0% found this document useful (0 votes)
85 views13 pages

ph217 Vip

This document provides lecture notes on quantum mechanics. It discusses quantum mechanical measurement and expectation values, including how the average value of energy is calculated. It also covers operators and expectation values, defining the position and momentum operators and how their expectation values are calculated by operating those operators on the wavefunction.

Uploaded by

Benson Shayo
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
85 views13 pages

ph217 Vip

This document provides lecture notes on quantum mechanics. It discusses quantum mechanical measurement and expectation values, including how the average value of energy is calculated. It also covers operators and expectation values, defining the position and momentum operators and how their expectation values are calculated by operating those operators on the wavefunction.

Uploaded by

Benson Shayo
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 13

Quantum Mechanics- Lecture Note

College of Science/Physics Department


Chapter Three First Semester 2018-2019

1. Quantum mechanical measurement and expectation values


When a normalized wavefunction is expanded in an orthonormal set, e.g.,

Ψ(𝑟, 𝑡) = ∑ 𝑐𝑛 (𝑡)𝜓𝑛 (𝑟)


𝑛
then the normalization integral requires that
+∞ +∞
∫ |𝛹(𝑟, 𝑡)|2 𝑑3 𝑟 = ∫ [∑ 𝑐𝑛∗ (𝑡)𝜓𝑛∗ (𝑟)] [∑ 𝑐𝑚 (𝑡)𝜓𝑚 (𝑟)] 𝑑3 𝑟
−∞ −∞ 𝑛 𝑚
If we look at the integral over the sums, we see that because of the orthogonality of the
basis functions, the only terms that will survive after integration will be for 𝑛 = 𝑚 , and
because of the orthonormality of the basis functions, the result from any such term in the
integration will simply be |𝑐𝑛 (𝑡)|2 . Hence, we have

∑|𝑐𝑛 (𝑡)|2 = 1
𝑛
In quantum mechanics, when we make a measurement on a small system with a large
measuring apparatus, of some quantity such as energy, we find the following behavior,
which is sometimes elevated to a postulate or hypothesis in quantum mechanics:
On measurement, the system collapses into an eigenstate of the quantity being measured,
with probability
𝑃𝑛 = |𝑐𝑛 |2
where 𝑐𝑛 is the expansion coefficient in the (orthonormal) eigenfunctions of the quantity
being measured.
Suppose now that we measure the energy of our system in such an experiment. We could
repeat the experiment many times and get a statistical distribution of results. Given the
probabilities, we would find in the usual way that the average value of energy 𝐸 that we
would measure would be

〈𝐸〉 = ∑ 𝐸𝑛 𝑃𝑛 = ∑ 𝐸𝑛 |𝑐𝑛 |2
𝑛 𝑛
where we are using the notation E to denote the average value of 𝐸, a quantity we call the
“expectation value of 〈𝐸〉 in quantum mechanics, the 𝐸𝑛 are the energy eigenvalues.)
For example, for the coherent state discussed above with parameter 𝑁, we have

𝑁 𝑛 exp(−𝑁)
〈𝐸〉 = ∑ 𝐸𝑛
𝑛!
𝑁=0

𝑁 𝑛 exp(−𝑁) 1
= ℏ𝜔 [∑ 𝑛 ] + ℏ𝜔
𝑛! 2
𝑁=0
1
= (𝑁 + ) ℏ𝜔
2

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


1
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

We can show that having an energy ≈ 𝑁ℏ𝜔 for the large 𝑁 implicit in a classical situation
corresponds very well to our notions of energy, frequency and oscillation amplitude in a
classical oscillator. Note that N is not restricted to being an integer – it can take on any real
value. Quite generally, the expectation value of the energy or any other quantum-mechanically
measurable quantity is not restricted to being one of a discrete set of values – it can take on any
real value within the physically possible range of the quantity in question.

2. Operators and expectation values

We can now show an important but simple relation between the Hamiltonian operator, the
wavefunction, and the expectation value of the energy. Consider the integral
̂ Ψ(𝑟, 𝑡)𝑑3 𝑟
𝐼 = ∫ 𝛹 ∗ (𝑟, 𝑡) 𝐻
Where Ψ(𝑟, 𝑡) is the wavefunction of some system of interest. We can expand this
wavefunction in energy eigenstates, as in Ψ(𝑟, 𝑡) = ∑𝑛 𝑐𝑛 (𝑡)𝜓𝑛 (𝑟). We know that, with 𝜓𝑛 (𝑟)
as the energy eigenstates (of the time-independent Schrödinger equation)
ℏ2 2 ℏ2 2
𝐻̂ 𝛹(𝑟, 𝑡) = [− ∇ + 𝑉(𝑟, 𝑡)] 𝛹(𝑟, 𝑡) = [− 𝛻 + 𝑉(𝑟, 𝑡)] ∑ 𝑐𝑛 (𝑡)𝜓𝑛 (𝑟)
2𝑚 2𝑚
𝑛

= ∑ 𝑐𝑛 (𝑡)𝐸𝑛 𝜓𝑛 (𝑟)
𝑛
And also
+∞
̂ Ψ(𝑟, 𝑡)𝑑3 𝑟 = ∫
∫ 𝛹 ∗ (𝑟, 𝑡) 𝐻 ∗ (𝑡)𝜓 ∗ (𝑟)
[∑ 𝑐𝑚 𝑚 ] [∑ 𝑐𝑛 (𝑡)𝐸𝑛 𝜓𝑛 (𝑟)] 𝑑 3 𝑟
−∞ 𝑚 𝑛
Given the orthonormality of the 𝜓𝑛 (𝑟) , we have
̂ Ψ(𝑟, 𝑡)𝑑3 𝑟 = ∑ 𝐸𝑛 |𝑐𝑛 |2
∫ 𝛹 ∗ (𝑟, 𝑡) 𝐻
𝑛
The expectation value of the energy is
̂ 𝛹(𝑟, 𝑡)𝑑 3 𝑟
〈𝐸〉 = ∫ 𝛹 ∗ (𝑟, 𝑡) 𝐻
The expectation value of the position is:
〈𝑥〉 = ∫ 𝑥 |𝛹(𝑥, 𝑡)|2 𝑑𝑥
+∞ +∞

〈𝑥〉 = ∫ 𝑥 𝛹 ∗ (𝑥, 𝑡) 𝛹(𝑥, 𝑡) 𝑑𝑥 = ∫ 𝛹 ∗ (𝑥, 𝑡) 𝑥 𝛹(𝑥, 𝑡) 𝑑𝑥


−∞ −∞

The expectation value of the momentum is


+∞
𝑑 𝑑
〈𝑝〉 = 𝑚 〈𝑥〉 = 𝑚 ∫ 𝑑𝑥 𝛹 ∗ (𝑥, 𝑡) 𝑥 𝛹(𝑥, 𝑡)
𝑑𝑡 𝑑𝑡
−∞

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


2
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

+∞
𝜕 𝛹 ∗ (𝑥, 𝑡) 𝜕𝛹(𝑥, 𝑡)
= 𝑚 ∫ 𝑑𝑥 ( 𝑥𝛹(𝑥, 𝑡) + 𝛹 ∗ (𝑥, 𝑡) 𝑥 )
𝜕𝑡 𝜕𝑡
−∞
𝑑〈𝑥〉
is the velocity of the expectation value of 𝑥, not the velocity of the particle.
𝑑𝑡
𝑑𝑥
Since classically 𝑝 = 𝑚𝑣 = 𝑚 𝑑𝑡 .

Note that there is no 𝑑𝑥⁄𝑑𝑡 under the integral sign. The only quantity that varies with time is
𝛹(𝑥, 𝑡), and it is this variation that gives rise to a change in 〈𝑥〉 with time. Use the Schrodinger
equation and its complex conjugate to evaluate the above and we have
+∞
ℏ 𝜕 2𝛹∗ ∗
𝜕 2𝛹
〈𝑝〉 = ∫ 𝑑𝑥 ( 2 𝑥𝛹 − 𝛹 𝑥 )
2𝑖 𝜕𝑥 𝜕𝑥 2
−∞

Now
𝜕2 𝛹∗ 𝜕 𝜕𝛹 ∗ 𝜕𝛹 ∗ 𝜕𝛹 ∗ 𝜕𝛹
𝑥𝛹 = 𝜕𝑥 [ 𝜕𝑥 𝑥𝛹] − 𝛹- 𝜕𝑥 𝑥 𝜕𝑥
𝜕𝑥 2 𝜕𝑥
𝜕 𝜕𝛹 ∗ 𝜕 𝜕𝛹 𝜕 𝜕𝛹 𝜕𝛹 𝜕2 𝛹
= 𝜕𝑥 [ 𝜕𝑥 𝑥𝛹] − 𝜕𝑥 (𝛹 ∗ 𝛹)-𝛹 ∗ − 𝜕𝑥 [𝛹 ∗ 𝑥 ] + 𝛹∗ + 𝛹 ∗𝑥
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥 2

This means that the integrand has the form


𝜕 2𝛹∗ 𝜕 2𝛹 𝜕 𝜕𝛹 ∗ 𝜕𝛹 𝜕𝛹
( 2 𝑥𝛹 − 𝛹 ∗ 𝑥 ) = ( 𝑥𝛹 − 𝛹 ∗
𝑥 ) + 2𝛹 ∗
𝜕𝑥 𝜕𝑥 2 𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥

Because the wave functions vanish at infinity, the first term does not contribute, and the
integral gives
+∞
ℏ 𝜕
〈𝑝〉 = ∫ 𝑑𝑥 𝛹 ∗ (𝑥, 𝑡) ( ) 𝛹(𝑥, 𝑡)
𝑖 𝜕𝑥
−∞
+∞
As the position expectation was represented by 〈𝑥〉 = ∫−∞ 𝛹 ∗ 𝑥 𝛹𝑑𝑥

This suggests that the momentum be represented by the differential operator


𝜕
𝑝̂ = −𝑖ℏ
𝜕𝑥
and the position operator be represented by 𝑥̂ ⟶ 𝑥
To calculate expectation values, operate the given operator on the wave function have a product
with the complex conjugate of the wave function and integrate.
+∞ +∞

〈𝑝〉 = ∫ 𝛹 ∗ 𝑝̂ 𝛹𝑑𝑥 and 〈𝑥〉 = ∫ 𝛹 ∗ 𝑥̂𝛹𝑑𝑥


−∞ −∞

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


3
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

Example: A particle of mass m is in the state


𝟐
𝛹(𝑥, 𝑡) = 𝑨𝒆−𝒂[(𝒎𝒙 ⁄ℏ)+𝒊𝒕]
where 𝐴 and a are positive real constants. (a) Find A. (b) Calculate the expectation values of
< 𝑥 >, < 𝑥 2 >, < 𝑝 >, and < 𝑝2 >. (d) Find 𝜎𝑥 and𝜎𝑝 . Is their product consistent with the
uncertainty principle?
−1⁄
∞ 𝟐 1 𝜋 𝜋ℏ 𝜋ℏ 4
(a) 1 = 2|𝐴| 2
∫0 𝒆−𝟐𝒂𝒎𝒙 ⁄ℏ 𝑑𝑥 =2|𝐴|2 2 √2𝑎𝑚⁄ℏ = |𝐴|2 √
2𝑎𝑚
; 𝐴 = (2𝑎𝑚)
+∞
(b) 〈𝑥〉 = ∫−∞ 𝑥|Ψ|2 𝑑𝑥 = 0 [odd integrand].

𝟐 ⁄ℏ 1 𝜋ℏ ℏ
〈𝑥 2 〉 = 2|𝐴| ∫ 𝑥 2 𝒆−𝟐𝒂𝒎𝒙
2
𝑑𝑥 = 2|𝐴|2 √ =
0 22 (2𝑎𝑚⁄ℏ) 2𝑎𝑚 4𝑎𝑚
𝑑〈𝑥〉
〈𝑝〉 = 𝑚 =0
𝑑𝑡

Remember:

2 1 𝜋 1⁄2
∫ 𝑒 −𝑎𝑥 𝑑𝑥 = ( )
0 2 𝑎

2 −𝑎𝑥 2
1 𝜋 1⁄2
∫ 𝑥 𝑒 𝑑𝑥 = ( )
0 4𝑎 𝑎

ℏ 𝜕 2 𝜕 2𝛹
〈𝑝2 〉 = ∫ 𝛹 ∗ ( ) 𝛹𝑑𝑥 = −ℏ2 ∫ 𝛹 ∗ 2 𝑑𝑥
𝑖 𝜕𝑥 𝜕𝑥
2𝑎𝑚 2𝑎𝑚𝑥 2
= −ℏ2 ∫ 𝛹 ∗ [− (1 − ) 𝛹] 𝑑𝑥 = 2𝑎𝑚ℏ{∫|Ψ|2 𝑑𝑥
ℏ ℏ
2𝑎𝑚
− ∫ 𝑥 2 |Ψ|2 𝑑𝑥}

2𝑎𝑚 2 2𝑎𝑚 ℏ 1
= 2𝑎𝑚ℏ (1 − 〈𝑥 〉) = 2𝑎𝑚ℏ (1 − ) = 2𝑎𝑚ℏ ( ) = 𝑎𝑚ℏ
ℏ ℏ 4𝑎𝑚 2
ℏ ℏ
(c) 𝜎𝑥2 = 〈𝑥 2 〉 − 〈𝑥 〉2 = 4𝑎𝑚 ⟹ 𝜎𝑥 = √4𝑎𝑚
𝜎𝑝2 = 〈𝑝2 〉 − 〈𝑝 〉2 = 𝑎𝑚ℏ ⟹ 𝜎𝑝 = √𝑎𝑚ℏ
ℏ ℏ
𝜎𝑥 𝜎𝑝 = √4𝑎𝑚 √𝑎𝑚ℏ = 2

3. Time evolution and the Hamiltonian operator


Taking Schrödinger’s time dependent equation
𝜕𝛹(𝑥, 𝑡)
̂ 𝛹(𝑥, 𝑡) = 𝑖ℏ
𝐻
𝜕𝑡
And rewriting as

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


4
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

𝜕𝛹(𝑥, 𝑡) ̂
𝑖𝐻
= − 𝛹(𝑥, 𝑡)
𝜕𝑡 ℏ
̂ does not depend explicitly on time (i.e., the potential 𝑉(𝑥)is constant)
And presuming 𝐻
could we somehow legally write
̂ (𝑡1 − 𝑡𝑜 )
𝑖𝐻
𝛹(𝑥, 𝑡1 ) = 𝑒𝑥𝑝 {= − 𝛹(𝑥, 𝑡𝑜 )}

̂ here was replaced by a constant number
Certainly, if the Hamiltonian operator 𝐻
we could perform such an integration of
𝜕𝛹(𝑥, 𝑡) ̂
𝑖𝐻
= − 𝛹(𝑥, 𝑡)
𝜕𝑡 ℏ
to get
̂ (𝑡1 − 𝑡𝑜 )
𝑖𝐻
𝛹(𝑥, 𝑡1 ) = 𝑒𝑥𝑝 {− 𝛹(𝑥, 𝑡𝑜 )}

If, with some careful definition, it was legal to do this then we would have an operator that
gives us the state at time 𝑡1 directly from that at time 𝑡𝑜 .
To think about this “legality” ,first we note that, because is a linear operator for any number a
̂ [𝑎𝛹(𝑟, 𝑡)] = 𝑎𝐻
𝐻 ̂ 𝛹(𝑟, 𝑡)

Since this works for any function, 𝛹(𝑟, 𝑡) we can write as a shorthand
̂ 𝑎 = 𝑎𝐻
𝐻 ̂

Next, we have to define what we mean by an operator raised to a power


̂ 2 we mean 𝐻
By 𝐻 ̂ 2 𝛹(𝑟, 𝑡) = 𝐻
̂ [𝐻
̂ 𝛹(𝑟, 𝑡)]

Specifically, for example, for the energy eigenfunction 𝜓𝑛 (𝑟)


̂ 2 𝜓𝑛 (𝑟) = 𝐻
𝐻 ̂ [𝐻
̂ 𝜓𝑛 (𝑟)] = 𝐻
̂ [𝐸𝑛 𝜓𝑛 (𝑟)] = 𝐸𝑛 𝐻
̂ 𝜓𝑛 (𝑟) = 𝐸𝑛2 𝜓𝑛 (𝑟)

We can proceed inductively to define all higher powers


̂ 𝑚+1 = 𝐻
𝐻 ̂ [𝐻
̂𝑚]
which will give, for the energy eigenfunction
̂ 𝑚 𝜓𝑛 (𝑟) = 𝐸𝑛𝑚 𝜓𝑛 (𝑟)
𝐻
Now let us look at the time evolution of some wavefunction 𝛹(𝑟, 𝑡) between times 𝑡𝑜 and 𝑡1
Suppose the wavefunction at time 𝑡𝑜 is 𝜓 (𝑟) which we expand in the energy
eigenfunctions 𝜓𝑛 (𝑟) as

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


5
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

𝜓 (𝑟) = ∑ 𝑎𝑛 𝜓𝑛 (𝑟)
𝑛
Then we know multiplying by the complex exponential factors for the time-evolution of each
basis function
𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 )
𝛹(𝑟, 𝑡1 ) = ∑ 𝑎𝑛 𝑒𝑥𝑝 [− ] 𝜓𝑛 (𝑟)

𝑛
𝑥2 𝑥3
noting that exp(𝑥) = 1 + 𝑥 + 2𝑖 + 3𝑖 + ⋯

we can write the exponentials as power series as


2
𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 ) 1 𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 )
𝛹(𝑟, 𝑡1 ) = ∑ 𝑎𝑛 [1 + [− ] + [− ] + ⋯ ] 𝜓𝑛 (𝑟)
ℏ 2𝑖 ℏ
𝑛

̂ 𝑚 𝜓𝑛 (𝑟) = 𝐸𝑛𝑚 𝜓𝑛 (𝑟)


because we showed that 𝐻
we can substitute to obtain

̂ (𝑡1 − 𝑡𝑜 )
𝑖𝐻 1 ̂ (𝑡1 − 𝑡𝑜 ) 2
𝑖𝐻
𝛹(𝑟, 𝑡1 ) = ∑ 𝑎𝑛 [1 + [− ] + [− ] + ⋯ ] 𝜓𝑛 (𝑟)
ℏ 2𝑖 ℏ
𝑛
̂ commute with scalar quantities (numbers) we can
because the operator and all its powers 𝐻
rewrite
2
𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 ) 1 𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 )
𝛹(𝑟, 𝑡1 ) = [1 + [− ] + [− ] + ⋯ ] ∑ 𝑎𝑛 𝜓𝑛 (𝑟)
ℏ 2𝑖 ℏ
𝑛
2
𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 ) 1 𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 )
𝛹(𝑟, 𝑡1 ) = [1 + [− ] + [− ] + ⋯ ] 𝛹(𝑟, 𝑡𝑜 )
ℏ 2𝑖 ℏ

So, provided we define the exponential of the operator in terms of a power series, i.e.,
2
𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 ) 𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 ) 1 𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 )
𝑒𝑥𝑝 [− ] = [1 + [− ] + [− ] +⋯]
ℏ ℏ 2𝑖 ℏ

then we can write our preceding expression as


𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 )
𝛹(𝑟, 𝑡1 ) = 𝑒𝑥𝑝 [− ] 𝛹(𝑟, 𝑡𝑜 )

Hence, we have established that there is a well-defined operator that given the quantum
mechanical wavefunction or “state” at time 𝑡𝑜 will tell us what the state is at a time 𝑡1

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


6
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

4. Dirac notation - Bra and Ket vectors – Vectors space.


The physical state of a system is represented in quantum mechanics by elements of a Hilbert
space; these elements are called state vectors. We can represent the state vectors in different
bases by means of function expansions. For instance, we can represent equivalently a vector
by its components in a Cartesian coordinate system, in a spherical coordinate system, or in a
cylindrical coordinate system. The meaning of a vector is, of course, independent of the
coordinate system chosen to represent its components. Similarly, the state of a microscopic
system has a meaning independent of the basis in which it is expanded.
To Free State vectors from coordinate meaning, Dirac introduced what was to become an
invaluable notation in quantum mechanics; it allows one to manipulate the formalism of
quantum mechanics with ease and clarity. He introduced the concepts of kets, bras, and bra-
kets, which will be explained below.
Kets: elements of a vector space
Dirac denoted the state vector 𝜓 by the symbol |𝜓⟩ , which he called a ket vector, or simply a
ket. Kets belong to the Hilbert (vector) space ℋ, or, in short, to the ket-space.
Bras: elements of a dual space
As mentioned above, we know from linear algebra that a dual space can be associated with
every vector space. Dirac denoted the elements of a dual space by the symbol ⟨ |, which he
called a bra vector, or simply a bra; for instance, the element ⟨𝜓|
represents a bra. Note: For every ket |𝜓⟩ there exists a unique bra ⟨𝜓|
and vice versa. Again, while kets belong to the Hilbert space ℋ, the corresponding bras belong
to its dual (Hilbert) space ℋ𝑑 .
Bra-ket: Dirac notation for the scalar product
Dirac denoted the scalar (inner) product by the symbol ⟨ | ⟩, which he called a a bra-ket. For
instance, the scalar product (𝜙, 𝜓) is denoted by the bra-ket ⟨𝜙|𝜓⟩:
(𝜙, 𝜓) → ⟨𝜙|𝜓⟩
Note: When a ket (or bra) is multiplied by a complex number, we also get a ket (or bra).
Remark: In wave mechanics we deal with wave functions 𝜓(𝑟⃗, 𝑡), but in the more general
formalism of quantum mechanics we deal with abstract kets
|𝜓⟩. Wave functions, like kets, are elements of a Hilbert space. We should note that, like a
wave function, a ket represents the system completely, and hence knowing
|𝜓⟩means knowing all its amplitudes in all possible representations. As mentioned above, kets
are independent of any particular representation. There is no reason to single out a particular
representation basis such as the representation in the position space. Of course, if we want to
know the probability of finding the particle at some position in space, we need to work out the
formalism within the coordinate representation. The state vector of this particle at time t will
be given by the spatial wave function ⟨𝑟⃗, 𝑡|𝜓⟩ = 𝜓(𝑟⃗, 𝑡). In the coordinate representation, the
scalar is given by

⟨𝜙|𝜓⟩ = ∫ 𝜙 ∗ (𝑟⃗, 𝑡)𝜓(𝑟⃗, 𝑡)𝑑 3 𝑟

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


7
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

Similarly, if we are considering the three-dimensional momentum of a particle, the ket


|𝜓⟩ will have to be expressed in momentum space. In this case the state of the particle will be
described by a wave function 𝜓(𝑝⃗, 𝑡), where 𝑝⃗, is the momentum of the particle.
Properties of kets, bras, and bra-kets
• Every ket has a corresponding bra
To every ket |𝜓⟩, there corresponds a unique bra ⟨𝜓| and vice versa:
|𝜓⟩ ↔ ⟨𝜓|
There is a one-to-one correspondence between bras and kets:
𝑎|𝜓⟩ + 𝑏|𝜓⟩ ↔ 𝑎∗ ⟨𝜓| + 𝑏 ∗ ⟨𝜓|
where a and b are complex numbers. The following is a common notation:
|𝑎𝜓⟩ = 𝑎|𝜓⟩, ⟨𝑏𝜓| = 𝑏⟨𝜓|
• Properties of the scalar product
In the case of function spaces, a “vector” element is given by a complex function and the
scalar product by integrals. That is, the scalar product of two functions 𝜓(𝑥) and 𝜙(𝑥) is
given by

(𝜓, 𝜙) = ∫ 𝜓 ∗ (𝑥) 𝜙(𝑥)𝑑𝑥


In quantum mechanics, since the scalar product is a complex number, the ordering matters a
lot. We must be careful to distinguish a scalar product from its complex conjugate ⟨𝜓|𝜙⟩; is
not the same thing as ⟨𝜙|𝜓⟩:
⟨𝜙|𝜓⟩∗ = ⟨𝜓|𝜙⟩
This property becomes clearer if we apply it to (2)

⟨𝜙|𝜓⟩∗ = (∫ 𝜙 ∗ (𝑟⃗, 𝑡) 𝜓(𝑟⃗, 𝑡)𝑑3 𝑟) = ∫ 𝜓 ∗ (𝑟⃗, 𝑡) 𝜙(𝑟⃗, 𝑡)𝑑 3 𝑟 = ⟨𝜓|𝜙⟩
When |𝜓⟩ and |𝜙⟩ are real, we would have ⟨𝜓|𝜙⟩ = ⟨𝜙|𝜓⟩. Let us list some additional
properties of the scalar product:
⟨𝜓|𝑎1 𝜓1 + 𝑎2 𝜓2 ⟩ = 𝑎1 ⟨𝜓|𝜓1 ⟩ + 𝑎2 ⟨𝜓|𝜓2 ⟩
⟨𝑎1 𝜙1 + 𝑎2 𝜙2 |𝜓⟩ = 𝑎1∗ ⟨𝜙1 |𝜓⟩ + 𝑎2∗ ⟨𝜙2 |𝜓⟩,
⟨𝑎1 𝜙1 + 𝑎2 𝜙2 |𝑏1 𝜓1 + 𝑏2 𝜓2 ⟩ = 𝑎1∗ 𝑏1 ⟨𝜙1 |𝜓1 ⟩ + 𝑎1∗ 𝑏2 ⟨𝜙1 |𝜓2 ⟩
+𝑎2∗ 𝑏1 ⟨𝜙2 |𝜓1 ⟩ + 𝑎2∗ 𝑏2 ⟨𝜙2 |𝜓2 ⟩.
• The norm is real and positive
For any state vector |𝜓⟩ of the Hilbert space ℋ, the norm ⟨𝜓|𝜓⟩ is real and positive; ⟨𝜓|𝜓⟩ is
equal to zero only for the case where |𝜓⟩ = 𝑂, where 𝑂 is the zero vector. If the state |𝜓⟩ is
normalized then ⟨𝜓|𝜓⟩ = 1.

• Schwarz inequality
For any two states|𝜓⟩and |𝜙⟩of the Hilbert space, we can show that
|⟨𝜓|𝜙⟩|2 ≤ ⟨𝜓|𝜓⟩⟨𝜙|𝜙⟩

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


8
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

If|𝜓⟩and |𝜙⟩are linearly dependent (i.e., proportional: |𝜓⟩ = 𝛼|𝜙⟩, where 𝛼 is a scalar), this
relation becomes an equality. The Schwarz inequality (12) is analogous to the following
relation of the real Euclidean space;
2 2 2
|𝐴⃗. 𝐵
⃗⃗ | ≤ |𝐴⃗| |𝐵
⃗⃗ | .

• Triangle inequality
√⟨𝜓 + 𝜙|𝜓 + 𝜙⟩ ≤ √⟨𝜓|𝜓⟩ + √⟨𝜙|𝜙⟩
If|𝜓⟩and |𝜙⟩are linearly dependent, |𝜓⟩ = 𝛼|𝜙⟩, and if the proportionality scalar 𝛼 is real and
positive, the triangle inequality becomes an equality. The counterpart of this inequality in
Euclidean space is given by |𝐴⃗ + 𝐵⃗⃗ | ≤ |𝐴⃗| + |𝐵
⃗⃗ |

• Orthogonal states
Two kets, |𝜓⟩and |𝜙⟩, are said to be orthogonal if they have a vanishing scalar product:
⟨𝜓|𝜙⟩ = 0
• Orthonormal states
Two kets, |𝜓⟩and |𝜙⟩, are said to be orthonormal if they are orthogonal and if each one of them
has a unit norm:
⟨𝜓|𝜙⟩ = 0, ⟨𝜓|𝜓⟩ = 1, ⟨𝜙|𝜙⟩ = 1

• Forbidden quantities
If |𝜓⟩and |𝜙⟩, belong to the same vector (Hilbert) space, products of the type |𝜓⟩|𝜙⟩ and
⟨𝜓|⟨𝜙|are forbidden. They are nonsensical, since |𝜓⟩|𝜙⟩ and ⟨𝜓|⟨𝜙| are neither kets nor bras
(an explicit illustration of this will be carried out in the example below and later on when we
discuss the representation in a discrete basis). If |𝜓⟩ and |𝜙⟩ belong, however, to different
vector spaces (e.g., |𝜓⟩ belongs to a spin space and |𝜙⟩to an orbital angular momentum space),
then the product |𝜓⟩|𝜙⟩written as|𝜓⟩⨂|𝜙⟩, represents a tensor product of|𝜓⟩ and|𝜙⟩. Only in
these typical cases are such products meaningful.

Example1: (Note: We will see later in this chapter that kets are represented by column matrices
and bras by row matrices; this example is offered earlier than it should because we need to
show some concrete illustrations of the formalism.) Consider the following two kets:
−3𝑖 2
|𝜓⟩ = (2 + 𝑖 ), |𝜙⟩ = ( −𝑖 )
4 2 − 3𝑖
(a) Find the bra ⟨𝜙|.
(b) Evaluate the scalar product ⟨𝜙|𝜓⟩.
(c) Examine why the products |𝜓⟩ |𝜙⟩and ⟨𝜙|⟨𝜓|do not make sense.
Solution:
(a) As will be explained later when we introduce the Hermitian adjoint of kets and bras,
we want to mention that the bra⟨𝜙| can be obtained by simply taking the complex
conjugate of the transpose of the ket |𝜙⟩:
⟨𝜙| = (2 𝑖 2 + 3𝑖)
(b) The product ⟨𝜙|𝜓⟩ can be calculated as follows:

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


9
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

−3𝑖
⟨𝜙|𝜓⟩ = (2 𝑖 2 + 3𝑖) (2 + 𝑖 )
4
= 2(−3𝑖) + 𝑖(2 + 𝑖) + 4(2 + 3𝑖) = 7 + 8𝑖
(c) First, the product |𝜓⟩ |𝜙⟩ cannot be performed because, from linear algebra, the product
of two column matrices cannot be performed. Similarly, since two row matrices cannot
be multiplied, the product ⟨𝜙|⟨𝜓| is meaningless.

Physical meaning of the scalar product


The scalar product can be interpreted in two ways. First, by analogy with the scalar product of
ordinary vectors in the Euclidean space, where; 𝐴⃗ ∙ 𝐵 𝐵 on⃗⃗⃗⃗
⃗⃗ represents the projection of ⃗⃗⃗⃗ 𝐴, the
product ⟨𝜙|𝜓⟩ also represents the projection of |𝜓⟩ onto |𝜙⟩. Second, in the case of normalized
states and according to Born’s probabilistic interpretation, the quantity ⟨𝜙|𝜓⟩ represents the
probability amplitude that the system’s state |𝜓⟩ will, after a measurement is performed on the
system, be found to be in another state|𝜙⟩.

Example 2 (Bra-ket algebra):


Consider the states |𝜓⟩ = 3𝑖|𝜙1 ⟩ − 7𝑖|𝜙2 ⟩ and |𝜒⟩ = −|𝜙1 ⟩ + 2𝑖|𝜙2 ⟩, where |𝜙1 ⟩ and |𝜙2 ⟩
are orthonormal.
(a) Calculate |𝜓 + 𝜒⟩ and ⟨𝜓 + 𝜒|.
(b) Calculate the scalar product ⟨𝜓|𝜒⟩ and ⟨𝜒|𝜓⟩.
(c) Show that the state |𝜓⟩ and |𝜒⟩ satisfy the Schwarz inequality.
(d) Show that the state |𝜓⟩ and |𝜒⟩ satisfy the triangle inequality.
Solution:
(a) The calculation of |𝜓 + 𝜒⟩ is straightforward:
|𝜓 + 𝜒⟩ = (3𝑖|𝜙1 ⟩ − 7𝑖|𝜙2 ⟩) + (−|𝜙1 ⟩ + 2𝑖|𝜙2 ⟩) = (−1 + 3𝑖)|𝜙1 ⟩ − 5𝑖|𝜙2 ⟩
This leads at once to the expression of ⟨𝜓 + 𝜒|:
⟨𝜓 + 𝜒| = (−1 + 3𝑖)∗ ⟨𝜙1 | + (−5𝑖)∗ ⟨𝜙2 | = (−1 − 3𝑖)⟨𝜙1 | + 5𝑖⟨𝜙2 |
(b) Since ⟨𝜙1 |𝜙1 ⟩ = ⟨𝜙2 |𝜙2 ⟩ = 1, ⟨𝜙1 |𝜙2 ⟩ = ⟨𝜙2 |𝜙1 ⟩ = 0, and since the bras
corresponding to the kets |𝜓⟩ = 3𝑖|𝜙1 ⟩ − 7𝑖|𝜙2 ⟩ and |𝜒⟩ = −|𝜙1 ⟩ + 2𝑖|𝜙2 ⟩ are given
by ⟨𝜓| = − 3𝑖⟨𝜙1 | + 7𝑖⟨𝜙2 | and ⟨𝜒| = −⟨𝜙1 | − 2𝑖⟨𝜙2 |, the scalar products are
⟨𝜓|𝜒⟩ = ( 3𝑖|𝜙1 ⟩ − 7𝑖|𝜙2 ⟩)( −|𝜙1 ⟩ − 2𝑖|𝜙2 ⟩)
= (−3𝑖)(−1)⟨𝜙1 |𝜙1 ⟩ + (7𝑖)(2𝑖)⟨𝜙2 |𝜙2 ⟩
= −14 + 3𝑖
⟨𝜒|𝜓⟩ =(−|𝜙1 ⟩ − 2𝑖|𝜙2 ⟩)( 3𝑖|𝜙1 ⟩ − 7𝑖|𝜙2 ⟩)
= (−1)(3𝑖)⟨𝜙1 |𝜙1 ⟩ + (−2𝑖)(−7𝑖)⟨𝜙2 |𝜙2 ⟩
= −14 − 3𝑖
We see that ⟨𝜓|𝜒⟩ is equal to the complex conjugate of ⟨𝜒|𝜓⟩
(c) Let us first calculate that ⟨𝜓|𝜓⟩ and ⟨𝜒|𝜒⟩:
⟨𝜓|𝜓⟩ = (− 3𝑖|𝜙1 ⟩ + 7𝑖|𝜙2 ⟩)( 3𝑖|𝜙1 ⟩ − 7𝑖|𝜙2 ⟩) = (−3𝑖)(3𝑖) + (7𝑖)(−7𝑖) = 58
⟨𝜒|𝜒⟩ = (−|𝜙1 ⟩ − 2𝑖|𝜙2 ⟩)( −|𝜙1 ⟩ + 2𝑖|𝜙2 ⟩) = (−1)(−1) + (−2𝑖)(2𝑖) = 5
Since ⟨𝜓|𝜒⟩ = −14 + 3𝑖 we have|⟨𝜓|𝜒⟩|2 = 142 + 32 = 205. Combing the values of
|⟨𝜓|𝜒⟩|2 , ⟨𝜓|𝜓⟩ and ⟨𝜒|𝜒⟩. We see that the Schwarz inequality is satisfied:
205 < (58)(5) ⇒ |⟨𝜓|𝜒⟩|2 < ⟨𝜓|𝜓⟩⟨𝜒|𝜒⟩
(d) First, let us use the equations in (a) to calculate ⟨𝜓 + 𝜒|𝜓 + 𝜒⟩:
⟨𝜓 + 𝜒|𝜓 + 𝜒⟩ = [(−1 − 3𝑖)⟨𝜙1 | + 5𝑖⟨𝜙2 |][(−1 + 3𝑖)|𝜙1 ⟩ − 5𝑖|𝜙2 ⟩]

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


10
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

= (−1 − 3𝑖)(−1 + 3𝑖) + (5𝑖)(−5𝑖)


= 35
Since ⟨𝜓|𝜓⟩ = 58 and ⟨𝜒|𝜒⟩ = 5, we infer that the triangle inequality is satisfied:
√35 < √58 + √5 ⇒ √⟨𝜓 + 𝜒|𝜓 + 𝜒⟩ < √⟨𝜓|𝜓⟩ + √⟨𝜒|𝜒⟩
Vectors
A vector space consists of a set of vectors (|𝛼⟩, |𝛽⟩, |𝛾⟩, … ), together with a set of scalars
(𝑎, 𝑏, 𝑐, … ), which are subject to two operations-vector addition and scalar multiplication.
Vector addition. The “sum” of any two vectors is another vector
|𝛼⟩ + |𝛽⟩ = |𝛾⟩,
Vector addition is commutative
|𝛼⟩ + |𝛽⟩ = |𝛽⟩ + |𝛼⟩,
And associative
|𝛼⟩ + (|𝛽⟩ + |𝛾⟩) = (|𝛽⟩ + |𝛼⟩) + |𝛾⟩,
There exists a zero (or null) vector, 0⟩ , with the property
|𝛼⟩ + |0⟩ = |𝛼⟩,
For any vector |𝛼⟩. And for every vector |𝛼⟩ there is an associated inverse vector (|−𝛼⟩), such
that
|𝛼⟩ + |−𝛼⟩ = |0⟩,
Scalar multiplication. The “product” of any scalar with any vector is another vector:
𝑎|𝛼⟩ = |𝛾⟩,
Scalar multiplication has the same properties such as vector addition (i.e. commutative,
distributive, and associative, etc…).

5. Time-dependent perturbation theory.


Time-dependent perturbation theory is one of the most useful techniques for understanding
how quantum mechanical systems respond in time to changes in their environment. It is
especially useful for understanding the consequences of periodic changes; a classic example is
understanding how a quantum mechanical system responds to light, which can often be usefully
approximated as a periodically oscillating electromagnetic field.
For time-dependent problems, we consider some time-dependent perturbation, 𝐻 ̂𝑝 (𝑡); and for
an unperturbed Hamiltonian, 𝐻 ̂𝑜 ; that is itself not dependent on time. The total Hamiltonian is
then
̂=𝐻
𝐻 ̂𝑜 +𝐻̂𝑝 (𝑡)
To deal with such a situation, we return to the time-dependent Schrödinger equation:
𝜕
𝑖ℏ |Ψ⟩ = 𝐻̂ |Ψ⟩
𝜕𝑡
where now the ket |Ψ⟩ is time-varying in general.

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


11
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

With |ψ𝑛 ⟩ and E𝑛 as the energy eigenfunctions and eigenvalues of the time independent
equation
̂𝑜 |ψ𝑛 ⟩ = E𝑛 |ψ𝑛 ⟩
𝐻
we expand the solution of the time-dependent Schrödinger equation |Ψ⟩ as:

|Ψ⟩ = ∑ 𝑎𝑛 (𝑡)𝑒𝑥𝑝(−𝑖 E𝑛 𝑡⁄ℏ)|ψ𝑛 ⟩


𝑛
Note we chose to include the time-dependent factor 𝑒𝑥𝑝(− iE𝑛 𝑡⁄ℏ) explicitly in the expansion.
We could have left that out, and merely included it in 𝑎𝑛 (𝑡) .
It is usually better to take out any major underlying time dependence leaving the time
dependence of 𝑎𝑛 (𝑡) to deal only with the additional changes.
Now we can substitute the expansion solution of the time-dependent Schrödinger equation into
the time-dependent Schrödinger equation and obtaining:

̂𝑜 +𝐻
∑(𝑖ℏ𝑎̇ 𝑛 + 𝑎𝑛 E𝑛 )𝑒𝑥𝑝(−𝑖 E𝑛 𝑡⁄ℏ)|ψ𝑛 ⟩ = ∑ 𝑎𝑛 (𝐻 ̂𝑝 (𝑡)) 𝑒𝑥𝑝(−𝑖 E𝑛 𝑡⁄ℏ)|ψ𝑛 ⟩
𝑛 𝑛
𝜕𝑎𝑛
where 𝑎̇ 𝑛 ≡ 𝜕𝑡

̂𝑜 |ψ𝑛 ⟩ with E𝑛 |ψ𝑛 ⟩ leads to the


Using the time-independent Schrödinger equation to replace 𝐻
cancellation of terms in E𝑛 |ψ𝑛 ⟩from the two sides.

Now premultiplying by ⟨𝜓𝑞 | on both sides of the equation above leads to

̂𝑝 (𝑡)|ψ𝑛 ⟩
𝑖ℏ𝑎̇ 𝑞 (𝑡)𝑒𝑥𝑝(− iE𝑞 𝑡⁄ℏ) = ∑ 𝑎𝑛 (𝑡)𝑒𝑥𝑝(−𝑖 E𝑛 𝑡⁄ℏ) ⟨𝜓𝑞 |𝐻
𝑛
We have made no approximations in going from the time-dependent Schrödinger equation to
the above equation, these are entirely equivalent equations.
Now we consider a perturbation series. We introduce the expansion parameter 𝛾 just as before,
̂𝑝
now writing our perturbation as 𝛾𝐻

As before, we can set this parameter to a value of 1 at the end. We presume that we can express
the expansion coefficients n a as a power series:
(0) (1) (2)
𝑎𝑛 = 𝑎𝑛 + 𝛾𝑎𝑛 + 𝛾 2 𝑎𝑛 + ⋯
and we substitute this expansion into the derived equation above:

̂𝑝 (𝑡)|ψ𝑛 ⟩
𝑖ℏ𝑎̇ 𝑞 (𝑡)𝑒𝑥𝑝(−𝑖 E𝑞 𝑡⁄ℏ) = ∑ 𝑎𝑛 (𝑡)𝑒𝑥𝑝(− iE𝑛 𝑡⁄ℏ) ⟨𝜓𝑞 |𝐻
𝑛
Equating powers of 𝛾 , we obtain for the zero order term
(0)
𝑎̇ 𝑞 (𝑡)=0
The zero order solution simply corresponds to the unperturbed solution, and hence there is no
change in the expansion coefficients in time. Repeating the relevant equations
(0) (1) (2)
𝑎𝑛 = 𝑎𝑛 + 𝛾𝑎𝑛 + 𝛾 2 𝑎𝑛 + ⋯

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


12
Quantum Mechanics- Lecture Note
College of Science/Physics Department
Chapter Three First Semester 2018-2019

̂𝑝 (𝑡)|ψ𝑛 ⟩
𝑖ℏ𝑎̇ 𝑞 (𝑡)𝑒𝑥𝑝(−𝑖 E𝑞 𝑡⁄ℏ) = ∑ 𝑎𝑛 (𝑡)𝑒𝑥𝑝(− iE𝑛 𝑡⁄ℏ) ⟨𝜓𝑞 |𝐻
𝑛
For the first order term, we have

(1) 1 (0)
𝑎̇ 𝑞 (𝑡) = ̂𝑝 (𝑡)|ψ𝑛 ⟩
∑ 𝑎𝑛 𝑒𝑥𝑝(𝑖𝜓𝑞𝑛 𝑡) ⟨𝜓𝑞 |𝐻
𝑖ℏ
𝑛
where we have introduced the notation

𝜓𝑞𝑛 = (E𝑞 − E𝑛 )⁄ℏ


(0)
Note here that the 𝑎𝑛 are all constants;
(0)
we deduced in (𝑎̇ 𝑞 (𝑡) = 0 )that they do not change in time.

They represent the “starting” state of the system at time 𝑡 = 0. We note now that, if we know
the starting state, and the perturbing potential and the unperturbed eigenvalues and
(1)
eigenfunctions, we can integrate (𝑎̇ 𝑞 (𝑡)) to obtain the first order, time-dependent correction
(1)
𝑎𝑞 (𝑡), to the expansion coefficients.

If we know the new approximate expansion coefficients,


(0) (1)
𝑎𝑞 = 𝑎𝑞 + 𝑎𝑞 (𝑡)

then we know the new wavefunction and can calculate the behavior of the system from this
new wavefunction. We can proceed to higher order in this time-dependent perturbation theory.
In general, equating powers of progressively higher order, we obtain

(𝑝+1) 1 (𝑝)
𝑎̇ 𝑞 (𝑡) = ̂𝑝 (𝑡)|ψ𝑛 ⟩
∑ 𝑎𝑛 𝑒𝑥𝑝(𝑖𝜓𝑞𝑛 𝑡) ⟨𝜓𝑞 |𝐻
𝑖ℏ
𝑛
We see that this perturbation theory is also a method of successive approximations, just like
the time-independent perturbation theory. We calculate each higher order correction from the
preceding correction.
Just as for the time-independent perturbation theory, the time-dependent theory is often most
useful for calculating some process to the lowest non-zero order.
Higher order time-dependent perturbation theory is very useful, for example, for understanding
nonlinear optical processes.
First order time-dependent perturbation theory gives the ordinary, linear optical properties of
materials. Higher order time-dependent perturbation theory is used to calculate processes such
as second harmonic generation and two photon absorption in nonlinear optics, processes that
are seen routinely with the high intensities of modern lasers.

Dr. Murtadha F. Sultan Dr. Abbas Albarazanghi


13

You might also like