of a single experiment (2)
No matter how many times you repeat the experiment, each and every
time the result will be one of the eigenvalues of the operator Â. This
will happen regardless of whether the system is in an eigenstate of Â.
If the system is in an eigenstate of A — say, one with eigenvalue
3.5 — then each and every repetition of the experiment will lead to
the same result: 3.5.
If the system is not in an eigenstate of Â, then the result may change
from one experiment to the next.
But the result of each and every experiment will still be an
eigenvalue of Â, just not always the same one.
For instance, the eigenvalue 3.5 may still come out now and then,
but so will other eigenvalues.
(Either way, the average result of repeated, identical
experiments will be given by the expectation value)
Slide 76 · QM principles and concepts · Expectation values and eigenvalues
The outcome of a single experiment (2)
No matter how many times you repeat the experiment, each and every
time the result will be one of the eigenvalues of the operator Â. This
will happen regardless of whether the system is in an eigenstate of Â.
If the system is in an eigenstate of A — say, one with eigenvalue
3.5 — then each and every repetition of the experiment will lead to
the same result: 3.5.
If the system is not in an eigenstate of Â, then the result may change
from one experiment to the next.
But the result of each and every experiment will still be an
eigenvalue of Â, just not always the same one.
For instance, the eigenvalue 3.5 may still come out now and then,
but so will other eigenvalues.
(Either way, the average result of repeated, identical
experiments will be given by the expectation value)
Slide 76 · QM principles and concepts · Expectation values and eigenvalues
The outcome of a single experiment (2)
No matter how many times you repeat the experiment, each and every
time the result will be one of the eigenvalues of the operator Â. This
will happen regardless of whether the system is in an eigenstate of Â.
If the system is in an eigenstate of A — say, one with eigenvalue
3.5 — then each and every repetition of the experiment will lead to
the same result: 3.5.
If the system is not in an eigenstate of Â, then the result may change
from one experiment to the next.
But the result of each and every experiment will still be an
eigenvalue of Â, just not always the same one.
For instance, the eigenvalue 3.5 may still come out now and then,
but so will other eigenvalues.
(Either way, the average result of repeated, identical
experiments will be given by the expectation value)
Slide 76 · QM principles and concepts · Expectation values and eigenvalues
The outcome of a single experiment (2)
No matter how many times you repeat the experiment, each and every
time the result will be one of the eigenvalues of the operator Â. This
will happen regardless of whether the system is in an eigenstate of Â.
If the system is in an eigenstate of A — say, one with eigenvalue
3.5 — then each and every repetition of the experiment will lead to
the same result: 3.5.
If the system is not in an eigenstate of Â, then the result may change
from one experiment to the next.
But the result of each and every experiment will still be an
eigenvalue of Â, just not always the same one.
For instance, the eigenvalue 3.5 may still come out now and then,
but so will other eigenvalues.
(Either way, the average result of repeated, identical
experiments will be given by the expectation value)
Slide 76 · QM principles and concepts · Expectation values and eigenvalues
The outcome of a single experiment (2)
No matter how many times you repeat the experiment, each and every
time the result will be one of the eigenvalues of the operator Â. This
will happen regardless of whether the system is in an eigenstate of Â.
If the system is in an eigenstate of A — say, one with eigenvalue
3.5 — then each and every repetition of the experiment will lead to
the same result: 3.5.
If the system is not in an eigenstate of Â, then the result may change
from one experiment to the next.
But the result of each and every experiment will still be an
eigenvalue of Â, just not always the same one.
For instance, the eigenvalue 3.5 may still come out now and then,
but so will other eigenvalues.
(Either way, the average result of repeated, identical
experiments will be given by the expectation value)
Slide 76 · QM principles and concepts · Expectation values and eigenvalues
The outcome of a single experiment (2)
No matter how many times you repeat the experiment, each and every
time the result will be one of the eigenvalues of the operator Â. This
will happen regardless of whether the system is in an eigenstate of Â.
If the system is in an eigenstate of A — say, one with eigenvalue
3.5 — then each and every repetition of the experiment will lead to
the same result: 3.5.
If the system is not in an eigenstate of Â, then the result may change
from one experiment to the next.
But the result of each and every experiment will still be an
eigenvalue of Â, just not always the same one.
For instance, the eigenvalue 3.5 may still come out now and then,
but so will other eigenvalues.
(Either way, the average result of repeated, identical
experiments will be given by the expectation value)
Slide 76 · QM principles and concepts · Expectation values and eigenvalues
Eigenstuff, example 1 (an analogy)

black white red green blue

f1 f2 f3 f4 f5
(“eigenstate” 1) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 4) (“eigenstate” 5)

• The “system” is a disk.


• The “observable” is colour.
• The “colour operator” has five characteristic values:
the “eigenvalues” black, white, red, green and blue.
• The “black eigenstate” is described by the wavefunction f1
(and the “white,” “red”, “green”, “blue” eigenstates by f2 , f3 , f4 , f5 ).
• Other “colour values” (say, yellow, grey, orange or pink)
are not characteristic values of the “colour operator.”
Slide 77 · QM principles and concepts · Expectation values and eigenvalues
Eigenstuff, example 1 (an analogy)

black white red green blue

f1 f2 f3 f4 f5
(“eigenstate” 1) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 4) (“eigenstate” 5)

• The “system” is a disk.


• The “observable” is colour.
• The “colour operator” has five characteristic values:
the “eigenvalues” black, white, red, green and blue.
• The “black eigenstate” is described by the wavefunction f1
(and the “white,” “red”, “green”, “blue” eigenstates by f2 , f3 , f4 , f5 ).
• Other “colour values” (say, yellow, grey, orange or pink)
are not characteristic values of the “colour operator.”
Slide 77 · QM principles and concepts · Expectation values and eigenvalues
Eigenstuff, example 1 (an analogy)

black white red green blue

f1 f2 f3 f4 f5
(“eigenstate” 1) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 4) (“eigenstate” 5)

• The “system” is a disk.


• The “observable” is colour.
• The “colour operator” has five characteristic values:
the “eigenvalues” black, white, red, green and blue.
• The “black eigenstate” is described by the wavefunction f1
(and the “white,” “red”, “green”, “blue” eigenstates by f2 , f3 , f4 , f5 ).
• Other “colour values” (say, yellow, grey, orange or pink)
are not characteristic values of the “colour operator.”
Slide 77 · QM principles and concepts · Expectation values and eigenvalues
Eigenstuff, example 1 (an analogy)

black white red green blue

f1 f2 f3 f4 f5
(“eigenstate” 1) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 4) (“eigenstate” 5)

• The “system” is a disk.


• The “observable” is colour.
• The “colour operator” has five characteristic values:
the “eigenvalues” black, white, red, green and blue.
• The “black eigenstate” is described by the wavefunction f1
(and the “white,” “red”, “green”, “blue” eigenstates by f2 , f3 , f4 , f5 ).
• Other “colour values” (say, yellow, grey, orange or pink)
are not characteristic values of the “colour operator.”
Slide 77 · QM principles and concepts · Expectation values and eigenvalues
Eigenstuff, example 1 (an analogy)

black white red green blue

f1 f2 f3 f4 f5
(“eigenstate” 1) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 4) (“eigenstate” 5)

• The “system” is a disk.


• The “observable” is colour.
• The “colour operator” has five characteristic values:
the “eigenvalues” black, white, red, green and blue.
• The “black eigenstate” is described by the wavefunction f1
(and the “white,” “red”, “green”, “blue” eigenstates by f2 , f3 , f4 , f5 ).
• Other “colour values” (say, yellow, grey, orange or pink)
are not characteristic values of the “colour operator.”
Slide 77 · QM principles and concepts · Expectation values and eigenvalues
Eigenstuff, example 1 (an analogy, cont.)

Case A: system is in a colour eigenstate

green = green

ψ f4
(“system state”) (“eigenstate” 4)

• System state is a colour eigenstate (green).


• System wavefunction is the “green eigenfunction”: ψ = f4 .
• Every single measurement of the “colour observable” will result in
the same value: green, the “eigenvalue” that is characteristic of
this state of the “system” (the disk).

Slide 78 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff, example 1 (an analogy, cont.)

Case B: system is not in a colour eigenstate

r r r
3 2 1
= white − red + blue
6 6 6

ψ f2 f3 f5
(“system state”) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 5)

• System state is not a colour eigenstate (it’s a superposition


of the white, red and blue “colour eigenstates”).
• System wavefunction isn’t a “colour eigenfunction” (it’s a linear
combination of the white, red and blue “colour eigenfunctions”).

Slide 79 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff, example 1 (an analogy, cont.)

Case B: system is not in a colour eigenstate

r r r
3 2 1
= white − red + blue
6 6 6

ψ f2 f3 f5
(“system state”) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 5)

• Each single measurement can result in any one of the superposed


“colour eigenvalues”. The disk can be found to be completely
white, completely red or completely blue.
• No single measurement can show the disk to be simultaneously
white, red and blue.

Slide 80 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff, example 1 (an analogy, cont.)

Case B: system is not in a colour eigenstate

r r r
3 2 1
= white − red + blue
6 6 6

ψ f2 f3 f5
(“system state”) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 5)

• No measurement can result in an eigenvalue that is absent from


the superposition. Disk will never be found to be black or green.
• No single measurement can result in a colour value that is not a
“colour eigenvalue”. Disk will never be found to be, say, pink or
purple.

Slide 81 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff, example 1 (an analogy, cont.)

Case B: system is not in a colour eigenstate

r r r
3 2 1
= white − red + blue
6 6 6

ψ f2 f3 f5
(“system state”) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 5)

• If measurement is repeated many times, disk will be found to be:


I Completely white in 3/6 = 50% of the experiments.
I Completely red in 2/6 = 33.33% of the experiments.
I Completely blue in 1/6 = 16.67% of the experiments.

Slide 82 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff, example 1 (an analogy, cont.)

Case B: system is not in a colour eigenstate

r r r
3 2 1
= white − red + blue
6 6 6

ψ f2 f3 f5
(“system state”) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 5)

• If measurement is repeated many times, the average


colour of the disk will be found to be a light purple
given by 0.5 white + 0.3333 red + 0.1667 blue.
light purple
• This “light purple” is the “expectation value” of the
“colour observable” when the disk “state” is as above.
It will not appear as the result of any single experiment.

Slide 83 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff, example 1 (an analogy, cont.)

Case B: system is not in a colour eigenstate

r r r
3 2 1
= white − red + blue
6 6 6

ψ f2 f3 f5
(“system state”) (“eigenstate” 2) (“eigenstate” 3) (“eigenstate” 5)

• If a single measurement results in, say, a red disk,


afterwards the “disk state” will be the “red eigenstate”.
red • All subsequent measurements of the disk colour will
then lead to a single result: a red disk.
• This is known as the collapse of the wavefunction.

Slide 84 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2

A molecule of mass m is trapped in a 1-D box stretching from x = 0


to x = L. Some questions about its translational states:
• are they position, momentum or kinetic energy eigenstates?
• if so, what are the eigenvalues?
Operators and wavefunction:

position operator: x̂ = x
momentum operator: p̂x = −i ħ ∂/∂x
kinetic energy operator: K̂x = −(ħ2 /2m) (∂ 2 /∂x 2 )
q
wavefunction: ψ̂n (x) = 2/L sin(nπx/L), n = 1, 2, 3, . . .

The calculations are in the maple worksheet eigen-1D.mws.


As for the results...

Slide 85 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2

A molecule of mass m is trapped in a 1-D box stretching from x = 0


to x = L. Some questions about its translational states:
• are they position, momentum or kinetic energy eigenstates?
• if so, what are the eigenvalues?
Operators and wavefunction:

position operator: x̂ = x
momentum operator: p̂x = −i ħ ∂/∂x
kinetic energy operator: K̂x = −(ħ2 /2m) (∂ 2 /∂x 2 )
q
wavefunction: ψ̂n (x) = 2/L sin(nπx/L), n = 1, 2, 3, . . .

The calculations are in the maple worksheet eigen-1D.mws.


As for the results...

Slide 85 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2

A molecule of mass m is trapped in a 1-D box stretching from x = 0


to x = L. Some questions about its translational states:
• are they position, momentum or kinetic energy eigenstates?
• if so, what are the eigenvalues?
Operators and wavefunction:

position operator: x̂ = x
momentum operator: p̂x = −i ħ ∂/∂x
kinetic energy operator: K̂x = −(ħ2 /2m) (∂ 2 /∂x 2 )
q
wavefunction: ψ̂n (x) = 2/L sin(nπx/L), n = 1, 2, 3, . . .

The calculations are in the maple worksheet eigen-1D.mws.


As for the results...

Slide 85 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

• x̂ψn (x) = xψn (x)


1
• thick line: ψn (x).
0.5
• thin line: x̂ψn (x).
0 • application of operator:
I does not multiply ψ (x) by a
−0.5 n
nonzero real number with a fixed
−1
value for this system and state.
I multiplies ψ (x) by x, a variable.
n
0 0.2 0.4 0.6 0.8 1 I distorts wavefunction.
x/L I changes state properties.

(electron in n = 3 state, L = 1 nm;


see worksheet eigen-1D.mws)

Slide 86 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

• x̂ψn (x) = xψn (x)


1
• thick line: ψn (x).
0.5
• thin line: x̂ψn (x).
0 • application of operator:
I does not multiply ψ (x) by a
−0.5 n
nonzero real number with a fixed
−1
value for this system and state.
I multiplies ψ (x) by x, a variable.
n
0 0.2 0.4 0.6 0.8 1 I distorts wavefunction.
x/L I changes state properties.

(electron in n = 3 state, L = 1 nm;


see worksheet eigen-1D.mws)

Slide 86 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

• x̂ψn (x) = xψn (x)


1
• thick line: ψn (x).
0.5
• thin line: x̂ψn (x).
0 • application of operator:
I does not multiply ψ (x) by a
−0.5 n
nonzero real number with a fixed
−1
value for this system and state.
I multiplies ψ (x) by x, a variable.
n
0 0.2 0.4 0.6 0.8 1 I distorts wavefunction.
x/L I changes state properties.

(electron in n = 3 state, L = 1 nm;


see worksheet eigen-1D.mws)

Slide 86 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

• x̂ψn (x) = xψn (x)


1
• thick line: ψn (x).
0.5
• thin line: x̂ψn (x).
0 • application of operator:
I does not multiply ψ (x) by a
−0.5 n
nonzero real number with a fixed
−1
value for this system and state.
I multiplies ψ (x) by x, a variable.
n
0 0.2 0.4 0.6 0.8 1 I distorts wavefunction.
x/L I changes state properties.

(electron in n = 3 state, L = 1 nm;


see worksheet eigen-1D.mws)

Slide 86 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

• x̂ψn (x) = xψn (x)


1
• thick line: ψn (x).
0.5
• thin line: x̂ψn (x).
0 • application of operator:
I does not multiply ψ (x) by a
−0.5 n
nonzero real number with a fixed
−1
value for this system and state.
I multiplies ψ (x) by x, a variable.
n
0 0.2 0.4 0.6 0.8 1 I distorts wavefunction.
x/L I changes state properties.

(electron in n = 3 state, L = 1 nm; ψn (x) not a position eigenstate.


see worksheet eigen-1D.mws)

Slide 86 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

1
• p̂x ψn (x) = −ai (2/L) 2 cos(nπx/L),
1 where a = nh/(2L).
0.5
• thick line: ψn (x).
 
• thin line: (1/a) Im p̂x ψn (x) .
0
• application of operator:
−0.5
I does not multiply ψ (x) by a
n
−1 nonzero real number with a fixed
value for this system and state.
I changes wavefunction.
0 0.2 0.4 0.6 0.8 1
x/L I distorts wavefunction.
I changes state properties.
(electron in n = 3 state, L = 1 nm;
see worksheet eigen-1D.mws)

Slide 87 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

1
• p̂x ψn (x) = −ai (2/L) 2 cos(nπx/L),
1 where a = nh/(2L).
0.5
• thick line: ψn (x).
 
• thin line: (1/a) Im p̂x ψn (x) .
0
• application of operator:
−0.5
I does not multiply ψ (x) by a
n
−1 nonzero real number with a fixed
value for this system and state.
I changes wavefunction.
0 0.2 0.4 0.6 0.8 1
x/L I distorts wavefunction.
I changes state properties.
(electron in n = 3 state, L = 1 nm;
see worksheet eigen-1D.mws)

Slide 87 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

1
• p̂x ψn (x) = −ai (2/L) 2 cos(nπx/L),
1 where a = nh/(2L).
0.5
• thick line: ψn (x).
 
• thin line: (1/a) Im p̂x ψn (x) .
0
• application of operator:
−0.5
I does not multiply ψ (x) by a
n
−1 nonzero real number with a fixed
value for this system and state.
I changes wavefunction.
0 0.2 0.4 0.6 0.8 1
x/L I distorts wavefunction.
I changes state properties.
(electron in n = 3 state, L = 1 nm;
see worksheet eigen-1D.mws)

Slide 87 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

1
• p̂x ψn (x) = −ai (2/L) 2 cos(nπx/L),
1 where a = nh/(2L).
0.5
• thick line: ψn (x).
 
• thin line: (1/a) Im p̂x ψn (x) .
0
• application of operator:
−0.5
I does not multiply ψ (x) by a
n
−1 nonzero real number with a fixed
value for this system and state.
I changes wavefunction.
0 0.2 0.4 0.6 0.8 1
x/L I distorts wavefunction.
I changes state properties.
(electron in n = 3 state, L = 1 nm;
see worksheet eigen-1D.mws)

Slide 87 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

1
• p̂x ψn (x) = −ai (2/L) 2 cos(nπx/L),
1 where a = nh/(2L).
0.5
• thick line: ψn (x).
 
• thin line: (1/a) Im p̂x ψn (x) .
0
• application of operator:
−0.5
I does not multiply ψ (x) by a
n
−1 nonzero real number with a fixed
value for this system and state.
I changes wavefunction.
0 0.2 0.4 0.6 0.8 1
x/L I distorts wavefunction.
I changes state properties.
(electron in n = 3 state, L = 1 nm;
see worksheet eigen-1D.mws)
ψn (x) not a momentum eigenstate.

Slide 87 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

• K̂x ψn (x) = aψn (x),


2 2
where a = (nh) /(8mL ).
1
• solid thick line: ψn (x).
0.5
• dashed thin line: (1/a) K̂x ψn (x).
0 • application of operator:
−0.5 I multiplies ψ (x) by a nonzero real
n
number with a fixed value for this
−1
system and state.
I scales wavefunction.
0 0.2 0.4 0.6 0.8 1 I does not distort wavefunction.
x/L I does not change state properties.

(electron in n = 3 state, L = 1 nm;


see worksheet eigen-1D.mws)

Slide 88 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

• K̂x ψn (x) = aψn (x),


2 2
where a = (nh) /(8mL ).
1
• solid thick line: ψn (x).
0.5
• dashed thin line: (1/a) K̂x ψn (x).
0 • application of operator:
−0.5 I multiplies ψ (x) by a nonzero real
n
number with a fixed value for this
−1
system and state.
I scales wavefunction.
0 0.2 0.4 0.6 0.8 1 I does not distort wavefunction.
x/L I does not change state properties.

(electron in n = 3 state, L = 1 nm;


see worksheet eigen-1D.mws)

Slide 88 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

• K̂x ψn (x) = aψn (x),


2 2
where a = (nh) /(8mL ).
1
• solid thick line: ψn (x).
0.5
• dashed thin line: (1/a) K̂x ψn (x).
0 • application of operator:
−0.5 I multiplies ψ (x) by a nonzero real
n
number with a fixed value for this
−1
system and state.
I scales wavefunction.
0 0.2 0.4 0.6 0.8 1 I does not distort wavefunction.
x/L I does not change state properties.

(electron in n = 3 state, L = 1 nm;


see worksheet eigen-1D.mws)

Slide 88 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

• K̂x ψn (x) = aψn (x),


2 2
where a = (nh) /(8mL ).
1
• solid thick line: ψn (x).
0.5
• dashed thin line: (1/a) K̂x ψn (x).
0 • application of operator:
−0.5 I multiplies ψ (x) by a nonzero real
n
number with a fixed value for this
−1
system and state.
I scales wavefunction.
0 0.2 0.4 0.6 0.8 1 I does not distort wavefunction.
x/L I does not change state properties.

(electron in n = 3 state, L = 1 nm;


see worksheet eigen-1D.mws)

Slide 88 · QM principles and concepts · Expectation values and eigenvalues


Eigenstuff: example 2 (cont.)

• K̂x ψn (x) = aψn (x),


2 2
where a = (nh) /(8mL ).
1
• solid thick line: ψn (x).
0.5
• dashed thin line: (1/a) K̂x ψn (x).
0 • application of operator:
−0.5 I multiplies ψ (x) by a nonzero real
n
number with a fixed value for this
−1
system and state.
I scales wavefunction.
0 0.2 0.4 0.6 0.8 1 I does not distort wavefunction.
x/L I does not change state properties.

(electron in n = 3 state, L = 1 nm; ψn (x) is kinetic energy eigenstate.


see worksheet eigen-1D.mws) Eigenvalue is a = (nh)2 /(8mL2 ).

Slide 88 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• Every single measurement of the observable A must result


in one of the eigenvalues of the operator Â.

• If the system is in an A eigenstate, then a measurement of A


can only result in the eigenvalue associated with that eigenstate.

• If the system is not in an A eigenstate, then a measurement of A


must still result in an  eigenvalue, but we can’t predict which one.

• This unpredictability does not come from experimental limitations.


It is a feature of nature.

Slide 89 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• Every single measurement of the observable A must result


in one of the eigenvalues of the operator Â.

• If the system is in an A eigenstate, then a measurement of A


can only result in the eigenvalue associated with that eigenstate.

• If the system is not in an A eigenstate, then a measurement of A


must still result in an  eigenvalue, but we can’t predict which one.

• This unpredictability does not come from experimental limitations.


It is a feature of nature.

Slide 89 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• Every single measurement of the observable A must result


in one of the eigenvalues of the operator Â.

• If the system is in an A eigenstate, then a measurement of A


can only result in the eigenvalue associated with that eigenstate.

• If the system is not in an A eigenstate, then a measurement of A


must still result in an  eigenvalue, but we can’t predict which one.

• This unpredictability does not come from experimental limitations.


It is a feature of nature.

Slide 89 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• Every single measurement of the observable A must result


in one of the eigenvalues of the operator Â.

• If the system is in an A eigenstate, then a measurement of A


can only result in the eigenvalue associated with that eigenstate.

• If the system is not in an A eigenstate, then a measurement of A


must still result in an  eigenvalue, but we can’t predict which one.

• This unpredictability does not come from experimental limitations.


It is a feature of nature.

Slide 89 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• If the system is in an A eigenstate, then a measurement of A


does not change the system state.

• If the system is not in an A eigenstate, then a measurement of A


inevitably changes the system state, placing it in the A eigenstate
associated with the eigenvalue that was observed.

• This “collapse of the wavefunction” is an unavoidable


consequence of each and every observation.
Observations inevitably affect the observed object.

• No one knows how or why wavefunctions collapse.

Slide 90 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• If the system is in an A eigenstate, then a measurement of A


does not change the system state.

• If the system is not in an A eigenstate, then a measurement of A


inevitably changes the system state, placing it in the A eigenstate
associated with the eigenvalue that was observed.

• This “collapse of the wavefunction” is an unavoidable


consequence of each and every observation.
Observations inevitably affect the observed object.

• No one knows how or why wavefunctions collapse.

Slide 90 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• If the system is in an A eigenstate, then a measurement of A


does not change the system state.

• If the system is not in an A eigenstate, then a measurement of A


inevitably changes the system state, placing it in the A eigenstate
associated with the eigenvalue that was observed.

• This “collapse of the wavefunction” is an unavoidable


consequence of each and every observation.
Observations inevitably affect the observed object.

• No one knows how or why wavefunctions collapse.

Slide 90 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• If the system is in an A eigenstate, then a measurement of A


does not change the system state.

• If the system is not in an A eigenstate, then a measurement of A


inevitably changes the system state, placing it in the A eigenstate
associated with the eigenvalue that was observed.

• This “collapse of the wavefunction” is an unavoidable


consequence of each and every observation.
Observations inevitably affect the observed object.

• No one knows how or why wavefunctions collapse.

Slide 90 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• The average result of many identical measurements of A can be


predicted. It is given by the expectation value of A.

• The expectation value depends on the system state:



hAi = Ψ
R
all ÂΨ d ω.
space

• These are terms whose meaning is very important: operator,


observable, expectation value, eigenvalue, eigenfunction,
eigenstate, eigenvalue equation. You must understand what they
mean and how they are related to each other.

Slide 91 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• The average result of many identical measurements of A can be


predicted. It is given by the expectation value of A.

• The expectation value depends on the system state:



hAi = Ψ
R
all ÂΨ d ω.
space

• These are terms whose meaning is very important: operator,


observable, expectation value, eigenvalue, eigenfunction,
eigenstate, eigenvalue equation. You must understand what they
mean and how they are related to each other.

Slide 91 · QM principles and concepts · Expectation values and eigenvalues


Summary: expectation values and eigenvalues

• The average result of many identical measurements of A can be


predicted. It is given by the expectation value of A.

• The expectation value depends on the system state:



hAi = Ψ
R
all ÂΨ d ω.
space

• These are terms whose meaning is very important: operator,


observable, expectation value, eigenvalue, eigenfunction,
eigenstate, eigenvalue equation. You must understand what they
mean and how they are related to each other.

Lots of fancy words! But sometimes...

Slide 91 · QM principles and concepts · Expectation values and eigenvalues


Sometimes one word says it all

Slide 92 · QM principles and concepts · Expectation values and eigenvalues


Part I — QM principles and concepts

1 Complex numbers

2 Questions and answers

3 Wavefunctions

Finding the wavefunction: 4 Observables and operators

5 Expectation values and eigenvalues


TI and TD 6 TI and TD Schrödinger equations

Schrödinger equations 7 Phase and wave-particle duality

8 Overlap and probability

9 Uncertainty

10 Superposition of wavefunctions

Slide 93 · QM principles and concepts · TI and TD Schrödinger equations


Schrödinger equation

The Schrödinger equation is at the heart of QM.


This is the equation that, when solved, gives not only the
all-important wavefunction of a system, but also its time evolution.
The Schrödinger equation comes in two flavours,
time-dependent and time-independent.
The time-dependent (TD) Schrödinger equation is the one that
must be used in the general case of a system whose total energy
is not “well defined” and changes with time.
The time-independent (TI) Schrödinger equation is less general, but
nevertheless very important. It is appropriate for the particular case of
a system whose total energy is “well defined” and does not change
with time.

Slide 94 · QM principles and concepts · TI and TD Schrödinger equations


Schrödinger equation

The Schrödinger equation is at the heart of QM.


This is the equation that, when solved, gives not only the
all-important wavefunction of a system, but also its time evolution.
The Schrödinger equation comes in two flavours,
time-dependent and time-independent.
The time-dependent (TD) Schrödinger equation is the one that
must be used in the general case of a system whose total energy
is not “well defined” and changes with time.
The time-independent (TI) Schrödinger equation is less general, but
nevertheless very important. It is appropriate for the particular case of
a system whose total energy is “well defined” and does not change
with time.

Slide 94 · QM principles and concepts · TI and TD Schrödinger equations


Schrödinger equation

The Schrödinger equation is at the heart of QM.


This is the equation that, when solved, gives not only the
all-important wavefunction of a system, but also its time evolution.
The Schrödinger equation comes in two flavours,
time-dependent and time-independent.
The time-dependent (TD) Schrödinger equation is the one that
must be used in the general case of a system whose total energy
is not “well defined” and changes with time.
The time-independent (TI) Schrödinger equation is less general, but
nevertheless very important. It is appropriate for the particular case of
a system whose total energy is “well defined” and does not change
with time.

Slide 94 · QM principles and concepts · TI and TD Schrödinger equations


Time-dependent (TD) Schrödinger equation

This is for the general case in which the total energy may not be well
defined — the total energy and other properties may change with time:

∂Ψ
time-dependent Schrödinger equation: ĤΨ = i ħ
∂t

where Ĥ is the Hamiltonian operator, t is the time,


and Ψ the wavefunction.

Note (i) that the TD Schrödinger equation is not an eigenvalue


equation, and (ii) that it involves a partial derivative.

Slide 95 · QM principles and concepts · TI and TD Schrödinger equations


The TD Schrödinger equation in plain-ish language

In a sense, the TD Schrödinger equation only gives us


indirect information about the wavefunction of a system.
The direct information we get from the TD Schrödinger equation
is not about the Ψ itself, but rather about how it changes with time.
In order to obtain the wavefunction at all times, we must first find
what it is like at a particular moment — say, at t = 0.
(Mathematicians call this an “initial value problem”)
If we do know Ψ(t = 0) and how Ψ changes with time
— we need to know both — then we can determine
what we want: Ψ(t) for all values of t.
The TD Schrödinger equation only gives us the second piece of
information. We cannot use it to work out Ψ(t) for a system about
which we know nothing: we need Ψ(t = 0).

Slide 96 · QM principles and concepts · TI and TD Schrödinger equations


The TD Schrödinger equation in plain-ish language

In a sense, the TD Schrödinger equation only gives us


indirect information about the wavefunction of a system.
The direct information we get from the TD Schrödinger equation
is not about the Ψ itself, but rather about how it changes with time.
In order to obtain the wavefunction at all times, we must first find
what it is like at a particular moment — say, at t = 0.
(Mathematicians call this an “initial value problem”)
If we do know Ψ(t = 0) and how Ψ changes with time
— we need to know both — then we can determine
what we want: Ψ(t) for all values of t.
The TD Schrödinger equation only gives us the second piece of
information. We cannot use it to work out Ψ(t) for a system about
which we know nothing: we need Ψ(t = 0).

Slide 96 · QM principles and concepts · TI and TD Schrödinger equations


The TD Schrödinger equation in plain-ish language

In a sense, the TD Schrödinger equation only gives us


indirect information about the wavefunction of a system.
The direct information we get from the TD Schrödinger equation
is not about the Ψ itself, but rather about how it changes with time.
In order to obtain the wavefunction at all times, we must first find
what it is like at a particular moment — say, at t = 0.
(Mathematicians call this an “initial value problem”)
If we do know Ψ(t = 0) and how Ψ changes with time
— we need to know both — then we can determine
what we want: Ψ(t) for all values of t.
The TD Schrödinger equation only gives us the second piece of
information. We cannot use it to work out Ψ(t) for a system about
which we know nothing: we need Ψ(t = 0).

Slide 96 · QM principles and concepts · TI and TD Schrödinger equations


The TD Schrödinger equation in plain-ish language

In a sense, the TD Schrödinger equation only gives us


indirect information about the wavefunction of a system.
The direct information we get from the TD Schrödinger equation
is not about the Ψ itself, but rather about how it changes with time.
In order to obtain the wavefunction at all times, we must first find
what it is like at a particular moment — say, at t = 0.
(Mathematicians call this an “initial value problem”)
If we do know Ψ(t = 0) and how Ψ changes with time
— we need to know both — then we can determine
what we want: Ψ(t) for all values of t.
The TD Schrödinger equation only gives us the second piece of
information. We cannot use it to work out Ψ(t) for a system about
which we know nothing: we need Ψ(t = 0).

Slide 96 · QM principles and concepts · TI and TD Schrödinger equations


The TD Schrödinger equation in plain-ish language

In a sense, the TD Schrödinger equation only gives us


indirect information about the wavefunction of a system.
The direct information we get from the TD Schrödinger equation
is not about the Ψ itself, but rather about how it changes with time.
In order to obtain the wavefunction at all times, we must first find
what it is like at a particular moment — say, at t = 0.
(Mathematicians call this an “initial value problem”)
If we do know Ψ(t = 0) and how Ψ changes with time
— we need to know both — then we can determine
what we want: Ψ(t) for all values of t.
The TD Schrödinger equation only gives us the second piece of
information. We cannot use it to work out Ψ(t) for a system about
which we know nothing: we need Ψ(t = 0).

Slide 96 · QM principles and concepts · TI and TD Schrödinger equations


Time-independent (TI) Schrödinger equation

This is for the less general but nevertheless very important case in
which the total energy is well defined — neither the total energy
nor any other observable property changes with time:

time-independent Schrödinger equation: Ĥψ = E ψ,

where Ĥ is the Hamiltonian operator, E the total energy and ψ the


time-independent wavefunction.

Note (i) that the TI Schrödinger equation is an eigenvalue equation,


and (ii) that I have used a lowcase ψ. The reason for the latter will be
clear when we look at actual solutions of the TD Schrödinger equation.

Slide 97 · QM principles and concepts · TI and TD Schrödinger equations


The TI Schrödinger equation in plain-ish language

In contrast to its TD counterpart, the TI Schrödinger equation


gives us direct information about the wavefunction of a system.
Provided the total energy of the system is well-defined,
we can use it to obtain the system wavefunctions from scratch.
In principle, all we need to know is what the system is.
Which finally puts us in a position to actually “do”
quantum mechanics.
In a nutshell...

Slide 98 · QM principles and concepts · TI and TD Schrödinger equations


The TI Schrödinger equation in plain-ish language

In contrast to its TD counterpart, the TI Schrödinger equation


gives us direct information about the wavefunction of a system.
Provided the total energy of the system is well-defined,
we can use it to obtain the system wavefunctions from scratch.
In principle, all we need to know is what the system is.
Which finally puts us in a position to actually “do”
quantum mechanics.
In a nutshell...

Slide 98 · QM principles and concepts · TI and TD Schrödinger equations


The TI Schrödinger equation in plain-ish language

In contrast to its TD counterpart, the TI Schrödinger equation


gives us direct information about the wavefunction of a system.
Provided the total energy of the system is well-defined,
we can use it to obtain the system wavefunctions from scratch.
In principle, all we need to know is what the system is.
Which finally puts us in a position to actually “do”
quantum mechanics.
In a nutshell...

Slide 98 · QM principles and concepts · TI and TD Schrödinger equations


The TI Schrödinger equation in plain-ish language

In contrast to its TD counterpart, the TI Schrödinger equation


gives us direct information about the wavefunction of a system.
Provided the total energy of the system is well-defined,
we can use it to obtain the system wavefunctions from scratch.
In principle, all we need to know is what the system is.
Which finally puts us in a position to actually “do”
quantum mechanics.
In a nutshell...

Slide 98 · QM principles and concepts · TI and TD Schrödinger equations


The TI Schrödinger equation in plain-ish language

In contrast to its TD counterpart, the TI Schrödinger equation


gives us direct information about the wavefunction of a system.
Provided the total energy of the system is well-defined,
we can use it to obtain the system wavefunctions from scratch.
In principle, all we need to know is what the system is.
Which finally puts us in a position to actually “do”
quantum mechanics.
In a nutshell...

Slide 98 · QM principles and concepts · TI and TD Schrödinger equations


QM in plain(ish) language

Here it is:
1. In order to obtain information about a chemical system
(state, properties, etc.) we must ask our “oracle,”
the system wavefunction.
2. In order to find that oracle, we have to solve the Schrödinger
equation associated with our system.
3. Once we have found our oracle, we know how to ask it questions
(using operators), and we know how to interpret the answers (the
expectation values and eigenvalues) given by the oracle.

Slide 99 · QM principles and concepts · TI and TD Schrödinger equations


QM in plain(ish) language

Here it is:
1. In order to obtain information about a chemical system
(state, properties, etc.) we must ask our “oracle,”
the system wavefunction.
2. In order to find that oracle, we have to solve the Schrödinger
equation associated with our system.
3. Once we have found our oracle, we know how to ask it questions
(using operators), and we know how to interpret the answers (the
expectation values and eigenvalues) given by the oracle.

Slide 99 · QM principles and concepts · TI and TD Schrödinger equations


QM in plain(ish) language

Here it is:
1. In order to obtain information about a chemical system
(state, properties, etc.) we must ask our “oracle,”
the system wavefunction.
2. In order to find that oracle, we have to solve the Schrödinger
equation associated with our system.
3. Once we have found our oracle, we know how to ask it questions
(using operators), and we know how to interpret the answers (the
expectation values and eigenvalues) given by the oracle.

Slide 99 · QM principles and concepts · TI and TD Schrödinger equations


Finding the wavefunction and getting information
Here’s the recipe for finding a system’s wavefunction...
1. Determine the Hamiltonian operator and boundary conditions
of the system.
2. Use the Hamiltonian to solve the TI Schrödinger equation at time
t = 0. This will give you the initial wavefunction of the system.
3. Use the Hamiltonian and the initial wavefunction to solve the
TD Schrödinger equation. This will give you the wavefunction
of the system at any moment in time.
...and here’s the recipe for extracting information from it:
1. Determine the operator corresponding to the observable
you want to know about.
2. Use that operator to find eigenvalues and/or expectation values.
3. Interpret the results.

Slide 100 · QM principles and concepts · TI and TD Schrödinger equations


Finding the wavefunction and getting information
Here’s the recipe for finding a system’s wavefunction...
1. Determine the Hamiltonian operator and boundary conditions
of the system.
2. Use the Hamiltonian to solve the TI Schrödinger equation at time
t = 0. This will give you the initial wavefunction of the system.
3. Use the Hamiltonian and the initial wavefunction to solve the
TD Schrödinger equation. This will give you the wavefunction
of the system at any moment in time.
...and here’s the recipe for extracting information from it:
1. Determine the operator corresponding to the observable
you want to know about.
2. Use that operator to find eigenvalues and/or expectation values.
3. Interpret the results.

Slide 100 · QM principles and concepts · TI and TD Schrödinger equations


Finding the wavefunction and getting information
Here’s the recipe for finding a system’s wavefunction...
1. Determine the Hamiltonian operator and boundary conditions
of the system.
2. Use the Hamiltonian to solve the TI Schrödinger equation at time
t = 0. This will give you the initial wavefunction of the system.
3. Use the Hamiltonian and the initial wavefunction to solve the
TD Schrödinger equation. This will give you the wavefunction
of the system at any moment in time.
...and here’s the recipe for extracting information from it:
1. Determine the operator corresponding to the observable
you want to know about.
2. Use that operator to find eigenvalues and/or expectation values.
3. Interpret the results.

Slide 100 · QM principles and concepts · TI and TD Schrödinger equations


Finding the wavefunction and getting information
Here’s the recipe for finding a system’s wavefunction...
1. Determine the Hamiltonian operator and boundary conditions
of the system.
2. Use the Hamiltonian to solve the TI Schrödinger equation at time
t = 0. This will give you the initial wavefunction of the system.
3. Use the Hamiltonian and the initial wavefunction to solve the
TD Schrödinger equation. This will give you the wavefunction
of the system at any moment in time.
...and here’s the recipe for extracting information from it:
1. Determine the operator corresponding to the observable
you want to know about.
2. Use that operator to find eigenvalues and/or expectation values.
3. Interpret the results.

Slide 100 · QM principles and concepts · TI and TD Schrödinger equations


Finding the wavefunction and getting information
Here’s the recipe for finding a system’s wavefunction...
1. Determine the Hamiltonian operator and boundary conditions
of the system.
2. Use the Hamiltonian to solve the TI Schrödinger equation at time
t = 0. This will give you the initial wavefunction of the system.
3. Use the Hamiltonian and the initial wavefunction to solve the
TD Schrödinger equation. This will give you the wavefunction
of the system at any moment in time.
...and here’s the recipe for extracting information from it:
1. Determine the operator corresponding to the observable
you want to know about.
2. Use that operator to find eigenvalues and/or expectation values.
3. Interpret the results.

Slide 100 · QM principles and concepts · TI and TD Schrödinger equations


Finding the wavefunction and getting information
Here’s the recipe for finding a system’s wavefunction...
1. Determine the Hamiltonian operator and boundary conditions
of the system.
2. Use the Hamiltonian to solve the TI Schrödinger equation at time
t = 0. This will give you the initial wavefunction of the system.
3. Use the Hamiltonian and the initial wavefunction to solve the
TD Schrödinger equation. This will give you the wavefunction
of the system at any moment in time.
...and here’s the recipe for extracting information from it:
1. Determine the operator corresponding to the observable
you want to know about.
2. Use that operator to find eigenvalues and/or expectation values.
3. Interpret the results.

Slide 100 · QM principles and concepts · TI and TD Schrödinger equations


Easy, isn’t it?

Slide 101 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• It is by solving the Schrödinger equation that we can determine


the wavefunctions describing the possible states of a system.
• The Schrödinger equation comes in two flavours: TD and TI.
• The TD Schrödinger is appropriate for all quantum systems,
regardless of whether they have a characteristic total energy.
• The TD Schrödinger equation is not an eigenvalue equation.
• Using the TD Schrödinger equation we can determine
how the wavefunction of a system changes with time.
• The “TD Schrödinger problem” is an initial value problem. In order
to solve it we need to know the “initial value” of the wavefunction.
• The “initial value” of the wavefunction can be its expression at any
given moment, in particular at t = 0.

Slide 102 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• It is by solving the Schrödinger equation that we can determine


the wavefunctions describing the possible states of a system.
• The Schrödinger equation comes in two flavours: TD and TI.
• The TD Schrödinger is appropriate for all quantum systems,
regardless of whether they have a characteristic total energy.
• The TD Schrödinger equation is not an eigenvalue equation.
• Using the TD Schrödinger equation we can determine
how the wavefunction of a system changes with time.
• The “TD Schrödinger problem” is an initial value problem. In order
to solve it we need to know the “initial value” of the wavefunction.
• The “initial value” of the wavefunction can be its expression at any
given moment, in particular at t = 0.

Slide 102 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• It is by solving the Schrödinger equation that we can determine


the wavefunctions describing the possible states of a system.
• The Schrödinger equation comes in two flavours: TD and TI.
• The TD Schrödinger is appropriate for all quantum systems,
regardless of whether they have a characteristic total energy.
• The TD Schrödinger equation is not an eigenvalue equation.
• Using the TD Schrödinger equation we can determine
how the wavefunction of a system changes with time.
• The “TD Schrödinger problem” is an initial value problem. In order
to solve it we need to know the “initial value” of the wavefunction.
• The “initial value” of the wavefunction can be its expression at any
given moment, in particular at t = 0.

Slide 102 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• It is by solving the Schrödinger equation that we can determine


the wavefunctions describing the possible states of a system.
• The Schrödinger equation comes in two flavours: TD and TI.
• The TD Schrödinger is appropriate for all quantum systems,
regardless of whether they have a characteristic total energy.
• The TD Schrödinger equation is not an eigenvalue equation.
• Using the TD Schrödinger equation we can determine
how the wavefunction of a system changes with time.
• The “TD Schrödinger problem” is an initial value problem. In order
to solve it we need to know the “initial value” of the wavefunction.
• The “initial value” of the wavefunction can be its expression at any
given moment, in particular at t = 0.

Slide 102 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• It is by solving the Schrödinger equation that we can determine


the wavefunctions describing the possible states of a system.
• The Schrödinger equation comes in two flavours: TD and TI.
• The TD Schrödinger is appropriate for all quantum systems,
regardless of whether they have a characteristic total energy.
• The TD Schrödinger equation is not an eigenvalue equation.
• Using the TD Schrödinger equation we can determine
how the wavefunction of a system changes with time.
• The “TD Schrödinger problem” is an initial value problem. In order
to solve it we need to know the “initial value” of the wavefunction.
• The “initial value” of the wavefunction can be its expression at any
given moment, in particular at t = 0.

Slide 102 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• It is by solving the Schrödinger equation that we can determine


the wavefunctions describing the possible states of a system.
• The Schrödinger equation comes in two flavours: TD and TI.
• The TD Schrödinger is appropriate for all quantum systems,
regardless of whether they have a characteristic total energy.
• The TD Schrödinger equation is not an eigenvalue equation.
• Using the TD Schrödinger equation we can determine
how the wavefunction of a system changes with time.
• The “TD Schrödinger problem” is an initial value problem. In order
to solve it we need to know the “initial value” of the wavefunction.
• The “initial value” of the wavefunction can be its expression at any
given moment, in particular at t = 0.

Slide 102 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• It is by solving the Schrödinger equation that we can determine


the wavefunctions describing the possible states of a system.
• The Schrödinger equation comes in two flavours: TD and TI.
• The TD Schrödinger is appropriate for all quantum systems,
regardless of whether they have a characteristic total energy.
• The TD Schrödinger equation is not an eigenvalue equation.
• Using the TD Schrödinger equation we can determine
how the wavefunction of a system changes with time.
• The “TD Schrödinger problem” is an initial value problem. In order
to solve it we need to know the “initial value” of the wavefunction.
• The “initial value” of the wavefunction can be its expression at any
given moment, in particular at t = 0.

Slide 102 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• The TI Schrödinger is appropriate only for systems


whose states have a characteristic total energy.

• The TI Schrödinger equation is an eigenvalue equation.

• Using the TI Schrödinger equation we can determine the


wavefunction of a system at a given moment in time.

• The “TI Schrödinger problem” is a boundary value problem.


In order to solve it we need to know the “boundary conditions”
of the system.

Slide 103 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• The TI Schrödinger is appropriate only for systems


whose states have a characteristic total energy.

• The TI Schrödinger equation is an eigenvalue equation.

• Using the TI Schrödinger equation we can determine the


wavefunction of a system at a given moment in time.

• The “TI Schrödinger problem” is a boundary value problem.


In order to solve it we need to know the “boundary conditions”
of the system.

Slide 103 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• The TI Schrödinger is appropriate only for systems


whose states have a characteristic total energy.

• The TI Schrödinger equation is an eigenvalue equation.

• Using the TI Schrödinger equation we can determine the


wavefunction of a system at a given moment in time.

• The “TI Schrödinger problem” is a boundary value problem.


In order to solve it we need to know the “boundary conditions”
of the system.

Slide 103 · QM principles and concepts · TI and TD Schrödinger equations


Summary: TI and TD Schrödinger equations

• The TI Schrödinger is appropriate only for systems


whose states have a characteristic total energy.

• The TI Schrödinger equation is an eigenvalue equation.

• Using the TI Schrödinger equation we can determine the


wavefunction of a system at a given moment in time.

• The “TI Schrödinger problem” is a boundary value problem.


In order to solve it we need to know the “boundary conditions”
of the system.

Slide 103 · QM principles and concepts · TI and TD Schrödinger equations


Part I — QM principles and concepts

1 Complex numbers

2 Questions and answers

3 Wavefunctions
How “time-independent” systems 4 Observables and operators
change with time: 5 Expectation values and eigenvalues

6 TI and TD Schrödinger equations


the importance of the phase
7 Phase and wave-particle duality
and wave-particle duality
8 Overlap and probability

9 Uncertainty

10 Superposition of wavefunctions

Slide 104 · QM principles and concepts · Phase and wave-particle duality


Do TI systems change with time?

Well, yes and no.

Yes, the total wavefunction of a TI system does change with time.

But no, observable properties of a TI system do not change with time.


Neither does its time-independent wavefunction.

The reason: the way total wavefunctions of TI systems change


with time does not have any effect on anything observable.

How come?

Slide 105 · QM principles and concepts · Phase and wave-particle duality


Solution of TD Schrödinger equation
when Ĥ does not depend on time

If the Hamiltonian does not depend on time, then the total


wavefunction is given by

 Et
Total wavefunction: Ψ(x, t) = exp −i ψ(x),
ħ

where ψ(x) is the time-independent wavefunction


and E the total energy.
Note the use of Ψ and ψ: Ψ for the total wavefunction,
ψ for its time-independent part (the “TI wavefunction”).
In order to obtain E and ψ(x), we must solve the TI Schrödinger
equation, Ĥψ(x) = E ψ(x).

Slide 106 · QM principles and concepts · Phase and wave-particle duality


TD wavefunction of electron in a 1D box

At this point, you should study the maple worksheet


overall phase.mws.

You’ll find illustrations showing the following:


• TI wavefunction of an electron in a 1D box.
• Time evolution of the total wavefunction of an electron in a 1D box.
• The probability density does not change with time.
• The phase of the total wavefunction (the “overall phase”) changes
with time, but the modulus of the total wavefunction does not.

Slide 107 · QM principles and concepts · Phase and wave-particle duality


Is the overall phase important?

The last item on the previous slide is important, but you must
remember it is valid for systems whose Hamiltonian does not depend
on time (that is, TI systems).

A couple of comments on the overall phase:


• The actual value of the overall phase of an isolated system is
completely irrelevant. It is not observable, and has no influence
on anything observable.
• The time it takes for the phase to get back to its initial value,
however, is very important. This time is the period of oscillation
of the system.

Slide 108 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (1)

Example: calculation of an expectation value.

Z
∗
hAi = Ψ(x, t) Â Ψ(x, t) d ω

as

Slide 109 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (1)

Example: calculation of an expectation value.

Z
∗
Z h i∗
hAi = Ψ(x, t) Â Ψ(x, t) d ω = e−i E t/ħ ψ(x) Â e−i E t/ħ ψ(x) d ω

as as

Slide 109 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (1)

Example: calculation of an expectation value.

Z h i∗
hAi = = e−i E t/ħ ψ(x) Â e−i E t/ħ ψ(x) d ω
as
Z
= ei E t/ħ ψ(x) Â e−i E t/ħ ψ(x) d ω
as

Slide 109 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (1)

Example: calculation of an expectation value.

hAi =
Z
= ei E t/ħ ψ(x) Â e−i E t/ħ ψ(x) d ω
as
Z
= ei E t/ħ e−i E t/ħ ψ(x) Â ψ(x) d ω
as

Slide 109 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (1)

Example: calculation of an expectation value.

hAi =

Z Z
i E t/ħ −i E t/ħ
= e e ψ(x) Â ψ(x) d ω = ei E t/ħ−i E t/ħ ψ(x) Â ψ(x) d ω
as as

Slide 109 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (1)

Example: calculation of an expectation value.

hAi =

Z
= ei E t/ħ−i E t/ħ ψ(x) Â ψ(x) d ω
as
Z
= 0
e ψ(x) Â ψ(x) d ω
as

Slide 109 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (1)

Example: calculation of an expectation value.

hAi =

Z Z
= e0 ψ(x) Â ψ(x) d ω = ψ(x) Â ψ(x) d ω
as as

Slide 109 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (1)

Example: calculation of an expectation value.

hAi =

Z
= ψ(x) Â ψ(x) d ω
as

Slide 109 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (1)

Example: calculation of an expectation value.

hAi =

Z
= ψ(x) Â ψ(x) d ω
as

The overall phase does not appear in the final expression.

Slide 109 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (2)

The overall phase does not appear in the final expression.

This means that the value of the overall phase is irrelevant


for the calculation of expectation values.

The same thing happens in the calculation of eigenvalues.


(You can have a go at proving this one yourself.)

Since observable results always have to do


with either expectation values or eigenvalues...

...and the overall phase makes no difference for those...

...the overall phase makes no difference for anything observable.

Slide 110 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (2)

The overall phase does not appear in the final expression.

This means that the value of the overall phase is irrelevant


for the calculation of expectation values.

The same thing happens in the calculation of eigenvalues.


(You can have a go at proving this one yourself.)

Since observable results always have to do


with either expectation values or eigenvalues...

...and the overall phase makes no difference for those...

...the overall phase makes no difference for anything observable.

Slide 110 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (2)

The overall phase does not appear in the final expression.

This means that the value of the overall phase is irrelevant


for the calculation of expectation values.

The same thing happens in the calculation of eigenvalues.


(You can have a go at proving this one yourself.)

Since observable results always have to do


with either expectation values or eigenvalues...

...and the overall phase makes no difference for those...

...the overall phase makes no difference for anything observable.

Slide 110 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (2)

The overall phase does not appear in the final expression.

This means that the value of the overall phase is irrelevant


for the calculation of expectation values.

The same thing happens in the calculation of eigenvalues.


(You can have a go at proving this one yourself.)

Since observable results always have to do


with either expectation values or eigenvalues...

...and the overall phase makes no difference for those...

...the overall phase makes no difference for anything observable.

Slide 110 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (2)

The overall phase does not appear in the final expression.

This means that the value of the overall phase is irrelevant


for the calculation of expectation values.

The same thing happens in the calculation of eigenvalues.


(You can have a go at proving this one yourself.)

Since observable results always have to do


with either expectation values or eigenvalues...

...and the overall phase makes no difference for those...

...the overall phase makes no difference for anything observable.

Slide 110 · QM principles and concepts · Phase and wave-particle duality


Irrelevance of the overall phase (2)

The overall phase does not appear in the final expression.

This means that the value of the overall phase is irrelevant


for the calculation of expectation values.

The same thing happens in the calculation of eigenvalues.


(You can have a go at proving this one yourself.)

Since observable results always have to do


with either expectation values or eigenvalues...

...and the overall phase makes no difference for those...

...the overall phase makes no difference for anything observable.

Slide 110 · QM principles and concepts · Phase and wave-particle duality


If it has energy, it oscillates

We have seen that quantum systems oscillate as time goes by


and that TI systems have a well-defined period of oscillation.

The temporal oscillation comes from the TD part of the total


wavefunction. In the case of TI systems, it is always the same:
 Et Et Et
exp −i = cos − i sin ,
ħ ħ ħ
and won’t go away unless the energy is zero.

Conclusion: everything that has energy oscillates.

Slide 111 · QM principles and concepts · Phase and wave-particle duality


If it oscillates, it’s a wave (or is it?)

In classical mechanics...
• Things that oscillate are waves.
• Things that have mass are particles.
• Particles do not oscillate.
• Waves do not have mass.
• Particles and waves are completely different things.

Slide 112 · QM principles and concepts · Phase and wave-particle duality


If it oscillates, it’s a wave (or is it?)

In classical mechanics...
• Things that oscillate are waves.
• Things that have mass are particles.
• Particles do not oscillate.
• Waves do not have mass.
• Particles and waves are completely different things.

Slide 112 · QM principles and concepts · Phase and wave-particle duality


If it oscillates, it’s a wave (or is it?)

In classical mechanics...
• Things that oscillate are waves.
• Things that have mass are particles.
• Particles do not oscillate.
• Waves do not have mass.
• Particles and waves are completely different things.

Slide 112 · QM principles and concepts · Phase and wave-particle duality


If it oscillates, it’s a wave (or is it?)

In classical mechanics...
• Things that oscillate are waves.
• Things that have mass are particles.
• Particles do not oscillate.
• Waves do not have mass.
• Particles and waves are completely different things.

Slide 112 · QM principles and concepts · Phase and wave-particle duality


If it oscillates, it’s a wave (or is it?)

In classical mechanics...
• Things that oscillate are waves.
• Things that have mass are particles.
• Particles do not oscillate.
• Waves do not have mass.
• Particles and waves are completely different things.

Slide 112 · QM principles and concepts · Phase and wave-particle duality


If it oscillates, it’s a wave (or is it?)

In classical mechanics...
• Things that oscillate are waves.
• Things that have mass are particles.
• Particles do not oscillate.
• Waves do not have mass.
• Particles and waves are completely different things.
Quantum methods, however, have shown that a particle
−31
(an electron with mass me = 9.1 × 10 kg) oscillates.

Slide 112 · QM principles and concepts · Phase and wave-particle duality


If it oscillates, it’s a wave (or is it?)

In classical mechanics...
• Things that oscillate are waves.
• Things that have mass are particles.
• Particles do not oscillate.
• Waves do not have mass.
• Particles and waves are completely different things.
Quantum methods, however, have shown that a particle
−31
(an electron with mass me = 9.1 × 10 kg) oscillates.
Particles not only oscillate, they also show all properties
classically associated with waves: interference, diffraction, etc.

Slide 112 · QM principles and concepts · Phase and wave-particle duality


If it oscillates, it’s a wave (or is it?)

In classical mechanics...
• Things that oscillate are waves.
• Things that have mass are particles.
• Particles do not oscillate.
• Waves do not have mass.
• Particles and waves are completely different things.
Quantum methods, however, have shown that a particle
−31
(an electron with mass me = 9.1 × 10 kg) oscillates.
Particles not only oscillate, they also show all properties
classically associated with waves: interference, diffraction, etc.
This wave-particle duality has been observed experimentally
many times and in many different ways.

Slide 112 · QM principles and concepts · Phase and wave-particle duality


The Planck equation (1)
The overall phase of a TI system oscillates with time:
 Et Et Et
exp −i = cos − i sin .
ħ ħ ħ

Slide 113 · QM principles and concepts · Phase and wave-particle duality


The Planck equation (1)
The overall phase of a TI system oscillates with time:
 Et Et Et
exp −i = cos − i sin .
ħ ħ ħ
The initial value of the phase (its value at time t = 0) is 1.

Slide 113 · QM principles and concepts · Phase and wave-particle duality


The Planck equation (1)
The overall phase of a TI system oscillates with time:
 Et Et Et
exp −i = cos − i sin .
ħ ħ ħ
The initial value of the phase (its value at time t = 0) is 1.
After one period of oscillation, the phase returns to its initial value.

Slide 113 · QM principles and concepts · Phase and wave-particle duality


The Planck equation (1)
The overall phase of a TI system oscillates with time:
 Et Et Et
exp −i = cos − i sin .
ħ ħ ħ
The initial value of the phase (its value at time t = 0) is 1.
After one period of oscillation, the phase returns to its initial value.
This happens when E t/ħ = 2π.

Slide 113 · QM principles and concepts · Phase and wave-particle duality


The Planck equation (1)
The overall phase of a TI system oscillates with time:
 Et Et Et
exp −i = cos − i sin .
ħ ħ ħ
The initial value of the phase (its value at time t = 0) is 1.
After one period of oscillation, the phase returns to its initial value.
This happens when E t/ħ = 2π. That is, when t = 2πħ/E = h/E .

Slide 113 · QM principles and concepts · Phase and wave-particle duality


The Planck equation (1)

This happens when E t/ħ = 2π. That is, when t = 2πħ/E = h/E .
This implies that the period of oscillation is given by τ = h/E .

Slide 113 · QM principles and concepts · Phase and wave-particle duality


The Planck equation (1)

This implies that the period of oscillation is given by τ = h/E .


Now, the period of oscillation (τ) is related to the frequency of
oscillation (ν) by ν = 1/τ.

Slide 113 · QM principles and concepts · Phase and wave-particle duality


The Planck equation (1)

This implies that the period of oscillation is given by τ = h/E .


Now, the period of oscillation (τ) is related to the frequency of
oscillation (ν) by ν = 1/τ.

This gives us ν = E/h and, after a simple rearrangement, the

Planck equation: E = hν

Slide 113 · QM principles and concepts · Phase and wave-particle duality


So, there is a point to it after all...

Slide 114 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (1)
(and a glimpse of how chemical bonds are formed)

By now you know that, in atoms or molecules, electrons arrange


themselves in (atomic or molecular) electronic orbitals.
We will later look at that carefully, but for now let us use
a very simple model for those orbitals.
Let us look at atoms containing one electron only, and model their
orbitals as one-dimensional boxes centred on the atomic nucleus.
The orbital size is determined by the length of the box, the orbital
shape by the quantum number. The energy is determined by both.
Because the approximation is so crude, this model won’t give us
rigourous or accurate results.
It will, however, give us some valuable insight
into how chemical bonds are formed.

Slide 115 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (1)
(and a glimpse of how chemical bonds are formed)

By now you know that, in atoms or molecules, electrons arrange


themselves in (atomic or molecular) electronic orbitals.
We will later look at that carefully, but for now let us use
a very simple model for those orbitals.
Let us look at atoms containing one electron only, and model their
orbitals as one-dimensional boxes centred on the atomic nucleus.
The orbital size is determined by the length of the box, the orbital
shape by the quantum number. The energy is determined by both.
Because the approximation is so crude, this model won’t give us
rigourous or accurate results.
It will, however, give us some valuable insight
into how chemical bonds are formed.

Slide 115 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (1)
(and a glimpse of how chemical bonds are formed)

By now you know that, in atoms or molecules, electrons arrange


themselves in (atomic or molecular) electronic orbitals.
We will later look at that carefully, but for now let us use
a very simple model for those orbitals.
Let us look at atoms containing one electron only, and model their
orbitals as one-dimensional boxes centred on the atomic nucleus.
The orbital size is determined by the length of the box, the orbital
shape by the quantum number. The energy is determined by both.
Because the approximation is so crude, this model won’t give us
rigourous or accurate results.
It will, however, give us some valuable insight
into how chemical bonds are formed.

Slide 115 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (1)
(and a glimpse of how chemical bonds are formed)

By now you know that, in atoms or molecules, electrons arrange


themselves in (atomic or molecular) electronic orbitals.
We will later look at that carefully, but for now let us use
a very simple model for those orbitals.
Let us look at atoms containing one electron only, and model their
orbitals as one-dimensional boxes centred on the atomic nucleus.
The orbital size is determined by the length of the box, the orbital
shape by the quantum number. The energy is determined by both.
Because the approximation is so crude, this model won’t give us
rigourous or accurate results.
It will, however, give us some valuable insight
into how chemical bonds are formed.

Slide 115 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (1)
(and a glimpse of how chemical bonds are formed)

By now you know that, in atoms or molecules, electrons arrange


themselves in (atomic or molecular) electronic orbitals.
We will later look at that carefully, but for now let us use
a very simple model for those orbitals.
Let us look at atoms containing one electron only, and model their
orbitals as one-dimensional boxes centred on the atomic nucleus.
The orbital size is determined by the length of the box, the orbital
shape by the quantum number. The energy is determined by both.
Because the approximation is so crude, this model won’t give us
rigourous or accurate results.
It will, however, give us some valuable insight
into how chemical bonds are formed.

Slide 115 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (1)
(and a glimpse of how chemical bonds are formed)

By now you know that, in atoms or molecules, electrons arrange


themselves in (atomic or molecular) electronic orbitals.
We will later look at that carefully, but for now let us use
a very simple model for those orbitals.
Let us look at atoms containing one electron only, and model their
orbitals as one-dimensional boxes centred on the atomic nucleus.
The orbital size is determined by the length of the box, the orbital
shape by the quantum number. The energy is determined by both.
Because the approximation is so crude, this model won’t give us
rigourous or accurate results.
It will, however, give us some valuable insight
into how chemical bonds are formed.

Slide 115 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (2)
(and a glimpse of how chemical bonds are formed)

Let us take a look at what happens when two of our “box” atoms
approach each other. Using two maple worksheets.

separate boxes.mws approaching boxes.mws


• Animation showing how the • Animation showing how the
total atomic wavefunctions total molecular wavefunction
of the separate atoms changes with time and as the
change with time. atoms approach each other.
• Plot of the (time-independent) • Animation showing how the
probability density associated probability density changes
with two separate “box” as the two atoms approach
atoms. each other.

Slide 116 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (3)
(and a glimpse of how chemical bonds are formed)

Using approaching boxes.mws we can see this:


• Constructive interference between the atomic wavefunctions
leads to formation of a bonding molecular orbital.
• Destructive interference between the atomic wavefunctions
leads to formation of an antibonding molecular orbital.
• The relative phases of the atomic wavefunctions determine
whether they interfere constructively of destructively.
• The relative phase values that lead to bond formation depend
on the quantum numbers (and shapes) of the atomic orbitals.
• Whether constructive of destructive, interference is stronger
when the atomic orbitals have similar oscillation periods.

Slide 117 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (3)
(and a glimpse of how chemical bonds are formed)

Using approaching boxes.mws we can see this:


• Constructive interference between the atomic wavefunctions
leads to formation of a bonding molecular orbital.
• Destructive interference between the atomic wavefunctions
leads to formation of an antibonding molecular orbital.
• The relative phases of the atomic wavefunctions determine
whether they interfere constructively of destructively.
• The relative phase values that lead to bond formation depend
on the quantum numbers (and shapes) of the atomic orbitals.
• Whether constructive of destructive, interference is stronger
when the atomic orbitals have similar oscillation periods.

Slide 117 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (3)
(and a glimpse of how chemical bonds are formed)

Using approaching boxes.mws we can see this:


• Constructive interference between the atomic wavefunctions
leads to formation of a bonding molecular orbital.
• Destructive interference between the atomic wavefunctions
leads to formation of an antibonding molecular orbital.
• The relative phases of the atomic wavefunctions determine
whether they interfere constructively of destructively.
• The relative phase values that lead to bond formation depend
on the quantum numbers (and shapes) of the atomic orbitals.
• Whether constructive of destructive, interference is stronger
when the atomic orbitals have similar oscillation periods.

Slide 117 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (3)
(and a glimpse of how chemical bonds are formed)

Using approaching boxes.mws we can see this:


• Constructive interference between the atomic wavefunctions
leads to formation of a bonding molecular orbital.
• Destructive interference between the atomic wavefunctions
leads to formation of an antibonding molecular orbital.
• The relative phases of the atomic wavefunctions determine
whether they interfere constructively of destructively.
• The relative phase values that lead to bond formation depend
on the quantum numbers (and shapes) of the atomic orbitals.
• Whether constructive of destructive, interference is stronger
when the atomic orbitals have similar oscillation periods.

Slide 117 · QM principles and concepts · Phase and wave-particle duality


Box model for atomic orbitals (3)
(and a glimpse of how chemical bonds are formed)

Using approaching boxes.mws we can see this:


• Constructive interference between the atomic wavefunctions
leads to formation of a bonding molecular orbital.
• Destructive interference between the atomic wavefunctions
leads to formation of an antibonding molecular orbital.
• The relative phases of the atomic wavefunctions determine
whether they interfere constructively of destructively.
• The relative phase values that lead to bond formation depend
on the quantum numbers (and shapes) of the atomic orbitals.
• Whether constructive of destructive, interference is stronger
when the atomic orbitals have similar oscillation periods.

Slide 117 · QM principles and concepts · Phase and wave-particle duality


Part I — QM principles and concepts

1 Complex numbers

2 Questions and answers

3 Wavefunctions
Measurements 4 Observables and operators
result in eigenvalues. 5 Expectation values and eigenvalues
But which ones? 6 TI and TD Schrödinger equations

7 Phase and wave-particle duality


(Overlap and probability)
8 Overlap and probability

9 Uncertainty

10 Superposition of wavefunctions

Slide 118 · QM principles and concepts · Overlap and probability


Measurement and eigenvalues

We’ve seen this:

The result of an individual measurement of the observable A


must be one of the eigenvalues of the operator Â.

Operators in general have several eigenvalues, each one associated


with a particular eigenfunction:

Âfn (x) = an fn (x), n = 1, 2, 3, . . . .

Slide 119 · QM principles and concepts · Overlap and probability


Measurement and eigenvalues

We’ve seen this:

The result of an individual measurement of the observable A


must be one of the eigenvalues of the operator Â.

Operators in general have several eigenvalues, each one associated


with a particular eigenfunction:

Âfn (x) = an fn (x), n = 1, 2, 3, . . . .

Question: what eigenvalue will actually come out


as the measurement result?

Slide 119 · QM principles and concepts · Overlap and probability


If it’s an eigenstate, it’s a certainty

We’ve seen this:

If the system is in an eigenstate of Â, then the result of


an individual measurement of the observable A can only be
the particular eigenvalue that is characteristic of that eigenstate.

If the wavefunction of the system is, say,

ψ(x) = f7 (x),

then there’s only one possible outcome: the measurement will result in
A = a7 . This is so because

Âψ(x) = Âf7 (x) = a7 f7 (x) = a7 ψ(x).

Slide 120 · QM principles and concepts · Overlap and probability


What if it’s not an eigenstate?

We’ve seen this:

If the system is not in an eigenstate of A, then the result of


an individual measurement of the observable A the measuring A
must be one of the eigenvalues of Â, but we can’t predict which one.

If the system wavefunction satisfies

ψ(x) 6= fn (x) whatever the value of n

then the result of measuring A can be any of the eigenvalues:


a1 , a7 , a12 or any other.

Slide 121 · QM principles and concepts · Overlap and probability


What if it’s not an eigenstate?

We’ve seen this:

If the system is not in an eigenstate of A, then the result of


an individual measurement of the observable A the measuring A
must be one of the eigenvalues of Â, but we can’t predict which one.

If the system wavefunction satisfies

ψ(x) 6= fn (x) whatever the value of n

then the result of measuring A can be any of the eigenvalues:


a1 , a7 , a12 or any other.

Question: does that mean that the result of our measurement


is completely unpredictable?

Slide 121 · QM principles and concepts · Overlap and probability


If it’s not an eigenstate,
is it completely unpredictable?

Yes, completely: we can’t foretell which particular eigenvalue


will pop out of our measurement.

No, not completely: we can foretell the average result of many


experiments — that’s the expectation value.

No, not completely: we can foretell the likelihood — the probability —


of obtaining a particular eigenvalue.

Slide 122 · QM principles and concepts · Overlap and probability


If it’s not an eigenstate,
is it completely unpredictable?

Yes, completely: we can’t foretell which particular eigenvalue


will pop out of our measurement.

No, not completely: we can foretell the average result of many


experiments — that’s the expectation value.

No, not completely: we can foretell the likelihood — the probability —


of obtaining a particular eigenvalue.

Slide 122 · QM principles and concepts · Overlap and probability


If it’s not an eigenstate,
is it completely unpredictable?

Yes, completely: we can’t foretell which particular eigenvalue


will pop out of our measurement.

No, not completely: we can foretell the average result of many


experiments — that’s the expectation value.

No, not completely: we can foretell the likelihood — the probability —


of obtaining a particular eigenvalue.

Slide 122 · QM principles and concepts · Overlap and probability


If it’s not an eigenstate,
is it completely unpredictable?

Yes, completely: we can’t foretell which particular eigenvalue


will pop out of our measurement.

No, not completely: we can foretell the average result of many


experiments — that’s the expectation value.

No, not completely: we can foretell the likelihood — the probability —


of obtaining a particular eigenvalue.

But we haven’t seen how to actually deal with this last one yet.

Now’s the time, but first we need to talk about overlap.

Slide 122 · QM principles and concepts · Overlap and probability


Overlap is related to resemblance...

Remember this: the overlap is a complex number whose squared


modulus quantifies how much two functions resemble each other.
We represent the overlap by S and calculate it with the formula
Z
Sf g = [f (x)]∗ g(x) dx
all
space

where f (x) and g(x) are the two functions we are comparing.

Slide 123 · QM principles and concepts · Overlap and probability


Overlap is related to resemblance...

Remember this: the overlap is a complex number whose squared


modulus quantifies how much two functions resemble each other.
We represent the overlap by S and calculate it with the formula
Z
Sf g = [f (x)]∗ g(x) dx
all
space

where f (x) and g(x) are the two functions we are comparing.
If |Sf g |2 is large, then f (x) and g(x) are very similar.
If |Sf g |2 is small, then f (x) and g(x) are very different.

Slide 123 · QM principles and concepts · Overlap and probability


Overlap is related to resemblance...

Remember this: the overlap is a complex number whose squared


modulus quantifies how much two functions resemble each other.
We represent the overlap by S and calculate it with the formula
Z
Sf g = [f (x)]∗ g(x) dx
all
space

where f (x) and g(x) are the two functions we are comparing.
If |Sf g |2 is large, then f (x) and g(x) are very similar.
If |Sf g |2 is small, then f (x) and g(x) are very different.
But... what do we mean by “large” and “small”? If we have, say,
|Sf g |2 = 0.99, are f (x) and g(x) very similar or very different?

Slide 123 · QM principles and concepts · Overlap and probability


...and normalisation sets the scale (1)

There’s a very natural way of defining a scale


for the resemblance between two functions.
If they are completely different, their resemblance is zero.
This implies that when two functions are completely different,
their overlap is zero.
But if two functions are exactly equal, what is their overlap?
Is it one, one hundred, one million, or infinity?
That’s where normalisation steps in: normalisation has to do with
how much a function resembles itself, and therefore with what is
the maximum possible resemblance between two functions.

Slide 124 · QM principles and concepts · Overlap and probability


...and normalisation sets the scale (1)

There’s a very natural way of defining a scale


for the resemblance between two functions.
If they are completely different, their resemblance is zero.
This implies that when two functions are completely different,
their overlap is zero.
But if two functions are exactly equal, what is their overlap?
Is it one, one hundred, one million, or infinity?
That’s where normalisation steps in: normalisation has to do with
how much a function resembles itself, and therefore with what is
the maximum possible resemblance between two functions.

Slide 124 · QM principles and concepts · Overlap and probability


...and normalisation sets the scale (1)

There’s a very natural way of defining a scale


for the resemblance between two functions.
If they are completely different, their resemblance is zero.
This implies that when two functions are completely different,
their overlap is zero.
But if two functions are exactly equal, what is their overlap?
Is it one, one hundred, one million, or infinity?
That’s where normalisation steps in: normalisation has to do with
how much a function resembles itself, and therefore with what is
the maximum possible resemblance between two functions.

Slide 124 · QM principles and concepts · Overlap and probability


...and normalisation sets the scale (2)

The squared modulus of the overlap between a function and itself


is called the “norm” of the function. It is given by
Z Z
Nf = ∗
[f (x)] f (x) dx = |f (x)|2 dx
all all
space space

As no function can resemble f (x) more than f (x) itself,


the norm of a function sets the maximum value for the overlap.
But... does N have a fixed value? As it turns out, in QM
we are free to choose the value of N to be whatever we like.
The best way of dealing with this is to use normalisation to unity.
That is, to define every wavefunction so that its norm is N = 1.
In this convention, the overlap between two identical functions is 1
and the resemblance of two identical functions is also 1.

Slide 125 · QM principles and concepts · Overlap and probability


...and normalisation sets the scale (2)

The squared modulus of the overlap between a function and itself


is called the “norm” of the function. It is given by
Z Z
Nf = ∗
[f (x)] f (x) dx = |f (x)|2 dx
all all
space space

As no function can resemble f (x) more than f (x) itself,


the norm of a function sets the maximum value for the overlap.
But... does N have a fixed value? As it turns out, in QM
we are free to choose the value of N to be whatever we like.
The best way of dealing with this is to use normalisation to unity.
That is, to define every wavefunction so that its norm is N = 1.
In this convention, the overlap between two identical functions is 1
and the resemblance of two identical functions is also 1.

Slide 125 · QM principles and concepts · Overlap and probability


...and normalisation sets the scale (2)

The squared modulus of the overlap between a function and itself


is called the “norm” of the function. It is given by
Z Z
Nf = ∗
[f (x)] f (x) dx = |f (x)|2 dx
all all
space space

As no function can resemble f (x) more than f (x) itself,


the norm of a function sets the maximum value for the overlap.
But... does N have a fixed value? As it turns out, in QM
we are free to choose the value of N to be whatever we like.
The best way of dealing with this is to use normalisation to unity.
That is, to define every wavefunction so that its norm is N = 1.
In this convention, the overlap between two identical functions is 1
and the resemblance of two identical functions is also 1.

Slide 125 · QM principles and concepts · Overlap and probability


...and normalisation sets the scale (2)

The squared modulus of the overlap between a function and itself


is called the “norm” of the function. It is given by
Z Z
Nf = ∗
[f (x)] f (x) dx = |f (x)|2 dx
all all
space space

As no function can resemble f (x) more than f (x) itself,


the norm of a function sets the maximum value for the overlap.
But... does N have a fixed value? As it turns out, in QM
we are free to choose the value of N to be whatever we like.
The best way of dealing with this is to use normalisation to unity.
That is, to define every wavefunction so that its norm is N = 1.
In this convention, the overlap between two identical functions is 1
and the resemblance of two identical functions is also 1.

Slide 125 · QM principles and concepts · Overlap and probability


Example 1: wavefunctions of electron in a 1D box
r
Electron in a 1-D box 2  mπx 
• ψm (x) = sin
L 2
r
8 2  nπx 
quantum number

• ψn (x) = sin
L 2
2 L  mπx   nπx 
Z
6 • Smn = sin sin dx
L 0 2 2

0 0.25 0.5 0.75 1


x (nm)

Slide 126 · QM principles and concepts · Overlap and probability


Example 1: wavefunctions of electron in a 1D box
r
Electron in a 1-D box 2  mπx 
• ψm (x) = sin
L 2
r
8 2  nπx 
quantum number

• ψn (x) = sin
L 2
2 L  mπx   nπx 
Z
6 • Smn = sin sin dx
L 0 2 2
2
• S3,3 = ?, |S3,3 | = ?
3

0 0.25 0.5 0.75 1


x (nm)

Slide 126 · QM principles and concepts · Overlap and probability


Example 1: wavefunctions of electron in a 1D box
r
Electron in a 1-D box 2  mπx 
• ψm (x) = sin
L 2
r
8 2  nπx 
quantum number

• ψn (x) = sin
L 2
2 L  mπx   nπx 
Z
6 • Smn = sin sin dx
L 0 2 2
2
• S3,3 = 1, |S3,3 | = 1
3

0 0.25 0.5 0.75 1


x (nm)

Slide 126 · QM principles and concepts · Overlap and probability


Example 1: wavefunctions of electron in a 1D box
r
Electron in a 1-D box 2  mπx 
• ψm (x) = sin
L 2
r
8 2  nπx 
quantum number

• ψn (x) = sin
L 2
2 L  mπx   nπx 
Z
6 • Smn = sin sin dx
L 0 2 2

3
• S3,6 = ?, |S3,6 |2 = ?
0 0.25 0.5 0.75 1
x (nm)

Slide 126 · QM principles and concepts · Overlap and probability


Example 1: wavefunctions of electron in a 1D box
r
Electron in a 1-D box 2  mπx 
• ψm (x) = sin
L 2
r
8 2  nπx 
quantum number

• ψn (x) = sin
L 2
2 L  mπx   nπx 
Z
6 • Smn = sin sin dx
L 0 2 2

3
• S3,6 = 0, |S3,6 |2 = 0
0 0.25 0.5 0.75 1
x (nm)

Slide 126 · QM principles and concepts · Overlap and probability


Example 1: wavefunctions of electron in a 1D box
r
Electron in a 1-D box 2  mπx 
• ψm (x) = sin
L 2
r
8 2  nπx 
quantum number

• ψn (x) = sin
L 2
2 L  mπx   nπx 
Z
6 • Smn = sin sin dx
L 0 2 2

• S6,8 = ?, |S6,8 |2 = ?
0 0.25 0.5 0.75 1
x (nm)

Slide 126 · QM principles and concepts · Overlap and probability


Example 1: wavefunctions of electron in a 1D box
r
Electron in a 1-D box 2  mπx 
• ψm (x) = sin
L 2
r
8 2  nπx 
quantum number

• ψn (x) = sin
L 2
2 L  mπx   nπx 
Z
6 • Smn = sin sin dx
L 0 2 2

• S6,8 = 0, |S6,8 |2 = 0
0 0.25 0.5 0.75 1
x (nm)

Slide 126 · QM principles and concepts · Overlap and probability


Example 1: wavefunctions of electron in a 1D box
r
Electron in a 1-D box 2  mπx 
• ψm (x) = sin
L 2
r
8 2  nπx 
quantum number

• ψn (x) = sin
L 2
2 L  mπx   nπx 
Z
6 • Smn = sin sin dx
L 0 2 2
2
• S3,3 = 1, |S3,3 | = 1
3
• S3,6 = 0, |S3,6 |2 = 0
• S6,8 = 0, |S6,8 |2 = 0
0 0.25 0.5 0.75 1
x (nm)

Wavefunctions that describe states of the same system


but are associated with different eigenvalues of the same operator
always have zero overlap.

Slide 126 · QM principles and concepts · Overlap and probability


Example 2: electrons in different boxes

Electrons in different boxes


• short box: L = 1 nm
8 • long box: L = 2 nm
quantum number

0 0.5 1 1.5 2
x (nm)

Slide 127 · QM principles and concepts · Overlap and probability


Example 2: electrons in different boxes

Electrons in different boxes


• short box: L = 1 nm
8 • long box: L = 2 nm
quantum number

• S3S,3L = ?
6
|S3S,3L |2 = ?

0 0.5 1 1.5 2
x (nm)

Slide 127 · QM principles and concepts · Overlap and probability


Example 2: electrons in different boxes

Electrons in different boxes


• short box: L = 1 nm
8 • long box: L = 2 nm
quantum number

p
• S3S,3L = − 32/9π ≈ −0.200
6
|S3S,3L |2 = 32/81π2 ≈ 0.040

0 0.5 1 1.5 2
x (nm)

Slide 127 · QM principles and concepts · Overlap and probability


Example 2: electrons in different boxes

Electrons in different boxes


• short box: L = 1 nm
8 • long box: L = 2 nm
quantum number

• S3S,6L = ?
3
|S3S,6L |2 = ?
0 0.5 1 1.5 2
x (nm)

Slide 127 · QM principles and concepts · Overlap and probability


Example 2: electrons in different boxes

Electrons in different boxes


• short box: L = 1 nm
8 • long box: L = 2 nm
quantum number

p
• S3S,6L = 1/ 2 ≈ 0.707
3
|S3S,6L |2 = 1/2 = 0.5
0 0.5 1 1.5 2
x (nm)

Slide 127 · QM principles and concepts · Overlap and probability


Example 2: electrons in different boxes

Electrons in different boxes


• short box: L = 1 nm
8 • long box: L = 2 nm
quantum number

p
• S3S,3L = − 32/9π ≈ −0.200
6
|S3S,3L |2 = 32/81π2 ≈ 0.040
p
• S3S,6L = 1/ 2 ≈ 0.707
3
|S3S,6L |2 = 1/2 = 0.5
0 0.5 1 1.5 2
x (nm)

(see maple worksheets overlap example 1.mws


and overlap example 2.mws for calculation
of integrals involved in this and the previous example)

Slide 127 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (1)

Slide 122 says that we can foretell the probability of obtaining any one
of several possible eigenvalues (a1 , a2 , a3 , etc) when measuring the
observable A.

The mathematical formulas needed for that may look somewhat


complicated, but their meaning is simple:

The probability of obtaining the eigenvalue an


in a measurement of the observable A
is given by the resemblance between
the eigenfunction fn (x) and the wavefunction ψ(x)

Slide 128 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (2)

The same thing in a more schematic way:

1. We will look at a system whose state is described


mathematically by the wavefunction ψ(x).

2. We will measure the observable A, which is described


mathematically by the operator Â.

3. Measurement of A can only lead to one of the


characteristic values of this observable: a1 , a2 , a3 , etc.

4. The characteristic values of the observable A


are also the eigenvalues of the operator Â.

Slide 129 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (2)

The same thing in a more schematic way:

1. We will look at a system whose state is described


mathematically by the wavefunction ψ(x).

2. We will measure the observable A, which is described


mathematically by the operator Â.

3. Measurement of A can only lead to one of the


characteristic values of this observable: a1 , a2 , a3 , etc.

4. The characteristic values of the observable A


are also the eigenvalues of the operator Â.

Slide 129 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (2)

The same thing in a more schematic way:

1. We will look at a system whose state is described


mathematically by the wavefunction ψ(x).

2. We will measure the observable A, which is described


mathematically by the operator Â.

3. Measurement of A can only lead to one of the


characteristic values of this observable: a1 , a2 , a3 , etc.

4. The characteristic values of the observable A


are also the eigenvalues of the operator Â.

Slide 129 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (2)

The same thing in a more schematic way:

1. We will look at a system whose state is described


mathematically by the wavefunction ψ(x).

2. We will measure the observable A, which is described


mathematically by the operator Â.

3. Measurement of A can only lead to one of the


characteristic values of this observable: a1 , a2 , a3 , etc.

4. The characteristic values of the observable A


are also the eigenvalues of the operator Â.

Slide 129 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (3)

5. Each characteristic value of A is associated


with a characteristic state of the system.

6. The characteristic states are described mathematically by


f1 (x), f2 (x), f3 (x), etc. — the eigenfunctions of the operator Â.

7. The probability P (an ) of obtaining an when measuring A


is determined by how much the system state resembles
the characteristic state associated with an .

8. This resemblance is described mathematically by the


squared modulus of the overlap between fn (x) and ψ(x).

Slide 130 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (3)

5. Each characteristic value of A is associated


with a characteristic state of the system.

6. The characteristic states are described mathematically by


f1 (x), f2 (x), f3 (x), etc. — the eigenfunctions of the operator Â.

7. The probability P (an ) of obtaining an when measuring A


is determined by how much the system state resembles
the characteristic state associated with an .

8. This resemblance is described mathematically by the


squared modulus of the overlap between fn (x) and ψ(x).

Slide 130 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (3)

5. Each characteristic value of A is associated


with a characteristic state of the system.

6. The characteristic states are described mathematically by


f1 (x), f2 (x), f3 (x), etc. — the eigenfunctions of the operator Â.

7. The probability P (an ) of obtaining an when measuring A


is determined by how much the system state resembles
the characteristic state associated with an .

8. This resemblance is described mathematically by the


squared modulus of the overlap between fn (x) and ψ(x).

Slide 130 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (3)

5. Each characteristic value of A is associated


with a characteristic state of the system.

6. The characteristic states are described mathematically by


f1 (x), f2 (x), f3 (x), etc. — the eigenfunctions of the operator Â.

7. The probability P (an ) of obtaining an when measuring A


is determined by how much the system state resembles
the characteristic state associated with an .

8. This resemblance is described mathematically by the


squared modulus of the overlap between fn (x) and ψ(x).

Slide 130 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (4)

The same thing yet again, now in a mathematical way:

 
ψ(x) is a solution of the
Ĥψ(x) = E ψ(x)
TI Schrödinger equation

Slide 131 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (4)

The same thing yet again, now in a mathematical way:

 
ψ(x) is a solution of the
Ĥψ(x) = E ψ(x)
TI Schrödinger equation

an is an eigenvalue, and
 
Âfn (x) = an fn (x)
fn (x) an eigenfunction of Â

Slide 131 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (4)

The same thing yet again, now in a mathematical way:

 
ψ(x) is a solution of the
Ĥψ(x) = E ψ(x)
TI Schrödinger equation

an is an eigenvalue, and
 
Âfn (x) = an fn (x)
fn (x) an eigenfunction of Â

Sfn ,ψ is the overlap


Z  
Sfn ,ψ = [fn (x)]∗ ψ(x) dx
all between fn (x) and ψ(x)
space

Slide 131 · QM principles and concepts · Overlap and probability


Overlap, resemblance & probability (4)

The same thing yet again, now in a mathematical way:

 
ψ(x) is a solution of the
Ĥψ(x) = E ψ(x)
TI Schrödinger equation

an is an eigenvalue, and
 
Âfn (x) = an fn (x)
fn (x) an eigenfunction of Â

Sfn ,ψ is the overlap


Z  
Sfn ,ψ = [fn (x)]∗ ψ(x) dx
all between fn (x) and ψ(x)
space

probability of obtaining
 
2
P (an ) = |Sfn ,ψ |
an is given by |Sfn ,ψ |2

Slide 131 · QM principles and concepts · Overlap and probability


Example: energy of a “messed-up”
electron in a 1D box (1)

Suppose we’ve somehow “messed” with an electron in a 1D box in a


way that its state has become
r
4 
F (x) = (1 + i )ψ1 (x) − i ψ3 (x) + 12 ψ6 (x) ,

13
where the ψn (x) are the eigenfunctions of the total energy operator
(i.e., the eigenfunctions of the Hamiltonian, and the solutions of the
TI Schrödinger equation).

Slide 132 · QM principles and concepts · Overlap and probability


Example: energy of a “messed-up”
electron in a 1D box (1)

Suppose we’ve somehow “messed” with an electron in a 1D box in a


way that its state has become
r
4 
F (x) = (1 + i )ψ1 (x) − i ψ3 (x) + 12 ψ6 (x) ,

13
where the ψn (x) are the eigenfunctions of the total energy operator
(i.e., the eigenfunctions of the Hamiltonian, and the solutions of the
TI Schrödinger equation).

What result(s) should we expect if we measure the energy of our


electron while it’s in the “messed-up” state?

Slide 132 · QM principles and concepts · Overlap and probability


Example: energy of a “messed-up”
electron in a 1D box (2)
What about a single energy measurement?

eigenvalue overlap probability



−19 2 13 8
E1 = 0.15 × 10 J S1 = (1 + i ) P1 = = 61.54%
13 13

−19 2i 13 4
E3 = 1.36 × 10 J S3 = − P3 = = 30.77%
13 13

−19 13 1
E7 = 7.38 × 10 J S7 = P7 = = 7.69%
13 13

RL ∗ 2
Formulae: Sn = Sψn ,F = 0 [ψn (x)] F (x) dx, Pn = |Sn |
(see Maple worksheet probability.mws)

Slide 133 · QM principles and concepts · Overlap and probability


Example: energy of a “messed-up”
electron in a 1D box (2)
What about a single energy measurement?

eigenvalue overlap probability



−19 2 13 8
E1 = 0.15 × 10 J S1 = (1 + i ) P1 = = 61.54%
13 13

−19 2i 13 4
E3 = 1.36 × 10 J S3 = − P3 = = 30.77%
13 13

−19 13 1
E7 = 7.38 × 10 J S7 = P7 = = 7.69%
13 13

RL ∗ 2
Formulae: Sn = Sψn ,F = 0 [ψn (x)] F (x) dx, Pn = |Sn |
(see Maple worksheet probability.mws)

Slide 133 · QM principles and concepts · Overlap and probability


Example: energy of a “messed-up”
electron in a 1D box (3)

What about the average of many energy measurements?

hE i = P1 E1 + P3 E3 + P7 E7

Slide 134 · QM principles and concepts · Overlap and probability


Example: energy of a “messed-up”
electron in a 1D box (3)

What about the average of many energy measurements?

hE i = P1 E1 + P3 E3 + P7 E7
 
8 4 1
= 0.15 + 1.36 + 7.38 × 10−19 J
13 13 13

Slide 134 · QM principles and concepts · Overlap and probability


Example: energy of a “messed-up”
electron in a 1D box (3)

What about the average of many energy measurements?

hE i = P1 E1 + P3 E3 + P7 E7
 
8 4 1
= 0.15 + 1.36 + 7.38 × 10−19 J
13 13 13
ZL
= [ψ(x)]∗ Ĥψ(x) dx
0

Slide 134 · QM principles and concepts · Overlap and probability


Example: energy of a “messed-up”
electron in a 1D box (3)

What about the average of many energy measurements?

hE i = P1 E1 + P3 E3 + P7 E7
 
8 4 1
= 0.15 + 1.36 + 7.38 × 10−19 J
13 13 13
ZL
= [ψ(x)]∗ Ĥψ(x) dx
0
= 0.93 × 10−19 J

Slide 134 · QM principles and concepts · Overlap and probability


Example: energy of a “messed-up”
electron in a 1D box (3)

What about the average of many energy measurements?

hE i = P1 E1 + P3 E3 + P7 E7
 
8 4 1
= 0.15 + 1.36 + 7.38 × 10−19 J
13 13 13
ZL
= [ψ(x)]∗ Ĥψ(x) dx
0
= 0.93 × 10−19 J

(see Maple worksheet probability.mws)

Slide 134 · QM principles and concepts · Overlap and probability


Just do the math, my friend

Slide 135 · QM principles and concepts · Overlap and probability


Part I — QM principles and concepts

1 Complex numbers

2 Questions and answers

3 Wavefunctions

One more thing 4 Observables and operators

about measurement: 5 Expectation values and eigenvalues

6 TI and TD Schrödinger equations

uncertainty 7 Phase and wave-particle duality

8 Overlap and probability

9 Uncertainty

10 Superposition of wavefunctions

Slide 136 · QM principles and concepts · Uncertainty


Uncertainty (the famous one)

When the system we are considering is not in an eigenstate of the


observable we want to measure, there is no certainty about the
experimental result.
This, however, isn’t the famous “quantum mechanical uncertainty.”
The kind of quantum mechanical uncertainty that became a celebrity in
both science and philosophy is the one related to the Heisenberg
uncertainty principle.

Slide 137 · QM principles and concepts · Uncertainty


Uncertainty (the famous one)

When the system we are considering is not in an eigenstate of the


observable we want to measure, there is no certainty about the
experimental result.
This, however, isn’t the famous “quantum mechanical uncertainty.”
The kind of quantum mechanical uncertainty that became a celebrity in
both science and philosophy is the one related to the Heisenberg
uncertainty principle. From the Oxford English Dictionary:
(Heisenberg’s) uncertainty principle, a principle of
quantum mechanics implying that certain pairs of
observables cannot both be precisely and simultaneously
known, and that as one of any pair is more exactly
defined, the other becomes more uncertain.
The observables in such pairs are said to be complementary.

Slide 137 · QM principles and concepts · Uncertainty


Complementary observables
have noncommuting operators

To understand the uncertainty principle, we need to


understand the fancy words in the title of this slide.
Here’s what we’ll see in order to get there:

1. What’s a commutator?
2. What are commuting or noncommuting operators?
3. Examples of the things above.
4. A very simple analogy.
5. Mathematical definition of the uncertainty principle.
6. Numerical examples of uncertainty.

Slide 138 · QM principles and concepts · Uncertainty


Complementary observables
have noncommuting operators

To understand the uncertainty principle, we need to


understand the fancy words in the title of this slide.
Here’s what we’ll see in order to get there:

1. What’s a commutator?
2. What are commuting or noncommuting operators?
3. Examples of the things above.
4. A very simple analogy.
5. Mathematical definition of the uncertainty principle.
6. Numerical examples of uncertainty.

Slide 138 · QM principles and concepts · Uncertainty


Complementary observables
have noncommuting operators

To understand the uncertainty principle, we need to


understand the fancy words in the title of this slide.
Here’s what we’ll see in order to get there:

1. What’s a commutator?
2. What are commuting or noncommuting operators?
3. Examples of the things above.
4. A very simple analogy.
5. Mathematical definition of the uncertainty principle.
6. Numerical examples of uncertainty.

Slide 138 · QM principles and concepts · Uncertainty


Complementary observables
have noncommuting operators

To understand the uncertainty principle, we need to


understand the fancy words in the title of this slide.
Here’s what we’ll see in order to get there:

1. What’s a commutator?
2. What are commuting or noncommuting operators?
3. Examples of the things above.
4. A very simple analogy.
5. Mathematical definition of the uncertainty principle.
6. Numerical examples of uncertainty.

Slide 138 · QM principles and concepts · Uncertainty


Complementary observables
have noncommuting operators

To understand the uncertainty principle, we need to


understand the fancy words in the title of this slide.
Here’s what we’ll see in order to get there:

1. What’s a commutator?
2. What are commuting or noncommuting operators?
3. Examples of the things above.
4. A very simple analogy.
5. Mathematical definition of the uncertainty principle.
6. Numerical examples of uncertainty.

Slide 138 · QM principles and concepts · Uncertainty


Complementary observables
have noncommuting operators

To understand the uncertainty principle, we need to


understand the fancy words in the title of this slide.
Here’s what we’ll see in order to get there:

1. What’s a commutator?
2. What are commuting or noncommuting operators?
3. Examples of the things above.
4. A very simple analogy.
5. Mathematical definition of the uncertainty principle.
6. Numerical examples of uncertainty.

Slide 138 · QM principles and concepts · Uncertainty


The commutator of two operators

The commutator of the operators  and B̂ is given by

commutator of  and B̂: [Â, B̂] = ÂB̂ − B̂Â

Slide 139 · QM principles and concepts · Uncertainty


The commutator of two operators

The commutator of the operators  and B̂ is given by

commutator of  and B̂: [Â, B̂] = ÂB̂ − B̂Â

If the commutator of  and B̂ is zero,


we say that the operators  and B̂ commute.
In other terms, that  and B̂ are commuting operators.

Slide 139 · QM principles and concepts · Uncertainty


The commutator of two operators

The commutator of the operators  and B̂ is given by

commutator of  and B̂: [Â, B̂] = ÂB̂ − B̂Â

If the commutator of  and B̂ is zero,


we say that the operators  and B̂ commute.
In other terms, that  and B̂ are commuting operators.

If the commutator of  and B̂ is not zero,


we say that the operators  and B̂ do not commute.
In other terms, that  and B̂ are noncommuting operators.

Slide 139 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

and calculate two commutators: [x̂, ŷ] and [x̂, p̂x ].

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

and calculate two commutators: [x̂, ŷ] and [x̂, p̂x ].

Calculation of [x̂, ŷ] = x̂ ŷ − ŷ x̂:

x̂ ŷΨ = xyΨ

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

and calculate two commutators: [x̂, ŷ] and [x̂, p̂x ].

Calculation of [x̂, ŷ] = x̂ ŷ − ŷ x̂:

x̂ ŷΨ = xyΨ
ŷ x̂Ψ = yxΨ

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

and calculate two commutators: [x̂, ŷ] and [x̂, p̂x ].

Calculation of [x̂, ŷ] = x̂ ŷ − ŷ x̂:

x̂ ŷΨ = xyΨ
ŷ x̂Ψ = yxΨ = xyΨ

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

and calculate two commutators: [x̂, ŷ] and [x̂, p̂x ].

Calculation of [x̂, ŷ] = x̂ ŷ − ŷ x̂:

x̂ ŷΨ = xyΨ
ŷ x̂Ψ = yxΨ = xyΨ
[x̂, ŷ]Ψ = (x̂ ŷ − ŷ x̂)Ψ

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

and calculate two commutators: [x̂, ŷ] and [x̂, p̂x ].

Calculation of [x̂, ŷ] = x̂ ŷ − ŷ x̂:

x̂ ŷΨ = xyΨ
ŷ x̂Ψ = yxΨ = xyΨ
[x̂, ŷ]Ψ = (x̂ ŷ − ŷ x̂)Ψ = x̂ ŷΨ − ŷ x̂Ψ

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

and calculate two commutators: [x̂, ŷ] and [x̂, p̂x ].

Calculation of [x̂, ŷ] = x̂ ŷ − ŷ x̂:

x̂ ŷΨ = xyΨ
ŷ x̂Ψ = yxΨ = xyΨ
[x̂, ŷ]Ψ = (x̂ ŷ − ŷ x̂)Ψ = x̂ ŷΨ − ŷ x̂Ψ = xyΨ − xyΨ

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

and calculate two commutators: [x̂, ŷ] and [x̂, p̂x ].

Calculation of [x̂, ŷ] = x̂ ŷ − ŷ x̂:

x̂ ŷΨ = xyΨ
ŷ x̂Ψ = yxΨ = xyΨ
[x̂, ŷ]Ψ = (x̂ ŷ − ŷ x̂)Ψ = x̂ ŷΨ − ŷ x̂Ψ = xyΨ − xyΨ = 0

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (1)

Let us consider the following operators:


• Position along x direction: x̂ = x
• Position along y direction: ŷ = y
• Momentum along x direction: p̂x = −i ħ ∂x

and calculate two commutators: [x̂, ŷ] and [x̂, p̂x ].

Calculation of [x̂, ŷ] = x̂ ŷ − ŷ x̂:

x̂ ŷΨ = xyΨ
ŷ x̂Ψ = yxΨ = xyΨ
[x̂, ŷ]Ψ = (x̂ ŷ − ŷ x̂)Ψ = x̂ ŷΨ − ŷ x̂Ψ = xyΨ − xyΨ = 0

This implies [x̂, ŷ] = 0.

Slide 140 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


 

x̂ p̂x Ψ = x −i ħ Ψ
∂x

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ
∂x ∂x

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x
 

p̂x x̂Ψ = −i ħ xΨ
∂x

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x
 
∂ ∂(xΨ)
p̂x x̂Ψ = −i ħ xΨ = −i ħ
∂x ∂x

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x
   
∂ ∂(xΨ) ∂Ψ
p̂x x̂Ψ = −i ħ xΨ = −i ħ = −i ħ Ψ + x
∂x ∂x ∂x

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x
   
∂ ∂(xΨ) ∂Ψ
p̂x x̂Ψ = −i ħ xΨ = −i ħ = −i ħ Ψ + x
∂x ∂x ∂x
∂Ψ
= −i ħΨ − i ħx
∂x

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x
   
∂ ∂(xΨ) ∂Ψ
p̂x x̂Ψ = −i ħ xΨ = −i ħ = −i ħ Ψ + x
∂x ∂x ∂x
∂Ψ
= −i ħΨ − i ħx
∂x
[x̂, p̂x ]Ψ = (x̂ p̂x − p̂x x̂)Ψ

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x
   
∂ ∂(xΨ) ∂Ψ
p̂x x̂Ψ = −i ħ xΨ = −i ħ = −i ħ Ψ + x
∂x ∂x ∂x
∂Ψ
= −i ħΨ − i ħx
∂x
[x̂, p̂x ]Ψ = (x̂ p̂x − p̂x x̂)Ψ = x̂ p̂x Ψ − p̂x x̂Ψ

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x
   
∂ ∂(xΨ) ∂Ψ
p̂x x̂Ψ = −i ħ xΨ = −i ħ = −i ħ Ψ + x
∂x ∂x ∂x
∂Ψ
= −i ħΨ − i ħx
∂x
[x̂, p̂x ]Ψ = (x̂ p̂x − p̂x x̂)Ψ = x̂ p̂x Ψ − p̂x x̂Ψ
 
∂Ψ ∂Ψ
= −i ħx − −i ħΨ − i ħx
∂x ∂x

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x
   
∂ ∂(xΨ) ∂Ψ
p̂x x̂Ψ = −i ħ xΨ = −i ħ = −i ħ Ψ + x
∂x ∂x ∂x
∂Ψ
= −i ħΨ − i ħx
∂x
[x̂, p̂x ]Ψ = (x̂ p̂x − p̂x x̂)Ψ = x̂ p̂x Ψ − p̂x x̂Ψ
 
∂Ψ ∂Ψ
= −i ħx − −i ħΨ − i ħx
∂x ∂x
∂Ψ ∂Ψ
= −i ħx + i ħΨ + i ħx
∂x ∂x

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x
   
∂ ∂(xΨ) ∂Ψ
p̂x x̂Ψ = −i ħ xΨ = −i ħ = −i ħ Ψ + x
∂x ∂x ∂x
∂Ψ
= −i ħΨ − i ħx
∂x
[x̂, p̂x ]Ψ = (x̂ p̂x − p̂x x̂)Ψ = x̂ p̂x Ψ − p̂x x̂Ψ
 
∂Ψ ∂Ψ
= −i ħx − −i ħΨ − i ħx
∂x ∂x
∂Ψ ∂Ψ
= −i ħx + i ħΨ + i ħx = i ħΨ
∂x ∂x

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (2)

Calculation of [x̂, p̂x ] = x̂ p̂x − p̂x x̂:


   
∂ ∂Ψ ∂Ψ
x̂ p̂x Ψ = x −i ħ Ψ = x −i ħ = −i ħx
∂x ∂x ∂x
   
∂ ∂(xΨ) ∂Ψ
p̂x x̂Ψ = −i ħ xΨ = −i ħ = −i ħ Ψ + x
∂x ∂x ∂x
∂Ψ
= −i ħΨ − i ħx
∂x
[x̂, p̂x ]Ψ = (x̂ p̂x − p̂x x̂)Ψ = x̂ p̂x Ψ − p̂x x̂Ψ
 
∂Ψ ∂Ψ
= −i ħx − −i ħΨ − i ħx
∂x ∂x
∂Ψ ∂Ψ
= −i ħx + i ħΨ + i ħx = i ħΨ
∂x ∂x
This implies [x̂, p̂x ] = i ħ.

Slide 141 · QM principles and concepts · Uncertainty


Example: position and momentum (3)

We have obtained [x̂, ŷ] = 0 and [x̂, p̂x ] = i ħ.

Slide 142 · QM principles and concepts · Uncertainty


Example: position and momentum (3)

We have obtained [x̂, ŷ] = 0 and [x̂, p̂x ] = i ħ.

These results imply the following:


• x̂ and ŷ are commuting operators.
Application of x̂ and then ŷ to a function leads to
the same result as application of ŷ and then x̂.
• x̂ and p̂x are noncommuting operators.
Application of x̂ and then p̂x to a function does not
lead to the same result as application of p̂x and then x̂.

Slide 142 · QM principles and concepts · Uncertainty


Example: position and momentum (3)

We have obtained [x̂, ŷ] = 0 and [x̂, p̂x ] = i ħ.

These results imply the following:


• x̂ and ŷ are commuting operators.
Application of x̂ and then ŷ to a function leads to
the same result as application of ŷ and then x̂.
• x̂ and p̂x are noncommuting operators.
Application of x̂ and then p̂x to a function does not
lead to the same result as application of p̂x and then x̂.

Slide 142 · QM principles and concepts · Uncertainty


Example: position and momentum (3)

We have obtained [x̂, ŷ] = 0 and [x̂, p̂x ] = i ħ.

These results imply the following:


• x̂ and ŷ are commuting operators.
Application of x̂ and then ŷ to a function leads to
the same result as application of ŷ and then x̂.
• x̂ and p̂x are noncommuting operators.
Application of x̂ and then p̂x to a function does not
lead to the same result as application of p̂x and then x̂.

These conclusions do not depend on what particular function is


considered. They are valid for all functions and wavefunctions.

Slide 142 · QM principles and concepts · Uncertainty


Can we measure x and y at the same time? (1)

What if we want to simultaneously measure x and y? Would the


average measurement result be given by hxyi or hyxi?

Calculating the two expectation values we can see


if there’s a difference:
• If hxyi = hyxi, problem solved. The order of the measurements
does not make any difference. One measurement then the other,
the other way around, or simultaneous measurements — it’s all
the same.
• If hxyi =
6 hyxi, we’re stuck. Different calculations give different
results, and we don’t know which one (if any) is right.
The order in which the measurements are carried out matters.

Slide 143 · QM principles and concepts · Uncertainty


Can we measure x and y at the same time? (1)

What if we want to simultaneously measure x and y? Would the


average measurement result be given by hxyi or hyxi?

Calculating the two expectation values we can see


if there’s a difference:
• If hxyi = hyxi, problem solved. The order of the measurements
does not make any difference. One measurement then the other,
the other way around, or simultaneous measurements — it’s all
the same.
• If hxyi =
6 hyxi, we’re stuck. Different calculations give different
results, and we don’t know which one (if any) is right.
The order in which the measurements are carried out matters.

Slide 143 · QM principles and concepts · Uncertainty


Can we measure x and y at the same time? (1)

What if we want to simultaneously measure x and y? Would the


average measurement result be given by hxyi or hyxi?

Calculating the two expectation values we can see


if there’s a difference:
• If hxyi = hyxi, problem solved. The order of the measurements
does not make any difference. One measurement then the other,
the other way around, or simultaneous measurements — it’s all
the same.
• If hxyi =
6 hyxi, we’re stuck. Different calculations give different
results, and we don’t know which one (if any) is right.
The order in which the measurements are carried out matters.

Slide 143 · QM principles and concepts · Uncertainty


Can we measure x and y at the same time? (2)

The commutator of x̂ and ŷ is zero: [x̂, ŷ] = 0.

Slide 144 · QM principles and concepts · Uncertainty


Can we measure x and y at the same time? (2)

The commutator of x̂ and ŷ is zero: [x̂, ŷ] = 0.


In other words: x̂ ŷ − ŷ x̂ = 0, implying x̂ ŷ = ŷ x̂ and hxyi = hyxi.

Slide 144 · QM principles and concepts · Uncertainty


Can we measure x and y at the same time? (2)

The commutator of x̂ and ŷ is zero: [x̂, ŷ] = 0.


In other words: x̂ ŷ − ŷ x̂ = 0, implying x̂ ŷ = ŷ x̂ and hxyi = hyxi.

Conclusion:

The fact that x̂ and ŷ commute means that the order in which the
measurements are carried out does not affect the observed result.
In particular, the x position and the y position
can be determined simultaneously without problem.

Measuring the two things simultaneously is not a problem.


Their simultaneous measurement will lead to no uncertainty.

Slide 144 · QM principles and concepts · Uncertainty


Can we measure x and px at the same time? (1)

What if we want to simultaneously measure x and px ? Would the


average measurement result be given by hxpx i or hpx xi?

Calculating the two expectation values we can see


if there’s a difference:
• If hxpx i = hpx xi, problem solved. The order of the measurements
does not make any difference. One measurement then the other,
the other way around, or simultaneous measurements — it’s all
the same.
• If hxpx i =
6 hpx xi, we’re stuck. Different calculations give different
results, and we don’t know which one (if any) is right.
The order in which the measurements are carried out matters.

Slide 145 · QM principles and concepts · Uncertainty


Can we measure x and px at the same time? (1)

What if we want to simultaneously measure x and px ? Would the


average measurement result be given by hxpx i or hpx xi?

Calculating the two expectation values we can see


if there’s a difference:
• If hxpx i = hpx xi, problem solved. The order of the measurements
does not make any difference. One measurement then the other,
the other way around, or simultaneous measurements — it’s all
the same.
• If hxpx i =
6 hpx xi, we’re stuck. Different calculations give different
results, and we don’t know which one (if any) is right.
The order in which the measurements are carried out matters.

Slide 145 · QM principles and concepts · Uncertainty


Can we measure x and px at the same time? (1)

What if we want to simultaneously measure x and px ? Would the


average measurement result be given by hxpx i or hpx xi?

Calculating the two expectation values we can see


if there’s a difference:
• If hxpx i = hpx xi, problem solved. The order of the measurements
does not make any difference. One measurement then the other,
the other way around, or simultaneous measurements — it’s all
the same.
• If hxpx i =
6 hpx xi, we’re stuck. Different calculations give different
results, and we don’t know which one (if any) is right.
The order in which the measurements are carried out matters.

Slide 145 · QM principles and concepts · Uncertainty


Can we measure x and px at the same time? (2)

The commutator of x̂ and p̂x is not zero: [x̂, p̂x ] = i ħ.

Slide 146 · QM principles and concepts · Uncertainty


Can we measure x and px at the same time? (2)

The commutator of x̂ and p̂x is not zero: [x̂, p̂x ] = i ħ.


In other words: x̂ p̂x − p̂x x̂ 6= 0, implying x̂ p̂x 6= p̂x x̂ and
hxpx i =
6 hpx xi.

Slide 146 · QM principles and concepts · Uncertainty


Can we measure x and px at the same time? (2)

The commutator of x̂ and p̂x is not zero: [x̂, p̂x ] = i ħ.


In other words: x̂ p̂x − p̂x x̂ 6= 0, implying x̂ p̂x 6= p̂x x̂ and
hxpx i =
6 hpx xi.

Conclusion:

The fact that x̂ and p̂x do not commute means that the order in which
the measurements are carried out affects the observed result.
In particular, simultaneous measurement of the x position
and the px momentum is problematic.

Measuring the two things simultaneously is a problem.


Their simultaneous measurement will lead to uncertainty.

Slide 146 · QM principles and concepts · Uncertainty


Compatible or complementary?

Compatible: what would you rather have?


Fish and chips, or chips and fish?

Complementary: what would you rather do?


Live before you die, or die before you live?

Slide 147 · QM principles and concepts · Uncertainty


Compatible or complementary?

Compatible: what would you rather have?


Fish and chips, or chips and fish?

If two observables are compatible, the order in which they are


measured makes no difference. They can be simultaneously
measured with no uncertainty.

Complementary: what would you rather do?


Live before you die, or die before you live?

Slide 147 · QM principles and concepts · Uncertainty


Compatible or complementary?

Compatible: what would you rather have?


Fish and chips, or chips and fish?

Complementary: what would you rather do?


Live before you die, or die before you live?

If two observables are complementary, the order in which they are


measured does make a difference. Their simultaneous
measurement will lead to uncertainty.

Slide 147 · QM principles and concepts · Uncertainty


Uncertainty principle (quantitative)

In slide 137 we have seen a plain-language


definition of the uncertainty principle.

Now a mathematical, quantitative one,


for a general pair of observables (A and B):

1
uncertainty principle: ∆A × ∆B ≥ | h [Â, B̂] i |
2

Slide 148 · QM principles and concepts · Uncertainty


Uncertainty principle (quantitative)

In slide 137 we have seen a plain-language


definition of the uncertainty principle.

Now a mathematical, quantitative one,


for a general pair of observables (A and B):

1
uncertainty principle: ∆A × ∆B ≥ | h [Â, B̂] i |
2

In particular:

| h [x̂, ŷ] i | = | h 0 i |

Slide 148 · QM principles and concepts · Uncertainty


Uncertainty principle (quantitative)

In slide 137 we have seen a plain-language


definition of the uncertainty principle.

Now a mathematical, quantitative one,


for a general pair of observables (A and B):

1
uncertainty principle: ∆A × ∆B ≥ | h [Â, B̂] i |
2

In particular:

| h [x̂, ŷ] i | = | h 0 i | = | 0 |

Slide 148 · QM principles and concepts · Uncertainty


Uncertainty principle (quantitative)

In slide 137 we have seen a plain-language


definition of the uncertainty principle.

Now a mathematical, quantitative one,


for a general pair of observables (A and B):

1
uncertainty principle: ∆A × ∆B ≥ | h [Â, B̂] i |
2

In particular:

| h [x̂, ŷ] i | = | h 0 i | = | 0 | = 0

Slide 148 · QM principles and concepts · Uncertainty


Uncertainty principle (quantitative)

In slide 137 we have seen a plain-language


definition of the uncertainty principle.

Now a mathematical, quantitative one,


for a general pair of observables (A and B):

1
uncertainty principle: ∆A × ∆B ≥ | h [Â, B̂] i |
2

In particular:

| h [x̂, ŷ] i | = | h 0 i | = | 0 | = 0 ∆x × ∆y ≥ 0.

Slide 148 · QM principles and concepts · Uncertainty


Uncertainty principle (quantitative)

In slide 137 we have seen a plain-language


definition of the uncertainty principle.

Now a mathematical, quantitative one,


for a general pair of observables (A and B):

1
uncertainty principle: ∆A × ∆B ≥ | h [Â, B̂] i |
2

In particular:

| h [x̂, p̂x ] i | = | h i ħ i |

Slide 148 · QM principles and concepts · Uncertainty


Uncertainty principle (quantitative)

In slide 137 we have seen a plain-language


definition of the uncertainty principle.

Now a mathematical, quantitative one,


for a general pair of observables (A and B):

1
uncertainty principle: ∆A × ∆B ≥ | h [Â, B̂] i |
2

In particular:

| h [x̂, p̂x ] i | = | h i ħ i | = | i ħ |

Slide 148 · QM principles and concepts · Uncertainty


Uncertainty principle (quantitative)

In slide 137 we have seen a plain-language


definition of the uncertainty principle.

Now a mathematical, quantitative one,


for a general pair of observables (A and B):

1
uncertainty principle: ∆A × ∆B ≥ | h [Â, B̂] i |
2

In particular:

| h [x̂, p̂x ] i | = | h i ħ i | = | i ħ | = ħ

Slide 148 · QM principles and concepts · Uncertainty


Uncertainty principle (quantitative)

In slide 137 we have seen a plain-language


definition of the uncertainty principle.

Now a mathematical, quantitative one,


for a general pair of observables (A and B):

1
uncertainty principle: ∆A × ∆B ≥ | h [Â, B̂] i |
2

In particular:

ħ
| h [x̂, p̂x ] i | = | h i ħ i | = | i ħ | = ħ ∆x × ∆px ≥ .
2

Slide 148 · QM principles and concepts · Uncertainty


Uncertainty principle (quantitative)

In slide 137 we have seen a plain-language


definition of the uncertainty principle.

Now a mathematical, quantitative one,


for a general pair of observables (A and B):

1
uncertainty principle: ∆A × ∆B ≥ | h [Â, B̂] i |
2

In particular:

| h [x̂, ŷ] i | = | h 0 i | = | 0 | = 0 ∆x × ∆y ≥ 0.
ħ
| h [x̂, p̂x ] i | = | h i ħ i | = | i ħ | = ħ ∆x × ∆px ≥ .
2
What’s the meaning of these formulae?
Slide 148 · QM principles and concepts · Uncertainty
Meaning of uncertainty principle
Simultaneous measurement of x and y

∆x is the uncertainty in the value of x,


while ∆y is the uncertainty in the value of y.
The minimum value for the product of the two uncertainties is
zero. It is therefore possible to have ∆x = 0 and ∆y = 0, for it is
true that 0 × 0 ≥ 0.
There’s no intrinsic limit to the precision we can hope for when
determining the values of these compatible observables.
Practical (experimental) errors may lead to nonzero uncertainties,
but there’s no intrinsic limit to the precision of the measurement.

Slide 149 · QM principles and concepts · Uncertainty


Meaning of uncertainty principle
Simultaneous measurement of x and p x

∆x is the uncertainty in the value of x,


while ∆px is the uncertainty in the value of px .
The minimum value for the product of the two uncertainties is ħ/2.
It is therefore impossible to have ∆x = 0 and ∆px = 0, for it is not
true that 0 × 0 ≥ ħ/2.
There’s an intrinsic limit to the precision we can hope for when
determining the values of these complementary observables.
Regardless of practical (experimental) errors,
there’s an intrinsic limit to the precision of the measurement.
We must always have ∆x × ∆px ≥ ħ/2. The more ∆x approaches
zero, the more ∆px approaches infinity. And vice-versa.

Slide 150 · QM principles and concepts · Uncertainty


Numerical example 1: football

Footballers can easily make a football (m ≈ 10−1 kg) travel at speeds


near 20 mph (roughly 10 m s−1 ). This implies

pball = mv ≈ (10−1 kg)(10 m s−1 ) = 1 kg m s−1 .

If we measure the position of the ball with an uncertainty


−12
∆x = 1.0 × 10 m, the uncertainty in a simultaneous momentum
measurement must satisfy

ħ 1.1 × 10−34 J s
∆p ≥ = = 5.5 × 10−23 kg m s−1 .
2∆x 2.0 × 10−12 m
∆x is tiny in comparison to the distances the ball travels
and so is ∆p in comparison to the ball’s momentum.

Slide 151 · QM principles and concepts · Uncertainty


Numerical example 1: football

Footballers can easily make a football (m ≈ 10−1 kg) travel at speeds


near 20 mph (roughly 10 m s−1 ). This implies

pball = mv ≈ (10−1 kg)(10 m s−1 ) = 1 kg m s−1 .

If we measure the position of the ball with an uncertainty


−12
∆x = 1.0 × 10 m, the uncertainty in a simultaneous momentum
measurement must satisfy

ħ 1.1 × 10−34 J s
∆p ≥ = = 5.5 × 10−23 kg m s−1 .
2∆x 2.0 × 10−12 m
∆x is tiny in comparison to the distances the ball travels
and so is ∆p in comparison to the ball’s momentum.

The limitations imposed by the uncertainty principle are negligible.

Slide 151 · QM principles and concepts · Uncertainty


Numerical example 2: chemical bond
At its lowest energy state, an electron in a chemical bond (length
−10
r ≈ 10 m) has a typical average momentum

pe ≈ 10−24 kg m s−1 .

If we measure the position of the electron with an uncertainty


∆x = 5.0 × 10−11 m, the uncertainty in a simultaneous momentum
measurement must satisfy
−34
ħ 1.1 × 10 Js
∆p ≥ = = 1.1 × 10−24 kg m s−1 .
2∆x 1.0 × 10 −10 m
∆x is quite large in comparison to the distances the electron travels
(roughly the length of the bond) and so is ∆p in comparison to the
electron’s momentum.

Slide 152 · QM principles and concepts · Uncertainty


Numerical example 2: chemical bond
At its lowest energy state, an electron in a chemical bond (length
−10
r ≈ 10 m) has a typical average momentum

pe ≈ 10−24 kg m s−1 .

If we measure the position of the electron with an uncertainty


∆x = 5.0 × 10−11 m, the uncertainty in a simultaneous momentum
measurement must satisfy
−34
ħ 1.1 × 10 Js
∆p ≥ = = 1.1 × 10−24 kg m s−1 .
2∆x 1.0 × 10 −10 m
∆x is quite large in comparison to the distances the electron travels
(roughly the length of the bond) and so is ∆p in comparison to the
electron’s momentum.

The limitations imposed by the uncertainty principle are drastic.


Slide 152 · QM principles and concepts · Uncertainty
Food for thought

What exactly does the uncertainty principle tell us?

• That experiments and calculations are limited and that this


restricts the knowledge we can obtain about the physical world?

• That the concepts we use are limited and that this restricts the
understanding we can obtain about the physical world?

• Or that the physical world itself is not a precisely defined thing?

Slide 153 · QM principles and concepts · Uncertainty


Food for thought

What exactly does the uncertainty principle tell us?

• That experiments and calculations are limited and that this


restricts the knowledge we can obtain about the physical world?

• That the concepts we use are limited and that this restricts the
understanding we can obtain about the physical world?

• Or that the physical world itself is not a precisely defined thing?

Slide 153 · QM principles and concepts · Uncertainty


Food for thought

What exactly does the uncertainty principle tell us?

• That experiments and calculations are limited and that this


restricts the knowledge we can obtain about the physical world?

• That the concepts we use are limited and that this restricts the
understanding we can obtain about the physical world?

• Or that the physical world itself is not a precisely defined thing?

Slide 153 · QM principles and concepts · Uncertainty


Food for thought

What exactly does the uncertainty principle tell us?

• That experiments and calculations are limited and that this


restricts the knowledge we can obtain about the physical world?

• That the concepts we use are limited and that this restricts the
understanding we can obtain about the physical world?

• Or that the physical world itself is not a precisely defined thing?

Unfortunately, such philosophical questions (they can’t be answered by


experimental measurements) aren’t included in the course material.

Slide 153 · QM principles and concepts · Uncertainty


Uncertainty principle — summary (1)

• The quantitative expression of the uncertainty principle,


∆A ∆B ≥ 21 | h [Â, B̂] i |, shows that the product of the uncertainties
in the simultaneous measurement of observables A and B is
related to the expectation value of the [Â, B̂] commutator.

Slide 154 · QM principles and concepts · Uncertainty


Uncertainty principle — summary (1)

• The quantitative expression of the uncertainty principle,


∆A ∆B ≥ 21 | h [Â, B̂] i |, shows that the product of the uncertainties
in the simultaneous measurement of observables A and B is
related to the expectation value of the [Â, B̂] commutator.

• If the operators  and B̂ do not commute, then the corresponding


observables are complementary (and vice-versa).
I Application of  and then B̂ to a function does not lead
to the same result as application of B̂ and then Â.
I Measurement of the value of A and then B is not the same
as measurement of the value of B and then A.

Slide 154 · QM principles and concepts · Uncertainty


Uncertainty principle — summary (2)

• The values of pairs of complementary observables cannot be


exactly and simultaneously known.

Slide 155 · QM principles and concepts · Uncertainty


Uncertainty principle — summary (2)

• The values of pairs of complementary observables cannot be


exactly and simultaneously known.

• In a simultaneous measurement of two complementary


observables, if the precision in the measurement of one of them
becomes very high (that is, the uncertainty becomes very small,
approaching zero), the precision in the measurement of the other
one will become very low (that is, the uncertainty will become very
large, approaching infinity).

Slide 155 · QM principles and concepts · Uncertainty


Uncertainty principle — summary (2)

• The values of pairs of complementary observables cannot be


exactly and simultaneously known.

• In a simultaneous measurement of two complementary


observables, if the precision in the measurement of one of them
becomes very high (that is, the uncertainty becomes very small,
approaching zero), the precision in the measurement of the other
one will become very low (that is, the uncertainty will become very
large, approaching infinity).

• The product of the uncertainties of complementary observables in


a simultaneous measurement is of the order of ħ. This is irrelevant
in the macroscopic world, but not in the molecular world.

Slide 155 · QM principles and concepts · Uncertainty


Part I — QM principles and concepts

1 Complex numbers

2 Questions and answers

3 Wavefunctions
Using simple things 4 Observables and operators
to describe complex ones: 5 Expectation values and eigenvalues

6 TI and TD Schrödinger equations


superposition
7 Phase and wave-particle duality
of wavefunctions 8 Overlap and probability

9 Uncertainty

10 Superposition of wavefunctions

Slide 156 · QM principles and concepts · Superposition of wavefunctions


Wavefunctions of complex systems

We’ve seen how, mixing the wavefunctions of electrons in separate


1D boxes, we can learn quite a few things about chemical bonds.
The principle is general. Mixing wavefunctions of simple systems,
we can learn lots of things about complex systems.
The reason: by superposing wavefunctions of simple systems, we can
obtain approximations to the wavefunctions of complex systems.

Slide 157 · QM principles and concepts · Superposition of wavefunctions


Wavefunctions of complex systems

We’ve seen how, mixing the wavefunctions of electrons in separate


1D boxes, we can learn quite a few things about chemical bonds.
The principle is general. Mixing wavefunctions of simple systems,
we can learn lots of things about complex systems.
The reason: by superposing wavefunctions of simple systems, we can
obtain approximations to the wavefunctions of complex systems.

Before moving on to “simple systems” other than a particle in a box,


let us look at two examples of wavefunction superposition:
• Hybrid states of particle in a 2D box.
• Time-dependent states of particle in a 1D box.

Slide 157 · QM principles and concepts · Superposition of wavefunctions


Example 1: states of a particle in a 2D box (1)

We will look at a particle of mass m in a square 2D box with sides


Lx = Ly = L. The energy levels and “natural” TI wavefunctions of this
system are given by
2
h
Enx ny = Enx + Eny = (n2x + n2y ),
8mL 2
2  n πx   ny πy 
x
ψnx ny (x, y) = ψnx (x) × ψny (y) = sin sin .
L L L
The system has degenerate states. Superposing the “natural”
wavefunctions of degenerate states we can obtain different (“hybrid”)
wavefunctions for the same system and energy level.

In this example we’ll consider natural and hybrid wavefunctions


associated with the first (and doubly degenerate) excited energy level.

Slide 158 · QM principles and concepts · Superposition of wavefunctions


Example 1: states of a particle in a 2D box (2)

ψ1,2 state (“natural”) ψ2,1 state (“natural”)

• ψ1,2 (x, y) = 2
L
sin( πx
L
) sin( 2πy
L
) • ψ2,1 (x, y) = 2
L
sin( 2πx
L
) sin( πy
L
)
• quantum numbers: nx =1, ny =2 • quantum numbers: nx =2, ny =1

well-defined quantum numbers


same well-defined energy: E+ = E− = 5h2 /8mL2
overlap between the two states is zero: Sψ1,2 ,ψ2,1 = 0
observable properties other than energy will differ
(note differing wavefunctions and probability densities)
Slide 159 · QM principles and concepts · Superposition of wavefunctions
Example 1: states of a particle in a 2D box (2)

ψ1,2 state (“natural”) ψ2,1 state (“natural”)

• ψ1,2 (x, y) = 2
L
sin( πx
L
) sin( 2πy
L
) • ψ2,1 (x, y) = 2
L
sin( 2πx
L
) sin( πy
L
)
• quantum numbers: nx =1, ny =2 • quantum numbers: nx =2, ny =1

well-defined quantum numbers


same well-defined energy: E+ = E− = 5h2 /8mL2
overlap between the two states is zero: Sψ1,2 ,ψ2,1 = 0
observable properties other than energy will differ
(note differing wavefunctions and probability densities)
Slide 159 · QM principles and concepts · Superposition of wavefunctions
Example 1: states of a particle in a 2D box (2)

ψ1,2 state (“natural”) ψ2,1 state (“natural”)

• ψ1,2 (x, y) = 2
L
sin( πx
L
) sin( 2πy
L
) • ψ2,1 (x, y) = 2
L
sin( 2πx
L
) sin( πy
L
)
• quantum numbers: nx =1, ny =2 • quantum numbers: nx =2, ny =1

well-defined quantum numbers


same well-defined energy: E+ = E− = 5h2 /8mL2
overlap between the two states is zero: Sψ1,2 ,ψ2,1 = 0
observable properties other than energy will differ
(note differing wavefunctions and probability densities)
Slide 159 · QM principles and concepts · Superposition of wavefunctions
Example 1: states of a particle in a 2D box (2)

ψ1,2 state (“natural”) ψ2,1 state (“natural”)

• ψ1,2 (x, y) = 2
L
sin( πx
L
) sin( 2πy
L
) • ψ2,1 (x, y) = 2
L
sin( 2πx
L
) sin( πy
L
)
• quantum numbers: nx =1, ny =2 • quantum numbers: nx =2, ny =1

ψ |ψ|2 ψ |ψ|2

well-defined quantum numbers


same well-defined energy: E+ = E− = 5h2 /8mL2
overlap between the two states is zero: Sψ1,2 ,ψ2,1 = 0
observable properties other than energy will differ
(note differing wavefunctions and probability densities)
Slide 159 · QM principles and concepts · Superposition of wavefunctions
Example 1: states of a particle in a 2D box (3)

ψ+ state (“hybrid”) ψ− state (“hybrid”)

• ψ+ (x, y) = ψ2,1 (x, y) + ψ1,2 (x, y) • ψ− (x, y) = ψ2,1 (x, y) − ψ1,2 (x, y)
• quantum numbers: ? • quantum numbers: ?

quantum numbers not well defined


same well-defined energy: E1,2 = E2,1 = 5h2 /8mL2
overlap between the two states is zero: Sψ+ ,ψ− = 0
observable properties other than energy will differ
(note differing wavefunctions and probability densities)
Slide 160 · QM principles and concepts · Superposition of wavefunctions
Example 1: states of a particle in a 2D box (3)

ψ+ state (“hybrid”) ψ− state (“hybrid”)

• ψ+ (x, y) = ψ2,1 (x, y) + ψ1,2 (x, y) • ψ− (x, y) = ψ2,1 (x, y) − ψ1,2 (x, y)
• quantum numbers: ? • quantum numbers: ?

quantum numbers not well defined


same well-defined energy: E1,2 = E2,1 = 5h2 /8mL2
overlap between the two states is zero: Sψ+ ,ψ− = 0
observable properties other than energy will differ
(note differing wavefunctions and probability densities)
Slide 160 · QM principles and concepts · Superposition of wavefunctions
Example 1: states of a particle in a 2D box (3)

ψ+ state (“hybrid”) ψ− state (“hybrid”)

• ψ+ (x, y) = ψ2,1 (x, y) + ψ1,2 (x, y) • ψ− (x, y) = ψ2,1 (x, y) − ψ1,2 (x, y)
• quantum numbers: ? • quantum numbers: ?

quantum numbers not well defined


same well-defined energy: E1,2 = E2,1 = 5h2 /8mL2
overlap between the two states is zero: Sψ+ ,ψ− = 0
observable properties other than energy will differ
(note differing wavefunctions and probability densities)
Slide 160 · QM principles and concepts · Superposition of wavefunctions
Example 1: states of a particle in a 2D box (3)

ψ+ state (“hybrid”) ψ− state (“hybrid”)

• ψ+ (x, y) = ψ2,1 (x, y) + ψ1,2 (x, y) • ψ− (x, y) = ψ2,1 (x, y) − ψ1,2 (x, y)
• quantum numbers: ? • quantum numbers: ?

ψ |ψ|2 ψ |ψ|2

quantum numbers not well defined


same well-defined energy: E1,2 = E2,1 = 5h2 /8mL2
overlap between the two states is zero: Sψ+ ,ψ− = 0
observable properties other than energy will differ
(note differing wavefunctions and probability densities)
Slide 160 · QM principles and concepts · Superposition of wavefunctions
Example 1: states of a particle in a 2D box (4)

• wavefunctions seen
from above (contour
map representation)
ψ1,2 ψ+ • all states have
the same energy
• change in shapes
of wavefunctions
• change in shapes
and orientations of
probability densities
ψ2,1 ψ− • next time you hear
about hybrid electronic
orbitals, will you know
where they come from?

Slide 161 · QM principles and concepts · Superposition of wavefunctions


Example 1: states of a particle in a 2D box (4)

• wavefunctions seen
from above (contour
map representation)
ψ1,2 ψ+ • all states have
the same energy
• change in shapes
of wavefunctions
• change in shapes
and orientations of
probability densities
ψ2,1 ψ− • next time you hear
about hybrid electronic
orbitals, will you know
where they come from?

Slide 161 · QM principles and concepts · Superposition of wavefunctions


Example 1: states of a particle in a 2D box (4)

• wavefunctions seen
from above (contour
map representation)
ψ1,2 ψ+ • all states have
the same energy
• change in shapes
of wavefunctions
• change in shapes
and orientations of
probability densities
ψ2,1 ψ− • next time you hear
about hybrid electronic
orbitals, will you know
where they come from?

Slide 161 · QM principles and concepts · Superposition of wavefunctions


Example 1: states of a particle in a 2D box (4)

• wavefunctions seen
from above (contour
map representation)
ψ1,2 ψ+ • all states have
the same energy
• change in shapes
of wavefunctions
• change in shapes
and orientations of
probability densities
ψ2,1 ψ− • next time you hear
about hybrid electronic
orbitals, will you know
where they come from?

Slide 161 · QM principles and concepts · Superposition of wavefunctions


Example 1: states of a particle in a 2D box (4)

• wavefunctions seen
from above (contour
map representation)
ψ1,2 ψ+ • all states have
the same energy
• change in shapes
of wavefunctions
• change in shapes
and orientations of
probability densities
ψ2,1 ψ− • next time you hear
about hybrid electronic
orbitals, will you know
where they come from?

Slide 161 · QM principles and concepts · Superposition of wavefunctions


The previous example... and the next one
In the previous example, we superposed degenerate wavefunctions.
The resulting (hybrid) wavefunctions had a well-defined energy.

Slide 162 · QM principles and concepts · Superposition of wavefunctions


The previous example... and the next one
In the previous example, we superposed degenerate wavefunctions.
The resulting (hybrid) wavefunctions had a well-defined energy.
Having a well-defined energy, the wavefunctions generated
by the superposition were time-independent.

Slide 162 · QM principles and concepts · Superposition of wavefunctions


The previous example... and the next one
In the previous example, we superposed degenerate wavefunctions.
The resulting (hybrid) wavefunctions had a well-defined energy.
Having a well-defined energy, the wavefunctions generated
by the superposition were time-independent.

superpositions of degenerate TI wavefunctions with energy E


lead to degenerate TI wavefunctions with energy E .

Slide 162 · QM principles and concepts · Superposition of wavefunctions


The previous example... and the next one
In the previous example, we superposed degenerate wavefunctions.
The resulting (hybrid) wavefunctions had a well-defined energy.
Having a well-defined energy, the wavefunctions generated
by the superposition were time-independent.

superpositions of degenerate TI wavefunctions with energy E


lead to degenerate TI wavefunctions with energy E .

Now we will superpose nondegenerate wavefunctions. The resulting


wavefunctions (wavepackets) will not have a well-defined energy.

Slide 162 · QM principles and concepts · Superposition of wavefunctions


The previous example... and the next one
In the previous example, we superposed degenerate wavefunctions.
The resulting (hybrid) wavefunctions had a well-defined energy.
Having a well-defined energy, the wavefunctions generated
by the superposition were time-independent.

superpositions of degenerate TI wavefunctions with energy E


lead to degenerate TI wavefunctions with energy E .

Now we will superpose nondegenerate wavefunctions. The resulting


wavefunctions (wavepackets) will not have a well-defined energy.
Not having a well-defined energy, these wavepackets will change with
time. They will be TD wavefunctions instead of TI ones.

Slide 162 · QM principles and concepts · Superposition of wavefunctions


The previous example... and the next one
In the previous example, we superposed degenerate wavefunctions.
The resulting (hybrid) wavefunctions had a well-defined energy.
Having a well-defined energy, the wavefunctions generated
by the superposition were time-independent.

superpositions of degenerate TI wavefunctions with energy E


lead to degenerate TI wavefunctions with energy E .

Now we will superpose nondegenerate wavefunctions. The resulting


wavefunctions (wavepackets) will not have a well-defined energy.
Not having a well-defined energy, these wavepackets will change with
time. They will be TD wavefunctions instead of TI ones.

superpositions of nondegenerate TI wavefunctions


lead to TD wavefunctions.

Slide 162 · QM principles and concepts · Superposition of wavefunctions


Example 2: 1D wavepacket (1)
Our system is an H atom in a 1D box of length L = 10 nm. Its energy
levels, TI wavefunctions and total wavefunctions are given by

n2 h2
r
2  nπx   E 
En = , ψn (x) = sin , Ψn (x, t) = exp −i t ψn (x).
8mL2 L L h

The superposition of total wavefunctions results in a wavepacket.


According to Atkins’ Quanta, a wavepacket is...

a superposition of wavefunctions that is usually strongly


peaked in one region of space and virtually zero elsewhere.
The peak of the wavepacket denotes the most likely location
of the particle; it occurs where the contributing
wavefunctions are in phase and interfere constructively.
Elsewhere the wavefunctions interfere destructively, and
the net probability density is small or zero.
Slide 163 · QM principles and concepts · Superposition of wavefunctions
Example 2: 1D wavepacket (2)

As the various components of a wavepacket have different energies,


their phases change with time at different rates.

The way they interfere with each other changes with time,
and so does the wavepacket shape.

From Atkins’ Quanta:

A wavepacket moves because all the component wavefunctions


change at different rates, and at different times the point
of maximum constructive interference is in different
locations. An important difference from classical physics
is that the wavepacket spreads with time.

Slide 164 · QM principles and concepts · Superposition of wavefunctions


Example 2: 1D wavepacket (3)
The expression used to generate the wavepacket we will look at was

X
Φ(x, t) = c1 Ψ1 (x, t) + c2 Ψ2 (x, t) + c3 Ψ3 (x, t) + · · · = cn Ψn (x, t),
n=1

where the expansion coefficients (the cn ’s) are the numbers that
quantify the contribution of each total wavefunction to the wavepacket.

|cn |2
The expansion coefficients
were chosen so that we get a
“Gaussian” wavepacket.

As for the wavepacket itself...


−21
En /10 J

Slide 165 · QM principles and concepts · Superposition of wavefunctions


Example 2: 1D wavepacket (4)

Probability density (squared modulus of wavepacket).


As time passes, H atom moves and wavepacket spreads.

Slide 166 · QM principles and concepts · Superposition of wavefunctions


Time-energy uncertainty relation
The time dependence associated with a mixture of different energies
has to do with the

time-energy uncertainty relation: τ∆E ≈ ħ,

where τ is the lifetime — a number with dimensions of time, whose


inverse describes how quickly a system changes with time.

Slide 167 · QM principles and concepts · Superposition of wavefunctions


Time-energy uncertainty relation
The time dependence associated with a mixture of different energies
has to do with the

time-energy uncertainty relation: τ∆E ≈ ħ,

where τ is the lifetime — a number with dimensions of time, whose


inverse describes how quickly a system changes with time. Note:
• If the energy uncertainty is very small, approaching zero, then
τ approaches infinity, which indicates no change with time.

Slide 167 · QM principles and concepts · Superposition of wavefunctions


Time-energy uncertainty relation
The time dependence associated with a mixture of different energies
has to do with the

time-energy uncertainty relation: τ∆E ≈ ħ,

where τ is the lifetime — a number with dimensions of time, whose


inverse describes how quickly a system changes with time. Note:
• If the energy uncertainty is very small, approaching zero, then
τ approaches infinity, which indicates no change with time.
I If the energy is well defined, the properties of the system do not
change with time; we are dealing with a TI state.

Slide 167 · QM principles and concepts · Superposition of wavefunctions


Time-energy uncertainty relation
The time dependence associated with a mixture of different energies
has to do with the

time-energy uncertainty relation: τ∆E ≈ ħ,

where τ is the lifetime — a number with dimensions of time, whose


inverse describes how quickly a system changes with time. Note:

• If the energy uncertainty is finite (i.e., not zero), then τ is also


finite, which indicates that observable things change with time.

Slide 167 · QM principles and concepts · Superposition of wavefunctions


Time-energy uncertainty relation
The time dependence associated with a mixture of different energies
has to do with the

time-energy uncertainty relation: τ∆E ≈ ħ,

where τ is the lifetime — a number with dimensions of time, whose


inverse describes how quickly a system changes with time. Note:

• If the energy uncertainty is finite (i.e., not zero), then τ is also


finite, which indicates that observable things change with time.
I If the energy is not well defined, the properties of the system
change with time; we are dealing with a TD state.

Slide 167 · QM principles and concepts · Superposition of wavefunctions


Summary: superpositions of wavefunctions

• By superposing wavefunctions of simple systems, we can


obtain approximations to the wavefunctions of complex systems.
• In superpositions of degenerate wavefunctions the energy is well
defined. Just like its components, a hybrid wavefunction describes
a TI state whose observable properties do not change with time.
• In superpositions of nondegenerate wavefunctions the energy is
not well defined. Unlike its components, a wavepacket describes
a TD state whose observable properties change with time.
• While they can be initially localised, wavepackets spread as time
passes. This leads to loss of the initial localisation.
• The time dependence associated with a mixture of different
energies is quantified by the time-energy uncertainty relation.
• The inverse of the lifetime, 1/τ, describes how quickly
the observable properties of a system change with time.

Slide 168 · QM principles and concepts · Superposition of wavefunctions


Summary: superpositions of wavefunctions

• By superposing wavefunctions of simple systems, we can


obtain approximations to the wavefunctions of complex systems.
• In superpositions of degenerate wavefunctions the energy is well
defined. Just like its components, a hybrid wavefunction describes
a TI state whose observable properties do not change with time.
• In superpositions of nondegenerate wavefunctions the energy is
not well defined. Unlike its components, a wavepacket describes
a TD state whose observable properties change with time.
• While they can be initially localised, wavepackets spread as time
passes. This leads to loss of the initial localisation.
• The time dependence associated with a mixture of different
energies is quantified by the time-energy uncertainty relation.
• The inverse of the lifetime, 1/τ, describes how quickly
the observable properties of a system change with time.

Slide 168 · QM principles and concepts · Superposition of wavefunctions


Summary: superpositions of wavefunctions

• By superposing wavefunctions of simple systems, we can


obtain approximations to the wavefunctions of complex systems.
• In superpositions of degenerate wavefunctions the energy is well
defined. Just like its components, a hybrid wavefunction describes
a TI state whose observable properties do not change with time.
• In superpositions of nondegenerate wavefunctions the energy is
not well defined. Unlike its components, a wavepacket describes
a TD state whose observable properties change with time.
• While they can be initially localised, wavepackets spread as time
passes. This leads to loss of the initial localisation.
• The time dependence associated with a mixture of different
energies is quantified by the time-energy uncertainty relation.
• The inverse of the lifetime, 1/τ, describes how quickly
the observable properties of a system change with time.

Slide 168 · QM principles and concepts · Superposition of wavefunctions


Summary: superpositions of wavefunctions

• By superposing wavefunctions of simple systems, we can


obtain approximations to the wavefunctions of complex systems.
• In superpositions of degenerate wavefunctions the energy is well
defined. Just like its components, a hybrid wavefunction describes
a TI state whose observable properties do not change with time.
• In superpositions of nondegenerate wavefunctions the energy is
not well defined. Unlike its components, a wavepacket describes
a TD state whose observable properties change with time.
• While they can be initially localised, wavepackets spread as time
passes. This leads to loss of the initial localisation.
• The time dependence associated with a mixture of different
energies is quantified by the time-energy uncertainty relation.
• The inverse of the lifetime, 1/τ, describes how quickly
the observable properties of a system change with time.

Slide 168 · QM principles and concepts · Superposition of wavefunctions


Summary: superpositions of wavefunctions

• By superposing wavefunctions of simple systems, we can


obtain approximations to the wavefunctions of complex systems.
• In superpositions of degenerate wavefunctions the energy is well
defined. Just like its components, a hybrid wavefunction describes
a TI state whose observable properties do not change with time.
• In superpositions of nondegenerate wavefunctions the energy is
not well defined. Unlike its components, a wavepacket describes
a TD state whose observable properties change with time.
• While they can be initially localised, wavepackets spread as time
passes. This leads to loss of the initial localisation.
• The time dependence associated with a mixture of different
energies is quantified by the time-energy uncertainty relation.
• The inverse of the lifetime, 1/τ, describes how quickly
the observable properties of a system change with time.

Slide 168 · QM principles and concepts · Superposition of wavefunctions


Summary: superpositions of wavefunctions

• By superposing wavefunctions of simple systems, we can


obtain approximations to the wavefunctions of complex systems.
• In superpositions of degenerate wavefunctions the energy is well
defined. Just like its components, a hybrid wavefunction describes
a TI state whose observable properties do not change with time.
• In superpositions of nondegenerate wavefunctions the energy is
not well defined. Unlike its components, a wavepacket describes
a TD state whose observable properties change with time.
• While they can be initially localised, wavepackets spread as time
passes. This leads to loss of the initial localisation.
• The time dependence associated with a mixture of different
energies is quantified by the time-energy uncertainty relation.
• The inverse of the lifetime, 1/τ, describes how quickly
the observable properties of a system change with time.

Slide 168 · QM principles and concepts · Superposition of wavefunctions


End of the conceptual part. What’s next?

Slide 169 · QM principles and concepts · Superposition of wavefunctions


Part II

The palette of quantum chemists:


models and approximations

Slide 170 · Models and approximations


What we’re after...

Things about the motion or internal structure of atoms and molecules


that are important for chemistry:
• translation (how atoms and molecules move from here to there);
• vibration (how atoms in molecules vibrate, stretching and
compressing chemical bonds and/or changing bond angles);
• rotation (how molecules rotate, how electrons move around nuclei,
spin);
• electronic structure (how electrons arrange themselves in atomic
and molecular orbitals, how they form chemical bonds).

Slide 171 · Models and approximations


...and how we’ll get there

We’ll describe those things using simple models.

These models allow us to understand lots of things about why


atoms and molecules are the way they are, and why they behave
the way they behave.

The models are so good, they even allow us to get quantitative details
about atoms and molecules.

And when the models aren’t good enough, we can combine


(superpose) their wavefunctions in order to obtain good
approximations for the wavefunctions of real, complex systems.

Slide 172 · Models and approximations


Part II — Models and approximations

11 Molecular translation
The model for 12 Molecular vibration
molecular translation: 13 Molecular rotation

14 Electronic structure of H atom


particle in a 3-D box
15 Spin

Slide 173 · Models and approximations · Molecular translation


Molecular translation

Let us consider the translation of molecules in a gas. That’s when they


are free to move in all three directions in space.
If our gas is in a cubic container of dimensions Lx , Ly and Lz ,
calculating the molecular energies is easy.
The molecules are particles in a 3D box.
Model is good enough to describe translational motion of molecules
not only in a cubic box, but also in macroscopic boxes of other shapes.
It’s not good enough for translation in liquids. That’s a more
difficult problem we won’t be looking at.

Slide 174 · Models and approximations · Molecular translation


Molecules in a 3D box

potential: V = ∞
• energy is purely kinetic: no
V =0 intermolecular interactions.
Free translation inside box.
• translations along x, y, z
independent of each other:
Ly E = Ex + Ey + Ez ,
ψ(x, y, z) = ψ(x)ψ(y)ψ(z).
• wavefunction must be zero
at wall and outside the box.
• for more details, see
Lx Energy Matters material.

Slide 175 · Models and approximations · Molecular translation


Part II — Models and approximations

11 Molecular translation
The model for 12 Molecular vibration
molecular vibration: 13 Molecular rotation

14 Electronic structure of H atom


harmonic oscillator 15 Spin

Slide 176 · Models and approximations · Molecular vibration


Molecular vibration

Let us consider the vibrations of molecules. A molecule with N atoms


has 3N − 6 vibrational modes (or vibrational degrees of freedom)
if it’s not linear, and 3N − 5 if it is.
Roughly speaking, each vibrational mode is a different way
in which the molecule can (and will) vibrate
around its equilibrium geometry.
Note that molecules will not vibrate in only one of those ways.
Instead, they will vibrate in all modes simultaneously.
The overall molecular vibration will be the result of a
superposition of different vibrational modes.

Slide 177 · Models and approximations · Molecular vibration


Vibrational modes of SO2 and CO2

In our website you’ll find a link to a Journal of Chemical Education


article with animations showing the vibrational modes of SO2 and CO2 .
SO2 is nonlinear, and has three modes of vibration: the symmetric
bond stretch, the asymmetric bond stretch and the bending
vibration.
CO2 is linear, and has four modes of vibration: symmetric and
asymmetric bond stretches as well as a pair of perpendicular
bending vibrations.
The animations show individual modes of vibration and one
possible superposition of them.
Remember, however, that it is impossible to find a molecule vibrating
along only one of its modes of vibration.

Slide 178 · Models and approximations · Molecular vibration


Vibration of a diatomic (1)

Potential energy curve, V (r): potential


De energy (that is, energy from interaction
between the two atoms) as a function of
the internuclear distance, r.
vibrational energy

• V isn’t the vibrational energy.


• No interaction when r = ∞,
V (r) but note that location of the zero
of energy is arbitrary.
• Equilibrium bond length, req :
the internuclear distance around
0
which the diatomic vibrates.
0 1 2 3 4 5 6 7 8 9 • Potential is attractive at r > req
internuclear distance and repulsive at r < req .

Slide 179 · Models and approximations · Molecular vibration


Vibration of a diatomic (1)

Potential energy curve, V (r): potential


De energy (that is, energy from interaction
between the two atoms) as a function of
the internuclear distance, r.
vibrational energy

• V isn’t the vibrational energy.


• No interaction when r = ∞,
but note that location of the zero
of energy is arbitrary.
• Equilibrium bond length, req :
the internuclear distance around
0
which the diatomic vibrates.
0 1 2 3 4 5 6 7 8 9 • Potential is attractive at r > req
internuclear distance and repulsive at r < req .

Slide 179 · Models and approximations · Molecular vibration


Vibration of a diatomic (1)

Potential energy curve, V (r): potential


De req energy (that is, energy from interaction
between the two atoms) as a function of
the internuclear distance, r.
vibrational energy

• V isn’t the vibrational energy.


• No interaction when r = ∞,
but note that location of the zero
of energy is arbitrary.
• Equilibrium bond length, req :
the internuclear distance around
0
which the diatomic vibrates.
0 1 2 3 4 5 6 7 8 9 • Potential is attractive at r > req
internuclear distance and repulsive at r < req .

Slide 179 · Models and approximations · Molecular vibration


Vibration of a diatomic (1)

Potential energy curve, V (r): potential


De energy (that is, energy from interaction
between the two atoms) as a function of
the internuclear distance, r.
vibrational energy

• V isn’t the vibrational energy.


• No interaction when r = ∞,
but note that location of the zero
of energy is arbitrary.
• Equilibrium bond length, req :
the internuclear distance around
0
which the diatomic vibrates.
0 1 2 3 4 5 6 7 8 9 • Potential is attractive at r > req
internuclear distance and repulsive at r < req .

Slide 179 · Models and approximations · Molecular vibration


Vibration of a diatomic (2)

De
• Minimum energy, Emin : the energy the
diatomic would have if fixed at r = req .
vibrational energy

• Vibrational energy, Ev : the total


energy of a vibrational level,
E =K +V.
• Classical turning points: where a
classical molecule would reach zero
kinetic energy and reverse its
0 Emin
vibration. Appear at internuclear
0 1 2 3 4 5 6 7 8 9 distances at which V (r) = Ev .
internuclear distance

Slide 180 · Models and approximations · Molecular vibration


Vibration of a diatomic (2)

De
• Minimum energy, Emin : the energy the
diatomic would have if fixed at r = req .
vibrational energy

• Vibrational energy, Ev : the total


energy of a vibrational level,
Ev
E =K +V.
• Classical turning points: where a
classical molecule would reach zero
kinetic energy and reverse its
0
vibration. Appear at internuclear
0 1 2 3 4 5 6 7 8 9 distances at which V (r) = Ev .
internuclear distance

Slide 180 · Models and approximations · Molecular vibration


Vibration of a diatomic (2)

inner
De turning
point • Minimum energy, Emin : the energy the
outer
turning diatomic would have if fixed at r = req .
vibrational energy

point
• Vibrational energy, Ev : the total
energy of a vibrational level,
E =K +V.
• Classical turning points: where a
classical molecule would reach zero
kinetic energy and reverse its
0
vibration. Appear at internuclear
0 1 2 3 4 5 6 7 8 9 distances at which V (r) = Ev .
internuclear distance

Slide 180 · Models and approximations · Molecular vibration


Vibration of a diatomic (3)

De dissociation limit

• Dissociation: there’s an upper limit


vibrational energy

for the vibrational energy. At higher


energies the molecule dissociates.
• De : the difference between the
energy at which the molecule
dissociates and Emin .
• The potential is steeper at r < req
0 than at r > req . It’s easier to stretch a
chemical bond than to compress it.
0 1 2 3 4 5 6 7 8 9
internuclear distance

Slide 181 · Models and approximations · Molecular vibration


Vibration of a diatomic (3)

De

• Dissociation: there’s an upper limit


vibrational energy

for the vibrational energy. At higher


energies the molecule dissociates.
De • De : the difference between the
energy at which the molecule
dissociates and Emin .
• The potential is steeper at r < req
0 than at r > req . It’s easier to stretch a
chemical bond than to compress it.
0 1 2 3 4 5 6 7 8 9
internuclear distance

Slide 181 · Models and approximations · Molecular vibration


Vibration of a diatomic (3)

De

• Dissociation: there’s an upper limit


ep
steeper
vibrational energy

ste

for the vibrational energy. At higher


so

energies the molecule dissociates.


not

• De : the difference between the


energy at which the molecule
dissociates and Emin .
• The potential is steeper at r < req
0 than at r > req . It’s easier to stretch a
chemical bond than to compress it.
0 1 2 3 4 5 6 7 8 9
internuclear distance

Slide 181 · Models and approximations · Molecular vibration


Vibration of a diatomic (4)

De v= = 23
v 22
v==2021
v v=1819
vv==17
v = 16
v = 15
• vibrational energy is quantised. The
vibrational energy

v = 14
v = 13
v = 12 quantum number is v = 0, 1, 2, 3, . . .
v = 11
v = 10
v =9 • spacing between vibrational energy
v =8
v =7 levels decreases with increasing
v =6
v =5 vibrational energy.
v =4
v =3 • it’s impossible to stop molecular
v =2
v =1 vibration: lowest-energy state has
v =0
0 zero-point vibrational energy,
ZPE = Ev=0 − Emin .
0 1 2 3 4 5 6 7 8 9
internuclear distance

Slide 182 · Models and approximations · Molecular vibration


Vibration of a diatomic (4)

De v= = 23
v 22
v==2021
v v=1819
vv==17
v = 16
v = 15
• vibrational energy is quantised. The
vibrational energy

v = 14
v = 13
v = 12 quantum number is v = 0, 1, 2, 3, . . .
v = 11
v = 10
v =9 • spacing between vibrational energy
v =8
v =7 levels decreases with increasing
v =6
v =5 vibrational energy.
v =4
v =3 • it’s impossible to stop molecular
v =2
v =1 vibration: lowest-energy state has
v =0
0 zero-point vibrational energy,
ZPE = Ev=0 − Emin .
0 1 2 3 4 5 6 7 8 9
internuclear distance

Slide 182 · Models and approximations · Molecular vibration


Vibration of a diatomic (4)

De v= = 23
v 22
v==2021
v v=1819
vv==17
v = 16
v = 15
• vibrational energy is quantised. The
vibrational energy

v = 14
v = 13
v = 12 quantum number is v = 0, 1, 2, 3, . . .
v = 11
v = 10
v =9 • spacing between vibrational energy
v =8
v =7 levels decreases with increasing
v =6
v =5 vibrational energy.
v =4
v =3 • it’s impossible to stop molecular
v =2
v =1 vibration: lowest-energy state has
v =0 zero-point
0 energy zero-point vibrational energy,
ZPE = Ev=0 − Emin .
0 1 2 3 4 5 6 7 8 9
internuclear distance

Slide 182 · Models and approximations · Molecular vibration


Harmonic oscillator

• restoring force proportional to


displacement from equilibrium:
v =6
F = −kf ∆r, where:
vibrational energy

v =5 I kf is the force constant. It


v =4
quantifies the stiffness of the bond.
I ∆r = r − req is the displacement
v =3
from the equilibrium internuclear
v =2 distance.
v =1 • parabolic potential: V = 12 kf (∆r)2 .
v =0
• evenly spaced
vibrational energy levels.

Slide 183 · Models and approximations · Molecular vibration


Harmonic oscillator

• restoring force proportional to


displacement from equilibrium:
v =6
F = −kf ∆r, where:
vibrational energy

v =5 I kf is the force constant. It


v =4
quantifies the stiffness of the bond.
I ∆r = r − req is the displacement
v =3
from the equilibrium internuclear
v =2 distance.
v =1 • parabolic potential: V = 12 kf (∆r)2 .
v =0
• evenly spaced
vibrational energy levels.

Slide 183 · Models and approximations · Molecular vibration


Harmonic oscillator

• restoring force proportional to


displacement from equilibrium:
v =6
F = −kf ∆r, where:
vibrational energy

v =5 I kf is the force constant. It


v =4
quantifies the stiffness of the bond.
I ∆r = r − req is the displacement
v =3
from the equilibrium internuclear
v =2 distance.
v =1 • parabolic potential: V = 12 kf (∆r)2 .
v =0
• evenly spaced
vibrational energy levels.

Slide 183 · Models and approximations · Molecular vibration


Harmonic v. anharmonic oscillator

De
vibrational energy

Harmonic potential:
• not steep enough at r < req ,
too steep at r > req
• does not allow for dissociation
• good model for low vibrational levels,
poor model for other vibrational levels
0

0 1 2 3 4 5 6 7 8 9
internuclear distance

Slide 184 · Models and approximations · Molecular vibration


Harmonic v. anharmonic oscillator

De
vibrational energy

Harmonic potential:
• not steep enough at r < req ,
too steep at r > req
• does not allow for dissociation
• good model for low vibrational levels,
poor model for other vibrational levels
0

0 1 2 3 4 5 6 7 8 9
internuclear distance

Slide 184 · Models and approximations · Molecular vibration


Harmonic v. anharmonic oscillator

De
vibrational energy

Harmonic potential:
• not steep enough at r < req ,
too steep at r > req
• does not allow for dissociation
• good model for low vibrational levels,
poor model for other vibrational levels
0

0 1 2 3 4 5 6 7 8 9
internuclear distance

Slide 184 · Models and approximations · Molecular vibration


Harmonic v. anharmonic oscillator

De
vibrational energy

Harmonic potential:
• not steep enough at r < req ,
too steep at r > req
• does not allow for dissociation
• good model for low vibrational levels,
poor model for other vibrational levels
0

0 1 2 3 4 5 6 7 8 9
internuclear distance

Slide 184 · Models and approximations · Molecular vibration


Energy levels of harmonic oscillator
It is possible to solve the TI Schrödinger equation for a harmonic
oscillator in an exact and analytical way.

The energy levels of a harmonic oscillator are given by

 1  1  1
Ev = v + ħω = v + hν = v + hcν̃, v = 0, 1, 2, . . . ,
2 2 2

where:
• v is the vibrational quantum number
• ν is the vibrational frequency
(the number of complete oscillations per second)
• ω is the angular frequency (the number of radians per second,
each complete oscillation corresponding to 2π radians), ω = 2πν
• ν̃ the vibrational wavenumber, ν̃ = ν/c

Slide 185 · Models and approximations · Molecular vibration


What vibrational energies tell us about molecules
The angular frequency of a harmonic oscillator is related to its force
constant by q
ω= kf /µ,
where µ is the reduced mass of the diatomic, related to the masses of
the two atoms (say, atom A and atom B) by µ = mA mB /[mA + mB ].

Using vibrational spectroscopy — that is, measuring the gaps between


various vibrational energy levels — we can learn about chemical
bonding and molecular structure.
Vibrational spectra allow us to determine force constants and
reduced masses. In other words, to figure out the stiffness of
chemical bonds and the isotopic compositions of molecules.
And if we know about the bonds in a given molecule, we can use
vibrational spectroscopy as a detection tool.
Slide 186 · Models and approximations · Molecular vibration
What vibrational energies tell us about molecules
The angular frequency of a harmonic oscillator is related to its force
constant by q
ω= kf /µ,
where µ is the reduced mass of the diatomic, related to the masses of
the two atoms (say, atom A and atom B) by µ = mA mB /[mA + mB ].

Using vibrational spectroscopy — that is, measuring the gaps between


various vibrational energy levels — we can learn about chemical
bonding and molecular structure.
Vibrational spectra allow us to determine force constants and
reduced masses. In other words, to figure out the stiffness of
chemical bonds and the isotopic compositions of molecules.
And if we know about the bonds in a given molecule, we can use
vibrational spectroscopy as a detection tool.
Slide 186 · Models and approximations · Molecular vibration
Things we can do with vibrational spectroscopy

• figure out how stiff a given chemical bond is;

• identify molecules in a mixture of many;

• identify the structure of complex molecules,


including their isotopic composition;

• characterize molecules that are difficult to isolate;

• figure out how anharmonic a given vibration is;

• learn about the potential energy and


the dissociation limits of molecules;

• learn about how the environment around a given molecule


affects its chemical bonds.

Slide 187 · Models and approximations · Molecular vibration


Wavefunctions of harmonic oscillator

Solution of the Schrödinger equation for a harmonic oscillator leads not


only to vibrational energies, it also leads to vibrational wavefunctions.

You’ll see the explicit expression for the vibrational wavefunctions


in the Physical Chemistry lab. We won’t consider them here.

We will, however, look at what they look like.

We will pay attention to one thing in particular: tunneling,


a classically-forbidden phenomenon that is very important
in chemistry.

Slide 188 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (1)

• kf = 527 N m−1
vibrational energy

• µ = 0.5mH
• ν̃ = 4230 cm−1
• Ev = v + 1

2
hcν̃
• ∆E = Ev+1 −Ev = hcν̃ = 8.4×10−20 J
∆E
• ZPE = 2
= 4.2 × 10−20 J

Slide 189 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (1)

• kf = 527 N m−1
vibrational energy

• µ = 0.5mH
• ν̃ = 4230 cm−1
• Ev = v + 1

2
hcν̃
• ∆E = Ev+1 −Ev = hcν̃ = 8.4×10−20 J
∆E
• ZPE = 2
= 4.2 × 10−20 J

Slide 189 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (1)

• kf = 527 N m−1
vibrational energy

• µ = 0.5mH
• ν̃ = 4230 cm−1
• Ev = v + 1

2
hcν̃
• ∆E = Ev+1 −Ev = hcν̃ = 8.4×10−20 J
∆E
• ZPE = 2
= 4.2 × 10−20 J

Slide 189 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (1)

v =3 • kf = 527 N m−1
vibrational energy

• µ = 0.5mH
v =2 • ν̃ = 4230 cm−1
• Ev = v + 1

2
hcν̃
v =1 • ∆E = Ev+1 −Ev = hcν̃ = 8.4×10−20 J
∆E
• ZPE = 2
= 4.2 × 10−20 J
v =0

Slide 189 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (1)

v =3 • kf = 527 N m−1
vibrational energy

∆E • µ = 0.5mH
v =2 • ν̃ = 4230 cm−1
• Ev = v + 1

∆E 2
hcν̃
v =1 • ∆E = Ev+1 −Ev = hcν̃ = 8.4×10−20 J
∆E
∆E • ZPE = 2
= 4.2 × 10−20 J
v =0

Slide 189 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (1)

v =3 • kf = 527 N m−1
vibrational energy

∆E • µ = 0.5mH
v =2 • ν̃ = 4230 cm−1
• Ev = v + 1

∆E 2
hcν̃
v =1 • ∆E = Ev+1 −Ev = hcν̃ = 8.4×10−20 J
∆E
∆E • ZPE = 2
= 4.2 × 10−20 J
v =0
ZPE

Slide 189 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (2)

• H2 (v = 0) state
• no nodes
• |ψ|2 maximum at r = re
• tunneling: molecule can be
found in classically-forbidden
regions (probability is 15.7%)

Slide 190 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (2)

• H2 (v = 0) state
• no nodes
• |ψ|2 maximum at r = re
• tunneling: molecule can be
found in classically-forbidden
regions (probability is 15.7%)

Slide 190 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (2)

ψ0 (r) 2
|ψ0 (r)|

• H2 (v = 0) state
• no nodes
• |ψ|2 maximum at r = re
• tunneling: molecule can be
found in classically-forbidden
regions (probability is 15.7%)

Slide 190 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (2)

ψ0 (r) 2
|ψ0 (r)|

• H2 (v = 0) state
• no nodes
• |ψ|2 maximum at r = re
• tunneling: molecule can be
found in classically-forbidden
regions (probability is 15.7%)

Slide 190 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (2)

ψ0 (r) 2
|ψ0 (r)|

• H2 (v = 0) state
• no nodes
• |ψ|2 maximum at r = re
• tunneling: molecule can be
tunneling
found in classically-forbidden
regions (probability is 15.7%)

Slide 190 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (3)

• H2 (v = 1) state
• 1 node
• |ψ|2 maxima shifted toward
turning points
• tunneling: probability of finding
molecule at classically-
forbidden regions decreases
(now 11.2%)

Slide 191 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (3)

ψ1 (r) |ψ1 (r)|2


• H2 (v = 1) state
• 1 node
• |ψ|2 maxima shifted toward
turning points
• tunneling: probability of finding
molecule at classically-
forbidden regions decreases
(now 11.2%)

Slide 191 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (3)

ψ1 (r) |ψ1 (r)|2


• H2 (v = 1) state
• 1 node
• |ψ|2 maxima shifted toward
turning points
• tunneling: probability of finding
molecule at classically-
forbidden regions decreases
(now 11.2%)

Slide 191 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (3)

ψ1 (r) |ψ1 (r)|2


• H2 (v = 1) state
• 1 node
• |ψ|2 maxima shifted toward
turning points
• tunneling: probability of finding
molecule at classically-
forbidden regions decreases
(now 11.2%)
tunneling

Slide 191 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (4)

• H2 (v = 2) state
• 2 nodes
• |ψ|2 maxima further shifted
toward turning points
• tunneling: probability of finding
molecule at classically-
forbidden regions decreases
further (now 9.5%)

Slide 192 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (4)

ψ2 (r) |ψ2 (r)|2


• H2 (v = 2) state
• 2 nodes
• |ψ|2 maxima further shifted
toward turning points
• tunneling: probability of finding
molecule at classically-
forbidden regions decreases
further (now 9.5%)

Slide 192 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (4)

ψ2 (r) |ψ2 (r)|2


• H2 (v = 2) state
• 2 nodes
• |ψ|2 maxima further shifted
toward turning points
• tunneling: probability of finding
molecule at classically-
forbidden regions decreases
further (now 9.5%)

Slide 192 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (4)

ψ2 (r) |ψ2 (r)|2


• H2 (v = 2) state
• 2 nodes
• |ψ|2 maxima further shifted
toward turning points
• tunneling: probability of finding
molecule at classically-
tunneling forbidden regions decreases
further (now 9.5%)

Slide 192 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (5)

• H2 (v = 3) state
• 3 nodes
• |ψ|2 maxima even further
shifted toward turning points
• tunneling: probability of finding
molecule at classically-
forbidden regions decreases
even further (now 8.5%)

Slide 193 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (5)

ψ3 (r) 2
|ψ3 (r)|
• H2 (v = 3) state
• 3 nodes
• |ψ|2 maxima even further
shifted toward turning points
• tunneling: probability of finding
molecule at classically-
forbidden regions decreases
even further (now 8.5%)

Slide 193 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (5)

ψ3 (r) 2
|ψ3 (r)|
• H2 (v = 3) state
• 3 nodes
• |ψ|2 maxima even further
shifted toward turning points
• tunneling: probability of finding
molecule at classically-
forbidden regions decreases
even further (now 8.5%)

Slide 193 · Models and approximations · Molecular vibration


Wavefunctions and probability densities of H2 (5)

ψ3 (r) 2
|ψ3 (r)|
• H2 (v = 3) state
• 3 nodes
• |ψ|2 maxima even further
shifted toward turning points
tunneling • tunneling: probability of finding
molecule at classically-
forbidden regions decreases
even further (now 8.5%)

Slide 193 · Models and approximations · Molecular vibration


Isotope effect: H2 v D2

Does replacement of H by D affect molecular vibration?

property H2 D2 changed?
−10
• kf and re values do not
re /10 m 0.741 0.741 no
change: consistent with the
kf / N m−1 527 527 no fact that V (r) does not depend
µ/mu 1/2 1 yes
on nuclear mass.
• µ value does change, and as a
ν̃/ cm−1 4230 2990 yes
consequence so do vibrational
Ev+1 −Ev −1
hc
/ cm 4230 2990 yes frequency, interlevel spacings
E0
/ cm
−1
2115 1495 yes and zero-point energy.
hc

Slide 194 · Models and approximations · Molecular vibration


Vibration of polyatomics
A nonlinear polyatomic molecule containing N atoms has 3N − 6
vibrational degrees of freedom.
In the case of low-lying vibrational states, each of the 3N − 6
vibrations is approximately harmonic.
Furthermore, they are, to a good approximation, independent of each
other. In other words, the 3N − 6 harmonic oscillations are uncoupled.
(Approximation ok for only for low-lying vibrational states.)

The uncoupled harmonic oscillations will involve combinations of


several types of vibrations:
• stretching/compressing of chemical bonds;
• angular vibrations;
• torsion of chemical bonds.
The SO2 /CO2 animations seen earlier illustrate the first two cases.
Slide 195 · Models and approximations · Molecular vibration
Vibration of polyatomics
A nonlinear polyatomic molecule containing N atoms has 3N − 6
vibrational degrees of freedom.
In the case of low-lying vibrational states, each of the 3N − 6
vibrations is approximately harmonic.
Furthermore, they are, to a good approximation, independent of each
other. In other words, the 3N − 6 harmonic oscillations are uncoupled.
(Approximation ok for only for low-lying vibrational states.)

The uncoupled harmonic oscillations will involve combinations of


several types of vibrations:
• stretching/compressing of chemical bonds;
• angular vibrations;
• torsion of chemical bonds.
The SO2 /CO2 animations seen earlier illustrate the first two cases.
Slide 195 · Models and approximations · Molecular vibration
Normal modes of vibration
The uncoupled, harmonic modes of vibration of a polyatomic molecule
are known as its normal modes of vibration.
The combinations of bond stretches, compressions, torsions and
angular vibrations involved in each normal mode are not arbitrary.
They are related to the symmetry of the molecule at its equilibrium
geometry, and can be determined with the help of group theory.
Just as x and y are coordinates associated with orthogonal
directions of motion that can be used to specify the position of a
point on a plane...
...normal coordinates are coordinates associated with orthogonal
directions of motion that can be uses to specify the positions of the
atoms of a molecule.
Note that in the latter we case we are dealing with a space whose
dimension is 3N − 6.
Slide 196 · Models and approximations · Molecular vibration
Normal modes of vibration
The uncoupled, harmonic modes of vibration of a polyatomic molecule
are known as its normal modes of vibration.
The combinations of bond stretches, compressions, torsions and
angular vibrations involved in each normal mode are not arbitrary.
They are related to the symmetry of the molecule at its equilibrium
geometry, and can be determined with the help of group theory.
Just as x and y are coordinates associated with orthogonal
directions of motion that can be used to specify the position of a
point on a plane...
...normal coordinates are coordinates associated with orthogonal
directions of motion that can be uses to specify the positions of the
atoms of a molecule.
Note that in the latter we case we are dealing with a space whose
dimension is 3N − 6.
Slide 196 · Models and approximations · Molecular vibration
Two simple examples of normal coordinates

The displacement of a molecule from the equilibrium geometry in a


normal vibration is measured by a normal coordinate.
Normal coordinates are usually represented by q.
Example: for a CO2 molecule the asymmetric stretch (as) and
symmetric stretch (ss) normal coordinates are related to the lengths
of the two C–O bonds (r1 and r2 ) by

qas = 2−1/2 (r1 − re ) − (r2 − re ) ,


 

qss = 2−1/2 (r1 − re ) + (r2 − re ) ,


 

where re is the equilibrium C–O bond length.

Note: the directions of motion specified dy differing values


of qas and qss are orthogonal.

Slide 197 · Models and approximations · Molecular vibration


Two simple examples of normal coordinates

The displacement of a molecule from the equilibrium geometry in a


normal vibration is measured by a normal coordinate.
Normal coordinates are usually represented by q.
Example: for a CO2 molecule the asymmetric stretch (as) and
symmetric stretch (ss) normal coordinates are related to the lengths
of the two C–O bonds (r1 and r2 ) by

qas = 2−1/2 (r1 − re ) − (r2 − re ) ,


 

qss = 2−1/2 (r1 − re ) + (r2 − re ) ,


 

where re is the equilibrium C–O bond length.

Note: the directions of motion specified dy differing values


of qas and qss are orthogonal.

Slide 197 · Models and approximations · Molecular vibration


Vibrational Hamiltonian of a polyatomic

The Hamiltonian describing the 3N − 6 normal vibrations of a nonlinear


polyatomic is a sum of 3N − 6 independent Hamiltonians:
3N−6
X
Ĥvib = Ĥ(q1 ) + Ĥ(q2 ) + · · · + Ĥ(q3N−6 ) = Ĥ(qn ),
n=1

where each Ĥ(qn ) Hamiltonian is associated with an independent


harmonic oscillator vibrating along the normal coordinate qn with force
constant kn :
2 2
ħ ∂ 1
Ĥ(qn ) = − + V (qn ), V (qn ) = k q2 .
2µn ∂qn2 2 n n

(normal coordinates are always defined so that the equilibrium


geometry of the molecule is given by q1 = q2 = q3 = · · · = q3N−6 = 0)

Slide 198 · Models and approximations · Molecular vibration


Vibrational energy of a polyatomic

In the harmonic approximation the total vibrational energy


of a polyatomic is given by
3N−6 3N−6
(total)
X X 1
Ev1 v2 ···v3N−6 = Ev1 + Ev2 + · · · + Ev3N−6 = Ev n = vn + hcν̃n ,
2
n=1 n=1
v1 = 1, 2, 3, . . .
v2 = 1, 2, 3, . . .
..
.
v3N−6 = 1, 2, 3, . . .
th
where Evn is the vibrational energy of the n normal vibration, and
vn and ν̃n are the corresponding quantum number and vibrational
wavenumber.

Slide 199 · Models and approximations · Molecular vibration


Part II — Models and approximations

11 Molecular translation
The model for 12 Molecular vibration
molecular rotation: 13 Molecular rotation

14 Electronic structure of H atom


rigid rotor
15 Spin

Slide 200 · Models and approximations · Molecular rotation


Rotation v translation: similarities (1)

Free translation Free rotation


• V = 0 (interactions among • V = 0 (interactions among
electrons and nuclei electrons and nuclei
do not change do not change
as molecule translates) as molecule rotates)
• energy is purely kinetic • energy is purely kinetic
• key quantities: • key quantities:
I mass I moment of inertia

I velocity I angular velocity

I (linear) momentum I angular momentum

According to the Merriam-Webster dictionary...

Slide 201 · Models and approximations · Molecular rotation


Rotation v translation: similarities (1)

Free translation Free rotation


• V = 0 (interactions among • V = 0 (interactions among
electrons and nuclei electrons and nuclei
do not change do not change
as molecule translates) as molecule rotates)
• energy is purely kinetic • energy is purely kinetic
• key quantities: • key quantities:
I mass I moment of inertia

I velocity I angular velocity

I (linear) momentum I angular momentum

According to the Merriam-Webster dictionary...

Slide 201 · Models and approximations · Molecular rotation


Rotation v translation: similarities (1)

Free translation Free rotation


• V = 0 (interactions among • V = 0 (interactions among
electrons and nuclei electrons and nuclei
do not change do not change
as molecule translates) as molecule rotates)
• energy is purely kinetic • energy is purely kinetic
• key quantities: • key quantities:
I mass I moment of inertia

I velocity I angular velocity

I (linear) momentum I angular momentum

According to the Merriam-Webster dictionary...

Slide 201 · Models and approximations · Molecular rotation


Rotation v translation: similarities (1)

Free translation Free rotation


• V = 0 (interactions among • V = 0 (interactions among
electrons and nuclei electrons and nuclei
do not change do not change
as molecule translates) as molecule rotates)
• energy is purely kinetic • energy is purely kinetic
• key quantities: • key quantities:
I mass I moment of inertia

I velocity I angular velocity

I (linear) momentum I angular momentum

According to the Merriam-Webster dictionary...

Slide 201 · Models and approximations · Molecular rotation


Rotation v translation: similarities (2)

mass, n. moment of inertia, n.


the property of a body or the property of a body or
system that is a measure system that is a measure
of its resistance to of its resistance to
acceleration along a angular acceleration.
straight line.

velocity, n. angular velocity, n.


a quickness of a quickness of
linear motion; rotational motion;
the rate of change of the rate of change of
the position the spatial orientation
of a body or system with of a body or system with
respect to time. respect to time.

Slide 202 · Models and approximations · Molecular rotation


Rotation v translation: similarities (2)

mass, n. moment of inertia, n.


the property of a body or the property of a body or
system that is a measure system that is a measure
of its resistance to of its resistance to
acceleration along a angular acceleration.
straight line.

velocity, n. angular velocity, n.


a quickness of a quickness of
linear motion; rotational motion;
the rate of change of the rate of change of
the position the spatial orientation
of a body or system with of a body or system with
respect to time. respect to time.

Slide 202 · Models and approximations · Molecular rotation


Rotation v translation: similarities (3)

(linear) momentum, n. angular momentum, n.


a property that a body or a property that a body or
system has by virtue of system has by virtue of
its mass its moment of inertia
and velocity and angular velocity
and that is equal to the and that is equal to the
product of the two; product of the two;
a property of a body or a property of a body or
system in translational system in rotational
motion that determines the motion that determines the
length of time required to length of time required to
bring it to rest when bring it to rest when
under the action of a under the action of a
constant force opposed to constant torque opposed to
its translation. its rotation.
Slide 203 · Models and approximations · Molecular rotation
Rotation v translation: similarities (3)

(linear) momentum, n. angular momentum, n.


a property that a body or a property that a body or
system has by virtue of system has by virtue of
its mass its moment of inertia
and velocity and angular velocity
and that is equal to the and that is equal to the
product of the two; product of the two;
a property of a body or a property of a body or
system in translational system in rotational
motion that determines the motion that determines the
length of time required to length of time required to
bring it to rest when bring it to rest when
under the action of a under the action of a
constant force opposed to constant torque opposed to
its translation. its rotation.
Slide 203 · Models and approximations · Molecular rotation
Rotation v translation: differences

Unlike the mass,the moment of inertia has a directional dependence.


This was demonstrated by and .
“Analogous” quantities for translational or rotational motion
have different units.
Unlike translational motion, rotational motion is periodic.
In translation, the (linear) momentum and motion
are along the same direction. In rotation, the angular momentum
is perpendicular to the plane of rotation.
There are more pairs of complementary observables in rotational
motion than in translational motion.

Slide 204 · Models and approximations · Molecular rotation


Classical rotation

At this point you should take a look at H2 CMrotor.mws and other


Maple animations showing the classical rotation of rigid rotors.
They are on our website.

Comparison between them should show the effects of:


• higher molecular mass (D2 CMrotor.mws);
• redistributed molecular mass (HD CMrotor.mws);
• longer chemical bond (long H2 CMrotor.mws);
• lower rotational energy (lowE H2 CMrotor.mws).

Time now to start thinking about quantum rotation.

Slide 205 · Models and approximations · Molecular rotation


Complementary observables in rotation (1)
We have already mentioned that, as far as complementary
observables are concerned, rotation is quite unlike translation.
As it turns out, there are more pairs of complementary
observables in rotation than in translation.
In translation, x and px (the position and momentum along the x
direction) form a pair of complementary observables. And so do
y and py , and z and pz .
In translation, pairs of observables referring to different directions
(say, x and y, or x and py , or px and py ) are not complementary.
In rotation, components of the rotational angular momentum along
different directions are complementary.
For instance, jx and jy are complementary. And so are jx and jz ,
and jy and jz . Quantitatively:
ħ ħ ħ
∆jx ∆jy ≥ | h j i |2 , ∆jy ∆jz ≥ | h j i |2 , ∆jz ∆jx ≥ | h j i |2 .
2 x 2 z 2 y
Slide 206 · Models and approximations · Molecular rotation
Complementary observables in rotation (1)
We have already mentioned that, as far as complementary
observables are concerned, rotation is quite unlike translation.
As it turns out, there are more pairs of complementary
observables in rotation than in translation.
In translation, x and px (the position and momentum along the x
direction) form a pair of complementary observables. And so do
y and py , and z and pz .
In translation, pairs of observables referring to different directions
(say, x and y, or x and py , or px and py ) are not complementary.
In rotation, components of the rotational angular momentum along
different directions are complementary.
For instance, jx and jy are complementary. And so are jx and jz ,
and jy and jz . Quantitatively:
ħ ħ ħ
∆jx ∆jy ≥ | h j i |2 , ∆jy ∆jz ≥ | h j i |2 , ∆jz ∆jx ≥ | h j i |2 .
2 x 2 z 2 y
Slide 206 · Models and approximations · Molecular rotation
Complementary observables in rotation (1)
We have already mentioned that, as far as complementary
observables are concerned, rotation is quite unlike translation.
As it turns out, there are more pairs of complementary
observables in rotation than in translation.
In translation, x and px (the position and momentum along the x
direction) form a pair of complementary observables. And so do
y and py , and z and pz .
In translation, pairs of observables referring to different directions
(say, x and y, or x and py , or px and py ) are not complementary.
In rotation, components of the rotational angular momentum along
different directions are complementary.
For instance, jx and jy are complementary. And so are jx and jz ,
and jy and jz . Quantitatively:
ħ ħ ħ
∆jx ∆jy ≥ | h j i |2 , ∆jy ∆jz ≥ | h j i |2 , ∆jz ∆jx ≥ | h j i |2 .
2 x 2 z 2 y
Slide 206 · Models and approximations · Molecular rotation
Complementary observables in rotation (2)

In QM, it’s impossible to determine the exact direction of the rotational


angular momentum of a system.
Given that the direction of the angular momentum is tied to the
plane of rotation, this implies that in QM it’s impossible to
determine the exact plane in which a rotating body rotates.
In rotation, the incompatibility between “position” (strictly speaking,
spatial orientation) and angular momentum is more complicated than
what we’ve seen in translation.
We won’t consider that in detail.
There are, however, three things things you must remember. They
might become easier to understand when we look at pictures, but right
now we’ll stick to text.

Slide 207 · Models and approximations · Molecular rotation


Complementary observables in rotation (2)

In QM, it’s impossible to determine the exact direction of the rotational


angular momentum of a system.
Given that the direction of the angular momentum is tied to the
plane of rotation, this implies that in QM it’s impossible to
determine the exact plane in which a rotating body rotates.
In rotation, the incompatibility between “position” (strictly speaking,
spatial orientation) and angular momentum is more complicated than
what we’ve seen in translation.
We won’t consider that in detail.
There are, however, three things things you must remember. They
might become easier to understand when we look at pictures, but right
now we’ll stick to text.

Slide 207 · Models and approximations · Molecular rotation


Complementary observables in rotation (3)

Three things to remember:


• If you know the exact spatial orientation of an object, then there’s
absolutely nothing you can simultaneously know about the
angular momentum of that object. The uncertainty in the angular
momentum is infinite.
• If you know that a given object is not rotating (ie, if you know that
its angular momentum is exactly zero), then there’s absolutely
nothing you can simultaneously know about the spatial orientation
of that object. The uncertainty in the spatial orientation is infinite.
• If you know the exact value of one of the components of its
angular momentum (say, jz ), then you cannot simultaneously
know the exact spatial orientation of a rotating object. The
uncertainty in the spatial orientation, however, is not infinite. You
can have some information about it.

Slide 208 · Models and approximations · Molecular rotation


Complementary observables in rotation (3)

Three things to remember:


• If you know the exact spatial orientation of an object, then there’s
absolutely nothing you can simultaneously know about the
angular momentum of that object. The uncertainty in the angular
momentum is infinite.
• If you know that a given object is not rotating (ie, if you know that
its angular momentum is exactly zero), then there’s absolutely
nothing you can simultaneously know about the spatial orientation
of that object. The uncertainty in the spatial orientation is infinite.
• If you know the exact value of one of the components of its
angular momentum (say, jz ), then you cannot simultaneously
know the exact spatial orientation of a rotating object. The
uncertainty in the spatial orientation, however, is not infinite. You
can have some information about it.

Slide 208 · Models and approximations · Molecular rotation


Complementary observables in rotation (3)

Three things to remember:


• If you know the exact spatial orientation of an object, then there’s
absolutely nothing you can simultaneously know about the
angular momentum of that object. The uncertainty in the angular
momentum is infinite.
• If you know that a given object is not rotating (ie, if you know that
its angular momentum is exactly zero), then there’s absolutely
nothing you can simultaneously know about the spatial orientation
of that object. The uncertainty in the spatial orientation is infinite.
• If you know the exact value of one of the components of its
angular momentum (say, jz ), then you cannot simultaneously
know the exact spatial orientation of a rotating object. The
uncertainty in the spatial orientation, however, is not infinite. You
can have some information about it.

Slide 208 · Models and approximations · Molecular rotation


Rotational and magnetic quantum numbers

Solving the TI Schrödinger equation for the rotational motion of a rigid


diatomic we can obtain the energies and wavefunctions associated
with its rotational states.
Since molecular rotation is subject to (periodic) boundary
conditions, it is quantised.
Because the motion described by them is two-dimensional, the
rotational wavefunctions of a rigid diatomic depend on two quantum
numbers: the rotational quantum number J, and the magnetic
quantum number M.
We’ll see that while the rotational energy only depends on the first
of these quantum numbers, the angular momentum depends on
both of them.

Slide 209 · Models and approximations · Molecular rotation


Rotational energy...

As emphasised in the animations showing classical rotation, angular


momentum is a vector quantity, defined by magnitude and direction.
The magnitude of the angular momentum determines
the speed of the rotational motion.
Since the energy associated with molecular rotations is purely kinetic,
this implies that the magnitude of the angular momentum and the
rotational energy are directly related to each other.

The rotational energy of a rigid diatomic is determined


by the rotational quantum number, J:
2
ħ
EJ = J(J + 1), J = 0, 1, 2, 3, . . .
2I
where I is the moment of inertia of the diatomic.

Slide 210 · Models and approximations · Molecular rotation


Rotational energy...

As emphasised in the animations showing classical rotation, angular


momentum is a vector quantity, defined by magnitude and direction.
The magnitude of the angular momentum determines
the speed of the rotational motion.
Since the energy associated with molecular rotations is purely kinetic,
this implies that the magnitude of the angular momentum and the
rotational energy are directly related to each other.

The rotational energy of a rigid diatomic is determined


by the rotational quantum number, J:
2
ħ
EJ = J(J + 1), J = 0, 1, 2, 3, . . .
2I
where I is the moment of inertia of the diatomic.

Slide 210 · Models and approximations · Molecular rotation


...rotational energy gaps...

Using the expression for the rotational energy we can obtain a formula
for the energy gap between consecutive rotational energy levels:
2 2
ħ ħ
EJ+1 − EJ = (J + 1)(J + 2) − J(J + 1)
2I 2I
2

= (J + 2) − J (J + 1)

2I
ħ2
= 2 (J + 1) = 2B(J + 1),
2I

Slide 211 · Models and approximations · Molecular rotation


...rotational energy gaps...

Using the expression for the rotational energy we can obtain a formula
for the energy gap between consecutive rotational energy levels:
2 2
ħ ħ
EJ+1 − EJ = (J + 1)(J + 2) − J(J + 1)
2I 2I
2

= (J + 2) − J (J + 1)

2I
ħ2
= 2 (J + 1) = 2B(J + 1),
2I

Slide 211 · Models and approximations · Molecular rotation


...rotational energy gaps...

Using the expression for the rotational energy we can obtain a formula
for the energy gap between consecutive rotational energy levels:
2 2
ħ ħ
EJ+1 − EJ = (J + 1)(J + 2) − J(J + 1)
2I 2I
2

= (J + 2) − J (J + 1)

2I
ħ2
= 2 (J + 1) = 2B(J + 1),
2I

Slide 211 · Models and approximations · Molecular rotation


...rotational energy gaps...

Using the expression for the rotational energy we can obtain a formula
for the energy gap between consecutive rotational energy levels:
2 2
ħ ħ
EJ+1 − EJ = (J + 1)(J + 2) − J(J + 1)
2I 2I
2

= (J + 2) − J (J + 1)

2I
ħ2
= 2 (J + 1) = 2B(J + 1),
2I

Slide 211 · Models and approximations · Molecular rotation


...rotational energy gaps...

Using the expression for the rotational energy we can obtain a formula
for the energy gap between consecutive rotational energy levels:
2 2
ħ ħ
EJ+1 − EJ = (J + 1)(J + 2) − J(J + 1)
2I 2I
2

= (J + 2) − J (J + 1)

2I
ħ2
= 2 (J + 1) = 2B(J + 1),
2I

where B = ħ2 /2I is an important number in rotational spectroscopy,


known as the rotational constant.

Slide 211 · Models and approximations · Molecular rotation


...and angular momentum magnitude

In translation, the (translational) kinetic energy is related to the (linear)


2
momentum p by E = p /2m.
In rotation, the (rotational) kinetic energy is related to the
2
angular momentum j by E = j /2I.
Using the expression for EJ in slide 210 we can see that
q
|j | = ħ J(J + 1), J = 0, 1, 2, . . .

a formula that expresses the quantisation of rotational angular


momentum.

Slide 212 · Models and approximations · Molecular rotation


Degeneracy of rotational energy levels
As emphasised in the animations showing classical rotation, angular
momentum is a vector quantity, defined by magnitude and direction.
The direction of the angular momentum determines plane and
direction of the rotational motion.
As the kinetic energy is independent of the plane and sense of
rotation, the rotational energy is independent of the angular
momentum direction.

To some extent, the angular momentum direction is determined by the


magnetic quantum number, M:
M = 0, ±1, ±2, ±3, . . . , ±J.
The rotational energy (see slide 210) does not depend on M. States
associated with the same J but different M values are degenerate.
As for a fixed value of J there are 2J + 1 possible M values,
the degeneracy of rotational energy levels is given by 2J + 1.
Slide 213 · Models and approximations · Molecular rotation
Degeneracy of rotational energy levels
As emphasised in the animations showing classical rotation, angular
momentum is a vector quantity, defined by magnitude and direction.
The direction of the angular momentum determines plane and
direction of the rotational motion.
As the kinetic energy is independent of the plane and sense of
rotation, the rotational energy is independent of the angular
momentum direction.

To some extent, the angular momentum direction is determined by the


magnetic quantum number, M:
M = 0, ±1, ±2, ±3, . . . , ±J.
The rotational energy (see slide 210) does not depend on M. States
associated with the same J but different M values are degenerate.
As for a fixed value of J there are 2J + 1 possible M values,
the degeneracy of rotational energy levels is given by 2J + 1.
Slide 213 · Models and approximations · Molecular rotation
Uncertainty in angular momentum direction
The magnetic quantum number only determines the angular
momentum direction “to some extent.”

More specifically, what M specifies is the value of the z component of


the angular momentum vector:

jz = Mħ, M = −J, −J + 1, −J + 2, . . . , J − 2, J − 1, J,

a formula that expresses what is known as “space quantisation”.

As also mentioned earlier, the operators associated with different


components of the angular momentum vector do not commute:

[̂x , ̂y ] = i ħ̂z [̂y , ̂z ] = i ħ̂x [̂z , ̂x ] = i ħ̂y ,

which implies that it’s impossible to know the exact direction of an


angular momentum vector.
Slide 214 · Models and approximations · Molecular rotation
Uncertainty in angular momentum direction
The magnetic quantum number only determines the angular
momentum direction “to some extent.”

More specifically, what M specifies is the value of the z component of


the angular momentum vector:

jz = Mħ, M = −J, −J + 1, −J + 2, . . . , J − 2, J − 1, J,

a formula that expresses what is known as “space quantisation”.

As also mentioned earlier, the operators associated with different


components of the angular momentum vector do not commute:

[̂x , ̂y ] = i ħ̂z [̂y , ̂z ] = i ħ̂x [̂z , ̂x ] = i ħ̂y ,

which implies that it’s impossible to know the exact direction of an


angular momentum vector.
Slide 214 · Models and approximations · Molecular rotation
Uncertainty in angular momentum direction
The magnetic quantum number only determines the angular
momentum direction “to some extent.”

More specifically, what M specifies is the value of the z component of


the angular momentum vector:

jz = Mħ, M = −J, −J + 1, −J + 2, . . . , J − 2, J − 1, J,

a formula that expresses what is known as “space quantisation”.

As also mentioned earlier, the operators associated with different


components of the angular momentum vector do not commute:

[̂x , ̂y ] = i ħ̂z [̂y , ̂z ] = i ħ̂x [̂z , ̂x ] = i ħ̂y ,

which implies that it’s impossible to know the exact direction of an


angular momentum vector.
Slide 214 · Models and approximations · Molecular rotation
Angular momentum components (1)

Angular momentum components in different directions are


complementary: jx and jy are complementary, jy and jz are
complementary, and so are jz and jx .
If we know the exact value of jz , then jx and jy become
completely undetermined.
Let us consider a specific example. Say, a rotational state of a diatomic
molecule characterised by the quantum numbers J = 15 and M = 10.
Note:√we know the magnitude of the angular momentum vector,
|j| = 240ħ, and the value of its z component , jz = 10ħ.
What about jx and jy ? Can we figure these out, and thus determine
the exact direction of the angular momentum vector?
Let us look at some pictures.

Slide 215 · Models and approximations · Molecular rotation


Angular momentum components (1)

Angular momentum components in different directions are


complementary: jx and jy are complementary, jy and jz are
complementary, and so are jz and jx .
If we know the exact value of jz , then jx and jy become
completely undetermined.
Let us consider a specific example. Say, a rotational state of a diatomic
molecule characterised by the quantum numbers J = 15 and M = 10.
Note:√we know the magnitude of the angular momentum vector,
|j| = 240ħ, and the value of its z component , jz = 10ħ.
What about jx and jy ? Can we figure these out, and thus determine
the exact direction of the angular momentum vector?
Let us look at some pictures.

Slide 215 · Models and approximations · Molecular rotation


Angular momentum components (1)

Angular momentum components in different directions are


complementary: jx and jy are complementary, jy and jz are
complementary, and so are jz and jx .
If we know the exact value of jz , then jx and jy become
completely undetermined.
Let us consider a specific example. Say, a rotational state of a diatomic
molecule characterised by the quantum numbers J = 15 and M = 10.
Note:√we know the magnitude of the angular momentum vector,
|j| = 240ħ, and the value of its z component , jz = 10ħ.
What about jx and jy ? Can we figure these out, and thus determine
the exact direction of the angular momentum vector?
Let us look at some pictures.

Slide 215 · Models and approximations · Molecular rotation


Angular momentum components (2)

We know the exact values of |j| and jz . What about jx and jy ? For
instance, we could have one of these:

√ √ √ √
|j| = 240ħ |j| = 240ħ |j| = 240ħ |j| = 240ħ
jz = 10ħ jz = 10ħ jz = 10ħ jz = 10ħ
jx > 0, jy = 0 jx = 0, jy > 0 jx < 0, jy = 0 jx = 0, jy < 0

Can we identify the actual angular momentum direction?

Slide 216 · Models and approximations · Molecular rotation


Angular momentum components (3)

Classical vector model Because of the


complementarity between pairs
of angular momentum
components, we can’t know the
actual angular momentum
direction.

There is no way of
distinguishing between the
directions compatible with the
known jz value. All of them are
equally probable.

Slide 217 · Models and approximations · Molecular rotation


Angular momentum components (3)

Classical vector model Because of the


complementarity between pairs
of angular momentum
components, we can’t know the
actual angular momentum
direction.

There is no way of
distinguishing between the
directions compatible with the
known jz value. All of them are
equally probable. All of them!

(Some people would even say that talk about the existence
of an “actual” angular momentum orientation is nonsense)
Slide 217 · Models and approximations · Molecular rotation
Observables related to rotation
Summary of results

Here’s a summary of our results regarding observables that are


important for rotational motion:
2
ħ
EJ = J(J + 1), EJ+1 − EJ = 2B(J + 1)
2I
q
|j | = ħ J(J + 1)
jz = Mħ
J = 0, 1, 2, 3, . . .
M = −J, −J + 1, −J + 2, . . . , J − 1, J − 2, J,
[jx , jy ] = i ħjz
ħ
∆jx ∆jy ≥ | h jz i |2
2

Slide 218 · Models and approximations · Molecular rotation


Rotational energy levels of a rigid diatomic

12B 3

rotational quantum number (J)


• J = 0, 1, 2, 3, ...
rotational energy (EJ )

• M = 0, ±1, ±2, . . . , ±J
ħ2
• EJ = J(J + 1) = BJ(J + 1)
6B 2 2I
• level degeneracy is 2J + 1
• energy gaps increase with J
2B 1 • rotational spectrum features
absorption lines at
0 0
hν = 2B, 4B, 6B, 8B, . . .
−3 −2 −1 0 +1 +2 +3

magnetic quantum number (M)

Slide 219 · Models and approximations · Molecular rotation


Rotational energy levels of a rigid diatomic

12B 3

rotational quantum number (J)


• J = 0, 1, 2, 3, ...
rotational energy (EJ )

• M = 0, ±1, ±2, . . . , ±J
ħ2
• EJ = J(J + 1) = BJ(J + 1)
6B 2 2I
• level degeneracy is 2J + 1
• energy gaps increase with J
2B 1 • rotational spectrum features
absorption lines at
0 0
hν = 2B, 4B, 6B, 8B, . . .
−3 −2 −1 0 +1 +2 +3

magnetic quantum number (M)

Slide 219 · Models and approximations · Molecular rotation


Rotational energy levels of a rigid diatomic

12B 3

rotational quantum number (J)


• J = 0, 1, 2, 3, ...
rotational energy (EJ )

• M = 0, ±1, ±2, . . . , ±J
ħ2
• EJ = J(J + 1) = BJ(J + 1)
6B 2 2I
• level degeneracy is 2J + 1
• energy gaps increase with J
2B 1 • rotational spectrum features
absorption lines at
0 0
hν = 2B, 4B, 6B, 8B, . . .
−3 −2 −1 0 +1 +2 +3

magnetic quantum number (M)

Slide 219 · Models and approximations · Molecular rotation


Rotational energy levels of a rigid diatomic

12B 3

rotational quantum number (J)


• J = 0, 1, 2, 3, ...
rotational energy (EJ )

∆E = 6B • M = 0, ±1, ±2, . . . , ±J
ħ2
• EJ = J(J + 1) = BJ(J + 1)
6B 2 2I
• level degeneracy is 2J + 1
∆E = 4B
• energy gaps increase with J
2B 1 • rotational spectrum features
∆E = 2B absorption lines at
0 0
hν = 2B, 4B, 6B, 8B, . . .
−3 −2 −1 0 +1 +2 +3

magnetic quantum number (M)

Slide 219 · Models and approximations · Molecular rotation


Rotational energy levels of a rigid diatomic

12B 3

rotational quantum number (J)


• J = 0, 1, 2, 3, ...
rotational energy (EJ )

∆E = 6B • M = 0, ±1, ±2, . . . , ±J
ħ2
• EJ = J(J + 1) = BJ(J + 1)
6B 2 2I
• level degeneracy is 2J + 1
∆E = 4B
• energy gaps increase with J
2B 1 • rotational spectrum features
∆E = 2B absorption lines at
0 0
hν = 2B, 4B, 6B, 8B, . . .
−3 −2 −1 0 +1 +2 +3

magnetic quantum number (M)

Slide 219 · Models and approximations · Molecular rotation


What rotational spectra tell us about molecules

The rotational constant of a rigid diatomic is given by B = ħ2 /2I,


where I is the moment of the inertia of the diatomic, related to its
2
reduced mass (µ) and bond length (r) by I = µr .

Using rotational spectroscopy — ie, measuring the gaps between


various rotational energy levels — we can learn about molecular
structure.
Rotational spectra allow us to determine moments of inertia and
bond lengths. In other words, to figure out the geometries and
isotopic compositions of molecules.
And if we know about the rotational constants of a given molecule, we
can use rotational spectroscopy as a detection tool.

Slide 220 · Models and approximations · Molecular rotation


What rotational spectra tell us about molecules

The rotational constant of a rigid diatomic is given by B = ħ2 /2I,


where I is the moment of the inertia of the diatomic, related to its
2
reduced mass (µ) and bond length (r) by I = µr .

Using rotational spectroscopy — ie, measuring the gaps between


various rotational energy levels — we can learn about molecular
structure.
Rotational spectra allow us to determine moments of inertia and
bond lengths. In other words, to figure out the geometries and
isotopic compositions of molecules.
And if we know about the rotational constants of a given molecule, we
can use rotational spectroscopy as a detection tool.

Slide 220 · Models and approximations · Molecular rotation


What rotational spectra tell us about molecules

The rotational constant of a rigid diatomic is given by B = ħ2 /2I,


where I is the moment of the inertia of the diatomic, related to its
2
reduced mass (µ) and bond length (r) by I = µr .

Using rotational spectroscopy — ie, measuring the gaps between


various rotational energy levels — we can learn about molecular
structure.
Rotational spectra allow us to determine moments of inertia and
bond lengths. In other words, to figure out the geometries and
isotopic compositions of molecules.
And if we know about the rotational constants of a given molecule, we
can use rotational spectroscopy as a detection tool.

Slide 220 · Models and approximations · Molecular rotation


Wavefunctions of rigid rotor (1)

Solution of the Schrödinger equation for a rigid rotor leads not only to
rotational energies, it also leads to rotational wavefunctions.
Rotational wavefunctions have a special name: they are known as
the spherical harmonics, and are represented by YJM .
Note that the spherical harmonics depend on J and M, the two
quantum numbers we have found when dealing with rigid rotors.
We will not consider the explicit mathematical expressions for the
spherical harmonics. We will, however, look at what they look like.
This will introduce a new problem.
Contrary to the TI wavefunctions we have seen so far,
the spherical harmonics are complex functions.
That is, their values are in general complex.

Slide 221 · Models and approximations · Molecular rotation


Wavefunctions of rigid rotor (1)

Solution of the Schrödinger equation for a rigid rotor leads not only to
rotational energies, it also leads to rotational wavefunctions.
Rotational wavefunctions have a special name: they are known as
the spherical harmonics, and are represented by YJM .
Note that the spherical harmonics depend on J and M, the two
quantum numbers we have found when dealing with rigid rotors.
We will not consider the explicit mathematical expressions for the
spherical harmonics. We will, however, look at what they look like.
This will introduce a new problem.
Contrary to the TI wavefunctions we have seen so far,
the spherical harmonics are complex functions.
That is, their values are in general complex.

Slide 221 · Models and approximations · Molecular rotation


Wavefunctions of rigid rotor (2)

Besides the (real and non-negative) probability density, we will look at


both real and imaginary parts of the spherical harmonics.
We will have separate pictures for each. You must remember,
however, that the real and imaginary parts belong to a single,
complex-valued wavefunction.
When looking at Re(YJM ) and Im(YJM ) we’ll use the following
convention:
• RED (DARK on the handouts) means POSITIVE;
• YELLOW (LIGHT on the handouts) means NEGATIVE.

Slide 222 · Models and approximations · Molecular rotation


Rotational wavefunctions and probability densities
Rigid diatomic, J = 0, M = 0

Y00 2
|Y00 |

• since J = 0, there’s no rotation. This complete certainty about the


angular momentum implies that we can have no information about
the spatial orientation of the molecule. The probability density is
the same for every direction.

Slide 223 · Models and approximations · Molecular rotation


Rotational wavefunctions and probability densities
Rigid diatomic, J = 1, M = 0

Y10 2
|Y10 |

• internuclear axis likely to be found near the Z axis


• wavefunction has a node on the X Y plane;
probability of finding molecular axis on this nodal plane is zero.

Slide 224 · Models and approximations · Molecular rotation


Rotational wavefunctions and probability densities
Rigid diatomic, J = 2, M = 0

Y20 2
|Y20 |

• internuclear axis likely to be found near the Z axis


(large probability) or approximately perpendicular to it
(not so large probability)
• two nodal surfaces

Slide 225 · Models and approximations · Molecular rotation


Rotational wavefunctions and probability densities
Rigid diatomic, J = 1, M = −1

Re(Y1 −1 ) Im(Y1 −1 ) |Y1 −1 |2

• internuclear axis likely to be found near the X Y plane


• wavefunction has a node on the Z axis;
probability of finding molecular axis along this direction is zero.

Slide 226 · Models and approximations · Molecular rotation


Rotational wavefunctions and probability densities
Rigid diatomic, J = 1, M = +1

Re(Y1 +1 ) Im(Y1 +1 ) |Y1 +1 |2

• except for sign of the real part of the spherical harmonic, no


change with regard to the previous slide. This will have important
consequences when we look at atomic and molecular orbitals.

Slide 227 · Models and approximations · Molecular rotation


Rotational wavefunctions and probability densities
Rigid diatomic, J = 2, M = −1

Re(Y2 −1 ) Im(Y2 −1 ) |Y2 −1 |2

• Z axis and XY plane are nodal;


internuclear axis likely to be found between those
• nodal regions of |YJM |2 do not coincide
with nodal regions of Re(YJM ) or Im(YJM )

Slide 228 · Models and approximations · Molecular rotation


Rotational wavefunctions and probability densities
Rigid diatomic, J = 2, M = +1

Re(Y2 +1 ) Im(Y2 +1 ) |Y2 +1 |2

• except for sign of the real part of the spherical harmonic, no


change with regard to the previous slide. This will have important
consequences when we look at atomic and molecular orbitals.

Slide 229 · Models and approximations · Molecular rotation


Rotational wavefunctions and probability densities
Rigid diatomic, J = 2, M = −2

Re(Y2 −2 ) Im(Y2 −2 ) |Y2 −2 |2

• although probability density doesn’t look too different from those


found with J = 1 and M = ±1, wavefunction does.

Slide 230 · Models and approximations · Molecular rotation


Rotational wavefunctions and probability densities
Rigid diatomic, J = 2, M = +2

Re(Y2 +2 ) Im(Y2 +2 ) |Y2 +2 |2

• except for sign of the imaginary part of the spherical harmonic, no


change with regard to the previous slide. Not too different from
what we found in the other cases involving M 6= 0.

Slide 231 · Models and approximations · Molecular rotation


Part II — Models and approximations

11 Molecular translation
The model for 12 Molecular vibration
electronic structure: 13 Molecular rotation

14 Electronic structure of H atom


hydrogen atom
15 Spin

Slide 232 · Models and approximations · Electronic structure of H atom


Electronic structure
Let us consider the electronic structure of atoms. Two things to note:
• As you probably guessed, the motion of electrons around nuclei
has a lot in common with rotational motion.
• This isn’t all there is to it.

Slide 233 · Models and approximations · Electronic structure of H atom


Electronic structure
Let us consider the electronic structure of atoms. Two things to note:
• As you probably guessed, the motion of electrons around nuclei
has a lot in common with rotational motion.
• This isn’t all there is to it.

Slide 233 · Models and approximations · Electronic structure of H atom


Electronic structure
Let us consider the electronic structure of atoms. Two things to note:
• As you probably guessed, the motion of electrons around nuclei
has a lot in common with rotational motion.
• This isn’t all there is to it.
What makes electronic motion similar to or different from
the rotational motion of a rigid rotor?

Slide 233 · Models and approximations · Electronic structure of H atom


Electronic structure
Let us consider the electronic structure of atoms. Two things to note:
• As you probably guessed, the motion of electrons around nuclei
has a lot in common with rotational motion.
• This isn’t all there is to it.
What makes electronic motion similar to or different from
the rotational motion of a rigid rotor?

The simplest example is the hydrogen atom. Note:


• it’s made up of two things only: the electron and the nucleus
• because the nucleus is much heavier than the electron, we can
assume that the position of the nucleus coincides with the center
of mass, and that the electron “orbits” the nucleus
• the interaction between the electron and the nucleus
is an electrostatic attraction

Slide 233 · Models and approximations · Electronic structure of H atom


Electronic structure
Let us consider the electronic structure of atoms. Two things to note:
• As you probably guessed, the motion of electrons around nuclei
has a lot in common with rotational motion.
• This isn’t all there is to it.
What makes electronic motion similar to or different from
the rotational motion of a rigid rotor?

The simplest example is the hydrogen atom. Note:


• it’s made up of two things only: the electron and the nucleus
• because the nucleus is much heavier than the electron, we can
assume that the position of the nucleus coincides with the center
of mass, and that the electron “orbits” the nucleus
• the interaction between the electron and the nucleus
is an electrostatic attraction

Slide 233 · Models and approximations · Electronic structure of H atom


Electronic structure
Let us consider the electronic structure of atoms. Two things to note:
• As you probably guessed, the motion of electrons around nuclei
has a lot in common with rotational motion.
• This isn’t all there is to it.
What makes electronic motion similar to or different from
the rotational motion of a rigid rotor?

The simplest example is the hydrogen atom. Note:


• it’s made up of two things only: the electron and the nucleus
• because the nucleus is much heavier than the electron, we can
assume that the position of the nucleus coincides with the center
of mass, and that the electron “orbits” the nucleus
• the interaction between the electron and the nucleus
is an electrostatic attraction

Slide 233 · Models and approximations · Electronic structure of H atom


Electronic structure
Let us consider the electronic structure of atoms. Two things to note:
• As you probably guessed, the motion of electrons around nuclei
has a lot in common with rotational motion.
• This isn’t all there is to it.
What makes electronic motion similar to or different from
the rotational motion of a rigid rotor?

The simplest example is the hydrogen atom. Note:


• it’s made up of two things only: the electron and the nucleus
• because the nucleus is much heavier than the electron, we can
assume that the position of the nucleus coincides with the center
of mass, and that the electron “orbits” the nucleus
• the interaction between the electron and the nucleus
is an electrostatic attraction

Slide 233 · Models and approximations · Electronic structure of H atom


Electrostatic interaction

The electrostatic interaction between an electron with nominal charge


−1 and a nucleus with nominal charge +Z is described by the potential
2
qZ e
VeN (r) = − , q= = 2.307 × 10−28 J m
r 4πε0
Note the following:
• if the distance between the electron and the nucleus is zero,
we have VeN (r = 0) = −∞;
• if the distance between the electron and the nucleus is infinite,
we have VeN (r = ∞) = 0;

Slide 234 · Models and approximations · Electronic structure of H atom


The attraction between electron and nucleus
The potential that describes the electron-nucleus attraction is
proportional to −1/r. We have VeN (r = ∞) = 0 and VeN (r = 0) = −∞.
In plain language: as the electron and the nucleus approach each
other, their attraction becomes stronger and stronger.
No matter how far apart they are, electron and nucleus will attract each
other until the distance between them is zero, and there’s no stopping
it. Or is there?

Slide 235 · Models and approximations · Electronic structure of H atom


The attraction between electron and nucleus
The potential that describes the electron-nucleus attraction is
proportional to −1/r. We have VeN (r = ∞) = 0 and VeN (r = 0) = −∞.
In plain language: as the electron and the nucleus approach each
other, their attraction becomes stronger and stronger.
No matter how far apart they are, electron and nucleus will attract each
other until the distance between them is zero, and there’s no stopping
it. Or is there?
In the classical (ie, “intuitive”) world there would be no stopping
the electron and the nucleus getting ever closer.
With no restrictions imposed by quantisation or uncertainty, electron
and nucleus would continuously attract each other. Until they were
joined in a single, indivisible, neutral particle.
In a classical world, there can be no atoms as we know them. And
no molecules, and no chemistry.
Slide 235 · Models and approximations · Electronic structure of H atom
The H atom as it is: quantum

In the quantum (ie, actual) world, quantisation and the uncertainty


principle prevent the collapse of electron and nucleus into a neutral
particle.
Quantitative details are different, but qualitatively it’s just like what
we’ve seen in the case of vibrations.

Slide 236 · Models and approximations · Electronic structure of H atom


The H atom as it is: quantum

In the quantum (ie, actual) world, quantisation and the uncertainty


principle prevent the collapse of electron and nucleus into a neutral
particle.
Quantitative details are different, but qualitatively it’s just like what
we’ve seen in the case of vibrations.
We’ll now see what the hydrogen atom is really like. We’ll look at its
(quantised) energies and wavefunctions.

In order to do that, besides things we’ve already seen,


we’ll need the following:
• spherical coordinates;
• radial wavefunctions.

Slide 236 · Models and approximations · Electronic structure of H atom


H atom in Cartesian coordinates

• nucleus at center of mass


• three coordinates needed to
specify the electron position:
x, y and z
• electron-nucleus
p distance:
r= x2 + y 2 + z2
p
• VeN = −qZ/ x 2 + y 2 + z 2
• inconvenient

Slide 237 · Models and approximations · Electronic structure of H atom


H atom in Cartesian coordinates

• nucleus at center of mass


• three coordinates needed to
specify the electron position:
x, y and z
• electron-nucleus
p distance:
r= x2 + y 2 + z2
p
• VeN = −qZ/ x 2 + y 2 + z 2
• inconvenient

Slide 237 · Models and approximations · Electronic structure of H atom


H atom in Cartesian coordinates

• nucleus at center of mass


• three coordinates needed to
specify the electron position:
x, y and z
• electron-nucleus
p distance:
r= x2 + y 2 + z2
p
• VeN = −qZ/ x 2 + y 2 + z 2
• inconvenient

Slide 237 · Models and approximations · Electronic structure of H atom


H atom in Cartesian coordinates

• nucleus at center of mass


• three coordinates needed to
specify the electron position:
x, y and z
• electron-nucleus
p distance:
r= x2 + y 2 + z2
p
• VeN = −qZ/ x 2 + y 2 + z 2
• inconvenient

Slide 237 · Models and approximations · Electronic structure of H atom


H atom in Cartesian coordinates

• nucleus at center of mass


• three coordinates needed to
specify the electron position:
x, y and z
• electron-nucleus
p distance:
r= x2 + y 2 + z2
p
• VeN = −qZ/ x 2 + y 2 + z 2
• inconvenient

Slide 237 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (1)

• nucleus at center of mass


• three coordinates needed to
specify the electron position:
r, θ and ϕ (radius, polar angle
and azimuthal angle)
• electron-nucleus distance: r
• VeN = −qZ/r
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 238 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (1)

• nucleus at center of mass


• three coordinates needed to
specify the electron position:
r, θ and ϕ (radius, polar angle
and azimuthal angle)
• electron-nucleus distance: r
• VeN = −qZ/r
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 238 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (1)

• nucleus at center of mass


• three coordinates needed to
specify the electron position:
r, θ and ϕ (radius, polar angle
and azimuthal angle)
• electron-nucleus distance: r
• VeN = −qZ/r
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 238 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (1)

• nucleus at center of mass


• three coordinates needed to
specify the electron position:
r, θ and ϕ (radius, polar angle
and azimuthal angle)
• electron-nucleus distance: r
• VeN = −qZ/r
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 238 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (1)

• nucleus at center of mass


• three coordinates needed to
specify the electron position:
r, θ and ϕ (radius, polar angle
and azimuthal angle)
• electron-nucleus distance: r
• VeN = −qZ/r
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 238 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (2)

• angular coordinates ( θ and ϕ)


change as electron rotates
around nucleus
• polar angle θ changes as electron
moves from “north pole” (θ = 0)
to “south pole” (θ = π = 180◦ )
• potential does not depend on
polar angle: V (θ) = 0
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 239 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (2)

• angular coordinates ( θ and ϕ)


change as electron rotates
around nucleus
• polar angle θ changes as electron
moves from “north pole” (θ = 0)
to “south pole” (θ = π = 180◦ )
• potential does not depend on
polar angle: V (θ) = 0
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 239 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (2)

• angular coordinates ( θ and ϕ)


change as electron rotates
around nucleus
• polar angle θ changes as electron
moves from “north pole” (θ = 0)
to “south pole” (θ = π = 180◦ )
• potential does not depend on
polar angle: V (θ) = 0
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 239 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (2)

• angular coordinates ( θ and ϕ)


change as electron rotates
around nucleus
• polar angle θ changes as electron
moves from “north pole” (θ = 0)
to “south pole” (θ = π = 180◦ )
• potential does not depend on
polar angle: V (θ) = 0
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 239 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (3)

• angular coordinates ( θ and ϕ)


change as electron rotates
around nucleus
• azimuthal angle ϕ changes as
electron moves
“eastward around the globe”
from “Greenwich meridian”
(ϕ = 0 and ϕ = 2π = 360◦ )
• potential does not depend on
azimuthal angle: V (ϕ) = 0
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 240 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (3)

• angular coordinates ( θ and ϕ)


change as electron rotates
around nucleus
• azimuthal angle ϕ changes as
electron moves
“eastward around the globe”
from “Greenwich meridian”
(ϕ = 0 and ϕ = 2π = 360◦ )
• potential does not depend on
azimuthal angle: V (ϕ) = 0
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 240 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (3)

• angular coordinates ( θ and ϕ)


change as electron rotates
around nucleus
• azimuthal angle ϕ changes as
electron moves
“eastward around the globe”
from “Greenwich meridian”
(ϕ = 0 and ϕ = 2π = 360◦ )
• potential does not depend on
azimuthal angle: V (ϕ) = 0
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 240 · Models and approximations · Electronic structure of H atom


H atom in spherical coordinates (3)

• angular coordinates ( θ and ϕ)


change as electron rotates
around nucleus
• azimuthal angle ϕ changes as
electron moves
“eastward around the globe”
from “Greenwich meridian”
(ϕ = 0 and ϕ = 2π = 360◦ )
• potential does not depend on
azimuthal angle: V (ϕ) = 0
• convenient

(see maple worksheet Animate sphcoord.mws)

Slide 240 · Models and approximations · Electronic structure of H atom


Hamiltonian for H atom
If we use spherical coordinates, the Hamiltonian for the hydrogen atom
takes the form

Ĥ(r, θ, ϕ) = K̂ (r) + K̂ (θ) + K̂ (ϕ) + V̂ (r),

where the K̂ ’s are the kinetic energy operators associated with motion
along each coordinate, and V̂ (r) is the potential energy operator.

Slide 241 · Models and approximations · Electronic structure of H atom


Hamiltonian for H atom
If we use spherical coordinates, the Hamiltonian for the hydrogen atom
takes the form

Ĥ(r, θ, ϕ) = K̂ (r) + K̂ (θ) + K̂ (ϕ) + V̂ (r),

where the K̂ ’s are the kinetic energy operators associated with motion
along each coordinate, and V̂ (r) is the potential energy operator.

This implies that the H atom Hamiltonian can be separated in two parts
that are independent of each other:

Ĥ(r, θ, ϕ) = Ĥ(r) + Ĥ(θ, ϕ),

where the two independent parts are given by

Ĥ(r) = K̂ (r) + V̂ (r)


Ĥ(θ, ϕ) = K̂ (θ) + K (ϕ) = K̂ (θ, ϕ).

Slide 241 · Models and approximations · Electronic structure of H atom


Solving the Schrödinger equation
The fact that the Hamiltonian is a sum of independent terms means
that the energy of the H atom is a sum of two independent energies,

E = Eradial + Eangular ,

that its wavefunction is a product of two independent wavefunctions,

ψ(r, θ, ϕ) = R(r) × Y (θ, ϕ),

and that we do not need to solve the 3-D Schrödinger equation

Ĥ(r, θ, ϕ)ψ(r, θ, ϕ) = E ψ(r, θ, ϕ).

Instead, we can separately solve two simpler Schrödinger equations:

Ĥ(r)R(r) = Eradial R(r),


Ĥ(θ, ϕ)Y (θ, ϕ) = Eangular Y (θ, ϕ).

Slide 242 · Models and approximations · Electronic structure of H atom


Solution of the angular problem
The angular problem concerns the motion of the electron along the
angular coordinates θ and ϕ.
The electron-nucleus attraction depends on the distance between
them but not on the direction along which we can find the electron.
The potential for motion along θ and ϕ is zero: V (θ, ϕ) = 0.
That is, the energy associated with angular motion is purely kinetic.

Slide 243 · Models and approximations · Electronic structure of H atom


Solution of the angular problem
The angular problem concerns the motion of the electron along the
angular coordinates θ and ϕ.
The electron-nucleus attraction depends on the distance between
them but not on the direction along which we can find the electron.
The potential for motion along θ and ϕ is zero: V (θ, ϕ) = 0.
That is, the energy associated with angular motion is purely kinetic.

Let us look at the problem in a systematic way:


• it involves rotational motion;

Slide 243 · Models and approximations · Electronic structure of H atom


Solution of the angular problem
The angular problem concerns the motion of the electron along the
angular coordinates θ and ϕ.
The electron-nucleus attraction depends on the distance between
them but not on the direction along which we can find the electron.
The potential for motion along θ and ϕ is zero: V (θ, ϕ) = 0.
That is, the energy associated with angular motion is purely kinetic.

Let us look at the problem in a systematic way:


• it involves rotational motion;
• the potential does not depend on the spatial orientation;

Slide 243 · Models and approximations · Electronic structure of H atom


Solution of the angular problem
The angular problem concerns the motion of the electron along the
angular coordinates θ and ϕ.
The electron-nucleus attraction depends on the distance between
them but not on the direction along which we can find the electron.
The potential for motion along θ and ϕ is zero: V (θ, ϕ) = 0.
That is, the energy associated with angular motion is purely kinetic.

Let us look at the problem in a systematic way:


• it involves rotational motion;
• the potential does not depend on the spatial orientation;
• the energy is purely kinetic;

Slide 243 · Models and approximations · Electronic structure of H atom


Solution of the angular problem
The angular problem concerns the motion of the electron along the
angular coordinates θ and ϕ.
The electron-nucleus attraction depends on the distance between
them but not on the direction along which we can find the electron.
The potential for motion along θ and ϕ is zero: V (θ, ϕ) = 0.
That is, the energy associated with angular motion is purely kinetic.

Let us look at the problem in a systematic way:


• it involves rotational motion;
• the potential does not depend on the spatial orientation;
• the energy is purely kinetic;
• the distance between the two particles is fixed.

Slide 243 · Models and approximations · Electronic structure of H atom


Solution of the angular problem
The angular problem concerns the motion of the electron along the
angular coordinates θ and ϕ.
The electron-nucleus attraction depends on the distance between
them but not on the direction along which we can find the electron.
The potential for motion along θ and ϕ is zero: V (θ, ϕ) = 0.
That is, the energy associated with angular motion is purely kinetic.

Let us look at the problem in a systematic way:


• it involves rotational motion;
• the potential does not depend on the spatial orientation;
• the energy is purely kinetic;
• the distance between the two particles is fixed.

This is a rigid rotor. No wonder the rigid rotor wavefunctions we’ve


seen looked so much like s, p and d atomic orbitals...
Slide 243 · Models and approximations · Electronic structure of H atom
Notation for electronic orbitals (1)

While electronic rotation around the hydrogen nucleus is no different


from that of a rigid rotor, the notation is:
1. We usually refer to the electronic motion as “orbital” rather than
rotational.
2. Because of the very large mass of the nucleus compared to that
of the electron, in the H atom case the only particle that moves in
the rotational motion is the electron. We say that the electron
“orbits” around the nucleus.
3. Because of (2), we refer to the angular momentum associated
with the rotational motion as the “orbital angular momentum”
of the electron.
4. The quantum number associated with the (orbital) angular
momentum of the electron is represented by ` rather than by J.

Slide 244 · Models and approximations · Electronic structure of H atom


Notation for electronic orbitals (1)

While electronic rotation around the hydrogen nucleus is no different


from that of a rigid rotor, the notation is:
1. We usually refer to the electronic motion as “orbital” rather than
rotational.
2. Because of the very large mass of the nucleus compared to that
of the electron, in the H atom case the only particle that moves in
the rotational motion is the electron. We say that the electron
“orbits” around the nucleus.
3. Because of (2), we refer to the angular momentum associated
with the rotational motion as the “orbital angular momentum”
of the electron.
4. The quantum number associated with the (orbital) angular
momentum of the electron is represented by ` rather than by J.

Slide 244 · Models and approximations · Electronic structure of H atom


Notation for electronic orbitals (1)

While electronic rotation around the hydrogen nucleus is no different


from that of a rigid rotor, the notation is:
1. We usually refer to the electronic motion as “orbital” rather than
rotational.
2. Because of the very large mass of the nucleus compared to that
of the electron, in the H atom case the only particle that moves in
the rotational motion is the electron. We say that the electron
“orbits” around the nucleus.
3. Because of (2), we refer to the angular momentum associated
with the rotational motion as the “orbital angular momentum”
of the electron.
4. The quantum number associated with the (orbital) angular
momentum of the electron is represented by ` rather than by J.

Slide 244 · Models and approximations · Electronic structure of H atom


Notation for electronic orbitals (1)

While electronic rotation around the hydrogen nucleus is no different


from that of a rigid rotor, the notation is:
1. We usually refer to the electronic motion as “orbital” rather than
rotational.
2. Because of the very large mass of the nucleus compared to that
of the electron, in the H atom case the only particle that moves in
the rotational motion is the electron. We say that the electron
“orbits” around the nucleus.
3. Because of (2), we refer to the angular momentum associated
with the rotational motion as the “orbital angular momentum”
of the electron.
4. The quantum number associated with the (orbital) angular
momentum of the electron is represented by ` rather than by J.

Slide 244 · Models and approximations · Electronic structure of H atom


Notation for electronic orbitals (2)

While electronic rotation around the hydrogen nucleus is no different


from that of a rigid rotor, the notation is:
5. The magnetic quantum number associated with the z component
of the (orbital) angular momentum of the electron is represented
by m` rather than by M.
6. When used for electronic orbital motion, the rigid rotor
wavefunctions — that is, the spherical harmonics Y`m` (θ, ϕ) —
are referred to as “atomic orbitals.”
7. Atomic orbitals associated with particular values of the orbital
angular momentum quantum number have special names.
Orbitals associated with ` = 0, 1, 2, 3 and 4 are called,
respectively, s, p, d , f and g orbitals.

Slide 245 · Models and approximations · Electronic structure of H atom


Notation for electronic orbitals (2)

While electronic rotation around the hydrogen nucleus is no different


from that of a rigid rotor, the notation is:
5. The magnetic quantum number associated with the z component
of the (orbital) angular momentum of the electron is represented
by m` rather than by M.
6. When used for electronic orbital motion, the rigid rotor
wavefunctions — that is, the spherical harmonics Y`m` (θ, ϕ) —
are referred to as “atomic orbitals.”
7. Atomic orbitals associated with particular values of the orbital
angular momentum quantum number have special names.
Orbitals associated with ` = 0, 1, 2, 3 and 4 are called,
respectively, s, p, d , f and g orbitals.

Slide 245 · Models and approximations · Electronic structure of H atom


Notation for electronic orbitals (2)

While electronic rotation around the hydrogen nucleus is no different


from that of a rigid rotor, the notation is:
5. The magnetic quantum number associated with the z component
of the (orbital) angular momentum of the electron is represented
by m` rather than by M.
6. When used for electronic orbital motion, the rigid rotor
wavefunctions — that is, the spherical harmonics Y`m` (θ, ϕ) —
are referred to as “atomic orbitals.”
7. Atomic orbitals associated with particular values of the orbital
angular momentum quantum number have special names.
Orbitals associated with ` = 0, 1, 2, 3 and 4 are called,
respectively, s, p, d , f and g orbitals.

Slide 245 · Models and approximations · Electronic structure of H atom


Notation for electronic orbitals (2)

While electronic rotation around the hydrogen nucleus is no different


from that of a rigid rotor, the notation is:
5. The magnetic quantum number associated with the z component
of the (orbital) angular momentum of the electron is represented
by m` rather than by M.
6. When used for electronic orbital motion, the rigid rotor
wavefunctions — that is, the spherical harmonics Y`m` (θ, ϕ) —
are referred to as “atomic orbitals.”
7. Atomic orbitals associated with particular values of the orbital
angular momentum quantum number have special names.
Orbitals associated with ` = 0, 1, 2, 3 and 4 are called,
respectively, s, p, d , f and g orbitals.

That’s it! Apart from the notation, nothing new. We’re ready to move on
to the radial problem. But before that...

Slide 245 · Models and approximations · Electronic structure of H atom


Food for thought

What keeps an electron in an s orbital away from the nucleus?

Slide 246 · Models and approximations · Electronic structure of H atom


Food for thought

What keeps an electron in an s orbital away from the nucleus?


Remember that the opposite charges of nucleus and electron
cause them to attract each other.

Slide 246 · Models and approximations · Electronic structure of H atom


Food for thought

What keeps an electron in an s orbital away from the nucleus?


Remember that the opposite charges of nucleus and electron
cause them to attract each other.
In the early days of QM, people thought that the centrifugal force
associated with its orbital motion was what prevented the electron
from collapsing into the atomic nucleus.

Slide 246 · Models and approximations · Electronic structure of H atom


Food for thought

What keeps an electron in an s orbital away from the nucleus?


Remember that the opposite charges of nucleus and electron
cause them to attract each other.
In the early days of QM, people thought that the centrifugal force
associated with its orbital motion was what prevented the electron
from collapsing into the atomic nucleus.
Trouble is, an electron in an s orbital has ` = 0. In other terms, its
orbital angular momentum is zero; electrons in s do not orbit the
nucleus.

Slide 246 · Models and approximations · Electronic structure of H atom


Food for thought

What keeps an electron in an s orbital away from the nucleus?


Remember that the opposite charges of nucleus and electron
cause them to attract each other.
In the early days of QM, people thought that the centrifugal force
associated with its orbital motion was what prevented the electron
from collapsing into the atomic nucleus.
Trouble is, an electron in an s orbital has ` = 0. In other terms, its
orbital angular momentum is zero; electrons in s do not orbit the
nucleus.
And if there’s no orbiting, there can be no centrifugal force.

Slide 246 · Models and approximations · Electronic structure of H atom


Food for thought

What keeps an electron in an s orbital away from the nucleus?


Remember that the opposite charges of nucleus and electron
cause them to attract each other.
In the early days of QM, people thought that the centrifugal force
associated with its orbital motion was what prevented the electron
from collapsing into the atomic nucleus.
Trouble is, an electron in an s orbital has ` = 0. In other terms, its
orbital angular momentum is zero; electrons in s do not orbit the
nucleus.
And if there’s no orbiting, there can be no centrifugal force.
So... what keeps an electron in an s orbital away from the
nucleus?

Slide 246 · Models and approximations · Electronic structure of H atom


Solution of the radial problem (1)
The radial problem concerns electronic motion along the radial
coordinate, r.
In contrast to the angular problem, the potential here is not zero.
Instead, we have V (r) = VeN (r) ∝ −1/r.
That is, the energy associated with radial motion has both kinetic and
potential components.

Slide 247 · Models and approximations · Electronic structure of H atom


Solution of the radial problem (1)
The radial problem concerns electronic motion along the radial
coordinate, r.
In contrast to the angular problem, the potential here is not zero.
Instead, we have V (r) = VeN (r) ∝ −1/r.
That is, the energy associated with radial motion has both kinetic and
potential components.

While the explicit expressions for the radial wavefunctions can be


complicated, their general form is not:
Zr
− na
Rn,` (r) = Nn,` × polynomialn,` (r) × e 0 ,

where the new symbols are:


• n, the principal quantum number. We can have n = 1, 2, 3, . . .
• a0 , the Bohr radius. Its value is a0 = 5.291 × 10−11 m.
• Nn,` , a normalisation constant.
Slide 247 · Models and approximations · Electronic structure of H atom
Solution of the radial problem (2)

Some important observations:


1. Radial wavefunctions introduce a new quantum number
(the principal quantum number, n).
2. Radial wavefunctions determine the radial probability density.
It tell us how likely we are to find the electron
at any given distance from the nucleus.
3. Because they are associated with the electron-nucleus distance
(the radius), one can say that radial wavefunctions determine
the size (but not the shape) of the atomic orbital.
4. Besides depending on the principal quantum number, n, radial
wavefunctions also depend on the angular momentum quantum
number `. See Slide 251 for the explanation.

Slide 248 · Models and approximations · Electronic structure of H atom


Solution of the radial problem (2)

Some important observations:


1. Radial wavefunctions introduce a new quantum number
(the principal quantum number, n).
2. Radial wavefunctions determine the radial probability density.
It tell us how likely we are to find the electron
at any given distance from the nucleus.
3. Because they are associated with the electron-nucleus distance
(the radius), one can say that radial wavefunctions determine
the size (but not the shape) of the atomic orbital.
4. Besides depending on the principal quantum number, n, radial
wavefunctions also depend on the angular momentum quantum
number `. See Slide 251 for the explanation.

Slide 248 · Models and approximations · Electronic structure of H atom


Solution of the radial problem (2)

Some important observations:


1. Radial wavefunctions introduce a new quantum number
(the principal quantum number, n).
2. Radial wavefunctions determine the radial probability density.
It tell us how likely we are to find the electron
at any given distance from the nucleus.
3. Because they are associated with the electron-nucleus distance
(the radius), one can say that radial wavefunctions determine
the size (but not the shape) of the atomic orbital.
4. Besides depending on the principal quantum number, n, radial
wavefunctions also depend on the angular momentum quantum
number `. See Slide 251 for the explanation.

Slide 248 · Models and approximations · Electronic structure of H atom


Solution of the radial problem (2)

Some important observations:


1. Radial wavefunctions introduce a new quantum number
(the principal quantum number, n).
2. Radial wavefunctions determine the radial probability density.
It tell us how likely we are to find the electron
at any given distance from the nucleus.
3. Because they are associated with the electron-nucleus distance
(the radius), one can say that radial wavefunctions determine
the size (but not the shape) of the atomic orbital.
4. Besides depending on the principal quantum number, n, radial
wavefunctions also depend on the angular momentum quantum
number `. See Slide 251 for the explanation.

Slide 248 · Models and approximations · Electronic structure of H atom


Solution of the radial problem (3)
Some important observations, cont.:
4. Solution of the radial problem completes the solution of the H atom
problem, and gives us the total energy of each electronic orbital:

2
hcR∞ Z
En = − , n = 1, 2, 3, · · ·
n2

where R∞ = 109, 737 cm−1 is the Rydberg constant.


5. Using this result we can calculate the energy gap between any
two orbitals of the H atom:
!
1 1
En2 − En1 = hcR∞ − ,
n21 n22

a formula you’ve met in previous courses.


Slide 249 · Models and approximations · Electronic structure of H atom
Solution of the radial problem (3)
Some important observations, cont.:
4. Solution of the radial problem completes the solution of the H atom
problem, and gives us the total energy of each electronic orbital:

2
hcR∞ Z
En = − , n = 1, 2, 3, · · ·
n2

where R∞ = 109, 737 cm−1 is the Rydberg constant.


5. Using this result we can calculate the energy gap between any
two orbitals of the H atom:
!
1 1
En2 − En1 = hcR∞ − ,
n21 n22

a formula you’ve met in previous courses.


Slide 249 · Models and approximations · Electronic structure of H atom
Solution of the radial problem (4)

Some important observations, cont.:


6. Contrary to what happened in the case of a rigid rotor, here there
is an upper limit for the angular momentum quantum number. We
have 0 ≤ ` ≤ n − 1. That is,

` = 0, 1, 2, 3, . . . n − 1.

7. Contrary to what happened in the case of a rigid rotor, here the


energy is independent of the angular momentum quantum
number. See Slide 251 for the explanation.

Slide 250 · Models and approximations · Electronic structure of H atom


Solution of the radial problem (4)

Some important observations, cont.:


6. Contrary to what happened in the case of a rigid rotor, here there
is an upper limit for the angular momentum quantum number. We
have 0 ≤ ` ≤ n − 1. That is,

` = 0, 1, 2, 3, . . . n − 1.

7. Contrary to what happened in the case of a rigid rotor, here the


energy is independent of the angular momentum quantum
number. See Slide 251 for the explanation.

Slide 250 · Models and approximations · Electronic structure of H atom


Dependence on quantum numbers
The fact that the total energy depends only on n, and not on `, does
not mean that orbital motion does not contribute to the total energy.
The key idea is this: the principal quantum number is not there to
determine what the radial motion and radial energy will be.
Instead, its role is to determine what the total energy is.
We know (from the rigid rotor results) what the energy due to orbital
motion is: E` = (ħ2 /2I)`(` + 1).
This means that the (kinetic + potential) energy in the radial
degree of freedom is given by En − E` .
Remember this:
• En is the total energy, not the energy associated with the radial
degree of freedom.
• Since the radial energy is En − E` , it’s natural that radial
wavefunctions depend not only on n, but also on `.
Slide 251 · Models and approximations · Electronic structure of H atom
Dependence on quantum numbers
The fact that the total energy depends only on n, and not on `, does
not mean that orbital motion does not contribute to the total energy.
The key idea is this: the principal quantum number is not there to
determine what the radial motion and radial energy will be.
Instead, its role is to determine what the total energy is.
We know (from the rigid rotor results) what the energy due to orbital
motion is: E` = (ħ2 /2I)`(` + 1).
This means that the (kinetic + potential) energy in the radial
degree of freedom is given by En − E` .
Remember this:
• En is the total energy, not the energy associated with the radial
degree of freedom.
• Since the radial energy is En − E` , it’s natural that radial
wavefunctions depend not only on n, but also on `.
Slide 251 · Models and approximations · Electronic structure of H atom
Dependence on quantum numbers
The fact that the total energy depends only on n, and not on `, does
not mean that orbital motion does not contribute to the total energy.
The key idea is this: the principal quantum number is not there to
determine what the radial motion and radial energy will be.
Instead, its role is to determine what the total energy is.
We know (from the rigid rotor results) what the energy due to orbital
motion is: E` = (ħ2 /2I)`(` + 1).
This means that the (kinetic + potential) energy in the radial
degree of freedom is given by En − E` .
Remember this:
• En is the total energy, not the energy associated with the radial
degree of freedom.
• Since the radial energy is En − E` , it’s natural that radial
wavefunctions depend not only on n, but also on `.
Slide 251 · Models and approximations · Electronic structure of H atom
Radial wavefunctions

As shown in slide 247, the radial wavefunctions are given (except for a
constant factor) by the product of a polynomial and an exponentially
decaying term.
At large enough r, the exponential decay always prevails, and the
radial wavefunction approaches zero.
At smaller r, the radial wavefunction can in general have positive and
negative values, with the number of nodes matching the degree of the
polynomial:
number of nodes: n − ` − 1.

Slide 252 · Models and approximations · Electronic structure of H atom


Example 1: radial wavefunctions
of a pair of s orbitals: R3s (r ) and R5s (r )

• 3s orbital: n = 3, ` = 0.
Nodes: n − ` − 1 = 2.
hri = 13.5a0 = 0.71 nm.

• 5s orbital: n = 5, ` = 0.
Nodes: n − ` − 1 = 4.
hri = 36.5a0 = 1.9 nm.

(note maxima at r = 0. Electron at nucleus?)

Slide 253 · Models and approximations · Electronic structure of H atom


Example 2: probability densities
2 2
of a pair of s orbitals: |R3s (r )| and |R5s (r )|

• 3s orbital: n = 3, ` = 0.
Nodes: n − ` − 1 = 2.
hri = 13.5a0 = 0.71 nm.

• 5s orbital: n = 5, ` = 0.
Nodes: n − ` − 1 = 4.
hri = 36.5a0 = 1.9 nm.

(note that profiles have become very different, and that


probability densities at nucleus are zero. How come?)
Slide 254 · Models and approximations · Electronic structure of H atom
Example 3: radial wavefunctions
of a pair of n = 3 orbitals: R3s (r ) and R3d (r )

• 3s orbital: n = 3, ` = 0.
Nodes: n − ` − 1 = 2.
hri = 13.5a0 = 0.71 nm.

• 3d orbital: n = 3, ` = 2.
Nodes: n − ` − 1 = 0.
hri = 10.5a0 = 0.56 nm.

(note that maximum at r = 0 is there only for s orbital)

Slide 255 · Models and approximations · Electronic structure of H atom


Example 4: probability densities
2 2
of a pair of n = 3 orbitals: |R3s (r )| and |R3d (r )|

• 3s orbital: n = 3, ` = 0.
Nodes: n − ` − 1 = 2.
hri = 13.5a0 = 0.71 nm.

• 3d orbital: n = 3, ` = 2.
Nodes: n − ` − 1 = 0.
hri = 10.5a0 = 0.56 nm.

(note again how profiles have become very different,


and how most probable radius changes with `)
Slide 256 · Models and approximations · Electronic structure of H atom
Summary of results

The H atom energy and wavefunction are given by:

hcR∞
En = − ,
n2
ψn,`,m` (r, θ, ϕ) = Rn,` (r)Y`,m` (θ, ϕ),
n = 1, 2, 3, . . .
` = 0, 1, 2, . . . , n − 1
m` = 0, ±1, ±2, . . . , ±`

Now an important question: how many degenerate states


are there for each energy level?

Slide 257 · Models and approximations · Electronic structure of H atom


Degeneracy

The energy of a given orbital depends on the principal quantum


number (n) but is independent of the orbital angular momentum (`)
and magnetic (m` ) quantum numbers.
For each n value, there are n possible values for `,
ranging from ` = 0 to ` = n − 1.
For each ` value, there are 2` + 1 possible values for m` ,
ranging from m` = −` to m` = `.
Conclusion: for each energy level (ie, each n value)
of the hydrogen atom, there are n2 degenerate orbitals.
n−1
X
(this comes from (2` + 1) = n2 )
`=0

Slide 258 · Models and approximations · Electronic structure of H atom


Degeneracy

The energy of a given orbital depends on the principal quantum


number (n) but is independent of the orbital angular momentum (`)
and magnetic (m` ) quantum numbers.
For each n value, there are n possible values for `,
ranging from ` = 0 to ` = n − 1.
For each ` value, there are 2` + 1 possible values for m` ,
ranging from m` = −` to m` = `.
Conclusion: for each energy level (ie, each n value)
of the hydrogen atom, there are n2 degenerate orbitals.
n−1
X
(this comes from (2` + 1) = n2 )
`=0

Slide 258 · Models and approximations · Electronic structure of H atom


Degeneracy

The energy of a given orbital depends on the principal quantum


number (n) but is independent of the orbital angular momentum (`)
and magnetic (m` ) quantum numbers.
For each n value, there are n possible values for `,
ranging from ` = 0 to ` = n − 1.
For each ` value, there are 2` + 1 possible values for m` ,
ranging from m` = −` to m` = `.
Conclusion: for each energy level (ie, each n value)
of the hydrogen atom, there are n2 degenerate orbitals.
n−1
X
(this comes from (2` + 1) = n2 )
`=0

Slide 258 · Models and approximations · Electronic structure of H atom


Energy levels of H atom

0
hcR∞ • energies proportional to hcR∞
− 9
• energies inversely
electronic energy, En

hcR∞
− 2
4
proportional to n
• energy gaps decrease with n
• ` values in the
0 ≤ ` ≤ n − 1 range
• 2` + 1 states per ` value
(not shown on diagram)

hcR∞ • level degeneracy is n2
1

• ionization limit at E∞ = 0

Slide 259 · Models and approximations · Electronic structure of H atom


Energy levels of H atom

0 ∞
• energies proportional to hcR∞

principal quantum number, n


4
hcR
− 9∞ 3
• energies inversely
electronic energy, En

hcR∞
− 2 2
4
proportional to n
• energy gaps decrease with n
• ` values in the
0 ≤ ` ≤ n − 1 range
• 2` + 1 states per ` value
(not shown on diagram)

hcR∞
1
• level degeneracy is n2
1

• ionization limit at E∞ = 0

Slide 259 · Models and approximations · Electronic structure of H atom


Energy levels of H atom

0 ∞
• energies proportional to hcR∞

principal quantum number, n


4
hcR
− 9∞ 3
• energies inversely
electronic energy, En

hcR∞
− 2 2
4
proportional to n
• energy gaps decrease with n
• ` values in the
0 ≤ ` ≤ n − 1 range
• 2` + 1 states per ` value
(not shown on diagram)

hcR∞
1
• level degeneracy is n2
1

• ionization limit at E∞ = 0

Slide 259 · Models and approximations · Electronic structure of H atom


Energy levels of H atom

0 ∞
• energies proportional to hcR∞

principal quantum number, n


4
hcR
− 9∞ 3
• energies inversely
electronic energy, En

hcR∞
− 2 2
4
proportional to n
• energy gaps decrease with n
• ` values in the
0 ≤ ` ≤ n − 1 range
• 2` + 1 states per ` value
(not shown on diagram)

hcR∞
1
• level degeneracy is n2
1
0 1 2 3 4 5 6 • ionization limit at E∞ = 0
angular momentum quantum number, `

Slide 259 · Models and approximations · Electronic structure of H atom


Energy levels of H atom

0 ∞
• energies proportional to hcR∞

principal quantum number, n


4
hcR
− 9∞ 3
• energies inversely
electronic energy, En

hcR∞
− 2 2
4
proportional to n
• energy gaps decrease with n
• ` values in the
0 ≤ ` ≤ n − 1 range
• 2` + 1 states per ` value
(not shown on diagram)

hcR∞
1
• level degeneracy is n2
1
0 1 2 3 4 5 6 • ionization limit at E∞ = 0
angular momentum quantum number, `

Slide 259 · Models and approximations · Electronic structure of H atom


Energy levels of H atom

0 ∞
• energies proportional to hcR∞

principal quantum number, n


4
hcR
− 9∞ 3
• energies inversely
electronic energy, En

hcR∞
− 2 2
4
proportional to n
• energy gaps decrease with n
• ` values in the
0 ≤ ` ≤ n − 1 range
• 2` + 1 states per ` value
(not shown on diagram)

hcR∞
1
• level degeneracy is n2
1
0 1 2 3 4 5 6 • ionization limit at E∞ = 0
angular momentum quantum number, `

Slide 259 · Models and approximations · Electronic structure of H atom


Energy levels of H atom

0 ∞
• energies proportional to hcR∞

principal quantum number, n


4
hcR
− 9∞ 3
• energies inversely
electronic energy, En

hcR∞
− 2 2
4
proportional to n
• energy gaps decrease with n
• ` values in the
0 ≤ ` ≤ n − 1 range
• 2` + 1 states per ` value
(not shown on diagram)

hcR∞
1
• level degeneracy is n2
1
0 1 2 3 4 5 6 • ionization limit at E∞ = 0
angular momentum quantum number, `

Slide 259 · Models and approximations · Electronic structure of H atom


Hybrid orbitals

If we have a set of N degenerate wavefunctions, we can always


combine these wavefunctions in ways that lead to
new, different sets of N degenerate wavefunctions
that are equally acceptable solutions of the Schrödinger equation.
When describing electronic orbitals, these new wavefunctions
are called hybrid orbitals.

Slide 260 · Models and approximations · Electronic structure of H atom


Hybrid orbitals

If we have a set of N degenerate wavefunctions, we can always


combine these wavefunctions in ways that lead to
new, different sets of N degenerate wavefunctions
that are equally acceptable solutions of the Schrödinger equation.
When describing electronic orbitals, these new wavefunctions
are called hybrid orbitals.
Let us examine some examples you’re already familiar with.
In each case, we will take a set of degenerate atomic orbitals and
combine them in order to end up with different atomic orbitals
having the same energy as those in the initial set.

Slide 260 · Models and approximations · Electronic structure of H atom


Example 1: px , py and pz orbitals (1)
As they are all associated with n = 2,
the 2p−1 , 2p0 and 2p+1 orbitals of hydrogen are degenerate.
They form a set of three degenerate atomic orbitals
with energy E2 = −hcR/4.
We can recombine them and obtain a set of new atomic orbitals
with energy E2 = −hcR/4.

Slide 261 · Models and approximations · Electronic structure of H atom


Example 1: px , py and pz orbitals (1)
As they are all associated with n = 2,
the 2p−1 , 2p0 and 2p+1 orbitals of hydrogen are degenerate.
They form a set of three degenerate atomic orbitals
with energy E2 = −hcR/4.
We can recombine them and obtain a set of new atomic orbitals
with energy E2 = −hcR/4. For example (but not only!) like this:

1
ψ2px = √ ψ2p−1 + ψ2p+1

2
i
ψ2py = √ ψ2p−1 − ψ2p+1

2
ψ2pz = ψ2p0

The reason for calling these new orbitals “2px ,” “2py ,” and “2pz ” will
become obvious when we look at their angular parts (next slide).
Slide 261 · Models and approximations · Electronic structure of H atom
Example 1: px , py and pz orbitals (2)

ψ2px ψ2py ψ2pz

• colour coding as for rotational eigenfunctions;


• all three orbitals are real.

Slide 262 · Models and approximations · Electronic structure of H atom


3
Example 2: sp orbitals (1)
As they are all associated with n = 2,
the 2s, 2px , 2py and 2pz orbitals of hydrogen are degenerate.
They form a set of three degenerate atomic orbitals
with energy E2 = −hcR/4.
We can recombine them and obtain a set of new atomic orbitals
with energy E2 = −hcR/4.

Slide 263 · Models and approximations · Electronic structure of H atom


3
Example 2: sp orbitals (1)
As they are all associated with n = 2,
the 2s, 2px , 2py and 2pz orbitals of hydrogen are degenerate.
They form a set of three degenerate atomic orbitals
with energy E2 = −hcR/4.
We can recombine them and obtain a set of new atomic orbitals
with energy E2 = −hcR/4. For example (but not only!) like this:
1
ψa = ψ2s + ψ2px + ψ2py + ψ2pz

2
1
ψb = ψ2s + ψ2px − ψ2py − ψ2pz

2
1
ψc = ψ2s − ψ2px + ψ2py − ψ2pz

2
1
ψd = ψ2s − ψ2px − ψ2py + ψ2pz

2
The next slide shows the angular part of these orbitals.
Slide 263 · Models and approximations · Electronic structure of H atom
3
Example 2: sp orbitals (2)

ψa ψb • colour coding as for


rotational
wavefunctions;
• all four orbitals are real.

ψc ψd

Slide 264 · Models and approximations · Electronic structure of H atom


Hybrid orbitals: food for thought

When we mix degenerate wavefunctions to generate hybrid orbitals,


the energy does not change. But...
• is there any change in the radial part of the orbitals?
• is there any change in the angular part of the orbitals?
• is there any change in angular momentum?
• is there any change in angular momentum projection on z?
• is there any change in the way electrons move?
• is there any change in the way the atom is likely to form chemical
bonds?

Slide 265 · Models and approximations · Electronic structure of H atom


A possible answer?

Slide 266 · Models and approximations · Electronic structure of H atom


Part II — Models and approximations

11 Molecular translation
Another type of 12 Molecular vibration
angular momentum: 13 Molecular rotation

14 Electronic structure of H atom


spin
15 Spin

Slide 267 · Models and approximations · Spin


Spin

There’s one last thing that affects the states of atoms and molecules
that we haven’t seen yet: spin.
The reason for leaving it to this late stage is that we will not
discuss the origin of spin — we won’t see the explanation for it.
Contrary to popular belief, spin has nothing to do with particles
“spinning” around themselves.
Instead, spin results from a combination of quantum mechanical
and relativistic effects that we aren’t equipped to consider.
There is, something we are equipped to consider.
This: spin is an angular momentum.

Slide 268 · Models and approximations · Spin


Spin v. other angular momenta: differences

• contrary to other angular momenta, the value of the spin angular


momentum is fixed. In other words, every particle has an intrinsic
spin that, just like its charge, cannot be changed.
• contrary to other angular momenta, spin eigenfunctions are not
spherical harmonics. As it turns out, there’s no analytical
expression for spin eigenfunctions. The reason for this is that spin
is not associated with rotation along an angular coordinate.
• contrary to rotational and orbital angular momenta, spin can be
associated with half-integer quantum numbers. Electronic spin, for
instance, is always described by an electronic spin quantum
number s = 1/2, with spin projection on a given axis being
described by quantum numbers ms = −1/2 (aka “spin down”) or
ms = +1/2 (aka “spin up”).

Slide 269 · Models and approximations · Spin


Spin v. other angular momenta: differences

• contrary to other angular momenta, the value of the spin angular


momentum is fixed. In other words, every particle has an intrinsic
spin that, just like its charge, cannot be changed.
• contrary to other angular momenta, spin eigenfunctions are not
spherical harmonics. As it turns out, there’s no analytical
expression for spin eigenfunctions. The reason for this is that spin
is not associated with rotation along an angular coordinate.
• contrary to rotational and orbital angular momenta, spin can be
associated with half-integer quantum numbers. Electronic spin, for
instance, is always described by an electronic spin quantum
number s = 1/2, with spin projection on a given axis being
described by quantum numbers ms = −1/2 (aka “spin down”) or
ms = +1/2 (aka “spin up”).

Slide 269 · Models and approximations · Spin


Spin v. other angular momenta: differences

• contrary to other angular momenta, the value of the spin angular


momentum is fixed. In other words, every particle has an intrinsic
spin that, just like its charge, cannot be changed.
• contrary to other angular momenta, spin eigenfunctions are not
spherical harmonics. As it turns out, there’s no analytical
expression for spin eigenfunctions. The reason for this is that spin
is not associated with rotation along an angular coordinate.
• contrary to rotational and orbital angular momenta, spin can be
associated with half-integer quantum numbers. Electronic spin, for
instance, is always described by an electronic spin quantum
number s = 1/2, with spin projection on a given axis being
described by quantum numbers ms = −1/2 (aka “spin down”) or
ms = +1/2 (aka “spin up”).

Slide 269 · Models and approximations · Spin


Spin v. other angular momenta: similarities

• the commutation rules involving spin components are exactly the


same as those for components of other angular momenta:

[ŝx , ŝy ] = i ħŝz , [ŝy , ŝz ] = i ħŝx , [ŝz , ŝx ] = i ħŝy .

• because of what’s above, the uncertainty relations involving spin


components are exactly the same as those for components of
other angular momenta:

ħ ħ ħ
∆sx ∆sy ≥ | hs i |2 , ∆sy ∆sz ≥ | hs i |2 , ∆sz ∆sx ≥ | hs i |2 .
2 z 2 x 2 y
If we measure the spin component along (say) the z direction, we
loose all information about the other spin components.

Slide 270 · Models and approximations · Spin


Spin v. other angular momenta: similarities

• the commutation rules involving spin components are exactly the


same as those for components of other angular momenta:

[ŝx , ŝy ] = i ħŝz , [ŝy , ŝz ] = i ħŝx , [ŝz , ŝx ] = i ħŝy .

• because of what’s above, the uncertainty relations involving spin


components are exactly the same as those for components of
other angular momenta:

ħ ħ ħ
∆sx ∆sy ≥ | hs i |2 , ∆sy ∆sz ≥ | hs i |2 , ∆sz ∆sx ≥ | hs i |2 .
2 z 2 x 2 y
If we measure the spin component along (say) the z direction, we
loose all information about the other spin components.

Slide 270 · Models and approximations · Spin


Spin v. other angular momenta: similarities

• the commutation rules involving spin components are exactly the


same as those for components of other angular momenta:

[ŝx , ŝy ] = i ħŝz , [ŝy , ŝz ] = i ħŝx , [ŝz , ŝx ] = i ħŝy .

• because of what’s above, the uncertainty relations involving spin


components are exactly the same as those for components of
other angular momenta:

ħ ħ ħ
∆sx ∆sy ≥ | hs i |2 , ∆sy ∆sz ≥ | hs i |2 , ∆sz ∆sx ≥ | hs i |2 .
2 z 2 x 2 y
If we measure the spin component along (say) the z direction, we
loose all information about the other spin components.
It is because of the commutation rules between its components that
we say that spin is an angular momentum.
Slide 270 · Models and approximations · Spin
If that’s the case...

Slide 271 · Models and approximations · Spin


Slide 272 · Models and approximations · Spin

You might also like