Notes Phys300
Notes Phys300
These lecture notes are just prepared to help the students of Phys 300. There are not original; they
contain lots of material from several different sources
In classical physics, there exist two distinct entities: particles and waves.
Particles are localized. They have definite position energy and momentum.
They are confined in space.
Imagine a classical particle of mass m, moving under a force ⃗𝑭, at any instant.
⃗ (𝒕) of the particle at any given
Its state is specified by the position (trajectory) 𝒓
time. Once we know that, we can figure out the any other physical quantities
⃗ (𝒕) = 𝒅𝒓
relevant to the system, like the velocity 𝒗 ⃗ /𝒅𝒕 , the momentum
⃗ (𝒕) = 𝑚𝑑𝑟/𝑑𝑡 , the kinetic energy 𝑇 = 𝑚𝑣 2 /2 ,etc.
𝒑
One needs an equation, or a set of equations of motion that will predict the
state 𝑟(𝑡) . These are usually differential equations.
The basic equation of the classical mechanics for the motion of a particle under
⃗ 𝑉 is
the influence of a conservative force 𝐹 = −∇
𝑑2 𝑟(𝑡) 𝑑𝑝
𝑚 = ⃗𝑉=
−∇ ∶ the Newton’s equation
𝑑𝑡 2 𝑑𝑡
and this equation also implies energy conservation,
𝑝2
𝐸= +𝑉 (∗)
2𝑚
wave function
Ψ(𝑟, 𝑡) = 𝐴 𝑒𝑥𝑝[𝑖(𝑘⃗ ∙ 𝑟 − 𝜔𝑡)]
where
ℎ
𝐸 = ℎ𝜈 = 2𝜋𝜈 => 𝐸 = ℏ𝜔,
2𝜋
Now,
𝑓𝑟𝑜𝑚 (𝑘⃗, 𝜔) 𝑑𝑒𝑠𝑐𝑟𝑖𝑝𝑡𝑖𝑜𝑛 𝑡𝑜 (𝑝, 𝐸 )
ℎ
𝜆=
𝑝
ℎ
=> 𝑝=𝑘 => 𝑝 = ℏ𝑘
2𝜋
2𝜋
𝜆=
𝑘
𝑝2
Since for a free particle 𝐸 = =>
2𝑚
𝐸 𝑝2
𝜔= =
ℏ 2𝑚ℏ
Yields
4
𝑝∙𝑟 𝑝2
Ψ(𝑟, 𝑡) = 𝐴 𝑒𝑥𝑝 [𝑖 ( − 𝑡)]
ℏ 2𝑚ℏ
Due to the wave-particle duality we can now assume that this wave
function must somehow be related to the wave properties of free
electrons (as observed in the electron diffraction experiments).
Note that, this wave function satisfies a differential equation
𝜕Ψ(𝑟, 𝑡) 𝑝2
𝑖ℏ = Ψ(𝑟, 𝑡) = 𝐸Ψ(𝑟, 𝑡)
𝜕𝑡 2𝑚
And also
ℏ2 2 𝑝2
− ∇ Ψ(𝑟, 𝑡) = Ψ(𝑟, 𝑡) = 𝐸Ψ(𝑟, 𝑡)
2𝑚 2𝑚
OR
𝝏𝚿(𝒓
⃗ , 𝒕) ℏ𝟐 𝟐
𝒊ℏ =− 𝛁 𝚿(𝒓
⃗ , 𝒕) ∶ 𝒇𝒐𝒓 𝒂 𝒇𝒓𝒆𝒆 𝒑𝒂𝒓𝒕𝒊𝒄𝒍𝒆
𝝏𝒕 𝟐𝒎
𝑉 (𝑟, 𝑡) ≠ 0
All the above attempts do not give us any clue about the answer.
However, comparison of this differential equation with the classical energy
equation (*) can give us the idea to try the equation below as a starting
point for the calculation of wave functions for particles moving in a
potential
𝑉 (𝑟, 𝑡):
𝝏𝚿(𝒓
⃗ , 𝒕) ℏ𝟐 𝟐
𝒊ℏ =− 𝛁 𝚿(𝒓
⃗ , 𝒕) + 𝑽(𝒓
⃗ , 𝒕 ) 𝚿( 𝒓
⃗ , 𝒕)
𝝏𝒕 𝟐𝒎
5
𝜕Ψ(𝑟, 𝑡) ℏ2 2
𝑖ℏ =− ∇ Ψ(𝑟, 𝑡) + 𝑉 (𝑟, 𝑡) Ψ(𝑟, 𝑡)
𝜕𝑡 2𝑚
ℎ
where 𝑖 = √−1 and ℏ = = 1.0545 × 10−34 J.s, where
2𝜋
𝚿( 𝒓
⃗ , 𝒕): wave function in coordinate space
𝑑3 𝑟
Or in one dimension
|Ψ(𝑥, 𝑡)|2 𝑑𝑥 ≡ 𝑃(𝑥, 𝑡) 𝑑𝑥
which gives the probability that the particle will be found at time t at the
point 𝑥, which is between (𝑥, 𝑥 + 𝑑𝑥 ). Thus,
|Ψ(𝑥, 𝑡)|2 ≡ 𝑃(𝑥, 𝑡) ∶ 𝑡ℎ𝑒 𝑝𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 𝑑𝑒𝑛𝑠𝑖𝑡𝑦
Such that
7
at time t.
Normalization
What is the probability of finding the particle anywhere in the x axis
(between −∞ and ∞ ) at time t.?
This probability is the area under the graph of |Ψ(𝑥, 𝑡)|2 .
∞
∫ |Ψ(𝑥, 𝑡)|2 𝑑𝑥 = probability of finding the particle
−∞
between − ∞ and ∞.
Now, the statistical interpretation of the wave function suggests that the total
probability of finding the particle somewhere in the x-axis must be equal to 1.
That is, the sum of the probabilities for all of the x-axis must be equal to one.
This is expressed by the integral
∞
∫ |Ψ(𝑥, 𝑡)|2 𝑑𝑥 = 1 (∗∗)
−∞
(which means that if the particle exits it must be somewhere at all times)
.Without this, the statistical interpretation would be nonsense.
Note that if Ψ(𝑥, 𝑡) is a solution to the Schrödinger equation (*),
𝜕Ψ(𝑥, 𝑡) ℏ2 𝜕 2 Ψ(𝑥, 𝑡)
𝑖ℏ =− + 𝑉 (𝑥, 𝑡) Ψ(𝑥, 𝑡)
𝜕𝑡 2𝑚 𝜕𝑥 2
so too is 𝐴 Ψ(𝑥, 𝑡), where 𝐴 is any complex constant. What we must do, then,
is to pick this undetermined factor so as to ensure that eq.(**) is satisfied. This
process is called normalizing the wave function and a wave function that
obeys this equation is said to be normalized. Non-normalizable solutions
cannot represent the particle and they must be rejected. A wave function must
be square integrable if it is a normalizable function.
∞
∫ |Ψ(𝑥, 𝑡)|2 𝑑𝑥 < ∞ => lim Ψ(𝑥, 𝑡) → 0
−∞ 𝑥→±∞
11
Suppose that we have normalized the wave function at time t=0. How do
we know that it will stay normalized as time goes on and 𝚿 evolves ?
Schrödinger equation has the property that it preserves the normalization
of the wave function.
Proof:
𝑑 ∞ 2
∞
𝜕
∫ |Ψ(𝑥, 𝑡)| 𝑑𝑥 = ∫ |Ψ(𝑥, 𝑡)|2 𝑑𝑥 = 0
𝑑𝑡 −∞ −∞ 𝜕𝑡
𝜕 2 ∗
ℏ 𝜕 2 Ψ(𝑥, 𝑡) 1
|Ψ(𝑥, 𝑡)| = Ψ [− + 𝑉 (𝑥, 𝑡) Ψ(𝑥, 𝑡)]
𝜕𝑡 2𝑚𝑖 𝜕𝑥 2 𝑖ℏ
ℏ 𝜕 2 Ψ∗ (𝑥, 𝑡) 1
+[ 2
− 𝑉 (𝑥, 𝑡) Ψ ∗ (𝑥, 𝑡)] Ψ
2𝑚𝑖 𝜕𝑥 𝑖ℏ
𝑖ℏ ∗
𝜕 2 Ψ(𝑥, 𝑡) 𝜕 2 Ψ∗ (𝑥, 𝑡)
= [Ψ −Ψ ]
2𝑚 𝜕𝑥 2 𝜕𝑥 2
𝜕 𝑖ℏ ∗
𝜕Ψ(𝑥, 𝑡) 𝜕Ψ ∗ (𝑥, 𝑡)
= [ (Ψ −Ψ )] (∗)
𝜕𝑥 2𝑚 𝜕𝑥 𝜕𝑥
12
∞ ∞
𝜕 2
𝜕 𝑖ℏ ∗
𝜕Ψ(𝑥, 𝑡) 𝜕Ψ ∗ (𝑥, 𝑡)
∫ |Ψ(𝑥, 𝑡)| 𝑑𝑥 = ∫ [ (Ψ −Ψ )] 𝑑𝑥
−∞ 𝜕𝑡 −∞ 𝜕𝑥 2𝑚 𝜕𝑥 𝜕𝑥
∞
𝑖ℏ ∗
𝜕Ψ(𝑥, 𝑡) 𝜕Ψ∗ (𝑥, 𝑡)
=[ (Ψ −Ψ )]
2𝑚 𝜕𝑥 𝜕𝑥 −∞
But Ψ(𝑥, 𝑡) must go to zero as x goes to ±∞; otherwise the wave function
would not be normalizable. (Such functions are said to be square integrable.) It
follows that
∞
𝜕
∫ |Ψ(𝑥, 𝑡)|2 𝑑𝑥 = 0
−∞ 𝜕𝑡
𝑑𝑄
𝐼=− ∶ Electric current (1)
𝑑𝑡
𝑄 = ∫ 𝜌 𝑑𝑉 (2)
𝑉
Therefore,
𝑑𝑄 𝑑 𝜕𝜌
= ∫ 𝜌 𝑑𝑉 = ∫ 𝑑𝑉 (3)
𝑑𝑡 𝑑𝑡 𝜕𝑡
𝑉 𝑉
𝜕𝜌 𝜕𝜌
∫ ⃗ ∙ 𝐽 𝑑𝑉 => ∫ ( + ∇
𝑑𝑉 = − ∫ ∇ ⃗ ∙ 𝐽) 𝑑𝑉 = 0 =>
𝜕𝑡 𝜕𝑡
𝑉 𝑉 𝑉
𝜕𝜌
( + ⃗∇ ∙ 𝐽) = 0
𝜕𝑡
Any variation in the total charge within a closed surface must be due to
charges that flow across the surface.
This is called the equation of continuity. It is a direct expression of the
local law of conservation of charge.
14
Now, in QM,
𝜕 𝜕
𝑃(𝑥, 𝑡) + 𝐽(𝑥, 𝑡) = 0 (∗∗∗): 𝑐𝑜𝑛𝑡𝑖𝑛𝑢𝑖𝑡𝑦 𝑒𝑞𝑢𝑎𝑡𝑖𝑜𝑛
𝜕𝑡 𝜕𝑥
The continuity equation expresses the fact that a change in the probability
density in a region, say 𝑎 ≤ 𝑥 ≤ 𝑏, with time is compensated by a net
change in probability current (flux) into the region.
Let’s integrate (***)
𝑏 𝑏
𝜕 𝜕
∫ 𝑃(𝑥, 𝑡) 𝑑𝑥 = − ∫ 𝐽(𝑥, 𝑡)𝑑𝑥 = 𝐽(𝑎, 𝑡) − 𝐽(𝑏, 𝑡)
𝜕𝑡 𝜕𝑥
⏟
𝑎 𝑎
𝑃𝑎𝑏
OR, in 3.dim,
𝜕
𝑃(𝑟, 𝑡) + ⃗∇ ∙ 𝐽(𝑟, 𝑡) = 0
𝜕𝑡
𝑃(𝑟, 𝑡) = |Ψ(𝑥, 𝑡)|2 ∶ 𝑃𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 𝑑𝑒𝑛𝑠𝑖𝑡𝑦
𝐽(𝑟, 𝑡) ∶ 𝑃𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 𝑐𝑢𝑟𝑟𝑒𝑛𝑡
The time rate of change of probability of finding the particle within the
finite volume must be the negative of the probability of the net outflow of
the particle.
15
∑ 𝑁𝑖 = 𝑁
𝑖=1
𝑥̅ = ∑ 𝑃𝑖 𝑥𝑖 ∶ 𝑎𝑣𝑒𝑟𝑎𝑔𝑒 𝑜𝑓 𝑥
𝑖
with
∑ 𝑃𝑖 = 1
𝑖
∑ → ∫
𝑖
The rules for the discrete distribution translate in the obvious way:
∞
∫ 𝜌(𝑥 )𝑑𝑥 = 1
−∞
∞
Consider now a quantum particle in a state Ψ(𝑥, 𝑡). Let’s consider the
particle’s position observable, 𝑥. In analogy with (*), Its expectation value
is written as
∞ ∞
〈𝑥 〉 = ∫ 𝑥𝑃(𝑥, 𝑡)𝑑𝑥 = ∫ 𝑥 |Ψ(𝑥, 𝑡)|2 𝑑𝑥 = (∗∗)
−∞ −∞
With
∞
〈𝑥 〉 = ∫ 𝑥 |Ψ(𝑥, 𝑡)|2 𝑑𝑥
−∞
(We will return to this relation in CH3 when we study the time dependence
of expectation values.)
This is an assumption at this point, which has to prove itself in the following.
It is the “velocity” of the expectation value of x, which is not the same as the
velocity of the particle. It is not even clear what velocity means in quantum
mechanics: if the particle doesn’t have a determinate position (prior to the
measurement), neither does it have a well-defined velocity. For our present
purposes it will be Ok to postulate that the expectation value of the velocity is
equal to the time derivative of the expectation value of the position:
Since Ψ = Ψ(𝑥, 𝑡), as time goes on, 〈𝑥 〉 will change.
Integration by parts=>
𝑑 〈𝑥 〉 𝑖ℏ ∞ 𝜕 ∗
𝜕Ψ 𝜕Ψ ∗ 𝑖ℏ ∞
= ∫ ⏟ 𝑥 (Ψ − Ψ) 𝑑𝑥 = ∫ 𝑢 𝑑𝑉
𝑑𝑡 2𝑚 −∞ 𝑢 𝜕𝑥 ⏟ 𝜕𝑥 𝜕𝑥 2𝑚 −∞
𝑉
19
∞ ∞
𝑖ℏ ∗
𝜕Ψ 𝜕Ψ∗ ∗
𝜕Ψ 𝜕Ψ∗
= {𝑥 (Ψ − Ψ)| − ∫ 𝑑𝑥 (Ψ − Ψ)}
2𝑚 ⏟ 𝜕𝑥 𝜕𝑥 −∞ −∞ 𝜕𝑥 𝜕𝑥
=0
𝑑 〈𝑥 〉 𝑖ℏ ∞ ∗
𝜕Ψ 𝜕Ψ∗
=− ∫ 𝑑𝑥 (Ψ − Ψ)
𝑑𝑡 2𝑚 −∞ 𝜕𝑥 𝜕𝑥
𝑖ℏ ∞ ∗
𝜕Ψ 𝑖ℏ ∞ 𝜕Ψ ∗
=− ∫ 𝑑𝑥 Ψ + ∫ 𝑑𝑥 Ψ
2𝑚 −∞ 𝜕𝑥 2𝑚 −∞ 𝜕𝑥
𝜕Ψ 𝜕Ψ ∗
u = Ψ , du = 𝑑𝑥 ; dV = 𝑑𝑥 , 𝑉 = Ψ∗
𝜕𝑥 𝜕𝑥
𝑑 〈𝑥 〉 𝑖ℏ ∞ 𝜕Ψ 𝑖ℏ ∞
𝜕Ψ
=− ∫ 𝑑𝑥 Ψ∗ + (Ψ ∗ Ψ)|∞
{⏟ −∞ − ∫ 𝑑𝑥 Ψ ∗
}
𝑑𝑡 2𝑚 −∞ 𝜕𝑥 2𝑚 −∞ 𝜕𝑥
=0
𝑑 〈𝑥 〉 𝑖ℏ ∞ 𝜕Ψ 1
= − ∫ 𝑑𝑥 Ψ∗ = 〈𝑝〉 =>
𝑑𝑡 𝑚 −∞ 𝜕𝑥 𝑚
∞
𝜕Ψ
〈𝑝〉 = −𝑖ℏ ∫ 𝑑𝑥 Ψ ∗
−∞ 𝜕𝑥
Let’s rewrite 〈𝑥 〉 and 〈𝑝〉 in the form below
∞
〈𝑥 〉 = ∫ Ψ ∗ (𝑥 ) Ψ 𝑑𝑥
−∞
∞
ℏ 𝜕
〈𝑝 〉 = ∫ Ψ∗ ( ) Ψ 𝑑𝑥
−∞ 𝑖 𝜕𝑥
ℏ 𝜕
The operator ( ) represents momentum
𝑖 𝜕𝑥
20
can in fact show that expectation value of the momentum is always real.
We write
∞ ∞
ℏ 𝜕 ℏ 𝜕
〈𝑝 〉 − 〈𝑝 〉∗ =∫ Ψ ( ∗
) Ψ 𝑑𝑥 − ∫ Ψ (− ) Ψ∗ 𝑑𝑥
−∞ 𝑖 𝜕𝑥 −∞ 𝑖 𝜕𝑥
ℏ ∞ ∗
𝜕Ψ 𝜕Ψ ∗ ℏ ∞ 𝜕 ℏ
= ∫ (Ψ +Ψ ) 𝑑𝑥 = ∫ (Ψ ∗ Ψ) 𝑑𝑥 = (Ψ∗ Ψ)∞
−∞ = 0
𝑖 −∞ 𝜕𝑥 𝜕𝑥 𝑖 −∞ 𝜕𝑥 𝑖
Last step follows from the square integrability of the wave function.
An operator whose expectation value for all admissible wave functions is real
operator.
21
𝐴̂ 𝜓 = 𝑎 𝜓
An operator is an instruction to do something to the function that follows
it.
EX: Some mathematical operators which are not necessarily connected to
physics.
1)
𝑑 𝑑𝑓(𝑥)
̂=
𝐷 such that ̂ 𝑓(𝑥) =
𝐷
𝑑𝑥 𝑑𝑥
If f.ex. 𝑓(𝑥 ) = 𝑥 2 then
𝑑𝑓 (𝑥 )
̂ 𝑓(𝑥 ) =
𝐷 =2𝑥
𝑑𝑥
If 𝑓(𝑥 ) = 𝑒 −𝑖𝜔𝑥 , where 𝜔 is constant, then;:
𝜕𝑓 (𝑥 )
̂ 𝑓(𝑥 ) =
𝐷 = −𝑖𝜔𝑒 −𝑖𝜔𝑥 = −𝑖𝜔𝑓 (𝑥 ): 𝑒𝑖𝑔𝑒𝑛𝑣𝑎𝑙𝑢𝑒 𝑒𝑞.
𝜕𝑥
2)
𝐼̂ ∶ identity operator such that𝐼̂𝑓(𝑥 ) = 𝑓(𝑥 )for any function 𝑓(𝑥)
-------------------------------------------------------------------------------------------
22
How to calculate the expectation values of other quantities? For this note that
and/or momentum:
𝑸(𝒙, 𝒑)
EX:
1 2
𝑝2
𝑇 = 𝑚𝑣 =
2 2𝑚
Angular Momentum:
𝐿⃗ = 𝑟 × 𝑝
𝑸(𝒙, 𝒑),
ℏ 𝜕
we simply replace every 𝑝 → 𝑝̂ = ( ) , insert the resulting operator
𝑖 𝜕𝑥
∞
ℏ 𝜕
〈𝑄(𝑥, 𝑝)〉 = ∫ Ψ ∗ 𝑄 (𝑥, ) Ψ 𝑑𝑥
−∞ 𝑖 𝜕𝑥
𝑝2 1 ∞ ∗ ℏ 𝜕 2 ℏ2 ∞ ∗ 𝜕 2 Ψ
〈𝑇 〉 = 〈 〉 = ∫ Ψ ( ) Ψ 𝑑𝑥 = − ∫ Ψ 𝑑𝑥
2𝑚 2𝑚 −∞ 𝑖 𝜕𝑥 2𝑚 −∞ 𝜕𝑥 2
23
Note that the probability of finding this electron anywhere along the x-axis
is the same:
|𝜓(𝑥, 𝑡)|2 = |𝐴|2 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡.
It is clear therefore that this plane wave can not represent a localized electron; it is just
an indefinite series of wave with the same amplitude A. Instead, a particle which is
localized within a certain region of space can be described by a wave whose amplitude
is large in that region and zero outside it. In another word, the matter wave must be
localized around the region of space within which the particle is confined. Such a
localized wave is called a wave packet. We can build up localized wave packets by
adding up these free particle wave functions. A wave packet therefore consists of a
group of waves of slightly different wavelengths, with phases and amplitudes so chosen
that they interfere constructively over a small region of space and destructively
elsewhere. Mathematically, we can carry out this superposition by means of Fourier
transforms. For simplicity, we are going to consider a one-dimensional wave packet
∞
1
Ψ(𝑥, 𝑡) = ∫ 𝜙(𝑘, 𝑡)𝑒 𝑖(𝑘𝑥−𝜔𝑡) 𝑑𝑘
√2𝜋
−∞
“It is impossible to know both the exact position and exact momentum
of an object at the same time. There is a minimum for the product of
uncertainties of these two measurements.”
“It is impossible to know both the exact position and exact momentum
of an object at the same time. There is a minimum for the product of
uncertainties of these two measurements.”
HUP
The more precisely the position is known the more uncertain the
momentum is and vice versa.
3
and then take the square root to find the standard deviation 𝜎:
𝜎 = √∑ 𝑃𝑖 (𝑥𝑖 − 〈𝑥 〉)2
𝑖
= ∑ 𝑃𝑖 𝑥𝑖2 − 2〈𝑥 〉 ∑ 𝑃𝑖 𝑥𝑖 + 〈𝑥 〉2 ∑ 𝑃𝑖
⏟𝑖 ⏟𝑖 ⏟𝑖
〈𝑥 2 〉 〈𝑥〉 1
Note that when the position data comprise a discrete set, then the expectation
value is given by
〈𝑥 〉 = ∑ 𝑃𝑖 𝑥𝑖
𝑖
with
∑ 𝑃𝑖 = 1
𝑖
Thus
4
𝜎 2 = 〈𝑥 2 〉 − 2 〈⏟
𝑥 〉〈𝑥 〉 + 〈𝑥 〉2
〈𝑥〉2
OR
𝜎 2 = 〈𝑥 2 〉 − 〈𝑥 〉2 : 𝑣𝑎𝑟𝑖𝑎𝑛𝑐𝑒
And
∞ ∞ 0 ∞
For I1:
0 ∞
Thus,
∞ ∞ 0 ∞
𝐴 = √𝜆
7
∞
𝑛 −𝑎𝑥
𝑛!
∫𝑥 𝑒 𝑑𝑥 =
𝑎𝑛+1
0
8
|Ψ|2 = 𝜆𝑒 −2𝜆|𝑥|
Note that 〈𝑥 〉 = 0.
9
10
11
12
1
Schrödinger equation,
𝜕Ψ(𝑥, 𝑡) ℏ2 𝜕 2 Ψ(𝑥, 𝑡)
𝑖ℏ =− + 𝑉(𝑥, 𝑡) Ψ(𝑥, 𝑡)
𝜕𝑡 2𝑚 𝜕𝑥 2
Stationary States
Ψ(𝑥, 𝑡) = ⏟(𝑥 )
𝝍 𝑻
⏟ (𝑡)
𝑜𝑛𝑙𝑦 𝑎 𝑓𝑢𝑛𝑐𝑡𝑖𝑜𝑛 𝑜𝑛𝑙𝑦 𝑎 𝑓𝑢𝑛𝑐𝑡𝑖𝑜𝑛
𝑜𝑓 𝑥 𝑜𝑓 𝑡
2
𝜕 2 Ψ(𝑥, 𝑡) 𝑑2 𝜓(𝑥 )
= 𝑇(𝑡)
𝜕𝑥 2 𝑑𝑥 2
𝜕Ψ(𝑥, 𝑡) 𝑑T(𝑡)
(
=𝜓 𝑥 )
𝜕𝑡 𝑑𝑡
Then,
Then, we get
1.
1 𝑑𝑇(𝑡) 𝑑𝑇(𝑡) 𝑖𝐸
𝑖ℏ =𝐸 ⟹ + 𝑇(𝑡) = 0
𝑇(𝑡) 𝑑𝑡 𝑑𝑡 ℏ
𝑇(𝑡) = 𝐴 𝑒 −𝑖𝐸𝑡/ℏ
2.
−ℏ2 1 𝑑2 𝜓(𝑥 )
+ 𝑉(𝑥) = 𝐸 ⟹
2𝑚 𝜓(𝑥 ) 𝑑𝑥 2
−ℏ2 𝑑 2
[ + 𝑉(𝑥)] 𝜓(𝑥)𝐸𝜓(𝑥) :Steady-state form of the Schrödinger Equation
2𝑚 𝑑𝑥 2
𝑝𝑟𝑜𝑣𝑒𝑑 𝑎𝑙𝑟𝑒𝑎𝑑𝑦
𝑖𝑛 (2) ∞
〈𝐻 〉Ψ(x,t) =
⏞ 〈𝐻 〉𝜓(𝑥) = ∫ ̂⏟𝜓(𝑥 ) 𝑑𝑥
𝜓 ∗ (𝑥 ) 𝐻
−∞ 𝐸𝜓(𝑥)
=1
∞ ⏞∞
∗(
= 𝐸∫ 𝜓 𝑥 ) 𝜓(𝑥 )𝑑𝑥 = 𝐸 ∫ |𝜓(𝑥 )|2 𝑑𝑥
−∞ −∞
〈𝐻 〉 = 𝐸
6
∆𝐻 = √〈𝐻 2 〉 − 〈𝐻 〉2
∞ ∞
〈𝐻 2〉
=∫ ∗( ̂𝐻
𝜓 𝑥) 𝐻 ̂⏟𝜓(𝑥 ) 𝑑𝑥 = 𝐸 ∫ ̂⏟𝜓(𝑥 ) 𝑑𝑥 = 𝐸 2
𝜓 ∗ (𝑥 ) 𝐻
−∞ 𝐸𝜓(𝑥) −∞ 𝐸𝜓(𝑥)
∆𝐻 = √𝐸 2 − 𝐸 2
∆𝐻 = 0
It means that the distribution has zero spread; so that every member of the
Conclusion: A steady state of the system has the property that every
𝑖𝐸1 𝑡 𝑖𝐸2 𝑡
Ψ1 (𝑥, 𝑡) = 𝜓1 (𝑥 ) 𝑒 − ℏ , Ψ2 (𝑥, 𝑡) = 𝜓2 (𝑥 ) 𝑒 − ℏ ,…
Or explicitly writing
𝑖𝐸1 𝑡 𝑖𝐸2 𝑡 𝑖𝐸3 𝑡
Ψ(𝑥, 𝑡) = 𝑐1 𝜓1 (𝑥 ) 𝑒 − ℏ + 𝑐2 𝜓2 (𝑥 ) 𝑒 − ℏ + 𝑐3 𝜓3 (𝑥 ) 𝑒 − ℏ +⋯
∞
𝑖𝐸𝑛 𝑡
= ∑ 𝑐𝑛 𝜓𝑛 (𝑥 ) 𝑒 − ℏ
𝑛=1
8
That is
∞ ∞
𝑖𝐸𝑛 𝑡
Ψ(𝑥, 𝑡) = ∑ 𝑐𝑛 Ψ𝑛 (𝑥, 𝑡) = ∑ 𝑐𝑛 𝜓𝑛 (𝑥 ) 𝑒 − ℏ
𝑛=1 𝑛=1
SUMMARY:
The generic problem is as follows:
You are given a time-independent potential 𝑉(𝑥 ), and the starting wave function
Ψ(𝑥, 0):
∞ ∞
𝑖𝐸 0
− 𝑛
Ψ(𝑥, 0) = ∑ 𝑐𝑛 Ψ𝑛 (𝑥, 0) = ∑ 𝑐𝑛 𝜓𝑛 (𝑥 ) 𝑒⏟ ℏ
𝑛=1 𝑛=1 1
Your job is to find the wave function Ψ(𝑥, 𝑡) for any subsequent time t.
To do this you must solve the time-dependent Schrödinger equation
𝜕Ψ(𝑥, 𝑡) ℏ2 𝜕 2 Ψ(𝑥, 𝑡)
𝑖ℏ =− + 𝑉 (𝑥 ) Ψ(𝑥, 𝑡)
𝜕𝑡 2𝑚 𝜕𝑥 2
The strategy is first to solve the time-independent equation (Due to the
technique of separation of variables)
−ℏ2 𝑑2
[ + 𝑉(𝑥)] 𝜓(𝑥 ) = 𝐸𝜓(𝑥 )
2𝑚 𝑑𝑥 2
This yields in general an infinite set of solution 𝜓1 (𝑥 ), 𝜓2 (𝑥 ), 𝜓3 (𝑥 ), …, each
with its associated energy, 𝐸1 , 𝐸2 , 𝐸3 , …,.(separation constant). Thus, there is a
different wave function for each allowed energy:
𝑖𝐸1 𝑡 𝑖𝐸2 𝑡
Ψ1 (𝑥, 𝑡) = 𝜓1 (𝑥 ) 𝑒 − ℏ , Ψ2 (𝑥, 𝑡) = 𝜓2 (𝑥 ) 𝑒 − ℏ ,…
𝑖𝐸𝑛 𝑡
Ψ𝑛 (𝑥, 𝑡) = 𝜓𝑛 (𝑥 ) 𝑒 − ℏ
Now, the time-dependent Schrödinger equation has the property that any
linear combination of the solutions is itself a solution. Then the most
general solution is
10
∞ ∞
𝑖𝐸𝑛 𝑡
Ψ(𝑥, 𝑡) = ∑ 𝑐𝑛 Ψ𝑛 (𝑥, 𝑡) = ∑ 𝑐𝑛 𝜓𝑛 (𝑥 ) 𝑒 − ℏ
𝑛=1 𝑛=1
To fit Ψ(𝑥, 0) you write down the general linear combination of these
solutions
∞
Ψ(𝑥, 0) = ∑ 𝑐𝑛 𝜓𝑛 (𝑥 )
𝑛=1
are stationary solutions, in the sense that all probabilities and expectation
values are independent of time.
However, this property is not shared by the general solution Ψ(𝑥, 𝑡); the
energies are different for different stationary states, and the exponentials do
not cancel when |Ψ(𝑥, 𝑡)|2 is calculated
11
2
𝑖𝐸𝑛 𝑡 𝑖𝐸1 𝑡 𝑖𝐸2 𝑡
Ψ(𝑥, 𝑡) = ∑ 𝑐𝑛 𝜓𝑛 (𝑥 ) 𝑒 − ℏ = 𝑐1 𝜓1 (𝑥 ) 𝑒 − ℏ + 𝑐2 𝜓2 (𝑥 ) 𝑒 − ℏ
𝑛=1
Such that at t=0, we recover the initial state of the particle given by
Ψ(𝑥, 0) = 𝑐1 𝜓1 (𝑥 ) + 𝑐2 𝜓2 (𝑥 )
Here, 𝐸1 and 𝐸2 are energies associated with 𝜓1 and 𝜓2 .
The probability density is
|Ψ(𝑥, 𝑡)|2 =
𝑖𝐸 𝑡 𝑖𝐸 𝑡 ∗ 𝑖𝐸 𝑡 𝑖𝐸 𝑡
− 1 − 2 − 1 − 2
(𝑐1 𝜓1 (𝑥 ) 𝑒 ℏ + 𝑐2 𝜓2 (𝑥 ) 𝑒 ℏ ) (𝑐1 𝜓1 (𝑥 ) 𝑒 ℏ + 𝑐2 𝜓2 (𝑥 ) 𝑒 ℏ )
𝑖𝐸1 𝑡 𝑖𝐸2 𝑡 𝑖𝐸1 𝑡 𝑖𝐸2 𝑡
= (𝑐1 𝜓1 (𝑥 ) 𝑒 ℏ + 𝑐2 𝜓2 (𝑥 ) 𝑒 ℏ ) (𝑐1 𝜓1 (𝑥 ) 𝑒− ℏ + 𝑐2 𝜓2 (𝑥 ) 𝑒 − ℏ )
(𝐸2 − 𝐸1 )𝑡
|Ψ(𝑥, 𝑡)|2 = 𝑐12 𝜓12 + 𝑐22 𝜓22 + 2𝑐1 𝑐2 𝜓1 𝜓2 cos
ℏ
12
This is the simplest model that can be used for illustrating the quantum
mechanics. It is a hypothetical example, used to illustrate the differences
between classical and quantum systems.
2L 2L 2m 2m
The period of motion is T L L
v p/m 2mE E
However, when the well becomes very narrow (on the scale of a
few nanometers), quantum effects become important. The problem
becomes very interesting when one attempts a quantum mechanical
solution with applying Schrodinger equation, since many fundamental
quantum mechanical concepts need to be introduced to find the solution.
Nevertheless, it remains a very simple and solvable problem.
The possible de Broglie wavelengths of the particle in the box are
determined by L.
14
∞ , 𝑥 ≤ 0, 𝑥 ≥ 𝐿 (DI,DIII)
𝑉(𝑥) = {
0 0 < 𝑥 < 𝐿 (DII)
−ℏ2 𝑑2
̂
𝐻 𝜓(𝑥 ) = 𝐸𝜓(𝑥 ) => [ + 𝑉 (𝑥 )] 𝜓(𝑥 ) = 𝐸𝜓(𝑥 ) (1)
2𝑚 𝑑𝑥 2
−ℏ2 𝑑 2
[ + ∞] 𝜓𝐼 (𝑥 ) = 𝐸𝐼 𝜓𝐼 (𝑥 ) when 𝑥 ≤ 0
2𝑚 𝑑𝑥 2
−ℏ2 𝑑2
[ + ∞] 𝜓𝐼𝐼𝐼 (𝑥 ) = 𝐸𝐼𝐼𝐼 𝜓𝐼𝐼𝐼 (𝑥 ) when 𝒙 ≥ 𝑳
2𝑚 𝑑𝑥 2
For finite 𝐸𝐼 , 𝐸𝐼𝐼𝐼 and 𝜓𝐼 (𝑥 ), 𝜓𝐼𝐼𝐼 (𝑥 ), RHS is finite. This implies that
LHS must be finite too. This may be possible if
𝜓𝐼 (𝑥 ) = 0
𝜓𝐼𝐼𝐼 (𝑥 ) = 0
15
DI and DIII are the examples for forbidden domains, where E<V.
(E≡ 𝐾 + 𝑉 𝑎𝑛𝑑 𝑖𝑓 𝐸 < 𝑉 𝑡ℎ𝑖𝑠 𝑚𝑒𝑎𝑛𝑠 𝐾 < 0 !)
−ℏ2 𝑑2
[ + 0] 𝜓𝐼𝐼 (𝑥 ) = 𝐸𝜓𝐼𝐼 (𝑥 ) when 0 < 𝑥 < 𝐿 (2)
2𝑚 𝑑𝑥 2
𝜕 2 𝜓𝐼𝐼 (𝑥 )
2
+ 𝑘 2 𝜓𝐼𝐼 (𝑥 ) = 0
𝜕𝑥
OR,
Here A,B (or C,D) are constants, and k can be any real number.
16
Boundary Conditions:
1. 𝜓𝐼 (𝑥 )|𝑥=0 = 𝜓𝐼𝐼 (𝑥 )|𝑥=0 = 0 ⟹ 𝐵 cos 𝑘0 = 0 ⟹ 𝐵 = 0
sin(−𝜃 ) = − sin(𝜃 ),
and this minuse sign can be absorbed into the constant A. (Also, −sin(𝜃 )
gives the same probability when squared)
𝑛𝜋
𝜓𝐼𝐼 (𝑥 ) ≡ 𝜓𝑛 (𝑥 ) = 𝐴 sin 𝑘𝑛 𝑥 with 𝑘𝑛 = , 𝑛 = 1,2,3, …
𝐿
Note that n=0 solution is ruled out because in this case 𝜓𝐼𝐼 (𝑥 ) = 0 and
the whole solution to Sch.eq, is zero. This corresponds to the case
where there is no the particle in the box !.
17
𝑥 sin(2𝛽𝑥)
Using ∫ sin2 (𝛽𝑥 ) 𝑑𝑥 = − , evaluate the integral above
2 4𝛽
𝐿
𝑥 𝑛𝜋 1 𝐿 2
|2
|𝐴 [ − sin (2 𝑥) ] = 1 ⟹ |𝐴|2 [ − 0] = 1 ⟹ 𝐴 = √
2 𝐿 4𝑛𝜋/𝐿 0 2 𝐿
2 𝑛𝜋
𝜓𝑛 (𝑥 ) = √ sin ( 𝑥) ∶ eigenfunction of H
𝐿 𝐿
ℏ2 𝑘𝑛2 ℏ2 𝜋 2 2
𝐸𝑛 = = 𝑛 , 𝑛 = 1,2,3, … ∶ energy eigenvalues of the H
2𝑚 2𝑚𝐿2
ℏ2 𝜋 2
𝐸1 = ∶ the ground state energy
2𝑚𝐿2
Then, in terms of ground state energy, the quantized energy levels are
written as 𝐸𝑛 = 𝐸1 𝑛2 , 𝑛 = 1,2,3, …
18
The solution to the Schrödinger eq. for the particle in the infinite well
problem reveals some quantum behavior of the particle that contrasts
sharply with the predictions of classical mechanics.
The waves are in the form of standing waves on a string of length L with
wavelength such that it fits into the space available (Figure). We see that
the wavelength of the wave is restricted to one of the values corresponding
to a whole number of half wavelengths fitting into the box:
2𝜋 𝑛𝜋 2𝐿
𝜆= and 𝑘𝑛 = => 𝜆𝑛 =
𝑘 𝐿 𝑛
This means that only these particular values of the wavelength are
allowed and, as the electron momentum is determined by the wavelength
through the de Broglie relation, the momentum is also restricted to a
particular set of values.
2𝐿 ℎ
𝜆𝑛 = =
𝑛 𝑝𝑛
𝐿
In all quantum states, 〈𝑥 〉 =
2
𝐿 𝐿
2 𝑛𝜋𝑥
〈𝑥 〉 = ∫ 𝑥 | 𝜓𝑛 (𝑥 )|2 𝑑𝑥 = ∫ 𝑥 sin2 ( ) 𝑑𝑥
0 𝐿 0 𝐿
𝐿 𝐿 𝐿
= + (1 − cos
⏟ 2𝑛𝜋 ) − sin 2𝑛𝜋
⏟
2 4(𝑛𝜋)2 1
2𝑛𝜋 0
𝐿
〈𝑥 〉 =
2
EX: Find the probability that a particle trapped in a box L wide can be
found between (0.45L,0.55L) for the ground state and for the 1.st excited
state.
In general,
𝑥2
2 𝑥2 2 𝑛𝜋𝑥
𝑃𝑛 (𝑥1 , 𝑥2 ) = ∫ 𝜓𝑛∗ (𝑥 )
𝜓𝑛 (𝑥 ) 𝑑𝑥 = ∫ sin ( ) 𝑑𝑥
𝑥1 𝐿 𝑥1 𝐿
𝑥2
2𝐿 𝑥 1 2𝜋𝑛
𝑃𝑛 (𝑥1 , 𝑥2 ) = [ − sin ( 𝑥)]
𝐿 2 𝐿 2𝑛𝜋 𝐿 𝑥1
0.55𝐿
𝑥 1 2𝜋
𝑃1 (0.45𝐿, 0.55𝐿) = [ − sin ( 𝑥)] = 0.198 ⟹
𝐿 2𝜋 𝐿 0.45𝐿
0.55𝐿
𝑥 1 4𝜋
𝑃2 (0.45𝐿, 0.55𝐿) = [ − sin ( 𝑥)] = 0.006 ⟹
𝐿 4𝜋 𝐿 0.45𝐿
Ψ(𝑥, 𝑡) = ∑ 𝑐𝑛 Ψ𝑛 (𝑥, 𝑡)
𝑛=1
The stationary states of the infinite square well are written from
𝑖𝐸 𝑡
− 𝑛
Ψ𝑛 (𝑥, 𝑡) = 𝜓𝑛 (𝑥 ) 𝑒 ℏ =>
as
2 𝑛𝜋 ℏ𝜋2 𝑛2
−𝑖 𝑡
Ψ𝑛 (𝑥, 𝑡) = √ sin ( 𝑥) 𝑒 2𝑚𝐿2
𝐿 𝐿
Proof:
𝐿
𝐿
2 𝑛𝜋 𝑚𝜋
∫ 𝜓𝑛∗ (𝑥 ) 𝜓𝑚 (𝑥 ) 𝑑𝑥 = ∫ sin ( 𝑥) sin ( 𝑥) 𝑑𝑥
0 𝐿 𝐿 𝐿
0
𝐿
12 (𝑚 − 𝑛)𝜋 (𝑚 + 𝑛)𝜋
= ∫ Cos ( 𝑥) − Cos ( 𝑥) 𝑑𝑥
2𝐿 𝐿 𝐿
0
23
𝐿
1 𝐿 (𝑚 − 𝑛)𝜋 𝐿 (𝑚 + 𝑛)𝜋
= { Sin ( 𝑥) − Sin ( 𝑥)}
𝐿 (𝑚 − 𝑛)𝜋 𝐿 (𝑚 + 𝑛)𝜋 𝐿 0
1 (𝑚 − 𝑛)𝜋 (𝑚 + 𝑛)𝜋
= {Sin ( ) − Sin ( )} = 0
𝜋 (𝑚 − 𝑛) (𝑚 + 𝑛)
where we used
1
sin(𝑎𝑥 ) sin(𝑏𝑥 ) = [cos(𝑎 − 𝑏) 𝑥 − cos(𝑎 + 𝑏) 𝑥 ]
2
When 𝑛 ≠ 𝑚, normalization tells us that integral is unity. In fact we can
combine orthogonality and normalization into a single statement:
𝐿
0 𝑖𝑓 𝑛 ≠ 𝑚
∫ 𝜓𝑛∗ (𝑥 ) 𝜓𝑚 (𝑥 ) 𝑑𝑥 = 𝛿𝑛𝑚 = { }
0
1 𝑖𝑓 𝑛 = 𝑚
𝜓𝑚 (𝑥 )′𝑠 are complete, in the sense that any other function 𝑓(𝑥) can be
expressed as a linear combination of them
∞ ∞
2 𝑛𝜋
( ) ( )
𝑓 𝑥 = ∑ 𝑐𝑛 𝜓𝑛 𝑥 = √ ∑ 𝑐𝑛 sin ( 𝑥)
𝐿 𝐿
𝑛=1 𝑛=1
This is nothing but the Fourier series for 𝑓(𝑥 ) and the fact that any
function can be expanded in this way is called Dirichlet’s theorem.
∗ ( )
The coefficients 𝑐𝑛 can be obtained by multiplying both sides by 𝜓𝑚 𝑥
and integrating:
∞ ∞
∗ ( ) ( ) ∗ ( )
∫ 𝜓𝑚 𝑥 𝑓 𝑥 𝑑𝑥 = ∑ 𝑐𝑛 ∫ 𝑑𝑥 𝜓𝑛 (𝑥 )𝜓𝑚 𝑥 = ∑ 𝑐𝑛 𝛿𝑛𝑚 = 𝑐𝑚
𝑛=1 𝑛=1
Note that 𝛿𝑛𝑚 kills every term in the sum except the one for which n=m.
Therefore
𝑐𝑛 = ∫ 𝜓𝑛∗ (𝑥 ) 𝑓(𝑥 )𝑑𝑥
------------------------------------------------------------------
24
If the initial wave function is given 𝚿(𝒙, 𝟎) and we are asked to find the
most general solution to the time dependent Schrödinger equation:
1. Find 𝑐𝑛 . For this start from (*) and get Ψ(𝑥, 0):
∞ ∞ ∞
2 𝑛𝜋
Ψ(𝑥, 0) = ∑ 𝑐𝑛 Ψ𝑛 (𝑥, 0) = ∑ 𝑐𝑛 𝜓𝑛 (𝑥 ) = √ ∑ 𝑐𝑛 sin ( 𝑥)
𝐿 𝐿
𝑛=1 𝑛=1 𝑛=1
Thus,
2 𝑛𝜋
𝑐𝑛 = ∫ 𝜓𝑛∗ (𝑥 ) Ψ(𝑥, 0)𝑑𝑥 = √ ∫ sin ( 𝑥) Ψ(𝑥, 0)𝑑𝑥 (∗∗)
𝐿 𝐿
Example 2.2. A particle in the infinite square well has the initial wave function.
At time t=0, a particle is represented by the wave function
𝐴 𝑥 (𝐿 − 𝑥 ) 𝑖𝑓 0 ≤ 𝑥 ≤ 𝐿
Ψ(𝑥, 0) = {
0 𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒
Find Ψ(𝑥, 𝑡).
Since
∞
2 𝑛𝜋 ℏ𝜋2 𝑛2
−𝑖 𝑡
Ψ(𝑥, 𝑡) = √ ∑ 𝑐𝑛 sin ( 𝑥) 𝑒 2𝑚𝐿2 (∗)
𝐿 𝐿
𝑛=1
We need to find 𝑐𝑛 .
First normalize it:
𝐿 𝐿
> 𝐴2 ∫ 𝑥 2 (𝑥 2 − 2𝐿𝑥 + 𝐿2 ) 𝑑𝑥 = 1
0
5 𝐿 𝐿 𝐿
2
𝑥 𝑥4 2
𝑥3 2
𝐿5 𝐿5 𝐿5
1 = 𝐴 { | − 2𝐿 | + 𝐿 | }=𝐴 { − + }
5 0 4 0 3 0 5 2 3
𝐿5 2
=𝐿
30
30
𝐴=√
𝐿5
2√15 𝐿 2 𝑛𝜋 𝑛𝜋 𝑛𝜋
= 3 {𝐿 [( ) (sin ( 𝑥) − 𝑥 Cos ( 𝑥))]
𝐿 𝑛𝜋 𝐿 𝐿 𝐿
𝐿
𝐿 3 𝑛𝜋 𝑛𝜋 𝑛𝜋 𝑛𝜋 2 𝑛𝜋
− ( ) (2 Cos ( 𝑥) + 2 𝑥 Sin ( 𝑥) − ( 𝑥) Cos ( 𝑥))}
𝑛𝜋 𝐿 𝐿 𝐿 𝐿 𝐿 0
2√15 𝐿 2 𝑛𝜋
= 3 {[𝐿 [( ) (0 − 𝐿 Cos(𝑛𝜋))]
𝐿 𝑛𝜋 𝐿
𝐿 3 𝑛𝜋 2
− ( ) (2 Cos(𝑛𝜋) + 0 − ( 𝐿) Cos(𝑛𝜋))]
𝑛𝜋 𝐿
𝐿 2 𝐿 3
− [𝐿 [( ) (0 − 0)] − ( ) (2 Cos(0) + 0 − 0)]}
𝑛𝜋 𝑛𝜋
2√15 𝐿3 𝐿 3 𝐿 3
= 3 { Cos(𝑛𝜋) + ( ) ((𝑛𝜋)2 − 2) Cos(𝑛𝜋) + ( ) 2 Cos(0)}
𝐿 𝑛𝜋 𝑛𝜋 𝑛𝜋
0 𝑖𝑓 𝑛 𝑖𝑠 𝑒𝑣𝑒𝑛
4√15
= {Cos(0) − Cos(𝑛𝜋)} => 𝑐𝑛 = { 8√15 }
(𝑛𝜋)3 𝑖𝑓 𝑛 𝑖𝑠 𝑜𝑑𝑑
(𝑛𝜋)3
Put into (*):
∞
2 8√15 1 𝑛𝜋 ℏ𝜋2 𝑛2
−𝑖 𝑡
Ψ(𝑥, 𝑡) = √ ∑ sin ( 𝑥) 𝑒 2𝑚𝐿2
𝐿 (𝜋)3 𝑛3 𝐿
𝑛=1,3,5,…
OR
∞
30 2 3 1 𝑛𝜋 −𝑖
ℏ𝜋2 𝑛2
𝑡
Ψ(𝑥, 𝑡) = √ ( ) ∑ sin ( 𝑥) 𝑒 2𝑚𝐿2
𝐿 𝜋 𝑛3 𝐿
𝑛=1,3,5,…
27
1
sin(𝑎𝑥 ) sin(𝑏𝑥 ) = [cos(𝑎 − 𝑏) 𝑥 − cos(𝑎 + 𝑏) 𝑥 ]
2
-------------------------------------------------------------------------------
The meaning of 𝑐𝑛 :
|𝑐𝑛 |2 :
𝑡ℎ𝑒 𝑝𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 𝑡ℎ𝑎𝑡 𝑎 𝑚𝑒𝑎𝑠𝑢𝑟𝑒𝑚𝑒𝑛𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑒𝑛𝑒𝑟𝑔𝑦 𝑤𝑜𝑢𝑙𝑑 𝑦𝑖𝑒𝑙𝑑
𝑡ℎ𝑒 𝑣𝑎𝑙𝑢𝑒 𝐸𝑛
∑ |𝑐𝑛 |2 = 1
𝑛=1
1 = ∫|Ψ(𝑥, 0)|2 𝑑𝑥 = ∫ (∑ 𝑐𝑛 𝜓𝑛 (𝑥 )) ( ∑ 𝑐𝑚 𝜓𝑚 (𝑥 )) 𝑑𝑥
𝑛=1 𝑚=1
28
∞ 𝐿 ∞ ∞
Moreover, the expectation value of the energy in the most general state
is calculated directly:
∗
∞ ∞
〈𝐻 〉 = ∫ Ψ ∗ 𝐻Ψdx = ∫ (∑ 𝑐𝑛 𝜓𝑛 (𝑥 )) 𝐻 ( ∑ 𝑐𝑚 𝜓𝑚 (𝑥 )) 𝑑𝑥
𝑛=1 𝑚=1
∞ 𝐿 ∞ 𝐿
= ∑ 𝑐𝑛 𝑐𝑚 ∫ 𝜓𝑛∗ (𝑥 ) 𝐻𝜓
∗
⏟ 𝑚 (𝑥 ) 𝑑𝑥 = ∑ 𝑐𝑛 𝑐𝑚 𝐸𝑚 ∫ 𝜓𝑛∗ (𝑥 )𝜓𝑚
∗
𝑑𝑥
𝑛,𝑚 0 =𝐸𝑚 𝜓𝑚 𝑛,𝑚
⏟0
𝛿𝑛𝑚
∞ ∞
Note that
|𝑐𝑛 |2 = probability of getting a particular energy is independent of time
〈𝐻 〉: independent of time
This is a manifestation of conservation of energy
29
2 𝑛𝜋 𝑑 2 𝑛𝜋 𝑛𝜋
𝜓𝑛 = 𝜓𝑛∗ = √ sin ( 𝑥) ; 𝜓𝑛 = √ Cos ( 𝑥)
𝐿 𝐿 𝑑𝑥 𝐿 𝐿 𝐿
2
2 𝑛𝜋 𝐿 𝑛𝜋 𝑛𝜋
〈𝑝〉 = −𝑖ℏ (√ ) ∫ sin ( 𝑥) Cos ( 𝑥) 𝑑𝑥
𝐿 𝐿 0 𝐿 𝐿
1
Using ∫ sin 𝛼𝑥 cos 𝛼𝑥 𝑑𝑥 = 𝑆𝑖𝑛2 𝛼𝑥 ,
2𝛼
2 𝑛𝜋 𝐿 2
𝑛𝜋 𝐿 −𝑖ℏ
〈𝑝〉 = −𝑖ℏ 𝑆𝑖𝑛 ( 𝑥) = [𝑆𝑖𝑛2 𝑛𝜋 − 0] = 0
𝐿 𝐿 2𝑛𝜋 𝐿 0 𝐿
〈𝑝 〉 = 0
----------------------------------------------------------------------
Note that for any real function 𝑅 ∗ (𝑥 ) = 𝑅 (𝑥 ),
∞
𝑑
〈𝑝〉 = ∫ 𝑅 ∗ (−𝑖ℏ ) 𝑅 𝑑𝑥
−∞ 𝑑𝑥
is imaginary (because of i), which is incompatible with the requirement
that 〈𝑝〉 = 〈𝑝〉∗ unless 〈𝑝〉 = 0
-------------------------------------------------------------------------
30
𝑝𝑛2 2
ℏ2 𝜋 2 2
𝐸𝑛 = , and 𝐸𝑛 = 𝐸1 𝑛 = 𝑛
2𝑚 2𝑚𝐿2
𝑛𝜋ℏ
𝑝𝑛 = ±√2𝑚𝐸𝑛 = ±
𝐿
𝑛𝜋ℏ 𝑛𝜋ℏ
(+ ) + (− )
𝑝̅ = 𝐿 𝐿 =0
2
31
Momentum Eigenvalue:
2 𝑛𝜋 ℏ2 𝜋 2 2
𝜓𝑛 (𝑥) = √ sin ( 𝑥) , 𝐸𝑛 = 𝑛 , 𝑛 = 1,2,3, ….
𝐿 𝐿 2𝑚𝐿2
𝑑 𝑑 2 𝑛𝜋
𝐿𝐻𝑆 = 𝑝̂ 𝜓𝑛 = −𝑖ℏ 𝜓𝑛 = −𝑖ℏ [√ sin ( 𝑥)]
𝑑𝑥 𝑑𝑥 𝐿 𝐿
2 𝑛𝜋 𝑛𝜋
= −𝑖ℏ√ Cos ( 𝑥) ≠ (a constant)𝜓𝑛
𝐿 𝐿 𝐿
2 𝑛𝜋 2 1 𝑖𝑛𝜋𝑥 1 −𝑖𝑛𝜋𝑥
𝜓𝑛 (𝑥 ) = √ sin ( 𝑥) = √ ( 𝑒 𝐿 − 𝑒 𝐿 ) ≡ 𝜓𝑛+ − 𝜓𝑛−
𝐿 𝐿 𝐿 2𝑖 2𝑖
𝜓𝑛± ≡
𝑑 + 𝑑 2 1 𝑖𝑛𝜋𝑥
𝑝̂ 𝜓𝑛+ = 𝑝𝑛+ 𝜓𝑛+ ⟹ −𝑖ℏ 𝜓𝑛 = −𝑖ℏ [√ 𝑒 𝐿 ]
𝑑𝑥 𝑑𝑥 𝐿 2𝑖
Similarly,
𝑑 − 𝑑 2 1 𝑖−𝑛𝜋𝑥
𝑝̂ 𝜓𝑛− = 𝑝𝑛− 𝜓𝑛− ⟹ −𝑖ℏ 𝜓𝑛 = −𝑖ℏ [−√ 𝑒 𝐿 ]
𝑑𝑥 𝑑𝑥 𝐿 2𝑖
−𝑛𝜋ℏ
⟹ 𝑝𝑛− =
𝐿
33
A study of the Simple Harmonic Oscillator (SHO) is important in QM. Since there are only
limited number of continuous potentials for which it is possible to obtain solutions to the
Schrödinger equation by analytical techniques, and SHO potential is one of them.
Coulomb potential : 𝑉(𝑟)~𝑟 −1
SHO potential : 𝑉(𝑟)~𝑟 2
In addition, SHO is the prototype for any system involving oscillations. The system may be an
object supported by a spring or a diatomic molecule
The condition for harmonic motion is the presence of a restoring force that acts to return the
system to its equilibrium configuration when it is disturbed.
Note that any restoring force 𝐹(𝑥) can be expressed in a Maclaurin’s series about the
equilibrium position (take x=0) as
𝑑𝐹 1 𝑑𝐹 2
𝐹(𝑥) = 𝐹(0) + ( ) 𝑥+ ( ) 𝑥 2 + ⋯.
𝑑𝑥 𝑥=0 2 𝑑𝑥 𝑥=0
Since x=0 is the equilibrium position, 𝐹(0) = 0. For small x, keep the 1.st order term only
(higher orders are smaller). Then,
𝑑𝐹
𝐹(𝑥) ≅ ( ) 𝑥
𝑑𝑥 𝑥=0
𝑑𝐹
which is the Hooke’s law when ( ) < 0.
𝑑𝑥 𝑥=0
𝐹(𝑥) = −𝑘𝑥
2
The system has continuous energy starting from zero. Energy can have any value since A is
arbitrary.
However, quantum mechanics predicts that the total energy 𝐸 can assume only a discrete set of
values because the particle is bound by the potential to a region of finite extent. To find the
allowed energy values predicted by Schrödinger quantum mechanics, the time-independent
Schrödinger equation for the SHO potential must be solved.
3
Classical probability of finding the particle in the interval (𝑥, 𝑥 + 𝑑𝑥) is proportional to time
𝑑𝑡 which it takes to pass through this interval :
𝑑𝑡
𝑃(𝑥)𝑑𝑥 =
𝑇
Where
2𝜋
𝑇= : period of oscillation
𝜔
Then,
𝑑𝑡 𝑑𝑡 𝑑𝑥 1 𝑑𝑥 𝜔 𝑑𝑥
𝑃(𝑥)𝑑𝑥 = = = => 𝑃(𝑥)𝑑𝑥 =
2𝜋/𝜔 2𝜋/𝜔 𝑑𝑥 2𝜋/𝜔 𝑑𝑥/𝑑𝑡 2𝜋 𝑣
To find the velocity, write total energy:
1 1 1
𝐸 = 𝑚𝑣 2 + 𝑚𝜔2 𝑥 2 = 𝑚𝜔2 𝐴2 => 𝑣 = 𝜔√(𝐴2 − 𝑥 2 )
2 2 2
Then
1 𝑑𝑥 1
𝑃(𝑥)𝑑𝑥 = => 𝑃(𝑥) =
2𝜋 √(𝐴2 − 𝑥 2 ) 2𝜋√(𝐴2 − 𝑥 2 )
𝑃(𝑥) is largest near the turning points 𝑥 = ±𝐴, where the speed of the particle vanishes.
4
In contrast to the infinite square well problem, in SHO problem, the differential equation
that needs to be solved is not so trivial, and one reason for discussing this problem is to learn
something about the techniques for solving such equations.
The introduction of operators brings in a new concern: the order they act is important.
Consider
𝜕 𝜕𝜓(𝑥)
𝒙
̂𝒑̂ 𝜓(𝑥) = 𝑥 (−𝑖ℏ ) 𝜓(𝑥) = −𝑖ℏ𝑥 (𝑎)
𝜕𝑥 𝜕𝑥
And
𝜕 𝜕𝜓(𝑥)
̂𝒙
𝒑 ̂ 𝜓(𝑥) = (−𝑖ℏ ) (𝑥𝜓(𝑥)) = −𝑖ℏ {𝑥 + 𝜓(𝑥)} (𝑏)
𝜕𝑥 𝜕𝑥
Note that
𝒙
̂𝒑̂ 𝜓(𝑥) ≠ 𝒑
̂𝒙̂ 𝜓(𝑥)
(𝒙
̂𝒑̂− 𝒑
̂𝒙̂)𝜓(𝑥) ≡ [𝑥̂, 𝑝̂ ]𝜓(𝑥) = 𝑖ℏ𝜓(𝑥)
Since this is true for all 𝜓(𝑥), we conclude that we have an operator relation
The difference between classical and quantum mechanics lies in that physical variables are
described by operators in quantum mechanics and these do not necessarily commute.
5
If [𝐴̂, 𝐵̂] = 𝟎 => 𝐴̂𝐵̂ = 𝐵̂𝐴̂ , two operators are said to commute (𝐴̂ and 𝐵̂ are compatible)
[𝐴̂, 𝐴̂] = 𝟎
[𝑎̂− , 𝑎̂+ ] ≡ 𝑎̂− 𝑎̂+ − 𝑎̂+ 𝑎̂− = 1 => 𝑎̂− 𝑎̂+ ≠ 𝑎̂+ 𝑎̂− (∗)
We can write 𝑥̂ and 𝑝̂ in terms of 𝑎̂ and 𝑎̂+ :
𝑎̂− + 𝑎̂+ 𝑚𝜔 (𝑎̂− − 𝑎̂+ )
𝑥̂ = , 𝑝̂ =
√2𝛽 𝑖 √2𝛽
Therefore,
2 2
1 𝑚𝜔 (𝑎̂− − 𝑎̂+ ) 1 𝑎̂− + 𝑎̂+
̂=
𝐻 [ 2
] + 𝑚𝜔 [ ]
2𝑚 𝑖 √2𝛽 2 √2𝛽
2
(𝑎̂− ± 𝑎̂+ )2 = 𝑎̂− 2 + 𝑎̂+ ± (𝑎̂− 𝑎̂+ + 𝑎̂+ 𝑎̂−)
From eq.(*)
𝑎̂− 𝑎̂+ = 1 + 𝑎̂+ 𝑎̂−
Then,
2
(𝑎̂− ± 𝑎̂+ )2 = 𝑎̂− 2 + 𝑎̂+ ± (1 + 𝑎̂+ 𝑎̂− )
𝑚2 𝜔2 (−1) − 2 +2
𝑚𝜔2 − 2 2
̂=
𝐻 + −
(𝑎̂ + 𝑎̂ − (1 + 𝑎̂ 𝑎̂ )) + (𝑎̂ + 𝑎̂+ + (1 + 𝑎̂+ 𝑎̂− ))
2𝑚2𝛽 2 2 2𝛽 2
𝑚𝜔2 −2 +2 + − −2 +2
= 2
̂
[(−𝑎 − 𝑎
̂ + 1 + 𝑎
̂ 𝑎
̂ ) + (𝑎
̂ + 𝑎
̂ + 1 + 𝑎̂+ 𝑎̂− )]
4𝛽
7
1 ℏ
̂ = 𝑚𝜔2 (
𝐻 ) 2(1 + 𝑎̂+ 𝑎̂− ) ⟹
4 𝑚𝜔
1
̂ = 𝜔ℏ (𝑎̂+ 𝑎̂− + )
𝐻
2
OR
1
̂ = 𝜔ℏ (𝑎̂− 𝑎̂+ − )
𝐻
2
̂ , 𝑎̂∓ ]:
Let’s calculate the commutators [𝐻
1 1
̂ , 𝑎̂− ] = 𝜔ℏ [𝑎̂+ 𝑎̂− + , 𝑎̂− ] = 𝜔ℏ ([𝑎̂+ 𝑎̂− , 𝑎̂− ] + [ , 𝑎̂− ])
[𝐻
2 ⏟2
0
̂ , 𝑎̂− ] ≡ 𝐻
[𝐻 ̂ 𝑎̂− − 𝑎̂− 𝐻
̂ = −𝜔ℏ𝑎̂− (∗)
That is
̂ (𝑎
𝐻 ⏟̂ + 𝜓) = (𝐸 + 𝜔ℏ)(𝑎̂+ 𝜓) ∶ also an eigenvalue eq.
𝑎 𝑚𝑒𝑤 𝑠𝑡𝑎𝑡𝑒
̂ but, with a new egenvalue (𝐸 + 𝜔ℏ) that is
The new state (𝑎̂+ 𝜓) is also an eigenfunction of 𝐻
greater than E of the original state 𝜓 by one quantum of energy 𝜔ℏ.
Let’s prove that the state (𝑎̂− 𝜓) satisfies the Schrödinger equation too with the energy
(𝐸 − 𝜔ℏ). That is, we can write a new energy eigenvalue equation
̂ (𝑎
𝐻 ⏟̂ − 𝜓) = (𝐸 − 𝜔ℏ)(𝑎̂− 𝜓)
𝑎 𝑛𝑒𝑤 𝑠𝑡𝑎𝑡𝑒
That is
̂ (𝑎
𝐻 ⏟̂ − 𝜓) = (𝐸 − 𝜔ℏ)(𝑎̂− 𝜓) ∶ also an eigenvalue eq.
𝑎 𝑛𝑒𝑤 𝑠𝑡𝑎𝑡𝑒
̂ but, with a new egenvalue (𝐸 − 𝜔ℏ) that is
The new state (𝑎̂− 𝜓) is also an eigenfunction of 𝐻
smaller than E of the original state 𝜓 by one quantum of energy 𝜔ℏ.
9
Therefore we refer to these operators as ladder operators because they take us up and down a
ladder of energy eigenstates, as shown in the figures.
The 𝑎̂− is called the lowering operator , and 𝑎̂+ is called the raising operator.
𝑎̂− (𝑎̂+ ) lowers (raises) the energy eigenvalue by the amount of 𝜔ℏ.
10
We now use this condition to find the energy of the ground state:
1 1
̂ 𝜓0 = 𝐸0 𝜓0 ⟹ 𝐻
𝐻 ̂ 𝜓0 = 𝜔ℏ (𝑎̂+ 𝑎̂− + ) 𝜓0 = 𝜔ℏ (𝑎̂+ 𝑎
̂ − 𝜓0 + 𝜓0 ) ⟹
⏟
2 2
0
.
1 1
̂ 𝜓0 = 𝜔ℏ 𝜓0 = 𝐸0 𝜓0 ⟹ 𝐸0 = 𝜔ℏ ∶ Ground state Energy
𝐻
2 2
Note that ground state does not have zero energy, in contrast to the classical harmonics
oscillator. Rather, quantum mechanical ground state has a zero-point energy of 𝜔ℏ/2, that is
also consistent with the uncertainty principle.
11
To generate the next energy eigenstate up the ladder of states, we act with the raising operator
𝑎̂+ on the ground state 𝜓0 , which produces a new energy eigenstate with the energy increased
by one quantum 𝜔ℏ:
𝑎̂+ 𝜓0 ~𝜓1 such that 𝐻 ̂ 𝜓1 = 𝐸1 𝜓1
where
1 3
𝐸1 = 𝐸0 + 𝜔ℏ = 𝜔ℏ + 𝜔ℏ = 𝜔ℏ
2 2
We repeat the action of the raising operator
𝑎̂+ 𝜓1 ~𝜓2 such that ̂ 𝜓2 = 𝐸2 𝜓2
𝐻
where
3 5
𝐸2 = 𝐸1 + 𝜔ℏ = 𝜔ℏ + 𝜔ℏ = 𝜔ℏ
2 2
Normalize
∞
𝑚𝜔 2 𝜋ℏ 𝑚𝜔 1/4
1= |𝐴|2 ∫ 𝑒− ℏ 𝑥 𝑑𝑥 = |𝐴|2 √ => 𝐴 = ( )
𝑚𝜔 𝜋ℏ
−∞
𝑚𝜔 1/4 −𝑚𝜔𝑥 2
𝜓0 = ( ) 𝑒 2ℏ
𝜋ℏ
13
𝑚𝜔 𝑑 1 𝑚𝜔 ℏ 𝑑
𝑎̂+ ≡ √ [(𝑥̂ − 𝑖 (−𝑖ℏ ) )] = √ [(𝑥̂ − )]
2ℏ 𝑑𝑥 𝑚𝜔 2ℏ 𝑚𝜔 𝑑𝑥
𝑚𝜔 ℏ 𝑑 𝑚𝜔 1/4 −𝑚𝜔𝑥 2
𝐴𝑎̂+ 𝜓0 = 𝐴 √ [(𝑥̂ − )] {( ) 𝑒 2ℏ }
⏟ 2ℏ 𝑚𝜔 𝑑𝑥 ⏟ 𝜋ℏ
𝑎̂+ 𝜓0
𝑑 −𝑚𝜔𝑥 2 𝑚𝜔 −𝑚𝜔𝑥 2
𝑒 2ℏ = − 𝑥𝑒 2ℏ
𝑑𝑥 ℏ
𝑚𝜔 1/4 𝑚𝜔 ℏ 𝑚𝜔 −
𝑚𝜔 2
𝑥 𝑚𝜔 1/4 2𝑚𝜔 𝑚𝜔 2
+
𝐴𝑎̂ 𝜓0 = ( ) √ [(𝑥̂ − [− 𝑥])] 𝑒 2ℏ = 𝐴( ) √ 𝑥𝑒 2ℏ 𝑥
−
𝜋ℏ 2ℏ ⏟ 𝑚𝜔 ℏ 𝜋ℏ ℏ
2𝑥
Thus,
𝑚𝜔 1/4 2𝑚𝜔 𝑚𝜔 2
𝜓1 = 𝐴 ( ) √ 𝑥𝑒 2ℏ 𝑥
−
𝜋ℏ ℏ
𝐴=1
14
Consider, for any two functions, 𝑓(𝑥) and 𝑔(𝑥). Then, we have
∞ ∞
𝑑𝑔
𝑢 = 𝑓∗ , 𝑑𝑣 =
𝑑𝑥
∞ ∞
𝛽 ℏ 𝑑 ∗ 𝛽 ℏ 𝑑 ∗
𝐼𝐼 = ⏟∗ 𝑔|∞
(𝑓 −∞ − ∫ 𝑓 𝑔𝑑𝑥) => 𝐼𝐼 = − ∫ 𝑓 𝑔𝑑𝑥
√2 𝑚𝜔 0
𝑑𝑥 √2 𝑚𝜔 𝑑𝑥
−∞ −∞
Then
∗
∞ ∞ ∞
∗ −
𝛽 ℏ 𝑑𝑓 ∗∗
𝛽 ℏ 𝑑
∫ 𝑓 (𝑎̂ 𝑔)𝑑𝑥 = ∫ (𝑥̂𝑓 𝑔 − 𝑔) 𝑑𝑥 = ∫ [ (𝑥̂ − ) 𝑓] 𝑔𝑑𝑥
√2 𝑚𝜔 𝑑𝑥 ⏟
√2 𝑚𝜔 𝑑𝑥
−∞ −∞ −∞
𝑎̂ +
∞ ∞
Similarly,(HW!)
∞ ∞
And,
𝑎̂− 𝜓0 = 0 , 𝑎̂− 𝜓1 ~𝜓0
𝑎̂− 𝜓𝑚 ~𝜓𝑚−1
Let
𝑎̂− 𝜓𝑚 = 𝑑𝑚 𝜓𝑚−1 (𝑏)
𝑐𝑚 , 𝑑𝑚 =?
Consider (a).
∞ ∞ ∞
∫ (𝑎̂+ 𝜓𝑚 )∗ (𝑎̂+ 𝜓𝑚 ) 𝑑𝑥 = ∫ (𝑐𝑚 𝜓𝑚+1 )∗ (𝑐𝑚 𝜓𝑚+1 ) 𝑑𝑥 = |𝑐𝑚 |2 ∫ (𝜓𝑚+1 )∗ 𝜓𝑚+1 𝑑𝑥 = |𝑐𝑚 |2 (∗)
−∞ −∞ −∞
∞ ∞
Then
1
̂
𝐻 1 𝐸 1 𝜔ℏ (𝑚 + ) 1
− +
𝑎̂ 𝑎̂ 𝜓𝑚 = ( + )𝜓 = (
𝑚
+ )𝜓 = ( 2 + ) 𝜓𝑚 = (𝑚 + 1)𝜓𝑚
𝜔ℏ 2 𝑚 𝜔ℏ 2 𝑚 𝜔ℏ 2
∞ ∞
(∗)
(𝑎+
∫ ̂ 𝜓𝑚 )∗ (𝑎+ )∗ ⏞ |𝑐𝑚 |2
̂ 𝜓𝑚 ) 𝑑𝑥 = (𝑚 + 1) ∫ (𝜓𝑚 𝜓𝑚 𝑑𝑥 = (𝑚 + 1) =
−∞ −∞
𝑐𝑚 = √𝑚 + 1
16
Therefore
𝑎̂+ 𝜓𝑚 = √𝑚 + 1𝜓𝑚+1
Similarly,
∞ ∞ ∞
Then
1
̂ 1
𝐻 𝐸 1 𝜔ℏ (𝑚 − ) 1
𝑎̂+ 𝑎̂− 𝜓𝑚 = ( − ) 𝜓𝑚 = (
𝑚
− ) 𝜓𝑚 = ( 2 + ) 𝜓 = 𝑚𝜓
𝜔ℏ 2 𝜔ℏ 2 𝜔ℏ 2 𝑚 𝑚
∞ ∞
(∗)
(𝑎−
∫ ̂ 𝜓𝑚 )∗ (𝑎− )∗ ⏞ |𝑑𝑚 |2 => 𝑑𝑚 = √𝑚
̂ 𝜓𝑚 ) 𝑑𝑥 = 𝑚 ∫ (𝜓𝑚 𝜓𝑚 𝑑𝑥 = 𝑚 =
−∞ −∞
Therefore
𝑎̂− 𝜓𝑚 = √𝑚𝜓𝑚−1
𝑎̂− 𝜓𝑚 = √𝑚𝜓𝑚−1
𝑎̂+ 𝜓𝑚 = √𝑚 + 1𝜓𝑚+1
17
Now:
1
𝜓𝑚+1 = 𝑎̂+ 𝜓𝑚
√𝑚 + 1
𝑚 = 0:
1
𝜓1 = 𝑎̂+ 𝜓0
√1
𝑚 = 1:
1 1
𝜓2 = 𝑎̂+ 𝜓1 = 𝑎̂+ 𝑎̂+ 𝜓0
√2 √2.1
𝑚 = 2:
1 1
𝜓3 = 𝑎̂+ 𝜓2 = 𝑎̂+ 𝑎̂+ 𝑎̂+ 𝜓0
√3 √3.2.1
Clearly
1
𝜓𝑚 = (𝑎̂+ )𝑚 𝜓0
√𝑚!
18
𝑎̂+ 𝑎
̂ − 𝜓𝑛 = √𝑛(𝑎̂+ 𝜓𝑛−1 ) =
⏟ ⏟ √𝑛(√𝑛 − 1 + 1𝜓𝑛−1+1 ) = √𝑛√𝑛𝜓𝑛
(𝑎)=> (𝑏)
√𝑛𝜓𝑛−1
Thus
𝑎̂+ 𝑎̂− 𝜓𝑛 = 𝑛𝜓𝑛 (𝑐)
Then
∞ ∞
∗ + − ∗
𝐼=∫ 𝜓𝑚 (⏟𝑎̂ 𝑎̂ )𝜓𝑛 𝑑𝑥 => 𝐼 = 𝑛 ∫ 𝜓𝑚 𝜓𝑛 𝑑𝑥 (∗)
−∞ 𝑛𝜓𝑛 −∞
∞
∗
𝐼 = 𝑚 ∫ 𝜓𝑚 𝜓𝑛 𝑑𝑥 (∗∗)
−∞
Compare (*) and (**)=>
∞
∗
∫ 𝜓𝑚 𝜓𝑛 𝑑𝑥 = 0 𝑖𝑓 𝑚 ≠ 𝑛
−∞
19
Orthonormality means that we can expand any function 𝑓(𝑥) 𝑎s a linear combination of
stationary states
∞
𝑓(𝑥) = ∑ 𝑐𝑛 𝜓𝑛 (𝑥)
𝑛=1
Ψ(𝑥, 0) = ∑ 𝑐𝑛 𝜓𝑛 (𝑥)
𝑛=1
With
|𝑐𝑛 |2 : 𝑡ℎ𝑒 𝑝𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 𝑡ℎ𝑎𝑡 𝑎 𝑚𝑒𝑎𝑠𝑢𝑟𝑒𝑚𝑒𝑚𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑒𝑛𝑒𝑟𝑔𝑦 𝑤𝑜𝑢𝑙𝑑 𝑦𝑖𝑒𝑙𝑑 𝑡ℎ𝑒 𝑣𝑎𝑙𝑢𝑒 𝐸𝑛
∞ ∞
𝑖𝐸𝑚 𝑡
Ψ(𝑥, 𝑡) = ∑ 𝑐𝑚 Ψ𝑚 (𝑥, 𝑡) = ∑ 𝑐𝑚 𝜓𝑚 (𝑥) 𝑒 − ℏ
𝑚=1 𝑚=1
20
Problem 2.10
a) Construct 𝜓2
From
1
𝜓𝑚+1 = 𝑎̂+ 𝜓𝑚
√𝑚 + 1
𝑚 = 1:
1
𝜓2 = 𝑎̂+ 𝜓1
√2
From Ex.2.4, we have
𝑚𝜔 1/4 2𝑚𝜔 𝑚𝜔 2
𝜓1 = ( ) √ 𝑥𝑒 − 2ℏ 𝑥
𝜋ℏ ℏ
𝑚𝜔 𝑑 1 𝑚𝜔 ℏ 𝑑
𝑎̂+ ≡ √ [(𝑥̂ − 𝑖 (−𝑖ℏ ) )] = √ [(𝑥̂ − )]
2ℏ 𝑑𝑥 𝑚𝜔 2ℏ 𝑚𝜔 𝑑𝑥
1 1 𝑚𝜔 ℏ 𝑑
𝜓2 = 𝑎̂+ 𝜓1 = √[(𝑥̂ − )] 𝜓1
√2 √2 2ℏ 𝑚𝜔 𝑑𝑥
1 𝑚𝜔 𝑚𝜔 1/4 2𝑚𝜔 ℏ 𝑚𝜔 2 𝑚𝜔 2
𝜓2 = √ ( ) √ [(𝑥̂𝑥 − {1 − 𝑥 })] 𝑒 − 2ℏ 𝑥
√2 2ℏ 𝜋ℏ ℏ 𝑚𝜔 ℏ
c)
∞ 𝑜𝑑𝑑
∗ (𝑥) ⏞
∫ 𝜓 ⏟0 𝜓1 𝑑𝑥 =0
−∞ 𝑒𝑣𝑒𝑛
∞ 𝑜𝑑𝑑
∗ (𝑥) ⏞
∫ 𝜓 ⏟2 𝜓1 𝑑𝑥 =0
−∞ 𝑒𝑣𝑒𝑛
𝑚𝜔 1/4 −𝑚𝜔𝑥 2
𝜓0 = ( ) 𝑒 2ℏ
𝜋ℏ
∞
1 𝑚𝜔 1/4 𝑚𝜔 1/4 ∞ −𝑚𝜔𝑥 2 2𝑚𝜔 2 𝑚𝜔 2
∫ 𝜓0∗ (𝑥)𝜓2 𝑑𝑥 = ( ) ( ) ∫ 𝑒 2ℏ [ 𝑥 − 1] 𝑒 − 2ℏ 𝑥 𝑑𝑥
−∞ √2 𝜋ℏ 𝜋ℏ −∞ ℏ
1 𝑚𝜔 1/2 ∞ 2𝑚𝜔 2 𝑚𝜔 2
= ( ) ∫ [ 𝑥 − 1] 𝑒 − ℏ 𝑥 𝑑𝑥 =
√2 𝜋ℏ −∞ ℏ
Problem 2.13
Solution
a)
∞
= |𝐴|2 9 ∫ |𝜓0 |2 𝑑𝑥
⏟
−∞
1
(𝑛𝑜𝑟𝑚𝑎𝑙𝑖𝑧𝑎𝑡𝑖𝑜𝑛)
∞
+|𝐴|2 16 ∫ |𝜓1 |2 𝑑𝑥
⏟
−∞
1
(𝑛𝑜𝑟𝑚𝑎𝑙𝑖𝑧𝑎𝑡𝑖𝑜)
∞
1
= |𝐴|2 (9 + 16) => 𝐴 =
5
23
b)
∞
𝑖𝐸𝑚 𝑡
Ψ(𝑥, 𝑡) = ∑ 𝑐𝑚 𝜓𝑚 (𝑥) 𝑒 − ℏ
𝑚=1
Here;
𝑐0 = 3/5, 𝑐1 = 4/5 , 𝑐𝑚 = 0 𝑤ℎ𝑒𝑛 𝑚 ≠ 0,1
1 𝑖𝐸0 𝑡 𝑖𝐸1 𝑡
Ψ(𝑥, 𝑡) = [3𝜓0 𝑒 − ℏ + 4𝜓1 𝑒 − ℏ ]
5
1 ∗
𝑖(𝐸0 −𝐸1 )𝑡
∗
−𝑖(𝐸0 −𝐸1 )𝑡
|Ψ(𝑥, 𝑡)|2 = |𝜓 | 2 | 2
[9 0 + 16|𝜓1 + 12𝜓0 𝜓1 𝑒 ℏ + 12𝜓1 𝜓0 𝑒 ℏ ]
25
𝜓𝑛 = 𝜓𝑛∗ , 𝑓𝑜𝑟 𝐻𝑂
1 𝑖(𝐸0 −𝐸1 )𝑡 −𝑖(𝐸0 −𝐸1 )𝑡
|Ψ(𝑥, 𝑡)|2 = [9|𝜓0 |2 + 16|𝜓1 |2 + 12𝜓1 𝜓0 (𝑒 ℏ +𝑒 ℏ )]
25
1 (𝐸0 − 𝐸1 )𝑡
|Ψ(𝑥, 𝑡)|2 = [9|𝜓0 |2 + 16|𝜓1 |2 + 24𝜓1 𝜓0 cos ( )]
25 ℏ
Note that
1 1
(𝐸0 − 𝐸1 ) = 𝜔ℏ (0 + ) − 𝜔ℏ (1 + ) = −𝜔ℏ
2 2
Then
1
|Ψ(𝑥, 𝑡)|2 = [9|𝜓0 |2 + 16|𝜓1 |2 + 24𝜓1 𝜓0 cos(𝜔𝑡)]
25
The probability density oscillates with an angular frequency 𝜔.
If 𝜓1 is replaced by 𝜓2 , then we get
1
|Ψ(𝑥, 𝑡)|2 = [9|𝜓0 |2 + 16|𝜓2 |2 + 24𝜓2 𝜓0 cos(2𝜔𝑡)]
25
since
1 1
(𝐸0 − 𝐸2 ) = 𝜔ℏ (0 + ) − 𝜔ℏ (2 + ) = −2𝜔ℏ
2 2
24
c)
∞
∗ (𝑥,
𝑎̂− + 𝑎̂+
〈𝑥〉 = ∫ Ψ 𝑡)𝑥̂ Ψ(𝑥, 𝑡)𝑑𝑥 , 𝑤ℎ𝑒𝑟𝑒 𝑥̂ =
−∞ √2𝛽
∞
1 ∞ ∗
〈𝑥〉 = ∫ ∫ (3𝜓0 𝑒 −𝑖𝐸0 𝑡/ℏ + 4𝜓1 𝑒 −𝑖𝐸1 𝑡/ℏ ) 𝑥̂ (3𝜓0 𝑒 −𝑖𝐸0 𝑡/ℏ + 4𝜓1 𝑒 −𝑖𝐸1 𝑡/ℏ )𝑑𝑥
25 −∞
−∞
∞ ∞ ∞
1 𝑖(𝐸0 −𝐸1 )𝑡 −𝑖(𝐸0 −𝐸1 )𝑡
= 9 ∫ 𝜓0 𝑥𝜓0 𝑑𝑥 + 16 ∫ 𝜓1 𝑥𝜓1 𝑑𝑥 + 12 (𝑒 ℏ +𝑒 ℏ ) ∫ 𝜓0 𝑥𝜓1 𝑑𝑥
25
⏟
−∞ ⏟
−∞ −∞
[ 0 0 ]
∞
24 1 (𝐸1 − 𝐸2 )𝑡
= cos ( ) ∫ 𝜓0 (𝑎̂− + 𝑎̂+ )𝜓1 𝑑𝑥
25 √2𝛽 ℏ
−∞
∞ ∞
24 1
= 𝑎̂− 𝜓1 𝑑𝑥 + ∫ 𝜓0 ⏟
cos(𝜔𝑡) [ ∫ 𝜓0 ⏟ 𝑎̂+ 𝜓1 𝑑𝑥]
25 √2𝛽
−∞ √1𝜓0 −∞ √2𝜓2
∞ ∞
24 1
= cos(𝜔𝑡) ∫ |𝜓0 |2 𝑑𝑥 + √2 ∫ 𝜓0∗ (𝑥)𝜓2 (𝑥)𝑑𝑥
25 √2𝛽
⏟
−∞ ⏟ −∞
1 0
[(𝑛𝑜𝑟𝑚𝑎𝑙𝑖𝑧𝑎𝑡𝑖𝑜𝑛) (𝑜𝑟𝑡𝑜𝑔𝑜𝑛𝑎𝑙𝑖𝑡𝑦) ]
24 ℏ
〈𝑥〉 = √ cos(𝜔𝑡)
25 2𝑚𝜔
𝑑 24 ℏ 24
〈𝑝〉 = 𝑚 〈𝑥〉 = −𝑚𝜔 √ sin(𝜔𝑡) => 〈𝑝〉 = − √𝑚𝜔ℏ/2 sin(𝜔𝑡)
𝑑𝑡 25 2𝑚𝜔 25
d)
32 9
𝑃(𝐸 = 𝐸0 ) = | | =
5 25
2
4 16
𝑃(𝐸 = 𝐸1 ) = | | =
5 25
25
Example2.5: Find the expectation value of the potential energy in the nth state of HO.,
∞
1 1
〈𝑉〉 = 𝑚𝜔2 〈𝑥̂ 2 〉 = 𝑚𝜔2 ∫ 𝜓𝑚
∗ 2
𝑥̂ 𝜓𝑚
2 2
−∞
2
(𝑎̂− + 𝑎̂+ )2 ℏ
𝑥̂ = = (𝑎̂− + 𝑎̂+ )2
2𝛽 2 2𝑚𝜔
2
(𝑎̂− + 𝑎̂+ )2 = 𝑎̂− 2 + 𝑎̂+ + 𝑎̂− 𝑎̂+ + 𝑎̂+ 𝑎̂−
Use
[𝑎̂− , 𝑎̂+ ] = 𝑎̂− 𝑎̂+ − 𝑎̂+ 𝑎̂− = 1
2 2
(𝑎̂− + 𝑎̂+ )2 = 𝑎̂− 2 + 𝑎̂+ + 𝑎̂− 𝑎̂+ + 𝑎
̂ + 𝑎̂− = 𝑎̂− 2 + 𝑎̂+ + 2𝑎̂− 𝑎̂+ − 1
⏟
𝑎̂− 𝑎̂+ −1
∞
1 ℏ 2
〈𝑉〉 = 𝑚𝜔2 ∗
∫ 𝜓𝑚 (𝑎̂− 2 + 𝑎̂+ + 2𝑎̂− 𝑎̂+ − 1)𝜓𝑚
2 2𝑚𝜔
−∞
∞ ∞
∗ −2 ∗
∫ 𝜓𝑚 𝑎̂ 𝜓𝑚 ~ ∫ 𝜓𝑚 𝜓𝑚−2 = 0
−∞ −∞
∞ ∞
∗ + 2 ∗
∫ 𝜓𝑚 𝑎̂ 𝜓𝑚 ~ ∫ 𝜓𝑚 𝜓𝑚+2 = 0
−∞ −∞
∞ ∞ ∞
∗ − + ∗ − ∗
2 ∫ 𝜓𝑚 𝑎̂ 𝑎̂ 𝜓𝑚 = 2√𝑚 + 1 ∫ 𝜓𝑚 𝑎̂ 𝜓𝑚+1 = 2√𝑚 + 1√𝑚 + 1 ∫ 𝜓𝑚 𝜓𝑚 = 2(𝑚 + 1)
−∞ −∞ −∞
∞
∗
− ∫ 𝜓𝑚 𝜓𝑚 = −1
−∞
1 ℏ 1 1 1
〈𝑉〉 = 𝑚𝜔2 (0 + 0 + 2(𝑚 + 1) − 1 = 𝜔ℏ (𝑚 + ) = 〈𝐻〉
2 2𝑚𝜔 2 2 2
26
Problem 2.11
a)
∞ ∞
〈𝑥〉0 = ∫ 𝑥 𝜓0∗ (𝑥)𝜓0 𝑑𝑥 = ∫ (𝑜𝑑𝑑) (𝑒𝑣𝑒𝑛)2 𝑑𝑥 = 0
−∞ −∞
∞ ∞
〈𝑥〉1 = ∫ 𝑥 𝜓1∗ (𝑥)𝜓1 𝑑𝑥 = ∫ (𝑜𝑑𝑑) ⏟
(𝑜𝑑𝑑)2 𝑑𝑥 = 0
−∞ −∞ 𝑒𝑣𝑒𝑛
𝑑
〈𝑝〉0,1 = 𝑚 〈𝑥〉 = 0
𝑑𝑡 0,1
〈𝑥〉𝑚 = 〈𝑝〉𝑚 = 0
n=0
∞
𝑚𝜔 1/2 ∞ 2 −𝑚𝜔𝑥 2 𝑚𝜔 1/2 1 ℏ 𝜋ℏ
〈𝑥 2 〉0 = ∫ 𝑥 2 𝜓0∗ (𝑥)𝜓0 𝑑𝑥 = ( ) ∫ 𝑥 𝑒 ℏ 𝑑𝑥 = ( ) √ =
−∞ 𝜋ℏ −∞ 𝜋ℏ 2 𝑚𝜔 𝑚𝜔
1 ℏ
〈𝑥 2 〉0 =
2 𝑚𝜔
∞
𝑑2
〈𝑝2 〉0 = −ℏ2 ∫ 𝜓0∗ (𝑥) 𝜓 𝑑𝑥
−∞ 𝑑𝑥 2 0
∞
𝑚𝜔 𝑚𝜔 1/2
2 −
𝑚𝜔 2
𝑥 𝑚𝜔 ∞ 2 −𝑚𝜔𝑥 2
=ℏ ( )( ) ∫ 𝑒 ℏ 𝑑𝑥 − ∫ 𝑥 𝑒 ℏ 𝑑𝑥
ℏ 𝜋ℏ ⏟−∞ ℏ ⏟−∞
{ √ 𝜋ℏ 1 ℏ √ 𝜋ℏ
}
𝑚𝜔 2𝑚𝜔 𝑚𝜔
27
𝑚𝜔 𝑚𝜔 1/2 𝜋ℏ 1 𝑚𝜔
〈𝑝2 〉0 = ℏ2 ( )( ) √ {1 − } = ℏ2 ( )
ℏ 𝜋ℏ 𝑚𝜔 2 2ℏ
𝑚𝜔ℏ
〈𝑝2 〉0 =
2
Short-cut Method:
From
𝑝̂ 2 1
̂=
𝐻 + 𝑚𝜔2 𝑥̂ 2
2𝑚 2
We write
n=1
2
∞ ∞
𝑚𝜔 1/4 2𝑚𝜔 𝑚𝜔 2
〈𝑥 2 〉1 =∫ 𝑥 2
𝜓1∗ (𝑥)𝜓1 𝑑𝑥 = [( ) √ ] ∫ (𝑥)4 𝑒 − 2ℏ 𝑥 𝑑𝑥
−∞ 𝜋ℏ ℏ ⏟−∞
3 ℏ 2 √ 𝜋ℏ
( )
4 𝑚𝜔 𝑚𝜔
∞ 1 1
𝑚𝜔 2+1−2−2 3 ℏ 3
〈𝑥 2 〉1 = ∫ 𝑥 2 𝜓1∗ (𝑥)𝜓1 𝑑𝑥 = 2( ) =
−∞ ℏ 4 𝑚𝜔 2
3 ℏ
〈𝑥 2 〉1 =
2 𝑚𝜔
∞
𝑑2
〈𝑝2 〉1 2
= −ℏ ∫ 𝜓1∗ (𝑥) 𝜓 𝑑𝑥
−∞ 𝑑𝑥 2 1
𝑚𝜔 1/4 2𝑚𝜔 𝑚𝜔 2
𝜓1 = ( ) √ 𝑥𝑒 − 2ℏ 𝑥
𝜋ℏ ℏ
28
𝑑2 𝑑 𝑑 𝑚𝜔 1/4 2𝑚𝜔 𝑑 𝑑 𝑚𝜔 2
2
𝜓1 = ( 𝜓1 ) = ( ) √ (𝑥𝑒 2ℏ 𝑥 )
−
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝜋ℏ ℏ 𝑑𝑥 𝑑𝑥
𝑚𝜔 1/4 2𝑚𝜔 𝑚𝜔 2 𝑚𝜔 𝑚𝜔 𝑚𝜔 2
=( ) √ {(1 − 𝑥 ) (− 𝑥) + (−2 𝑥)} 𝑒 − 2ℏ 𝑥
𝜋ℏ ℏ ℏ ℏ ℏ
2
𝑚𝜔 1/4 2𝑚𝜔 𝑚𝜔 ∞ 2 𝑚𝜔 2 −𝑚𝜔𝑥 2
〈𝑝2 〉1 = −ℏ [(2
) √ ] (− ) ∫ 𝑥 (3 − 𝑥 ) 𝑒 2ℏ 𝑑𝑥
𝜋ℏ ℏ ℏ −∞ ℏ
2
∞
𝑚𝜔 1/4 2𝑚𝜔 𝑚𝜔 𝑚𝜔 2
2 − ℏ 𝑥
𝑚𝜔 ∞ 𝑚𝜔 2
2
= −ℏ [( ) √ ] (− ) 3∫ 𝑥 𝑒 𝑑𝑥 − ( ) ∫ (𝑥)4 𝑒 − 2ℏ 𝑥 𝑑𝑥
𝜋ℏ ℏ ℏ ⏟−∞ ℏ ⏟−∞
1 ℏ √ 𝜋ℏ 3 ℏ 2 √ 𝜋ℏ
{ ( ) }
2𝑚𝜔 𝑚𝜔 4 𝑚𝜔 𝑚𝜔
1
𝑚𝜔 1/2 2𝑚𝜔 𝑚𝜔 𝑚𝜔 −1−1/2 3 𝑚𝜔 −2−2+1 3
〈𝑝2 〉1 =ℏ ( 2
) ( )( ) {( ) √𝜋 − ( ) √𝜋}
𝜋ℏ ℏ ℏ ℏ 2 ℏ 4
3𝑚𝜔ℏ
〈𝑝2 〉1 = ( )
2
Short-cut Method:
From
̂ 〉𝑛 − (𝑚𝜔)2 〈𝑥 2 〉𝑛
〈𝑝2 〉𝑛 = 2𝑚〈𝐻
n=1
3 ℏ 1 3 ℏ 3
〈𝑝2 〉1 = 2𝑚𝐸1 − (𝑚𝜔)2 ( ) = 2𝑚 (𝜔ℏ (1 + )) − (𝑚𝜔)2 ( ) = 3𝑚𝜔ℏ − 𝑚𝜔ℏ
2 𝑚𝜔 2 2 𝑚𝜔 2
3
〈𝑝2 〉1 = 𝑚𝜔ℏ
2
29
b) n=0
1 ℏ
𝜎𝑥 = √〈𝑥 2 〉0 − 〈𝑥〉20 = √〈𝑥 2 〉0 − 0 = √
2 𝑚𝜔
𝑚𝜔ℏ
𝜎𝑝 = √〈𝑝2 〉0 − 〈𝑝〉20 = √〈𝑝2 〉0 − 0 = √
2
1 ℏ 𝑚𝜔ℏ ℏ
𝜎𝑥 𝜎𝑝 = √ =
2 𝑚𝜔 2 2
n=1
3 ℏ
𝜎𝑥 = √〈𝑥 2 〉1 − 〈𝑥〉12 = √〈𝑥 2 〉1 − 0 = √
2 𝑚𝜔
3𝑚𝜔ℏ
𝜎𝑝 = √〈𝑝2 〉1 − 〈𝑝〉12 = √〈𝑝2 〉1 − 0 = √
2
3 ℏ 3𝑚𝜔ℏ ℏ
𝜎𝑥 𝜎𝑝 = √ =3
2 𝑚𝜔 2 2
30
∞ (2𝑛)!
2 𝜋
∫ 𝑥 2𝑛 𝑒 −𝑎𝑥 𝑑𝑥 = 2𝑛
√ 2𝑛+1 𝑓𝑜𝑟 𝑎 > 0
−∞ 𝑛! 2 𝑎
1
I) Analytic Method:
We return to Schrödinger eq. for the harmonic oscillator:
𝑑 2 𝜓 2𝑚 1
+ (𝐸 − 𝑚𝜔2 𝑥 2 ) 𝜓 = 0
𝑑𝑥 2 ℏ 2 ⏟
2
𝑉(𝑥)
𝑑 2 𝜓 2𝑚𝐸 𝑚𝜔 2 2 𝑚𝜔 1/2
+ 𝜓 − ( ) 𝑥 𝜓 = 0 ; 𝛽 = ( )
𝑑𝑥 2 ℏ2 ⏟ℏ ℏ
𝛽4
𝑑 2 𝜓 2𝑚𝐸
2
+ 2 𝜓 − 𝛽 4 𝑥 2 𝜓 = 0 (∗)
𝑑𝑥 ℏ
We will solve this equation directly by the series method.
Now, introduce the dimensionless quantity:
𝑦 ≡ 𝛽𝑥 ,
To write eq.(*) in terms of 𝑦, find
𝑑 2 𝜓 𝑑 2 𝜓 𝑑𝑦 2 𝑑 2 𝜓 2 𝑑 2 𝜓 2 2𝑚𝐸 4
𝑦2
= ( ) = 𝛽 => 𝛽 + 2 𝜓−𝛽 2𝜓 = 0
𝑑𝑥 2 𝑑𝑦 2 𝑑𝑥 𝑑𝑦 2 𝑑𝑦 2 ℏ 𝛽
OR,
𝑑 2 𝜓 2𝑚𝐸 1
+ 𝜓 − 𝑦2𝜓 = 0
𝑑𝑦 2 ⏟ℏ2 𝛽 2
2𝐸
ℏ𝜔
2𝐸
Let 𝛼 ≡ . Note that 𝛼 is dimensionless too.
ℏ𝜔
2𝐸 𝐸
𝛼≡ =
ℏ𝜔 1 ℏ𝜔
2
1
So that it is the energy in units of ℏ𝜔. Simplify the above eq.;
2
𝑑2𝜓
2
+ (𝛼 − 𝑦 2 )𝜓 = 0 (∗∗)
𝑑𝑦
To find the solution we first analyze the behavior of 𝜓(𝑦) in the asymptotic region |𝑦| → ∞.
In the limit |𝑦| → ∞ , for any finite value of the energy, the quantity 𝛼 becomes negligible with
respect to 𝑦 2 term In this case, the DE. is approximately given by
2
𝑑2𝜓
2
− 𝑦2𝜓 ≈ 0
𝑑𝑦
with the solution
𝜓(𝑦) ≈ 𝐴 𝑒𝑥𝑝[−𝑦 2 /2] + 𝐵 𝑒𝑥𝑝[𝑦 2 /2]
However,
𝑒𝑥𝑝[𝑦 2 /2] → ∞ 𝑎𝑠 𝑦 → ±∞
Thus the 2.nd term is not normalizable. Therefore, we do not take it and
𝑦2
𝜓(𝑦) ≈ 𝑒𝑥𝑝 [− ]
2
This suggest that we make a guess about the solution as if we separate out the exponential part
𝜓(𝑦) = ℎ(𝑦)𝑒𝑥𝑝[−𝑦 2 /2]
To rewrite (**) in terms of ℎ(𝑦), note that
𝑑𝜓 𝑑H 2 2 𝑑ℎ 2
= [ 𝑒 −𝑦 /2 + ℎ(−𝑦𝑒 −𝑦 /2 )] = [ − ℎ𝑦] 𝑒 −𝑦 /2
𝑑𝑦 𝑑𝑦 𝑑𝑦
𝑑2𝜓 𝑑 2 ℎ 𝑑ℎ −𝑦 2 /2
𝑑ℎ 2
2
= [ 2
− 𝑦 − ℎ] 𝑒 + [ − ℎ𝑦] (−𝑦𝑒 −𝑦 /2 )
𝑑𝑦 𝑑𝑦 𝑑𝑦 𝑑𝑦
𝑑2𝜓 𝑑2ℎ 𝑑ℎ 2 −𝑦 2 /2
= [ − 2 𝑦 − ℎ(1 − 𝑦 )] 𝑒
𝑑𝑦 2 𝑑𝑦 2 𝑑𝑦
Plugging this into the DE (**), we get
𝑑2ℎ 𝑑ℎ 2 2
[ 2 − 2 𝑦 − ℎ(1 − 𝑦 2 )] 𝑒 −𝑦 /2 + (𝛼 − 𝑦 2 ) ℎ 𝑒 −𝑦 /2 = 0
𝑑𝑦 𝑑𝑦
𝑑2ℎ 𝑑ℎ
− 2 𝑦 + (𝛼 − 1) ℎ = 0
𝑑𝑦 2 𝑑𝑦
which is called the Hermite equation.
We look for solutions to this equation for ℎ(𝑦) in the form of a power series in 𝑦:
∞
ℎ(𝑦) = ∑ 𝐶𝑛 𝑦 𝑛 (∗∗∗)
𝑛=0
(Note that according to Taylor’s theorem, any reasonable well-behaved function can be
expressed as a power series, so eq. (***) involves no loss of generality.)
3
∞ ∞
𝑑ℎ 𝑛−1
𝑑2ℎ
= ∑ 𝐶𝑛 𝑛𝑦 , 2
= ∑ 𝐶𝑛 𝑛(𝑛 − 1)𝑦 𝑛−2
𝑑𝑦 𝑑𝑦
𝑛=0 𝑛=0
∞ ∞ ∞
Simplifying
∞ ∞
It follows that coefficient of each power of 𝑦 must vanish. So, we arrive at the recursion
relation:
(2𝑚 + 1 − 𝛼)
𝐶𝑚+2 = 𝐶
(𝑚 + 2)(𝑚 + 1) 𝑚
This recursion formula generates all the even-numbered coefficients, starting with 𝐶0 ;
(1 − 𝛼) (5 − 𝛼) (5 − 𝛼)(1 − 𝛼)
𝐶2 = 𝐶0 ; 𝐶4 = 𝐶2 = 𝐶0 ;
2 12 12 ∙ 2
(9 − 𝛼) (9 − 𝛼)(5 − 𝛼)(1 − 𝛼)
𝐶6 = 𝐶4 = 𝐶0 ; . 𝑒𝑡𝑐
30 30 ∙ 12 ∙ 2
(3 − 𝛼) (7 − 𝛼) (7 − 𝛼)(3 − 𝛼)
𝐶3 = 𝐶1 ; 𝐶5 = 𝐶3 = 𝐶1 ;
6 20 20 ∙ 6
ℎ𝑒𝑣𝑒𝑛 (𝑦) = ∑ 𝐶𝑚 𝑦 𝑚 = 𝐶0 + 𝐶2 𝑦 2 + 𝐶4 𝑦 4 + ⋯
𝑚:𝑒𝑣𝑒𝑛
∞
ℎ𝑜𝑑𝑑 (𝑦) = ∑ 𝐶𝑚 𝑦 𝑚 = 𝐶1 𝑦 + 𝐶3 𝑦 3 + 𝐶5 𝑦 5 + ⋯
𝑚:𝑜𝑑𝑑
These equations together with the recursion relation determine ℎ(𝑦) in terms of two arbitrary
constants 𝐶0 and 𝐶1 , which is just what we would expect for a second-order differential
equation.
In theory we have solved the problem. However, there is a problem here: Not all the solutions
so obtained are normalizable.
In the limit of large m, we find that
𝐶𝑚+2 2
lim →
𝑚→∞ 𝐶𝑚 𝑚
2
It is possible to show that this is the behavior exhibited by 𝑒 𝑦 as 𝑦 gets larger:
𝑦2
𝑦4 𝑦6
2
𝑦𝑚 𝑦 𝑚+2
𝑒 =1+𝑦 + + + ⋯+ 𝑚 + +⋯
2! 3! (2)! ( 𝑚 + 2
!
2 )
Thus
1 1
𝐶𝑚 = 𝑚 , 𝐶𝑚+2 = 𝑚
(2)! ( 2 + 1) !
1
( 𝑚 ) 𝑚 𝑚
𝐶𝑚+2 ( 2 + 1) ! (2)! (2)! 𝑓𝑜𝑟 𝑙𝑎𝑟𝑔𝑒 𝑚 2
lim = = 𝑚 = 𝑚 𝑚 →
𝑚→∞ 𝐶𝑚 𝑚
1 ( 2 + 1) ! ( 2 + 1) ( 2 ) !
( 𝑚 )
(2)!
5
From the requirement that the power series converge, and therefore truncate, we find that the
energy levels for the harmonic oscillator are quantized.
6
Wave Functions:
The recursion relation
(2𝑚 + 1 − 𝛼)
𝐶𝑚+2 = 𝐶
(𝑚 + 2)(𝑚 + 1) 𝑚
For the allowed values of 𝛼 = 2𝑛 + 1, it takes the form
2(𝑚 − 𝑛)
𝐶𝑚+2 = 𝐶
(𝑚 + 2)(𝑚 + 1) 𝑚
This recursion formula generates all the even-numbered coefficients, starting with 𝐶0 ;
−2𝑛
𝑚 = 0; 𝐶2 = 𝐶
2 0
2(1 − 𝑛)
𝑚=1; 𝐶3 = 𝐶1
6
Therefore, when n= even, ℎ(𝑦) = ℎ𝑒𝑣𝑒𝑛 (𝑦), survives and we separately insist that 𝑐1 = 0 to
kill ℎ𝑜𝑑𝑑 (𝑦).
Similarly, when n= odd, ℎ(𝑦) = ℎ𝑜𝑑𝑑 (𝑦), survives and we separately insist that 𝑐0 = 0 to kill
ℎ𝑒𝑣𝑒𝑛 (𝑦).
7
Let us construct some solutions for this. Note that 𝑛 ≡ 𝑚𝑚𝑎𝑥 and
𝑚 = 0; 𝐶2 = 0
𝜓0 = 𝐶0 𝑒𝑥𝑝[−𝑦 2 /2]
With
𝑚𝜔 1/2
𝑦=( ) 𝑥
ℏ
𝜓0 (𝑥) = 𝐶0 𝑒𝑥𝑝[−𝑚𝜔𝑥 2 /2ℏ]
Normalization=>
∞ ∞
2 /ℏ 𝜋ℏ 𝑚𝜔 1/4
1=∫ 𝜓0∗ (𝑥) 𝜓0 (𝑥)𝑑𝑥 = 2
𝐶0 ∫ 𝑒 −𝑚𝜔𝑥 𝑑𝑥 = 𝐶02 √ => 𝐶0 = ( )
−∞ −∞ 𝑚𝜔 𝜋ℏ
2(𝑚 − 1)
𝐶𝑚+2 = 𝐶 => 𝑚 = 1 => 𝐶3 = 0
(𝑚 + 2)(𝑚 + 1) 𝑚
𝜓1 = 𝐶1 𝑦 𝑒𝑥𝑝[−𝑦 2 /2]
8
OR
𝑚𝜔 1/2 2
𝜓1 = 𝐶1 ( ) 𝑥 𝑒 −𝑚𝜔𝑥 /2ℏ
ℏ
Normalization=>
∞
𝑚𝜔 ∞ 2 −𝑚𝜔𝑥 2 𝑚𝜔 1 ℏ 𝜋ℏ
1=∫ 𝜓1∗ (𝑥) 𝜓1 (𝑥)𝑑𝑥 = 𝐶12 ( ) ∫ 𝑥 𝑒 ℏ 𝑑𝑥 = 𝐶12 ( ) ( )√
−∞ ℏ −∞ ℏ 2 𝑚𝜔 𝑚𝜔
𝜋ℏ 1/2 1 𝑚𝜔 1/4
= 𝐶12 ( ) => 𝐶1 = ( ) √2
𝑚𝜔 2 𝜋ℏ
𝑚𝜔 1/4 𝑚𝜔 1/2 2
𝜓1 = ( ) (2 ) 𝑥 𝑒 −𝑚𝜔𝑥 /2ℏ
𝜋ℏ ℏ
4 1/4 𝑚𝜔 3/4 2
𝜓1 = ( ) ( ) 𝑥 𝑒 −𝑚𝜔𝑥 /2ℏ
𝜋 ℏ
n=2. (even)
𝜓2 = 𝐶0 (2 − 4𝑦 2 ) 𝑒𝑥𝑝[−𝑦 2 /2]
n=3 (odd)
3
2(𝑚 − 3)
𝐶𝑚+2 = 𝐶
(𝑚 + 2)(𝑚 + 1) 𝑚
2(1 − 3) −4
𝑚 = 1: 𝐶3 = 𝐶1 => 𝐶3 = 𝐶1 ,
(1 + 2)(1 + 1) 6
2 1
ℎ(𝑦) = ℎ𝑜𝑑𝑑 (𝑦) = 𝐶1 (𝑦 − 𝑦 3 ) = 𝐶1 (12𝑦 − 8𝑦 3 )
3 12
1
𝜓3 = 𝐶1 (12𝑦 − 8𝑦 3 )
12
In general ℎ(𝑦) will be a polynomial of degree n in y, involving even powers only, if n is an even
integer, and odd powers only, if n is an odd integer. Apart from the overall factor (𝐶0 or 𝐶1 ), they
are the so called Hermite polynomials , 𝐻𝑛 (𝑦).
By tradition, the arbitrary multiplicative factor is chosen so that the coefficient of the highest
power of y is 2𝑛 . With this convention, the normalized stationary states for the HO are
𝑚𝜔 1/4 1
𝜓𝑛 = ( ) 𝐻𝑛 (𝑦) 𝑒𝑥𝑝[−𝑦 2 /2]
𝜋ℏ 𝑛
√2 𝑛!
10
At large n, we begin to see some resemblance to the classical case. Classical probability of
finding the particle in the interval (x,x+dx) is proportional to time dt which it takes to pass
through this interval :
dt 2
P(x) dx where T : the period of oscillation.
T
dt dt dx dx dx
Then, P(x) dx
T 2/ dx 2 dx/dt 2 v
1 1 1 k
To find v, E m v 2 k x 2 = k A 2 v 2 (A 2 x 2 ) v 2 (A 2 x 2 )
2 2 2 m
dx 1
Then, P(x) dx P(x)
2 A 2 x 2 2 A 2 x 2
12
http://hyperphysics.phy-
astr.gsu.edu/hbase/quantum/hosc6.html#:~:text=The%20fact%20that%20the%20overall,is%20called%20the%20correspondence%20
principle.&text=The%20quantum%20probabilities%20do%20extend,exponentially%20decaying%20into%20that%20region.
13
It has been superimposed the classical position distribution on the quantum one (for
n=100). If you smoothed out the bumps, the two would fit pretty well.
The fact that the overall picture of probability of finding the oscillator at a given
value of x converges for the quantum and classical pictures is called
the correspondence principle.
14
15
Problem 2.17:
a) The Rodrigues formula says that
2 𝑑 𝑚 −𝑦 2
𝐻𝑚 (𝑦) = (−1)𝑚 𝑒 𝑦 𝑒
𝑑𝑦 𝑚
2 𝑑 0 −𝑦 2
𝐻0 (𝑦) = (−1)0 𝑒 𝑦 𝑒 =1
𝑑𝑦 0
2 𝑑 −𝑦 2 2 2
𝐻1 (𝑦) = (−1)1 𝑒 𝑦 𝑒 = −𝑒 𝑦 (−2𝑦)𝑒 −𝑦 = 2𝑦
𝑑𝑦
2 𝑑2 −𝑦 2 𝑦2
𝑑 2 2 2 2
𝐻2 (𝑦) = (−1)2 𝑒 𝑦 𝑒 = 𝑒 (−2𝑦𝑒 −𝑦 ) = 𝑒 𝑦 (−2𝑒 −𝑦 − 2𝑦(−2𝑦)𝑒 −𝑦 )
𝑑𝑦 2 𝑑𝑦
𝐻2 (𝑦) = −2 + 4𝑦 2
2 𝑑3 −𝑦 2 3 𝑦2
𝑑2 −𝑦 2 3 (−2)𝑒 𝑦 2
𝑑1 2
𝐻3 (𝑦) = (−1)3 𝑒 𝑦 𝑒 = (−1) 𝑒 [(−2𝑦)𝑒 ] = (−1) [(1−2𝑦 2 )𝑒 −𝑦 ]
𝑑𝑦 3 𝑑𝑦 2 𝑑𝑦 1
2 2
= 2𝑒 𝑦 [(−4𝑦) + (1−2𝑦 2 )(−2𝑦)]𝑒 −𝑦
𝐻3 (𝑦) = −12𝑦 + 8𝑦 3
Similarly,
2 𝑑 4 −𝑦 2
𝐻4 (𝑦) = (−1)4 𝑒 𝑦 4
𝑒 = 16𝑦 4 − 48𝑦 2 + 12
𝑑𝑦
b)
OR
𝑑2 𝜓(𝑥 ) 2𝑚𝐸
+ 2 𝜓(𝑥 ) = 0
𝑑𝑥 2 ⏟ℏ
𝑘2
Solution
𝜓(𝑥 ) = 𝐴𝑒 𝑖𝑘𝑥 + 𝐵𝑒 −𝑖𝑘𝑥
Unlike the infinite well, there are no boundary conditions to restrict the possible
values of k and hence E: The fee particle can carry any (+) energy.
Tacking on the standard time dependence
𝑖𝐸𝑡 𝑖𝐸𝑡
− 𝑖𝑘𝑥 −𝑖𝑘𝑥 −
Ψ(𝑥, 𝑡) = 𝜓(𝑥 )𝑒 ℏ = [𝐴𝑒 + 𝐵𝑒 ]𝑒 ℏ
and
Ψ(𝑥, 𝑡) = 𝐴𝑒 𝑖𝑘𝑥 𝑒 −𝑖𝐸𝑡/ℏ + 𝐵𝑒 −𝑖𝑘𝑥 𝑒 −𝑖𝐸𝑡/ℏ
OR with
ℏ2 𝑘 2
𝐸=
2𝑚
2
ℏ𝑘 ℏ𝑘
Ψ(𝑥, 𝑡) = 𝐴 𝑒𝑥𝑝 [𝑖𝑘 (𝑥 − 𝑡)] + 𝐵𝑒𝑥𝑝 [−𝑖𝑘 (𝑥 + 𝑡)]
2𝑚 2𝑚
The structure of this solution is characteristic of a propagating wave.
Any function of 𝑥 and 𝑡 that depends on these variables in the special combination
ℏ𝑘
(𝑥 ∓ 𝑣𝑡) , where 𝑣 = = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡, represents a wave of fixed profile travelling
2𝑚
Since every point on the waveform is moving along with the same velocity, its shape
doesn’t change as it propagates.
𝑥 → 𝑥 + ∆𝑥 ≡ 𝑥 + 𝑣∆𝑡
Since Ψ+ (𝑥, 𝑡) and Ψ− (𝑥, 𝑡) only differ by the sign of k, we might as well write
ℏ𝑘 2
𝑖(𝑘𝑥− 𝑡)
Ψk (𝑥, 𝑡) = 𝐴 𝑒 2𝑚
with
Conclusion:
Stationary states of the free particle are propagating waves.
Their wavelength is
2𝜋
𝜆=
|𝑘|
From De Broglie formula
ℎ
𝜆=
𝑝
Therefore, we have
𝑝 = ℏ𝑘
ℏ𝑘 1 𝑝
𝑣𝑞𝑢𝑎𝑛𝑡𝑢𝑚 = =
2𝑚 2 𝑚
On the other hand, the classical speed of a free particle with mass m is
𝑝
𝑣𝑐𝑙𝑎𝑠𝑠𝑖𝑐𝑎𝑙 = = 2 𝑣𝑞𝑢𝑎𝑛𝑡𝑢𝑚 : Discrepancy‼
𝑚
We will return this paradox soon.
In addition, the plane wave solution is not normalizable.
∞ ∞
Then, in case of the free particle the stationary state solutions do not represent
physically realizable states.
A free particle cannot exist in a stationary state; or
There is no such thing as a free particle with a definite energy.
But this doesn’t mean that the stationary state solutions are of no use to us, for they
play a mathematical role that is entirely independent of their physical interpretation.
The way out of this difficulty is to write the most general solution of the time-
dependent Schrödinger equation for the free particle as a superposition of stationary
state solutions:
1 1 ℏ𝑘 2
𝑖(𝑘𝑥− 2𝑚 𝑡)
Ψ(𝑥, 𝑡) = ∫ 𝑑𝑘 𝜙(𝑘)Ψk (𝑥, 𝑡) = ∫ 𝑑𝑘 𝜙(𝑘)𝑒 (∗)
√2𝜋 √2𝜋
1
The factor is for convenience.
√2𝜋
Compare this with the one for ∞ well:
∞
Ψ(𝑥, 𝑡) = ∑ 𝑐𝑛 Ψ𝑛 (𝑥, 𝑡)
𝑛=0
∑ → ∫ 𝑑𝑘
𝑛
𝑐𝑛 → 𝜙(𝑘)
The most general solution of the free particle necessarily carries a range of k’s and
hence a range of energies and speeds. We call it a wave packet; it is a superposition
of sinusoidal functions whose amplitudes are modulated by 𝜙(𝑘).
1 𝜀 𝜀
𝑤ℎ𝑒𝑛 − <𝑥<
𝛿 (𝜀) (𝑥 ) = { 𝜀 2
𝜀
2 (1)
0 𝑓𝑜𝑟 |𝑥 | >
2
The Dirac delta function 𝛿(𝑥) can be defined as the limit of 𝛿 (𝜀) (𝑥) as 𝜀 → 0:
Dirac Delta function is a very peculiar kind of function. Actually, it is not a function
of the usual mathematical sense (it is not finite at x=0!) , but it is rather “a generalized
function” or “a distribution” , It is an infinitesimally narrow spike at the origin. But,
it is an extremely useful construct in theoretical physics.
6
∞ ∞
∫ 𝛿 (𝑥 − 𝑥0 ) 𝑓(𝑥0 )𝑑𝑥 = ∫ 𝛿 (𝑥 − 𝑥0 ) 𝑓(𝑥 )𝑑𝑥
−∞ −∞
∞
𝐿𝐻𝑆 = 𝑓 (𝑥0 ) ∫ 𝛿 (𝑥 − 𝑥0 ) 𝑑𝑥 => 𝐿𝐻𝑆 = 𝑓(𝑥0 )
⏟−∞
=1
Therefore we get
∞
∫ 𝛿 (𝑥 − 𝑥0 ) 𝑓(𝑥 )𝑑𝑥 = 𝑓(𝑥0 ) (5)
−∞
This is the most important property of Dirac delta function. Of couse the limits of
the integral need not go from −∞ to +∞; all that matters is that the domain of
integration include the point 𝑥0 , so 𝑥0 − ∞ to 𝑥0 +∞ would do for any 𝜖 > 0
Now let’s prove that
∞
1
𝛿 (𝑥 ) = ∫ 𝑑𝑥 𝑒 𝑖𝑘𝑥
2𝜋
−∞
F(k) is called the Fourier transform of f(x); f(x) is the inverse Fourier transform
of F(x).
Let
𝑓 (𝑥 ) = 𝛿 (𝑥 )
7
OR,
∞
1
𝛿 (𝑘) = ∫ 𝑑𝑥 𝑒 −𝑖𝑘𝑥 (𝑏)
2𝜋
−∞
--------------------------------------------------------------------------------------
The most general solution of the time-dependent Schrödinger equation for the free
particle as a superposition of the stationary state solutions:
1 1 ℏ𝑘 2
𝑖(𝑘𝑥− 𝑡)
Ψ(𝑥, 𝑡) = ∫ 𝑑𝑘 𝜙(𝑘)Ψk (𝑥, 𝑡) = ∫ 𝑑𝑘 𝜙(𝑘)𝑒 2𝑚 (∗)
√2𝜋 √2𝜋
Note that
∞ ∞
1 1
∫ 𝑑𝑥 𝑒 −𝑖(𝑘𝑥−𝜔𝑡) 𝑒 𝑖(𝑘′𝑥−𝜔′𝑡) = 𝑒 𝑖(𝜔−𝜔′) ∫ 𝑑𝑥 𝑒 𝑖(𝑘′−𝑘)𝑥
2𝜋 2𝜋
−∞ −∞
8
Since
∞
1
∫ 𝑑𝑥 𝑒 𝑖(𝑘′−𝑘)𝑥 = 𝛿(𝑘 ′ − 𝑘)
2𝜋
−∞
Then
∞
∫ 𝑑𝑥 |Ψ(𝑥, 𝑡)|2
∞
𝑘=𝑘 ′
𝜔=𝜔′
= ∫ 𝑑𝑘 ∫ 𝑑𝑘′ 𝜙(𝑘′)𝜙 ∗ (𝑘)𝛿(𝑘 − 𝑘′)𝑒 𝑖(𝜔−𝜔′) =
⏞ ∫ 𝑑𝑘 𝜙(𝑘)𝜙 ∗ (𝑘)
∞ ∞
In a generic quantum problem, we are given Ψ(𝑥, 0) and we are asked to find
Ψ(𝑥, 𝑡). For the free particle it is given by (*). But what is 𝜙(𝑘) there to match with
the initial packet ?
We write from (*)
1
⏟(𝑥, 0) =
Ψ ∫ 𝑑𝑘 𝜙(𝑘)𝑒 𝑖𝑘𝑥
𝑔𝑖𝑣𝑒𝑛 √2𝜋
To find 𝜙(𝑘)
1
Ψ(𝑥, 0)𝑒 −𝑖𝑘′𝑥 = ∫ 𝑑𝑘 𝜙(𝑘)𝑒 𝑖𝑘𝑥 𝑒 −𝑖𝑘′𝑥
√2𝜋
Then integrate
9
∞ ∞
1
∫ 𝑑𝑥 Ψ(𝑥, 0)𝑒 −𝑖𝑘′𝑥 = ∫ 𝑑𝑘 𝜙(𝑘) ∫ 𝑑𝑥 𝑒 𝑖𝑘𝑥 𝑒 −𝑖𝑘′𝑥
√2𝜋 ⏟
−∞ −∞
2𝜋𝛿(𝑘−𝑘′)
2𝜋
= ∫ 𝑑𝑘 𝜙(𝑘)𝛿(𝑘 − 𝑘′)
√2𝜋
∞
′𝑥
∫ 𝑑𝑥 Ψ(𝑥, 0)𝑒 −𝑖𝑘 = √2𝜋𝜙(𝑘 ′ ) =>
−∞
∞
1
𝜙(𝑘′) = ∫ 𝑑𝑥 Ψ(𝑥, 0)𝑒 −𝑖𝑘′𝑥
√2𝜋
−∞
So the solution to the generic QM problem for the free particle is to replace 𝜙 in (*).
𝜙(𝑘) is called the Fourier transform of Ψ(𝑥, 0); Ψ(𝑥, 0)is the inverse Fourier
transform of 𝜙(𝑘).
10
𝑝
𝑣𝑐𝑙𝑎𝑠𝑠𝑖𝑐𝑎𝑙 = = 2 𝑣𝑞𝑢𝑎𝑛𝑡𝑢𝑚 : Discrepancy‼
𝑚
Example: For the wave function of a free particle, we can show that 𝑣𝑔 = 2𝑣𝑝 . To
show this, we determine the group velocity of a wave packet with the general form
1
Ψ(𝑥, 𝑡) = ∫ 𝑑𝑘 𝜙(𝑘)𝑒 𝑖(𝑘𝑥−𝜔(𝑘)𝑡)
√2𝜋
Let us assume that 𝜙(𝑘) is narrowly peak about some particular point 𝑘0 .
We may then Taylor expand 𝜔(𝑘) about 𝑘0 , since the integrand is negligible
except at the vicinity of 𝑘0 .
𝑑𝜔(𝑘)
𝜔(𝑘) ≅ 𝜔0 + | (𝑘 − 𝑘0 )
𝑑𝑘 𝑘=𝑘0
𝑑𝜔(𝑘)
𝜔0′ ≡ |
𝑑𝑘 𝑘=𝑘0
1 ′
Ψ(𝑥, 𝑡) ≅ ∫ 𝑑𝑘 𝜙(𝑘, 𝑘0 ) 𝑒 𝑖(𝑘𝑥−[𝜔0 +𝜔0 (𝑘− 𝑘0 )]𝑡) 𝑒 −𝑖𝑘0 𝑥 𝑒 𝑖𝑘0 𝑥
√2𝜋
∞
1 𝑑𝜔𝑘
Ψ(𝑥, 𝑡) ≅ 𝑒
⏟𝑖(𝑘0 𝑥−𝜔0 𝑡)
∫ 𝜙(𝑘, 𝑘0 ) 𝑒 𝑖(𝑘−𝑘0 )(𝑥− 𝑑𝑘 𝑡)
𝑑𝑘
√2𝜋 𝑎 𝑤𝑎𝑣𝑒 𝑜𝑓 ⏟
−∞
𝑓𝑟𝑒𝑞𝑢𝑒𝑛𝑐𝑦
𝜔0 𝑤𝑎𝑣𝑒 𝑛𝑢𝑚𝑏𝑒𝑟 ⏟ 𝐴 𝑚𝑜𝑑𝑢𝑙𝑎𝑡𝑖𝑛𝑔 𝑓𝑎𝑐𝑡𝑜𝑟
⏟ 𝑘0 𝑑𝜔
𝐹(𝑥− 𝑑𝑘 𝑡)
𝜔
𝑓(𝑥− 𝑘 0 𝑡)
0
𝜔0 𝑑𝜔
Ψ(𝑥, 𝑡) ≅ 𝑓 (𝑥 − 𝑡) 𝐹 (𝑥 − 𝑡)
𝑘0 𝑑𝑘
Apart from the phase factor, which won’t affect |Ψ|2 , the wave packet moves along
𝑑𝜔𝑘
at a speed
𝑑𝑘
𝑑𝜔(𝑘)
𝑣𝑔𝑟𝑜𝑢𝑝 = 𝜔0′ = |
𝑑𝑘 𝑘=𝑘0
The ordinary phase velocity is
𝜔0 ℏ𝑘 2
𝑣𝑝ℎ𝑎𝑠𝑒 = ; 𝜔=
𝑘0 2𝑚
12
Thus
ℏ𝑘0
𝑣𝑝ℎ𝑎𝑠𝑒 =
2𝑚
And
𝑑𝜔(𝑘) ℏ𝑘0
𝑣𝑔𝑟𝑜𝑢𝑝 = | = = 2𝑣𝑝ℎ𝑎𝑠𝑒
𝑑𝑘 𝑘=𝑘0 𝑚
This confirms that it is the group velocity of the wave packet , not the phase
velocity of the stationary states that matches the classical velocity:,
Example 2.6: A free particle which is initially localized in the range −𝑎 < 𝑥 < 𝑎 is
released at t=0
𝐴 −𝑎 < 𝑥 < 𝑎
𝚿(𝒙, 𝟎) = {
0 𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒
If the spread in position (~𝒂) is small, the spread in momentum must be large.
𝑎 sin(𝑘𝑎)
𝜙(𝑘) = √ ( )
𝜋 (𝑘𝑎)
sin(𝑧)
𝑀𝑎𝑥 [ ]| =1
(𝑧) 𝑧=0
sin(𝑧)
=0 𝑎𝑡 𝑘𝑎 = ±𝜋
(𝑧)
Therefore, for large a, 𝜙(𝑘) is a sharp spike about k=0 (b). This time it’s got a well
defined momentum but an ill-defined position.
15
In classical mechanics:
In Fig. (a), let [𝑎, 𝑏] represents two classical turning
points.
If
𝑉 (𝑥 ) ≤ 𝐸 𝑓𝑜𝑟 𝑥 ∈ [𝑎, 𝑏]
𝑉 (𝑥 ) > 𝐸 𝑓𝑜𝑟 𝑥 ∉ [𝑎, 𝑏]
then the system is in a bound state for 𝑥 ∈ [𝑎, 𝑏]
The particle is stuck in the potential well; it rocks back and forth between two
turning points.
If, on one side or everywhere (b)
𝑉(𝑥 ) ≤ 𝐸
The particle comes in from infinity,
slows down or speeds up due to V(x).
It can’t get trapped in the potential, but
returns to infinity. This is called
a scattering state.
In quantum domain, the only thing that matters is the potential at infinity (because
of the phenomenon of tunneling that allows the particle to leak through any finite
potential barrier)
17
In reality, the most potentials go to zero at infinity and this further simplifies
criterion:
18
Delta-Function Potential
Consider a potential of the form
𝑉 (𝑥 ) = −𝛼 𝛿(𝑥)
where 𝛼 is a positive constant.
Here,
∞ 𝑓𝑜𝑟 𝑥=0
𝛿 (𝑥 ) = {
0 𝑓𝑜𝑟 𝑥≠0
and
∞
∫ 𝛿 (𝑥 ) 𝑑𝑥 = 1
−∞
This is an artificial potential, but it is very simple to work with and illuminates the
basic theory with a minimum mathematics.
Solution:
The Schrödinger equation reads
ℏ2 𝑑2 𝜓
− + 𝑉(𝑥)𝜓 = 𝐸𝜓
2𝑚 𝑑𝑥 2
But the
lim 𝐴𝑒 −𝜅𝑥 → ∞
𝑥→−∞
But the
lim 𝐺𝑒 𝜅𝑥 → ∞
𝑥→∞
The energy can be obtained from the discontinuity condition of the first
derivative of the wave function. It can be obtained by integrating the
Schrödinger eq,
ℏ2 𝑑2 𝜓
− + 𝑉(𝑥)𝜓 = 𝐸𝜓
2𝑚 𝑑𝑥 2
from −𝜖 to +𝜖, and then take the limit 𝜖 → 0: (𝜖 > 0, but arbitrarily small
which also will enable us to evaluate the discontinuity)
+𝜖 +𝜖 +𝝐
2 2
ℏ 𝑑 𝜓
− ∫ 𝑑𝑥 + ∫ 𝑉(𝑥)𝜓(𝑥) 𝑑𝑥 = 𝐸 ∫ 𝝍(𝒙) 𝒅𝒙
2𝑚 𝑑𝑥 2
−𝜖 −𝜖 −𝝐
RHS:
+𝜖
lim ∫ 𝜓(𝑥) 𝑑𝑥 = 0
𝜖→0
−𝜖
Since it is the area of a sliver with vanishing width and finite height.
Thus,
+𝜖
𝑑𝜓 𝑑𝜓 𝑑𝜓 2𝑚
| − | ≡ Δ ( ) = 2 lim ∫ 𝑉(𝑥)𝜓(𝑥) 𝑑𝑥
𝑑𝑥 +𝜖 𝑑𝑥 −𝜖 𝑑𝑥 ℏ 𝜖→0
−𝜖
𝒅𝝍 𝒅𝝍 2𝑚𝛼
| − | = − 2 𝐵 (∗)
𝒅𝒙 +𝝐 𝒅𝒙 −𝝐 ℏ
𝑑𝜓 𝑑𝜓𝐼𝐼 𝑑𝜓
When 𝑥 > 0, = = −𝐵𝐾𝑒 −𝜅𝑥 => | = −𝐵𝐾
𝑑𝑥 𝑑𝑥 𝑑𝑥 +𝜖→0
𝑑𝜓 𝑑𝜓𝐼 𝑑𝜓
When 𝑥 < 0, = = 𝐵𝐾𝑒 𝜅𝑥 => | = +𝐵𝐾
𝑑𝑥 𝑑𝑥 𝑑𝑥 −𝜖→0
(∗)
𝑑𝜓 𝑑𝜓 2𝑚𝛼 𝑚𝛼
| − | = −2𝐾𝐵 =⏞ − 2 𝐵 => 𝐾 = 2
𝑑𝑥 +𝜖 𝑑𝑥 −𝜖 ℏ ℏ
The allowed energy is
ℏ2 𝐾 2 𝑚𝛼 2
𝐸=− => 𝐸 = − 2
2𝑚 2ℏ
There is, therefore, only one bound state solution. As for the excited states, all of
them are unbound.
Normalize 𝝍:
+∞ 0 ∞ ∞
−1
2 √𝑚𝛼
= 2𝐵 ( )
0 − 1 => 𝐵 = √𝐾 =
2𝐾 ℏ
√𝑚𝛼 𝑚𝛼 𝑥
𝜓𝐼 (𝑥 ) = 𝑒 ℏ2 𝑥<0
𝜓 (𝑥 ) = ℏ ;
√𝑚𝛼 −𝑚𝛼 𝑥
𝜓𝐼𝐼 (𝑥 ) = 𝑒 ℏ2 𝑥>0
{ ℏ
𝑚𝛼 2
𝐸 = − 2 ∶ the energy of the single bound state
2ℏ
23
Note that:
𝒊𝑬 𝒊𝑬
𝒊𝒌(𝒙− 𝒕) 𝒊𝒌(𝒙− 𝒕)
𝑨𝒆 𝒌ℏ and 𝑭𝒆 𝒌ℏ represent the wave of amplitude 𝐴 and 𝐹 incident from
Boundary conditions:
1. The continuity of 𝝍(𝒙) at x=0, says
𝒅𝝍 𝒅𝝍 2𝑚𝛼
| − | = − 2 𝐵 (∗)
𝒅𝒙 +𝝐 𝒅𝒙 −𝝐 ℏ
𝑑𝜓 𝑑𝜓𝐼𝐼 𝑑𝜓
When 𝑥 > 0, = = 𝑖𝑘𝐹𝑒 𝑖𝑘𝑥 => | = 𝑖𝑘𝐹
𝑑𝑥 𝑑𝑥 𝑑𝑥 +𝜖→0
𝑑𝜓 𝑑𝜓𝐼 𝑑𝜓
When 𝑥 < 0, = = 𝑖𝑘(𝐴𝑒 𝑖𝑘𝑥 − 𝐵𝑒 −𝑖𝑘𝑥 ) => | = 𝑖𝑘(𝐴 − 𝐵)
𝑑𝑥 𝑑𝑥 𝑑𝑥 −𝜖→0
(∗)
𝑑𝜓 𝑑𝜓 2𝑚𝛼
| − | = 𝑖𝑘 (𝐹 − 𝐴 + 𝐵) =⏞− 2 𝜓
⏟ (0) =>
𝑑𝑥 +𝜖 𝑑𝑥 −𝜖 ℏ
𝐴+𝐵 𝑂𝑅 𝐹
2𝑚𝛼 2𝑚𝛼
𝑖𝑘(𝐹 − 𝐴 + 𝐵) = − (𝐴 + 𝐵 ) => 𝐹 = 𝑖 (𝐴 + 𝐵 ) + 𝐴 − 𝐵
ℏ2 𝑘ℏ2
𝑚𝛼
𝛽≡
𝑘ℏ2
𝑖𝛽 1
𝐵= 𝐴 ;𝐹= 𝐴
1 − 𝑖𝛽 1 − 𝑖𝛽
25
The probability of finding the particle at a specified location is |𝜓|2 . However, this
is not a normalizable wave function, so the absolute probability of finding the
particle at a particular location is not well defined; nevertheless, the ratio of
probabilities for the incident and reflected waves is meaningful:
Introduce
𝑩𝟐 𝑖𝛽 2 𝛽2 1
𝑹≡| | =| | = =
𝑨 1 − 𝑖𝛽 1 + 𝛽 2 1 + 1/𝛽 2
Now,
𝑚𝛼 2𝑚 𝑚𝛼
𝛽≡ , 𝑘 = √ 𝐸 => 𝛽 ≡
𝑘ℏ2 ℏ2 ℏ√2𝑚𝐸
1 1
𝑇= , 𝑅 =
1 + 𝑚𝛼 2 /2𝐸ℏ2 1 + 2𝐸ℏ2 /𝑚𝛼 2
Schrödinger eq:
𝑑2 𝜓 2𝑚
+ (𝐸 − 𝑉 (𝑥 ))𝜓 = 0
𝑑𝑥 2 ℏ2
𝑑2 𝜓𝐼 2𝑚 2
2𝑚
− |𝐸 | 𝜓 𝐼 = 0; 𝜅 ≡ |𝐸 | > 0
𝑑𝑥 2 ⏟2
ℏ ℏ2
𝜅2 >0
lim 𝐺𝑒 𝜅𝑥 → ∞ => 𝐺 = 0
𝑥→∞
𝜓𝐼 (𝑥 ) = 𝐵𝑒 𝜅𝑥 𝑓𝑜𝑟 𝑥 < −𝑎
𝜓(𝑥 ) = {𝜓𝐼𝐼 (𝑥 ) = 𝐶 sin ℓ𝑥 + 𝐷 cos ℓ𝑥 𝑓𝑜𝑟 − 𝑎 < 𝑥 < 𝑎
𝜓𝐼𝐼𝐼 (𝑥 ) = 𝐹𝑒 −𝜅𝑥 𝑓𝑜𝑟 𝑥 > 𝑎
3
Note that:
BC at 𝑥 = −𝑎 gives
𝐵𝐾𝑒 −𝜅𝑎 = 𝐶𝑙 cos(−ℓ𝑎) − 𝑙𝐷 sin(−ℓ𝑎)
BC at 𝑥 = 𝑎 gives
−𝐹𝐾𝑒 −𝜅𝑎 = 𝐶𝑙 cos(ℓ𝑎) − 𝑙𝐷 sin(ℓ𝑎)
Divide them
𝐵 𝐶𝑙 cos(ℓ𝑎) + 𝑙𝐷 sin(ℓ𝑎)
− =
𝐹 𝐶𝑙 cos(ℓ𝑎) − 𝑙𝐷 sin(ℓ𝑎)
For even solutions, where 𝐶 = 0 => 𝐵 = 𝐹
𝜓𝐼 (𝑥 ) = 𝐹𝑒 𝜅𝑥 𝑓𝑜𝑟 𝑥 < −𝑎
𝜓𝑒𝑣𝑒𝑛 (𝑥 ) = { 𝜓𝐼𝐼 (𝑥 ) = 𝐷 cos ℓ𝑥 𝑓𝑜𝑟 − 𝑎 < 𝑥 < 𝑎
𝜓𝐼𝐼𝐼 (𝑥 ) = 𝐹𝑒 −𝜅𝑥 𝑓𝑜𝑟 𝑥 > 𝑎
For odd solutions, where 𝐷 = 0 => 𝐵 = −𝐹
𝜓𝐼 (𝑥 ) = −𝐹𝑒 𝜅𝑥 𝑓𝑜𝑟 𝑥 < −𝑎
𝜓𝑜𝑑𝑑 (𝑥 ) = { 𝜓𝐼𝐼 (𝑥 ) = 𝐶 sin ℓ𝑥 𝑓𝑜𝑟 − 𝑎 < 𝑥 < 𝑎
𝜓𝐼𝐼𝐼 (𝑥 ) = 𝐹𝑒 −𝜅𝑥 𝑓𝑜𝑟 𝑥 > 𝑎
4
It is sufficient to match the solutions 𝝍𝒆𝒗𝒆𝒏 and 𝝍𝒐𝒅𝒅 only at 𝒙 = 𝒂 since the
same conditions will obtain when matching is done at 𝒙 = −𝒂. (Due to the
symmetry under reflection)
𝑑𝜓𝐼𝐼𝐼 (𝑥 = 𝑎) 𝑑𝜓𝐼𝐼 (𝑥 = 𝑎)
= => −𝜅𝐹𝑒 −𝜅𝑎 = −ℓ𝐷 sin ℓ𝑎 (𝑏)
𝑑𝑥 𝑑𝑥
Now define
2𝑚𝑎2
𝑧0 ≡ √ 2 |𝑉0 |
ℏ
Since
2
2𝑚𝑎2
(𝜅𝑎) = |𝐸 |
ℏ2
5
Then
2𝑚𝑎2 2𝑚𝑎2
𝒛 ≡ ℓ𝑎 = √ 2 |𝑉0 | − |𝐸 | => 𝑧 = √𝑧02 − (𝜅𝑎)2
⏟ℏ ⏟ℏ2
𝑧02 (𝜅𝑎)2
𝜿𝒂 = √𝑧02 − 𝑧 2
√𝑧02 − 𝑧 2 𝑧0 2
𝑡𝑎𝑛 𝑧 = √
= ( ) −1
𝑧 𝑧
Let
𝑧0 2 𝑧0 2
𝑦(𝑧) ≡ √( ) − 1 . Note that 𝑦(𝑧0 ) = √( ) − 1 = 0
𝑧 𝑧0
𝑦(𝑧) = 𝑡𝑎𝑛 𝑧
This is a transcendental eq. for z as a function of 𝑧0 . We can plot both sides of
this equation on the same graph for various values of to get an idea of
what happens.
6
z0 8
10
6
y z tanz
4
0
0 2 4 6 8 10
z0 50
40 z0 50
40
30 30
y z
20
20
10
10
tanz
0 3 5 7 9 11
0 2 2
2 2
3 2
4 2
5 2
6
0
0 10 20 30 40 50
7
In these plots, we show what happens for three different values of 𝑧0 . The blue
𝑧 2
curves show the plot of 𝑡𝑎𝑛 𝑧 ; the red curves that of 𝑦(𝑧) ≡ √( 0 ) − 1.
𝑧
1. In the first graph, with 𝑧0 = 2, we get only one intersection between the two
plots, around 𝑧 = 1 . Thus for 𝑧0 = 2, there is only one bound state, with an
energy that can be worked out from
1 2
𝑎 1 ℏ⏞
𝑧
𝒛 ≡ ℓ𝑎 = √2𝑚(−|𝐸| + |𝑉0 |) ≈ 1 => |𝐸 | ≈ |𝑉0 | − ( )
ℏ 2𝑚 𝑎
2. In the second graph, with 𝑧0 = 8, we get three intersections between the two
plots, around 𝑧1 = 1.40, 𝑧2 = 4.15, 𝑧3 = 6.80 . Thus for 𝑧0 = 8, there are three
bound states, with an energy that can be worked out from
𝑎
𝑧 ≡ ℓ𝑎 = √2𝑚(−|𝐸| + |𝑉0 |) ≈ 𝑧𝑛 𝑓𝑜𝑟 𝑛 = 1,2,3
ℏ
1 ℏ𝑧𝑛 2
|𝐸𝑛 | ≈ |𝑉0 | − ( ) ; 𝑧1 = 1.40, 𝑧2 = 4.15, 𝑧3 = 6.80
2𝑚 𝑎
3. In the 3.rd graph, with 𝑧0 = 50, we get 16 intersections between the two plots,
8
We can see that no matter how small we make 𝑧0 , we will always have
at least one bound state (since the 𝑡𝑎𝑛 𝑧 graph starts off from the origin).
Note that
When 0 ≤ 𝑧0 < 𝜋 => there is 1 bound state
𝑧0 2
lim 𝑦(𝑧) ≡ (√( ) − 1) → ∞
𝑧0 →∞, 𝑧
so that means that the entire curve gets higher, so the intersections
with the tangent curve will occur at higher locations. The tangent is
𝜋
asymptotic to the vertical lines 𝑛 for odd n, so we would expect the
2
intersection points to eventually become
𝜋
𝑧𝑛 = 𝑛 , 𝑤ℎ𝑒𝑟𝑒 𝑛 = 𝑜𝑑𝑑
2
This means that
𝟐
2𝑚 𝜋 2
𝒛 ≡ (ℓ𝑎 )2 2
= 𝑎 ( 2 (−|𝐸 | + |𝑉0 |)) ≈ (𝑛 )
ℏ 2
𝑛2 𝜋 2 ℏ2
|⏟
𝑉0 | − |𝐸𝑛 | =
2𝑚 (2𝑎)2
>0
Since |𝑉0 | − |𝐸𝑛 | is the height of the bound state above the bottom of the well,
we can see that this formula does indeed give us the expected energy levels
for an infinite square well of width 2𝑎, or at least those corresponding to odd
n.
10
𝑑𝜓𝐼𝐼𝐼 (𝑥 = 𝑎) 𝑑𝜓𝐼𝐼 (𝑥 = 𝑎)
= => −𝜅𝐹𝑒 −𝜅𝑎 = ℓ𝐶 cos ℓ𝑎 (𝑏)
𝑑𝑥 𝑑𝑥
𝑧0 2
√
−𝑐𝑜𝑡 𝑧 = ( ) − 1
𝑧
In the plots below, we show what happens for three different values of 𝑧0 . The
orange curves show the plot of −𝑐𝑜𝑡 𝑧 ; the green curves that of 𝑦(𝑧) ≡
2
√(𝑧0 ) − 1. (Together with 𝑡𝑎𝑛 𝑧, blue curves)
𝑧
11
z0=2 z0=8
10 10
8 8
6 6
4 4
2 2
0 z 0
3 3 5
0
2 2 0 2 3
2 2 2
z0=50
40
30
20
10
0
3 5 7 9 11
0 2 3 4 5 6
2 2 2 2 2 2
12
Note that
𝜋
When 𝑧0 < No odd bound state; only one even bound state
2
𝜋 𝑎 𝜋 1 ℏ𝜋 2
𝑧0 < => √2𝑚|𝑉0 | < => 𝐼𝑓 |𝑉0 | < ( ) 𝑡ℎ𝑒𝑛 𝑛𝑜 𝑜𝑑𝑑 𝑏𝑜𝑢𝑛𝑑 𝑠𝑡𝑎𝑡𝑒
2 ℏ 2 2𝑚 2𝑎
𝜋
In the 1.st graph: ≤ (𝑧0 = 2 ≅ 0.64𝜋) < 𝜋 => 2 𝑏𝑠
2
When 𝜋 ≤ 𝑧0 < 3𝜋/2 => there are 3 bound states (2 even , one odd)
Generalize this:
𝝅 𝝅
When 𝒏 ≤ 𝒛𝟎 < (𝒏 + 𝟏) => there are (n+1) bound states
𝟐 𝟐
5𝜋
In the 2.nd graph: ≤ (𝑧0 = 8 ≅ 2.55𝜋) < 3𝜋 => there are 6 bound states
2
3 5 7 9 11 13 15 17 19 21 23 25 27 29 31
2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2
14
1 ℏ𝑧𝑛 2 10
|𝐸𝑛 | ≈ |𝑉0 | − ( )
2𝑚 𝑎
8
𝜓𝐼 (𝑥) = 𝐹𝑒 𝜅𝑥 𝑓𝑜𝑟 𝑥 < −𝑎
𝜓𝑒𝑣𝑒𝑛 (𝑥) = {𝜓𝐼𝐼 (𝑥) = 𝐷 cos ℓ𝑥 𝑓𝑜𝑟 − 𝑎 < 𝑥 < 𝑎
6
𝜓𝐼𝐼𝐼 (𝑥) = 𝐹𝑒 −𝜅𝑥 𝑓𝑜𝑟 𝑥 > 𝑎
0
3 5
0 2 3
2 2 2
Inside the well, in the classically allowed region, we have E>V(x) and the
differential equation admits only sinusoidal solutions (sin ℓ𝑥 or cos ℓ𝑥)
characterized by the wave vector ℓ . The oscillatory part of the wave function
(inside the well) has a characteristic wavelength
2𝜋 2𝜋
𝜆ℓ = =
ℓ 2𝑚
√ (|𝑉0 | − |𝐸 |)
ℏ2
So the larger the energy difference between the eigenvalue and the potential energy,
the smaller the wavelength, which is seen in fig. above)
16
Outside the well, in the classically forbidden region, we have E<V(x) the
differential equation admits only real exponential solutions:
𝜓𝐼 (𝑥 ) = 𝐹𝑒 𝜅𝑥 𝑎𝑛𝑑 𝜓𝐼𝐼𝐼 (𝑥 ) = 𝐹𝑒 −𝜅𝑥
with a decay length of 1/𝜅, which is zero for for the infinite square well.
In the finite potential well, since the wave function is not zero in the classically
forbidden region, the quantum mechanical particle can be found where the classical
particle cannot:
|𝜓𝐼 (𝑥 )|2 = |𝐹 |2 𝑒 2𝜅𝑥 𝑎𝑛𝑑 |𝜓𝐼𝐼𝐼 (𝑥 )|2 = |𝐹 |2 𝑒 −2𝜅𝑥
This penetration of the wave function into the potential energy barrier leads to the
phenomenon of tunneling, which we explore in the problem 2.32.
The wave function plots in the above figure indicate that the barrier penetration is
more pronounced for higher energy levels and can become large for energies close
to the top of the well. To see this let’s calculate the decay length from uncertainty
relation:
1 1
Δ𝑥Δ𝑝~ℏ => Δ𝑥Δ(ℏ𝜅 )~ℏ => Δ𝑥~ =
𝜅 2𝑚
√ |𝐸 |
ℏ2
Since 𝜅 in the forbidden region decreases as the energy increases, which means that
the decay length Δ𝑥 becomes larger, so more of the wave function is outside the
well.
In comparing the finite and infinite well energies, we also note that a given finite
well energy eigenvalue 𝐸𝑛 lies below the corresponding infinite well energy
eigenvalue. This is in consistent with the longer wavelength of the finite well
eigenstate compared to the corresponding infinite well state.
17
Problem 2.40
∞ 𝑥<0
𝑉 (𝑥 ) = {−32ℏ2 /𝑚𝑎2 0<𝑥<𝑎
0 𝑥>𝑎
Let 𝐸 = −|𝐸 |
Schrödinger eq:
𝑑2 𝜓 2𝑚
+ (𝐸 − 𝑉 (𝑥 ))𝜓 = 0
𝑑𝑥 2 ℏ2
lim 𝐺𝑒 𝜅𝑥 → ∞ => 𝐺 = 0
𝑥→∞
0 𝑥<0
𝜓(𝑥 ) = {𝐶𝑠𝑖𝑛ℓ𝑥 + 𝐷 cos ℓ𝑥 0<𝑥<𝑎
𝐹𝑒 −𝜅𝑥 𝑥>𝑎
𝑑𝜓𝐼𝐼 (𝑥 = 𝑎) 𝑑𝜓𝐼𝐼𝐼 (𝑥 = 𝑎)
= => ℓ𝐶 cos ℓ𝑎 = −𝜅𝐹𝑒 −𝜅𝑎 (𝑏)
𝑑𝑥 𝑑𝑥
OR
−𝜅𝑎 = ℓ𝑎 cot ℓ𝑎
Introduce the notation:
𝑎 𝑎
𝒛 ≡ ℓ𝑎 = √2𝑚(|𝑉0 | − |𝐸 |) , 𝑧0 ≡ √2𝑚|𝑉0 |
ℏ ℏ
𝑎
𝜅𝑎 = √2𝑚|𝐸| = √𝑧02 − 𝑧 2
ℏ
So
√𝑧02 − 𝑧 2 𝑧0 2
−𝜅𝑎 = ℓ𝑎 cot ℓ𝑎 => − 𝑐𝑜𝑡 𝑧 = √
= ( ) −1
𝑧 𝑧
Let
𝑧0 2
𝑦(𝑧) ≡ √( ) − 1
𝑧
𝑎 32ℏ2 𝑎 32ℏ2
𝑧0 ≡ √2𝑚|𝑉0 | , |𝑉0 | ≡ => 𝑧0 ≡ √ 2𝑚 => 𝑧0 = 8
ℏ 𝑚𝑎2 ℏ 𝑚𝑎2
20
10
𝜋
5 ≅ 7.85
8 2
The number of bound states is 3
0 3 5 7
z
0 2 3
2 2 2 2
1 ℏ𝑧𝑛 2 32ℏ2 ℏ2 2
|𝐸𝑛 | ≈ |𝑉0 | − ( ) = − 𝑧
2𝑚 𝑎 𝑚𝑎2 2𝑚𝑎2 𝑛
Let
ℏ2
Δ≡
2𝑚𝑎2
Then
|𝐸𝑛 | = Δ(64 − 𝑧𝑛2 )
𝑛 𝑧𝑛 |𝐸𝑛 |/ Δ
1 2.785 56.2438
2 5.522 33.5075
3 7.9573 0.6813
21
b)
+∞ +∞ +∞
2
𝑎 1 2𝜅𝑎 2
𝑒 −2𝜅𝑎
= 𝐶 [( − 𝑠𝑖𝑛ℓ𝑎 𝑐𝑜𝑠ℓ𝑎) + 𝑒 𝑆𝑖𝑛 ℓ𝑎 ]
2 2ℓ 2𝜅
22
2
𝑎 1 𝑆𝑖𝑛2 ℓ𝑎
= 𝐶 [( − 𝑠𝑖𝑛ℓ𝑎 𝑐𝑜𝑠ℓ𝑎) + ]
2 2ℓ 2𝜅
𝐶2 𝜅
= [(𝑎 − 𝑠𝑖𝑛ℓ𝑎 𝑐𝑜𝑠ℓ𝑎) + 𝑆𝑖𝑛2 ℓ𝑎]
2𝜅 ⏟
ℓ
−𝐶𝑜𝑡ℓ𝑎
𝐶2 2 2
𝐶2 2𝜅
= [(𝑎𝜅 + 𝑐𝑜𝑠 ℓ𝑎) + 𝑆𝑖𝑛 ℓ𝑎] = (1 + 𝑎𝜅 ) = 1 => 𝐶 = √
2𝜅 2𝜅 (1 + 𝑎𝜅 )
2𝜅
𝐹 = 𝐶𝑒 𝜅𝑎 𝑠𝑖𝑛ℓ𝑎 = √ 𝑒 𝜅𝑎 𝑠𝑖𝑛ℓ𝑎
(1 + 𝑎𝜅 )
1 −2𝜅𝑎 2𝜅 1
𝑃(𝑥 > 𝑎) = |𝐹 |2 𝑒 = 𝑒 2𝜅𝑎 𝑠𝑖𝑛2 ℓ𝑎 𝑒 −2𝜅𝑎
2𝜅 (1 + 𝑎𝜅 ) 2𝜅
𝑠𝑖𝑛2 ℓ𝑎
𝑃(𝑥 > 𝑎) =
(1 + 𝑎𝜅 )
𝐹2 𝐵2
𝑇≡| | , 𝑅≡| |
𝐴 𝐴
----------------------------------------------------------------------
25
𝑑𝜓𝐼𝐼 (𝑥 = 𝑎) 𝑑𝜓𝐼𝐼𝐼 (𝑥 = 𝑎)
= =>
𝑑𝑥 𝑑𝑥
ℓ [𝐶 cos(ℓ𝑎) − 𝐷 Sin(ℓ𝑎) ] = 𝑖𝑘𝐹𝑒 𝑖𝑘𝑎 (4)
We use (3) and (4) to eliminate C and D :
Multiply (3) with sin(ℓ𝑎) and (4) with cos(ℓ𝑎):
𝐶 sin2 (ℓ𝑎) +𝐷sin(ℓ𝑎) cos(ℓ𝑎) = 𝐹𝑒 𝑖𝑘𝑎 sin(ℓ𝑎)
ℓ [𝐶 cos2 (ℓ𝑎) −𝐷 Sin(ℓ𝑎) cos(ℓ𝑎) ] = 𝑖𝑘𝐹𝑒 𝑖𝑘𝑎 cos(ℓ𝑎)
Add them:
𝑘
𝐶 = 𝐹𝑒 𝑖𝑘𝑎 (sin(ℓ𝑎) + 𝑖 cos(ℓ𝑎)) (5)
ℓ
26
𝑘
𝐴𝑒 −𝑖𝑘𝑎 + 𝐵𝑒 𝑖𝑘𝑎 = 𝐹𝑒 𝑖𝑘𝑎 {cos(2ℓ𝑎) − 𝑖 sin(2ℓ𝑎)} (7)
ℓ
Put (5) and (6) into (2):
𝑖𝑘[𝐴𝑒 −𝑖𝑘𝑎 − 𝐵𝑒 𝑖𝑘𝑎 ]
𝑘
= ℓ [𝐹𝑒 𝑖𝑘𝑎 (sin(ℓ𝑎) + 𝑖 cos(ℓ𝑎)) cos(ℓ𝑎)
ℓ
𝑘
+ 𝐹𝑒 𝑖𝑘𝑎 (cos(ℓ𝑎) − 𝑖 sin(ℓ𝑎)) Sin(ℓ𝑎) ]
ℓ
𝑘
= ℓ𝐹𝑒 𝑖𝑘𝑎 [2
⏟ cos(ℓ𝑎) sin(ℓ𝑎) + 𝑖 ⏟ 2 (ℓ𝑎) + cos2 (ℓ𝑎))]
(−sin
ℓ cos(2ℓ𝑎)
sin(2ℓ𝑎)
ℓ
[𝐴𝑒 −𝑖𝑘𝑎 − 𝐵𝑒 𝑖𝑘𝑎 ] = 𝐹𝑒 𝑖𝑘𝑎 [cos(2ℓ𝑎) − 𝑖 sin(2ℓ𝑎)] (8)
𝑘
27
(7)+(8)=>
ℓ 𝑘
2𝐴𝑒 −𝑖𝑘𝑎 = 𝐹𝑒 𝑖𝑘𝑎 [2 cos(2ℓ𝑎) − 𝑖 ( + ) sin(2ℓ𝑎)]
𝑘 ℓ
𝑖2𝑘𝑎
ℓ2 + 𝑘 2
𝐴 = 𝐹𝑒 [cos(2ℓ𝑎) − 𝑖 ( ) sin(2ℓ𝑎)] (9)
2𝑘ℓ
Put this into (7)
ℓ 𝑘
2𝐵𝑒 𝑖𝑘𝑎 = 𝐹𝑒 𝑖𝑘𝑎 [𝑖 ( − ) sin(2ℓ𝑎)]
𝑘 ℓ
ℓ2 − 𝑘 2
𝐵 = 𝐹 [𝑖 ( ) sin(2ℓ𝑎)]
2𝑘ℓ
Thus
sin(2ℓ𝑎) 2
𝐵=𝑖 (ℓ − 𝑘 2 )𝐹
(2𝑘ℓ)
And from (9)=>
𝑒 −2𝑖𝑘𝑎
𝐹= 𝐴
(ℓ2 + 𝑘 2 )
cos(2ℓ𝑎) − 𝑖 sin(2ℓ𝑎)
(2𝑘ℓ)
𝐹 𝑒 −2𝑖𝑘1 𝑎
=
𝐴 (ℓ2 + 𝑘 2 )
(cos(2ℓ𝑎) − 𝑖 sin(2ℓ𝑎))
(2𝑘ℓ)
28
Thus,
|𝐹 |2 1
𝑇= =
|𝐴|2 2
1 ℓ2 + 𝑘 2
⏟ (2ℓ𝑎) + 4 ( 𝑘ℓ ) sin2 (2ℓ𝑎))
(cos 2
1−sin2 (2ℓ𝑎)
1
=
2 1 𝑘 4 + ℓ4 + 2𝑘 2 ℓ2
(1 + sin (2ℓ𝑎) [−1 + ( )])
4 𝑘 2 ℓ2
1
𝑇= 2 (∗∗)
2 1 ℓ2 − 𝑘 2
(1 + sin (2ℓ𝑎) ( ) )
4 𝑘ℓ
2
2 2 2
2𝑚 2𝑚 2 2𝑚|𝑉0 |
(ℓ − 𝑘 ) = ( 2 (𝐸 + |𝑉0 |) − 2 𝐸) = ( ) (𝑎)
ℏ ℏ ℏ2
2𝑚 2𝑚
(𝑘ℓ)2 = ( 2 𝐸 2 (𝐸 + |𝑉0 |)) (𝑏)
ℏ ℏ
(𝑎) 𝑉02 𝑉02 1
= = =
(𝑏) 𝐸(𝐸 + |𝑉0 |) |𝑉0 |𝐸(1 + 𝐸/|𝑉0 |) 𝐸/|𝑉0 |(1 + 𝐸/|𝑉0 |)
Let
𝝐 ≡ 𝑬/|𝑽𝟎 |
2𝑚 2𝑚 2𝑚|𝑉0 |
2ℓ𝑎 = 2𝑎√ (𝐸 + |𝑉0 |) = 2𝑎 √ (𝐸 + |𝑉0 |) = 2𝑎 √ (1 + 𝜖 )
ℏ2 ℏ2 ℏ2
Therefore
−1
1 2𝑚|𝑉0 |
𝑇 = {1 + sin2 (2𝑎√ (1 + 𝜖 ))} (∗)
4𝜖(1 + 𝜖) ℏ2
29
Total Transmission
−1
1 2𝑚|𝑉0 |
𝑇 = {1 + sin2 (2𝑎√ (1 + 𝜖 ))}
4𝜖(1 + 𝜖) ℏ2
Note that we have total transmission at very high energies and/or at weak potential
well (potential well is completely transparent to the incoming particle):
𝑬 ≫ 𝑽𝟎 => 𝝐 ≫ 𝟏 ⟹ 𝑻 → 𝟏 𝒂𝒏𝒅 𝑹 → 𝟎
This means that particles with certain energies will be completely transmitted.
However, remember that the wave function associated with a particular, well
defined energy is not normalizable. This means that a wave packet whose energy is
peaked near the proper value will be mostly transmitted.
30
2𝑚 2𝑚
sin2 (2𝑎√ (𝐸 + |𝑉0 |)) = 0 => 2𝑎 √ (𝐸 + |𝑉0 |) = 𝑛𝜋 , 𝑛 = 0,1,2, …
ℏ2 ⏟ ℏ2
ℓ
The energies for which the finite potential well becomes transparent is given by
ℏ2 𝜋 2
𝐸𝑛 = 2
𝑛2 − |𝑉0 |, 𝑛 = 0,1,2, … : requirement for the perfect transition
2𝑚(2𝑎)
The right-hand side is the energy of the n-th bound state of the infinite square well
of width 2a.
We get full transmission for those energies 𝐸𝑛 > 0 that are in the spectrum of the
infinite-square-well extension of our finite square well. Since infinite square well
bound states are characterized by fitting an integer number of half wavelengths, we
have a resonance type situation in which perfect transmission is happening when
the scattering waves fit perfectly inside the finite square well. The phenomenon we
have observed is called resonant transmission which do not occur in classical
physic.
In figure below, transmission coefficient in Eq. (*), is plotted as a function of the
energy showing the positions 𝐸𝑅 of the resonances.
31
This means that the transmission is boosted whenever the distance 4a back and
forth that wave traverses inside the well due to reflection is equal to an integer
number of its de Broglie wave length.
The scattering cross section for the scattering of electrons by a rare gas of krypton
atoms exhibits a low-energy minimum at 𝐸 ≅ 0.9 𝑒𝑉. Assuming that the diameter
of the atomic well seen by the electrons is 1 Bohr radius (= 5.29 × 10−11 𝑚),
From
ℏ2 𝜋 2 2
ℏ2 𝜋 2
𝐸𝑛 + |𝑉0 | = 2
𝑛 => |𝑉0 | = 2
𝑛2 − 𝐸𝑛
2𝑚(2𝑎) 2𝑚(2𝑎)
(ℏ𝑐)2 𝜋 2 2
(197 × 10−9 𝑒𝑉. 𝑚)2 𝜋 2
|𝑉0 | = 2 2
𝑛 − 𝐸𝑛 = 6 −11 2
12 − 0.9 𝑒𝑉
2(𝑚𝑐 )(2 𝑎) 2(0.5 × 10 𝑒𝑉)(5.29 × 10 𝑚)
|𝑉0 | = 136 𝑒𝑉
33
+1
∫ 𝛿 (𝑥 − (−2) ⏟ 3 − 3𝑥 2 + 2𝑥 − 1) 𝑑𝑥 = 𝑓(𝑥0 ) = 𝑓(−2)
⏟ ) (𝑥
−3 𝑥0 𝑓(𝑥)
b) Since
0<𝜋<∞
+∞
∫ 𝛿 (𝑥 − 𝜋 (𝐶𝑜𝑠3𝑥 + 2) 𝑑𝑥 = 𝑓(𝑥0 ) = 𝑓(𝜋) = 𝐶𝑜𝑠3𝜋 + 2 = −1 + 2
⏟) ⏟
0 𝑥0 𝑓(𝑥)
=1
c) the domain of integration (-1,+1) does not include the point 𝑥0 = 2
+1
∫ 𝛿 (𝑥 − ⏟ (𝑒 |𝑥|+3 ) 𝑑𝑥 = 0
2)⏟
−1 𝑥0 𝑓(𝑥)
1
E>0
𝑉(𝑥) = 0 𝑥 < −𝑎 (I)
𝑉(𝑥) = 𝑉0 −𝑎 ≤ 𝑥 ≤ 𝑎 (II)
𝑉(𝑥) = 0 𝑥 > 𝑎 (III)
Solution to this problem can be directly written from the solution to the problem of
scattering from the finite potential well.
2
3
Since the Schrödinger equation for the finite barrier shown above is
identical in form to that for a finite well, the solutions must be identical in
form too.
The only difference in the two solutions is that in region II, 𝓵 → 𝒒;
otherwise the solutions must be identical.
Now to find the transition coefficient for the rectangular barrier scattering
(RBS) we make the replacement :
2 −1
2
𝑞2 − 𝑘 2
In eq (1) replace ℓ → 𝑞 => 𝑇𝑅𝐵𝑆 = {1 + sin (2𝑞𝑎) ( ) } (2)
2𝑘𝑞
4
Now,
2𝑚
𝑞2 = (𝐸 − 𝑉0 )
ℏ2
−1
2 2 2
𝑞 − 𝑘
𝑇𝑅𝐵𝑆 = {1 + sin2 (2𝑞𝑎) ( ) } (2)
2𝑘𝑞
2 2 )2
2𝑚 2𝑚𝐸 2 2𝑚𝑉0 2
(𝑞 − 𝑘 = ( 2 (𝐸 − 𝑉0 ) − 2 ) = ( 2 ) (𝑎)
ℏ ℏ ℏ
2𝑚𝐸 2𝑚
(𝑘𝑞 )2 = ( 2 2 (𝐸 − 𝑉0 )) (𝑏)
ℏ ℏ
(𝑎) 𝑉02 𝑉0
= =
(𝑏) 𝐸(𝐸 − 𝑉0 ) 𝐸(𝐸/𝑉0 − 1)
Therefore
−1
2
1 𝑉0
𝑇𝑅𝐵𝑆 = {1 + sin (2𝑞𝑎) }
4 𝐸(𝐸/𝑉0 − 1)
1
𝑇𝑅𝐵𝑆 = ,𝑅 = 1 − 𝑇
1 1 2𝑚
[1 + sin2 (2𝑎√ 2 (𝐸 − 𝑉0 ))]
4 (𝐸/𝑉0 )(𝐸/𝑉0 − 1) ℏ
2𝑚
2𝑎√ (𝐸 − 𝑉0 ) = 𝑛𝜋 =>
ℏ2
ℏ2 𝜋 2
𝐸 = 𝑉0 + 2
𝑛2 , 𝑛 = 0,1,2, … : requirement for the perfect transition
2𝑚(2𝑎)
6
E<V0
𝑑2 𝜓 2𝑚
Schrödinger eq: + (𝐸 − 𝑉 (𝑥 ))𝜓 = 0
𝑑𝑥 2 ℏ2
Rewrite it
𝑑2 𝜓𝐼𝐼 2𝑚 2𝑚
− 2 (𝑉0 − 𝐸 ) 𝜓𝐼𝐼 = 0 𝑤𝑖𝑡ℎ 𝜅 2 ≡ (𝑉 − 𝐸 ) > 0
𝑑𝑥 2 ⏟
ℏ ℏ2 0
𝜅2 >0
𝑞 → −𝑖𝜅
So that it becomes
𝐶𝑠𝑖𝑛𝑞𝑥 + 𝐷 cos 𝑞𝑥 → 𝑀𝑒 𝑖(−𝑖𝜅)𝑥 + 𝑁𝑒 −𝑖(−𝑖𝜅)𝑥 = 𝑀𝑒 𝜅𝑥 + 𝑁𝑒 −𝜅𝑥
8
−1
2 2 2
𝑞 −𝑘
𝑇𝑅𝐵𝑆 = {1 + sin2 (2𝑞𝑎) ( ) } (2)
2𝑘𝑞
9
Now, we can use this result to find the transition coefficient of the
rectangular barrier penetration (RBP) from the transition coefficient of
the rectangular barrier scattering (RBS) in eq. (2) by making the
replacement
𝑞 → −𝑖𝜅
−1
2 2 2
(−𝑖𝜅) − 𝑘
𝑇𝑅𝐵𝑆 = {1 + sin2 (2(−𝑖𝜅)𝑎) ( ) } → 𝑇𝑅𝐵𝑃
2𝑘(−𝑖𝜅)
Using
sin(𝑖𝑧) = 𝑖 sinh(𝑧)
We get the transition coefficient of the rectangular barrier penetration
−1
2 2 2
𝜅 +𝑘
𝑇𝑅𝐵𝑃 = {1 + sinh2 (2𝜅𝑎) ( ) } (3)
2𝑘𝜅
10
Special cases:
1. 𝐸 ≪ 𝑉0 which implies that
2𝑚𝑎2
𝜅𝑎 ≡ √ 2 (𝑉0 − 𝐸 ) ≫ 1,
ℏ
Then
1 2𝜅𝑎 1
sinh(2𝜅𝑎) = (𝑒 − 𝑒 −2𝜅𝑎 ) ≅ 𝑒 2𝜅𝑎
2 2
So that
−1
2 2
1 2
𝜅 +𝑘 2
4𝑘𝜅 2 −4𝜅𝑎
𝑇𝑅𝐵𝑃 ≅ {( 𝑒 2𝜅𝑎 ) ( ) } => 𝑇𝑅𝐵𝑃 ≅( 2 ) 𝑒
2 2𝑘𝜅 𝜅 + 𝑘2
𝑇𝑅𝐵𝑃 ≠ 0
Classically it was zero. We see that even though the energy is very much
below the top of the barrier there is transmission. This is a wave
phenomenon and in quantum mechanics it is also exhibited by particles
2. When 𝐸 ≈ 𝑉0 => 𝜅𝑎 → 0. Then,
sinh(2𝜅𝑎) ≅ 2𝜅𝑎
Then,
2 −1
2
𝜅2 + 𝑘2
𝑇𝑅𝐵𝑃 ≅ {1 + sinh (2𝜅𝑎) ( ) }
2𝑘𝜅
2 2 )2 −1 2 2 )2 −1
(𝜅 + 𝑘 (𝜅 + 𝑘
𝑇𝑅𝐵𝑃 ≅ [1 + (2𝜅 )2 𝑎2 ] ≅ [1 + 𝑎2 ]
(2𝜅 )2 𝑘 2 𝑘2
11
2 2 )2
2𝑚 2𝑚 2 2𝑚 2
(𝜅 + 𝑘 = ( 2 (𝑉0 − 𝐸 ) + 2 𝐸) = ( 2 𝑉0 ) (𝑎)
ℏ ℏ ℏ
2𝑚
𝑘2 = 2 𝐸 (𝑏)
ℏ
𝐸≈𝑉
(𝑎) 2𝑚 𝑉02 0
2𝑚
=> 2 ⏞
≅ 𝑉
(𝑏) ℏ 𝐸 ℏ2 0
Thus,
𝐸≈𝑉0 −1
2𝑚𝑎2 𝑉0
𝑇𝑅𝐵𝑃 ⏞ [1 +
≅ ]
ℏ2
Let us analyze an apparent difficulty here: The wave function does not vanish
inside the barrier and thus there appears to be probability of finding a particle
with negative kinetic energy . How can this make sense?
We look to the uncertainty relation to remove an apparent paradox that arises
from “too classical” a description of the process.
An experiment to study the particle inside the potential barrier must be able to
localize it within an accuracy
Δ𝑥 ≪ 2𝑎
This measurement will transfer to the particle momentum, with an uncertainty
ℏ
Δ𝑝 ≫
2𝑎
which corresponds to a transfer of energy
Δ𝑝2 1 ℏ 2
Δ𝐸 = ≫ ( ) (𝑎)
2𝑚 2𝑚 2𝑎
In order to observe the negative kinetic energy 𝐾𝐸 = 𝐸 − 𝑉0 , the uncertainty
Δ𝐸 must be much less than |𝐾𝐸 |
ℏ2 𝜅 2
Δ𝐸 ≪ |𝐾𝐸 | = |𝐸 − 𝑉0 | = (𝑏)
2𝑚
From (a) and (b) we get
ℏ2 𝜅 2 1 ℏ 2
≫ Δ𝐸 ≫ ( ) => 2𝜅𝑎 ≫ 1
2𝑚 2𝑚 2𝑎
However, under this circumstances, the quantity to be measured 𝑇𝑅𝐵𝑃
4𝑘𝜅 2 −4𝜅𝑎
𝑇𝑅𝐵𝑃 ≅( 2 ) 𝑒
𝜅 + 𝑘2
is vanishingly small. For example, for 𝜅𝑎 = 10, 𝑇𝑅𝐵𝑃 ≈ 𝑒 −4𝜅𝑎 ≈ 10−18 ,
which means that 𝑇𝑅𝐵𝑃 is immeasurable and we have a total reflection.
Therefore there is no particle in region II to measure the KE.
14
𝜶 − 𝑫𝑬𝑪𝑨𝒀
The Rutherford’s experiment that led to the discovery of the atomic nucleus:
α-particles were used to hit thin foils (atoms in a crystal), most of them went straight
through the foil as atoms are pretty empty, a few of them got scattered right back (by
electrostatic repulsion) when they were heading for the core, which could be
explained by an atom with a very small, dense, positively-charged nucleus at its
center
EX:
𝟐𝟑𝟎 𝟐𝟐𝟔 𝟒
𝟗𝟎𝑻𝒉 → 𝟖𝟖𝑹𝒂 + 𝟐𝑯𝒆
⏟
𝜶−𝒑𝒂𝒓𝒕𝒊𝒄𝒍𝒆
15
The nuclear force is very strong and extremely short range,≈ 10−15 𝑚. α-particle is
confined in the parent nucleus by the nuclear force (when 0<r<R), and it is
insignificant outside the nucleus (when r>R). Beyond the nuclear force range, r>R,
the α-particle feels only the Coulomb potential U(r) which increases closer to the
nucleus.
We can imagine that the alpha particle “moves” about the small space in the
nucleus with some energy (𝐸𝛼 ~9 𝑀𝑒𝑉 ) that is less than the energy required to
overcome the Coulomb barrier (~30 𝑀𝑒𝑉.) Therefore the alpha particle is trapped
in a finite potential well.
Classically, such an α-particle with 𝐸𝛼 ~9 𝑀𝑒𝑉 initially bound to the nucleus can
not escape. However, the explanation of emission of α-particles is possible in
terms of the tunnel effect. According to quantum mechanics, the α-particle, with
its wave attributes, may tunnel through the barrier to appear on the outside. We
shall not go through the detailed arguments used to estimate T in this case, but we
shall note that they involve the approximation
2𝑚(𝑉(𝑥)−𝐸)
−2 ∫𝑏𝑎𝑟𝑟𝑖𝑒𝑟 𝑑𝑥 √
𝑇≅ 𝑒 ℏ2
𝑇 ≅ 𝐴 𝑒 −𝐵(𝑍−2)/√𝐸𝛼
The nuclear force is very strong and extremely short range,≈ 10−15 𝑚. α-particle is
confined in the parent nucleus by the nuclear force (when 0<r<R), and it is
insignificant outside the nucleus (when r>R). Beyond the nuclear force range, r>R,
the α-particle feels only the Coulomb potential U(r) which increases closer to the
nucleus.
We can imagine that the alpha particle “moves” about the small space in the
nucleus with some energy (𝐸𝛼 ~9 𝑀𝑒𝑉 ) that is less than the energy required to
overcome the Coulomb barrier (~30 𝑀𝑒𝑉.) Therefore the alpha particle is trapped
in a finite potential well.
Classically, such an α-particle with 𝐸𝛼 ~9 𝑀𝑒𝑉 initially bound to the nucleus can
not escape. However, the explanation of emission of α-particles is possible in
terms of the tunnel effect. According to quantum mechanics, the α-particle, with
its wave attributes, may tunnel through the barrier to appear on the outside. We
shall not go through the detailed arguments used to estimate T in this case, but we
shall note that they involve the approximation
2𝑚(𝑉(𝑥)−𝐸)
−2 ∫𝑏𝑎𝑟𝑟𝑖𝑒𝑟 𝑑𝑥 √
𝑇≅ 𝑒 ℏ2
𝑇 ≅ 𝐴 𝑒 −𝐵(𝑍−2)/√𝐸𝛼
CH0
HISTORICAL REVIEW: EXPERIMENTS AND THEORIES
The nature and behavior of matter and energy at the atomic and subatomic level is referred to as
Quantum mechanics was initially invented (at the beginning of the 20th century) because the classical
theories provided no means to explain the properties of atoms, electrons, and electromagnetic radiation..
The physics of the 19th century and before is called the classical physics. It consisted
essentially of
At the turn of the 20th century, with the development of new experimental techniques to probe atomic and
subatomic structures, it appeared that classical physics fails miserably in providing the proper explanation
for several newly discovered phenomena.
Some of the phenomena that could not be explained by classical physics at the beginning of the 20th
century are
It became evident that the validity of classical physics breaks off at the microscopic level and new
concepts radically different from the concepts of classical physics had to be invoked to describe the
structure of atoms and molecules and how light interacts with them.
The problem is just about how heated bodies radiate. All objects emit EM radiation to their surroundings
and also absorb radiation from them continuously, whatever their temperatures are.
This is the Thermal radiation which is generated by the thermal motion of charged particles in matter.
It represents a conversion of thermal energy into electromagnetic energy.
If a gas of particular element is heated to very high temperature it glows, i.e., sends electromagnetic (EM) radiation.
If this EM radiation is examined in a spectroscope, it is seen that only a discrete set of wavelengths –a line spectrum
– is emitted.
On the other hand, solids (such as the tungsten filament of an electric light bulb) when heated up to similarly high
T, emit a continuous range of wavelengths.
An object in thermal equilibrium with its surroundings radiates as much energy as it absorbs. It thus
follows that a blackbody is a perfect absorber as well as a perfect emitter of radiation.
The spectrum of the blackbody radiation is the ideal spectrum which is a characteristic of the
temperature T but completely independent of the nature and structure of the body producing it.
To understand the radiation inside the cavity, one needs simply to analyze the spectral distribution of the
radiation coming out of the hole.
The first quantitative guess based on experimental observation of hole radiation was:
Stefan’s Law (1879):
Josef Stefan found experimentally that total emitted power
P=E/t, from a unit area of a hot body at temperature T at all
frequencies is given by
𝑃 = 𝜎𝑇 4
where
2𝜋 5 𝑘 4
𝜎= = 5.67 × 10−8 𝑊𝑚−2 𝐾 −4
15ℎ3 𝑐 2
∶ 𝑆𝑡𝑒𝑓𝑎𝑛 − 𝐵𝑜𝑙𝑡𝑧𝑚𝑎𝑛𝑛 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡
This experimental result was confirmed theoretically by Boltzmann five years later using a combination of
the laws of thermodynamics and Maxwell’s eqs
Until the end of the 19.th century a great volume of experimental data on this subject was cumulated. They
found a radiation intensity/frequency curve close to this:
𝑢(𝜈, 𝑇)𝑑𝜈 : the radiant energy emitted per time in frequency range (𝜈, 𝜈 + 𝑑𝜈) from the unit area of the
surface at temperature T.
From these curves, you can tell what color or wavelength of light is being emitted by the blackbody. This
is done by looking for the peak wavelength using Wien's displacement law:
U(T) : the total radiant energy per unit volume in the cavity at temperature T
∞
𝑈(𝑇) = ∫ 𝑢(𝜈, 𝑇)𝑑𝜈
0
It only depends on temperature T. The actual value can be determined from the Stefan-Boltzmann law
The relation between the emitted power, 𝑷, which Stefan-Boltmann law gives, and
the total stored energy in the cavity, U(T), is given by (see Liboff page 33-34 or
explanation below)
𝑐
𝑃 = 𝑈(𝑇)
4
5
EX: The sun has 𝑇~5000 𝐾. Power radiated from unit surface of sun is
Why does the blackbody spectrum have the shape shown in the figure?
J. Rayleigh (1900) and J. Jeans (1905) derived the spectral distribution of thermal radiation on the basis of the
classical electromagnetic theory.
Since each standing wave originates from an oscillating electric charge, they
(each standing wave ) are considered to be equivalent to a one-dimensional
oscillator (SHO). When the cavity is in thermal equilibrium, the electromagnetic
energy density inside the cavity is equal to the energy density of the charged
particles in the walls of the cavity and can be calculated by finding
Assuming that the body is in the form of a cube with edge length L , by solving 3-dimensional wave equation , they
calculated
4𝜋𝜈 2
𝑁(𝜈) = 2 (2)
𝑐3
6
To calculate average energy per standing wave (𝑎 𝑆𝐻𝑂), , classical kinetic theory is called on. According to its
law of equipartition of energy, for a system in thermal equilibrium, each degree of freedom has average energy
1
𝑘𝑇, where 𝑘 = 1.38 × 10−23 J/K : Boltzman constant. Thus a one-dimensional simple harmonic oscillator (a
2
standing wave) has total energy
1 1
𝑘𝑇 = 𝑘𝑇 + 𝑘𝑇
2⏟ 2⏟
𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑃𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙
𝑒𝑛𝑒𝑟𝑔𝑦 𝑒𝑛𝑒𝑟𝑔𝑦
The origin of the equipartition law arises, basically , from a more complete result of classical statistical mechanics
called the Boltzmann distribution:
𝑒 −𝜀/𝑘𝑇
𝑃(𝜀) =
𝑘𝑇
such that
𝑷(𝜺)𝒅𝜺: The probability of finding a given entity (standing wave ) with energy in the interval , d .
The average energy per standing wave, , can be obtained from P ( ) by using
∞
∫0 𝜀 𝑃(𝜀)𝑑𝜀
𝜀̅ = ∞ (3)
∫0 𝑃(𝜀)𝑑𝜀
The denominator of Eq (3) is equal to unity, since it is the probability of finding the standing wave with any
energy.
Using the integral
n ax n!
x e dx
0 a n1
One can easily show that
𝜀̅ = 𝑘𝑇 (4)
Now, substituting the parts, we find
8𝜋𝜈 2
𝑢𝑅𝐽 (𝜈) = ( 3 ) 𝑘𝑇 ∶ 𝑅𝑎𝑦𝑙𝑒𝑖𝑔ℎ − 𝐽𝑒𝑎𝑛𝑠 𝐹𝑜𝑟𝑚𝑢𝑙𝑎
𝑐
Rayleigh-Jeans formula contains everything that classical physics can say about the spectrum of blackbody
radiation. However, it can not possibly be correct !
the oscillators in the Rayleigh-Jeans classical approach not to radiate energy continuously, as the classical
theory would demand, but they could only lose or gain energy in integer multiples of 𝒉𝝂, for an oscillator of
frequency 𝝂.
ℎ = 6.626 × 1034 𝑗. 𝑠
Thus, according to Planck’s Postulate, the energy of each oscillator is given by
𝜀𝑛 = 𝑛ℎ𝜈 , 𝑛 = 0,1,2, …
So, assuming that the energy of an oscillator is quantized, the correct thermodynamic relation for the average energy
be obtained by replacing the integration of (3)—that corresponds to an energy continuum—by a discrete summation
corresponding to the discreteness of the oscillators’ energies
∞ 𝑛ℎ𝜈
∫0 𝜀 𝑒 −𝜀/𝑘𝑇 𝑑𝜀 ∑∞ −
0 𝑛ℎ𝜈𝑒 𝑘𝑇
𝜀̅ = ∞ ⟼ 𝑛ℎ𝜈
∫0 𝑒 −𝜀/𝑘𝑇 𝑑𝜀 ∑∞ − 𝑘𝑇
0 𝑒
⇊ . ⇊
ℎ𝜈
(𝑘𝑇) ⟼ ( ℎ𝜈 )
𝑒 −1
𝑘𝑇
which gives
h
e h / kT 1
Then the energy density takes the form
8𝜋ℎ 𝜈3
𝑢(𝜈) = 3 ℎ𝜈 (6)
𝑐
𝑒 𝑘𝑇 − 1
This is known as Planck’s distribution. It gives an exact fit to the various experimental radiation distributions, as
displayed in Figure . The numerical value of h is obtained by fitting (6) with the experimental data .
At high frequencies ,
At low frequencies, ℎ𝜈 ≪ 𝑘𝑇, where RJ formula is a good approximation to the data, use
x2
ex 1 x ... and when x 1 , e x 1 x..
2!
h 1 1 kT
Since 1 then h / kT .
kT e 1 h h
1 1
kT
8 h 3 kT 8 kT 2
u( )d ( ) d d
c3 h c3
which is RJ formula!
We can find the total energy density in terms of Stefan-Boltzmann’s total power per unit surface area
∞
8𝜋ℎ ∞ 𝜈3
𝑈(𝑇) = ∫ 𝑢(𝜈, 𝑇)𝑑𝜈 = ∫ ℎ𝜈 𝑑𝜈
0 𝑐3 0
𝑒 𝑘𝑇 −1
Use a change of variable
ℎ𝜈
𝑥=
𝑘𝑇
=𝜋4 /15
8𝜋(𝑘𝑇)4 ⏞∞ 𝑥 3 8𝜋 5 𝑘 4 4 4 4
𝑈(𝑇) = ∫ 𝑑𝑥 = 𝑇 = 𝜎𝑇
(ℎ𝑐)3 0 𝑒 𝑥 − 1 15(ℎ𝑐)3 𝑐
where
2𝜋 5 𝑘 4
𝜎= = 5.67 × 10−8 𝑊𝑚−2 𝐾 −4 ∶ 𝑆𝑡𝑒𝑓𝑎𝑛 − 𝐵𝑜𝑙𝑡𝑧𝑚𝑎𝑛𝑛 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡
15(ℎ𝑐)3
Therefore,
4
𝑈(𝑇) = 𝑃
𝑐
In this way, Planck’s relation (6) leads to a finite total energy density of the radiation emitted from a blackbody, and
hence avoids the ultraviolet catastrophe.
9
1 −ℎ𝜈
lim ( ℎ𝜈 ) → 𝑒 𝑘𝑇
ℎ𝜈
≫1 𝑒 𝑘𝑇 − 1
𝑘𝑇
Then,
8𝜋ℎ𝜈 3 −ℎ𝜈
𝑢(𝜈)𝑑𝜈 ≈ 𝑒 𝑘𝑇 𝑑𝜈
𝑐3
Problem 2.1 To see that this relation is in the form of Wien’s, write
𝑐 1
𝑑𝜈 = |𝑑 ( )| = 𝑐 2 𝑑𝜆
𝜆 𝜆
Thus,
8𝜋ℎ𝑐 3 −ℎ𝑐 1 8𝜋ℎ𝑐 −ℎ𝑐
𝑢(𝜈)𝑑𝜈 → 3 3 𝑒 𝑘𝑇𝜆 (𝑐 2 𝑑𝜆) => 𝑢(𝜆)𝑑𝜆 = 5 𝑒 𝑘𝑇𝜆 𝑑𝜆
𝑐 𝜆 𝜆 𝜆
Or with
−ℎ𝑐
𝑊(𝜆𝑇) ≡ 8𝜋ℎ𝑐 𝑒 𝑘𝑇𝜆
We get
𝑊(𝜆𝑇)
𝑢(𝜆)𝑑𝜆 𝑑𝜆 ∶ 𝑊𝑖𝑒𝑛′ 𝑠𝑟𝑒𝑙𝑎𝑡𝑖𝑜𝑛
𝜆5
10
Experimental results
I) Curves show how the current i varies as V and light intensity I for a fixed value of frequency.
1
K max mv 2 eVStop. : Energy of the fastest electrons
2
11
The existence of a cut-off frequency 𝜈0 below which no photoelectric effect is observed can not be
explained by the classical EM theory of light, since according to this theory photoelectric effect
should occur at any 𝜈 of the incident light provided that light is intensive enough.
In 1905, A. Einstein resolved the paradox through the formulation of the photon theory and he received
the Nobel Prize in 1921 for this work.
He generalized the Planck’s notation that it was the absorption and emission of radiation that occurred in
quanta. Instead, Einstein proposed that radiation itself consisted of quanta of energy.
EM radiation consists of zero rest mass particles called photons that travel at the speed of light c, with
energy 𝜀 = ℎ𝜈.
Energy content of EM radiation of frequency 𝜈 in a radiant source can only be an integral multiple of
the quantity ℎ𝜈: 𝐸𝑛 = 𝑛ℎ𝜈 .
Keep the frequency of light constant, but Increase the intensity of light:
=> increase the number of photons => number of photoelectrons will increase too
=> photoelectric current 𝒊 increases
Keep the intensity of light constant, but increase the frequency of light:
=> the energy of each photon 𝜀 = ℎ𝜈 will increase
=> the photoelectrons will have more energy
Through the photon concept, the interaction between surface electrons and incident EM
radiation is a particle-particle interaction.
If ℎ𝜈 > Φ then electron absorb the photon and partially use the photon energy to overcome the work
function. Then rest of the energy appears as the kinetic energy for the electron.
If ℎ𝜈 < Φ , implies that 𝐾𝑚𝑎𝑥 < 0 , electrons will never be ejected from the surface: Photoelectric
effect is not observed below a certain cut-off frequency 𝜈0 .
The Photoelectric effect Experiment provides an important confirmation of the particle nature of radiation
13
ATOMIC SPECTRA
When an atomic gas is suitably “exited” it gives off light. The emitted light is then passed through a prism
which separates it into its constituent wavelengths. The result is a spectrum which contains certain specific
wavelengths only. A spectrum of this sort is called a line spectrum.
Spectral Series
In 1885, J.J. Balmer discovered that the wavelengths in the visible part of the hydrogen spectrum showed
regularity.
These regular sets are now called spectral series. The figure below shows the Balmer series.
The line with the longest 𝜆 = 656.2 𝑛𝑚(= 656.2 × 10−9 𝑚) is specified H𝛼 , the next with 𝜆 =
486.1 𝑛𝑚 (= 486.1 × 10−9 𝑚) is specified H𝛽 , and so on.
Balmer’s formula for the wavelengths of this (which were obtained by trial and error) is
1 1 1
= 𝑅 ( 2 − 2) 𝑛 = 3,4,5, …
𝜆 2 𝑛
where
𝑅 = 1.097 × 107 𝑚−1 ∶ Rydberg constant
The H𝛼 line corresponds to n=3, the H𝛽 line to n=4, and so on. The series limit corresponds to n=∞, so that
it occurs at a 𝜆 = 4/𝑅, in agreement with experiment.
Balmer series contains 𝜆′s in the visible portion of the H-spectrum. The spectral lines of H in the ultraviolet
and infrared regions fall into several other series.
In the ultraviolet the Lyman series contains the 𝜆′𝑠 given by the formula
1 1 1
= 𝑅 ( 2 − 2 ) 𝑛 = 2,3,4, … Lyman
𝜆 1 𝑛
In the infrared,
1 1 1
= 𝑅 ( 2 − 2 ) 𝑛 = 4,5,6 … Paschen
𝜆 3 𝑛
1 1 1
= 𝑅 ( 2 − 2 ) 𝑛 = 5,6,7, … Brackett
𝜆 4 𝑛
14
The Bohr Model: Niels Bohr (1913) developed a model of atomic structure to explain all these features of atomic
spectra of hydrogen atom by combining
Rutherford’s planetary model + Planck’s quantum hypothesis + Einstein’s photon concept
+ some postulates
As long as an electron remains in a stationary orbit (despite the fact that it is constantly
accelerating,) it does not radiate electromagnetic energy, thus, its total energy e remains
constant. Emission or absorption of electromagnetic radiation can take place only when an
electron, initially moving in an orbit of total energy Ei, jumps to another orbit with total energy
Ef. .The frequency of the emitted or absorbed radiation 𝜈 is equal to
𝐸𝑓 – 𝐸𝑖
𝜈=
ℎ
15
This is a quantized expression for the radius. Here 𝑟𝑛 represents the radius of the nth orbit. Integer n
is called the quantum number of the orbit.
The radius of the innermost orbit is customarily called the Bohr radius of the H-atom and is denoted
by the symbol 𝑎0
ℎ 2 𝜖0
𝑎0 = 𝑟1 ≡ = 5.292 × 10−11 𝑚
𝜋𝑚𝑒 2
Then, the other radii can be written as
𝑟𝑛 = 𝑛2 𝑎0
As for the total energy of the electron in H-atom;
1 2
𝑒2
𝐸 = 𝐾 + 𝑉 = 𝑚𝑒 𝑣 −
2 4𝜋𝜖0 𝑟𝑛
In deriving this relation, we have assumed that the nucleus, i.e., the proton, is infinitely heavy
compared with the electron and hence it can be considered at rest; that is, the energy of the electron–
proton system consists of the kinetic energy of the electron plus the electrostatic potential energy.
Substituting 𝑣 gives
1 𝑒2
𝐸𝑛 = −
2 4𝜋𝜖0 𝑟𝑛
Substituting 𝑟𝑛 , we find
𝑚𝑒 4 1 𝐸1
𝐸𝑛 = − 2 2
( 2) = 2 , 𝑛 = 1,2,3, …
8𝜖0 ℎ 𝑛 𝑛
Where
16
𝑚𝑒 4
𝐸1 = −
8𝜖0 2 ℎ2
Note that
Since the radius is quantized so does energy.
E<0. This holds for every atomic electron and reflects the fact that it is bound to nucleus.
The energy 𝐸𝑛 of each state of the atom is determined by the value of the quantum number n.
The energies specified by this equation are called the energy levels of the H-atom and are plotted in
figure below, (where, by convention, the energy levels are shown as horizontal lines).
The energies specified by this equation are called the energy levels of the H-atom and are plotted in figure
below, (where, by convention, the energy levels are shown as horizontal lines).
which is indeed the same as the Rydberg constant R. Therefore we can rewrite the hydrogen spectrum that
follows from the Bohr’s postulate
1 1 1
= 𝑅 ( 2 − 2)
𝜆 𝑛𝑓 𝑛𝑖
To produce the Balmer series, we take 𝑛𝑓 = 2, and
1 1 1
𝑛𝑓 = 2 ∶ = 𝑅 ( 2 − 2 ) 𝑛𝑖 = 3,4,5 … Balmer
𝜆 2 𝑛𝑖
Wave-Particle Duality
of electromagnetic radiation (light).
Luis De Broglie suggested that this wave–particle duality was not restricted to light, but must be universal:
All material particles should also display a dual wave–particle behavior.
That is, the wave–particle duality present in light must also occur in matter:
Just as a photon has a light wave associated with it that governs its motion, so a material particle (e.g. an
electron) has an associated matter wave that govern its motion.
From Relativity
𝑟𝑒𝑠𝑡
𝑚𝑎𝑠𝑠=0
𝐸
𝐸 = 𝛾𝑚𝑐 2 = (𝑝𝑐)2 + ⏟
⏟ (𝑚𝑐 2 )2 ⏞
⟹ 𝑝= ∶ momentum of a photon (*)
√𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑅𝑒𝑠𝑡 𝑚𝑎𝑠𝑠
𝑐
𝑒𝑛𝑒𝑟𝑔𝑦 𝑒𝑛𝑒𝑟𝑔𝑦
𝑡𝑒𝑟𝑚 𝑡𝑒𝑟𝑚
ℎ
From (∗)and (∗∗) ⟹ we get 𝜆=
∶ for a photon
𝑝
? 𝒉 𝒉
𝐓𝐡𝐞 𝐝𝐞 𝐁𝐫𝐨𝐠𝐥𝐢𝐞 𝐡𝐲𝐩𝐨𝐭𝐡𝐞𝐬𝐢𝐬 ∶ 𝝀 =
⏞ = ∶ for an electron
𝒑 𝒎𝒗
19
In order to observe the interference and diffraction effects, which are characteristics of the wave properties,
some geometric parameters of the instruments such as slits must have dimensions comparable to the
wavelength of the electrons.
Then
𝑚
ℎ𝑐 (6.626 × 10−34 𝐽. 𝑠)/(1.6 × 10−19 𝐽/𝑒𝑉)(3 × 108 𝑠 )
𝜆= =
√2(𝑚𝑒 𝑐 2 )𝐾 √2(0.511 × 106 𝑒𝑉)𝐾
≡1 Å
−7
12.3 × 10 𝑚 𝑒𝑉 12.3 × ⏞
10−10 𝑚 12.3 Å
𝜆= = √𝑒𝑉 => 𝜆 = √𝑒𝑉
√106 𝑒𝑉 𝐾 √𝐾 √𝐾
K(eV) 1 10 100 1000
𝜆(𝐴° ) 12.3 3.89 1.23 0.39
The values of 𝜆’s are comparable to the spacing of atoms in a crystal lattice. Thus, a
crystal can be used as a grating. Experiments of this type were performed in 1927 by
Davisson and Germer by using electron beams
20
The Experiment
What Davisson and Germer found was that, although the electrons are scattered in all directions from the
crystal, the intensity reaches a maximum for electrons with an energy of 54 eV at an angle of 𝜃 =50o .
This can only result from a constructive interference of the scattered electrons.
Davisson and Germer concluded that set of regularly spaced atomic
planes within the crystal work as a grating so that the intensity peak is
due constructive interference of the scattered electrons from this
grating.
In fact, the intensity maximum of the scattered electrons in the
Davisson–Germer experiment corresponds to the first maximum (n = 1)
of the Bragg formula, follows from
12.3 Å 12.3 Å
𝜆𝑑𝑏 = √𝑒𝑉 = √𝑒𝑉 ⇒ 𝜆𝑑𝑏 = 1.65Å
√𝐾 √54 𝑒𝑉
Thus, the Davisson and Germer experiment directly verifies de Broglie’s hypothesis of the wave nature
of moving bodies.
Inspired by de Broglie’s hypothesis, verified by the Davisson and Germer experiment, Schrödinger
constructed the theory of wave mechanics which deals with the dynamics of microscopic particles.
He described the motion of particles by means of a wave function (𝒓 ⃗ , 𝒕) , which corresponds to the de
Broglie wave of the particle.
We will deal with the physical interpretation of 𝜓(𝑟, 𝑡) in the chapter 1.
These lecture notes are not original; they contain lots of material from several different sources
1
CH 3
FORMALISM
The purpose of this chapter is to cast the quantum theory in a more powerful and
general form. There is not much here that is truly new, the idea rather is to
generalize what we have already discovered in particular cases.
EX:
𝜕 𝜕𝑓(𝑥 ) 𝜕ℎ(𝑥)
̂=
𝐷 such that ̂ 𝑓 (𝑥 ) =
𝐷 ̂ ℎ(𝑥) =
and 𝐷
𝜕𝑥 𝜕𝑥 𝜕𝑥
is a linear operator such that
𝜕𝑓(𝑥 ) 𝜕ℎ(𝑥 )
̂ (𝑓(𝑥 ) + ℎ(𝑥) =
𝐷 + =𝐷̂ 𝑓(𝑥) + 𝐷
̂ ℎ(𝑥)
𝜕𝑥 𝜕𝑥
̂ (𝑐𝑓(𝑥 )) = 𝑐𝐷
𝐷 ̂ 𝑓(𝑥)
The formalism of quantum mechanics deals with operators that are linear and
wave functions that belong to an abstract Hilbert space.
Thus, the structure and mathematical properties of Hilbert spaces and the linear
operators are essential for a proper understanding of the formalism of quantum
mechanics and the natural language of quantum mechanics is therefore the linear
algebra.
For this, we are going to review briefly the properties of Hilbert spaces and those of
linear operators. Then we will introduce a new notation: Dirac’s bra-ket notation
3
Hilbert Space
A Hilbert space ℋ consists of a set of vectors 𝜙1 , 𝜙2 , … and a set of scalars 𝑎, 𝑏, 𝑐, …
which satisfy the following properties:
(a) 𝓗 is a linear space:
If 𝜙1 , 𝜙2 ∈ ℋ, and 𝑎 is scalar
𝑎 𝜙1 ∈ ℋ
(𝜙1 + 𝜙2 ) ∈ ℋ
Commutativity: 𝜙1 + 𝜙2 = 𝜙2 + 𝜙1
‖𝜙𝑛 − 𝜙𝑚 ‖ ⟶ 0 𝑎𝑠 𝑛, 𝑚 ⟶ ∞
4
EX:
𝑖𝑘𝑥
𝑒 𝑖𝑘𝑥 − 𝑒 −𝑖𝑘𝑥 𝑎1 𝑎1
𝑎1 𝑆𝑖𝑛𝑘𝑥 + 𝑎2 𝑒 = 0 = 𝑎1 ( ) + 𝑎2 𝑒 𝑖𝑘𝑥 = 𝑒 𝑖𝑘𝑥 ( + 𝑎2 ) + 𝑖𝑒 −𝑖𝑘𝑥 = 0
2𝑖 2𝑖 2
=> 𝑎1 , 𝑎2 = 0
𝜌 = ∑ 𝑎𝑖 𝜙𝑖 (∗)
𝑖=1
The expansion coefficients 𝑎𝑖 are called the components of the vector 𝜌 in the basis
{𝜙𝑖 }. Each component is given by the scalar product of 𝜌 with the corresponding
base vector:
𝑎𝑖 = ⟨𝜙𝑖 |𝜌⟩
6
(1) One of the most familiar example of a Hilbert space is the Euclidean vector
space consisting of three-dimensional vectors, denoted by R3, and equipped with
the scalar product. (A finite dimensional
Euclidean space is in many ways like an infinite
dimensional Hilbert space but there are some ways in
which the infinite dimensionality leads to subtle
differences we need to be aware of.)
Any vector in this space can be expanded in terms of the three unit vectors 𝑒̂𝑥 ,𝑒̂𝑦 , 𝑒̂𝑧 ,
(or 𝑖̂, 𝑗̂, 𝑘̂ ) which is called a basis with the property:
0 𝑖𝑓 𝑖 ≠ 𝑗
𝑒̂𝑖 ∙ 𝑒̂𝑗 = 𝛿𝑖𝑗 = {
1 𝑖𝑓 𝑖 = 𝑗
The vectors 𝑒̂𝑥 ,𝑒̂𝑦 , 𝑒̂𝑧 are said to span the vector space.
Any vector 𝐴⃗ of the Euclidean space can be written in terms of the base vectors as
𝐴⃗ = 𝐴𝑥 𝑒̂𝑥 + 𝐴𝑦 𝑒̂𝑦 + 𝐴𝑧 𝑒̂𝑧 = ∑ 𝐴𝑖 𝑒̂𝑖
𝑖=𝑥,𝑦,𝑧
where 𝐴𝑥 , 𝐴𝑦 , 𝐴𝑧 are the components of 𝐴⃗ in the basis (𝑒̂𝑥 , 𝑒̂𝑦 , 𝑒̂𝑧 ). They are
determined as
𝐴𝑖 = 𝐴⃗ ∙ 𝑒̂𝑖
𝐴⃗ ∙ 𝐵
⃗⃗ = 𝐴𝑥 𝐵𝑥 + 𝐴𝑦 𝐵𝑦 + 𝐴𝑧 𝐵𝑧 = ∑ 𝐴𝑖 𝐵𝑖
𝑖=𝑥,𝑦,𝑧
Note that the scalar product in Euclidian space is real and symmetric.
The length of the vector is
|𝐴⃗| = √𝐴⃗ ∙ 𝐴⃗
if 𝐴⃗ and 𝐵
⃗⃗ are orthogonal then 𝐴⃗ ∙ 𝐵
⃗⃗ = |𝐴⃗||𝐵
⃗⃗ | cos(𝐴⃗, 𝐵
⃗⃗ ) = 0
8
(2) The second example of a Hilbert space is the space of the entire square
integrable complex functions 𝑓(𝑥 ).
It is easy to verify that the space of square-integrable functions possesses the
properties of a Hilbert space. For instance, any linear combination of square-
integrable functions is also a square-integrable.
A good example of square-integrable functions is the wave function of quantum
mechanics, 𝜓(𝑥, 𝑡). We have seen that, according to Born’s probabilistic
interpretation of 𝜓(𝑥, 𝑡) the quantity |𝜓(𝑥, 𝑡) |2 𝑑𝑥 represents the probability of
finding, at time t, the particle in dx centered around the point x . The probability of
finding the particle somewhere in space must then be equal to 1:
∞
⟨𝜓|𝜓⟩ = ∫ |𝜓(𝑥, 𝑡) |2 𝑑𝑥 = 1
−∞
hence the wave functions of quantum mechanics are square-integrable. Wave
functions satisfying above eq. are said to be normalized or square-integrable. As
wave mechanics deals with square-integrable functions, any wave function which is
not square-integrable has no physical meaning in quantum mechanics.
𝑓 (𝑥 ) = ∑ 𝑎𝑖 𝜙𝑖 (∗)
𝑖=1
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.
10
DIRAC NOTATION
To free state vectors from coordinate meaning, Dirac introduced an economical and
powerful 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.
Each state function 𝜓(𝑥) is associated with a ket |𝜓⟩:
𝜓
⏟ (𝑥 ) ⟶ |𝜓
⏟⟩
𝑎 𝑠𝑡𝑎𝑡𝑒 𝑎 𝑘𝑒𝑡
𝑓𝑢𝑛𝑐𝑖𝑜𝑛
Each complex conjugate state vector 𝜓 ∗ (𝑥) is associated with a bra ⟨𝜓|.
⏟∗ (𝑥 ) ⟶ ⟨⏟
𝜓 𝜓|
𝑎 𝑐𝑜𝑚𝑝𝑙𝑒𝑥 𝑎 𝑏𝑟𝑎
𝑐𝑜𝑛𝑗𝑢𝑔𝑎𝑡𝑒
𝑠𝑡𝑎𝑡𝑒
𝑣𝑒𝑐𝑡𝑜𝑟
Dirac denoted the scalar (inner) product by the symbol, ⟨ | ⟩, which he called a
bra-ket.
⟨⏟ | ⟩ ∶ 𝑎 𝑏𝑟𝑎-𝑘𝑒𝑡
𝑠𝑐𝑎𝑙𝑎𝑟
𝑝𝑟𝑜𝑑𝑢𝑐𝑡
For instance, the scalar product of two states vectors, 𝜓(𝑥) and 𝜙(𝑥) is denoted by
the bra-ket ⟨𝜓|𝜙⟩.
⟨𝜓|𝜙⟩: the scalar product of two states vectors, 𝜓(𝑥) and 𝜙(𝑥)
⟨𝜙|𝜓⟩∗ = ⟨𝜓|𝜙⟩
∞ ∗ ∞
⟨𝜙|𝜓 ⟩∗ = (∫ 𝜙 𝑥 )𝜓(𝑥 )𝑑𝑥 ) = ∫ 𝜓 ∗ (𝑥 )𝜙(𝑥 )𝑑𝑥 = ⟨𝜓|𝜙⟩
∗(
−∞ −∞
Thus,
⟨𝑎1 𝜓1 + 𝑎2 𝜓2 |𝑏1 𝜙1 + 𝑏2 𝜙2 ⟩
= 𝑎1∗ 𝑏1 ⟨𝜓1 |𝜙1 ⟩ + 𝑎2∗ 𝑏2 ⟨𝜓2 |𝜙2 ⟩ + 𝑎1∗ 𝑏2 ⟨𝜓1 |𝜙2 ⟩ + 𝑎2∗ 𝑏1 ⟨𝜓2 |𝜙1 ⟩
12
A state vector 𝜙𝑖 (𝑥) is said to be normalized if its scalar product with itseif is
equal to unity:
∞ ∞
⟨ 𝜙𝑖 | 𝜙𝑖 ⟩ = ∫ 𝜙𝑖∗ (𝑥 )𝜙𝑖 (𝑥 )𝑑𝑥 = ∫ |𝜙𝑖 (𝑥 )|2 𝑑𝑥 = 1
−∞ −∞
Two state vectors 𝜙𝑖 (𝑥) and 𝜙𝑗 (𝑥) are said to be orthogonal if their scalar
product is zero:
∞
⟨𝜙𝑖 |𝜙𝑗 ⟩ = ∫ 𝜙𝑖∗ (𝑥 )𝜙𝑗 (𝑥 )𝑑𝑥 = 0
−∞
A set of state vectors {𝜙𝑛 } are said to be orthonormal if they are normalized
and mutually orthogonal
Moreover, a set of state vectors {𝜙𝑛 } is said to be complete if it spans the entire
space; that is, there is no need to introduce any additional base vector, such that
any other function in this ∞-dimensional Hilbert space, 𝜒(𝑥) can be expanded
as a linear combination:
𝜒(𝑥) = ∑ 𝑐𝑛 𝜙𝑛 (𝑥)
𝑛=1
The expansion coefficients 𝒄𝒏 above are called the components of the vector 𝜒 in the
basis {𝜙𝑛 } and given by the scalar product of 𝜒 with the corresponding base
vector, 𝑐𝑖 = ⟨𝜙𝑖 |𝜒⟩.
Projection ⃗⃗ ∙ 𝑒̂𝑖
𝑉𝑖 = 𝑉 𝑐𝑖 = ⟨𝜑𝑖 |𝜓⟩
Norm 1/2 1/2
√𝑉 ⃗⃗ = (∑ 𝑉𝑖2 )
⃗⃗ ∙ 𝑉 ⟨𝜓|𝜓⟩1/2 = (∑|𝑐𝑖 |2 )
𝑖 𝑖
14
HERMITIAN OPERATORS
〈𝑄 〉 = ∫ 𝜓 ∗ 𝑄̂ 𝜓 𝑑𝑥 = ⟨𝜓|𝑄̂ 𝜓⟩
Observables are represented by operators whose expectation values are real, since the
outcome of a measurement has got to be real:
!
⏞ 〈𝑄 〉∗ (∗)
〈𝑄 〉 =
But the complex conjugate of a scalar product reverses the order:
∗
〈𝑄 〉∗ = ⟨𝜓|𝑄̂ 𝜓⟩ = ⟨𝑄̂ 𝜓|𝜓⟩
From (*) we have that if 𝑄̂ represents an observable then it satisfies the condition
But, it is possible to show that these two definitions are equivalent (See Probl.3.3)
𝑄̂ † = 𝑄̂
15
1.st way:
2.nd way:
𝜕
̂=
EX.(2) Consider the operator 𝐷 ̂ † =?
defined in a ℋ. 𝐷
𝜕𝑥
̂ † 𝑓|𝑔⟩ = ⟨𝑓|𝐷
⟨𝐷 ̂ 𝑔⟩
∞
𝜕
̂ 𝑔⟩ = ∫ ⏟
𝑅𝐻𝑆 = ⟨𝑓|𝐷 𝑓∗ 𝑔 𝑑𝑥
⏟
𝜕𝑥
−∞ 𝑢 𝑑𝑣
∞ ∞
∗
𝜕 𝜕 𝜕
⏟∗ 𝑔|∞
=𝑓 −∞ − ∫ (
̂ 𝑓|𝑔⟩
𝑓 ∗ ) 𝑔 𝑑𝑥 = ∫ (− 𝑓) 𝑔 𝑑𝑥 = ⟨− 𝑓|𝑔⟩ = ⟨−𝐷
𝜕𝑥 𝜕𝑥 𝜕𝑥
=0 −∞ −∞
̂ † 𝑓|𝑔⟩ = ⟨−𝐷
𝐿𝐻𝑆 = 𝑅𝐻𝑆 => ⟨𝐷 ̂ 𝑓|𝑔⟩ => 𝐷
̂ † = −𝐷
̂
16
†
EX: (3) Given a product operator 𝐶̂ ≡ 𝐴̂𝐵̂ . What is 𝐶̂ † = (𝐴̂𝐵̂) ?
†
We should find an object (𝐴̂𝐵̂) that fits in
†
⟨(𝐴̂𝐵̂) 𝑓|𝑔⟩ = ⟨𝑓|(𝐴̂𝐵̂)𝑔⟩
= ⟨𝑝̂ 𝑓|𝑔⟩
Problem 3.4
(a) Consider two Hermitian operators
𝐴̂ † =𝐴̂
𝐴̂+ = 𝐴̂ => ⟨𝑓|𝐴̂𝑔⟩ = ⟨𝐴̂† 𝑓|𝑔⟩ =
⏞ ⟨𝐴̂𝑓|𝑔⟩
𝐵̂+ = 𝐵̂ => ⟨𝑓|𝐵̂𝑔⟩ = ⟨𝐵̂† 𝑓|𝑔⟩ =
⏟ ⟨𝐵̂𝑓|𝑔⟩
𝐵̂ † =𝐵̂
Their sum
⟨𝑓|(𝐴̂ + 𝐵̂)𝑔⟩ = ⟨𝑓|𝐴̂𝑔⟩ + ⟨𝑓|𝐵̂𝑔⟩ = ⟨𝐴̂𝑓|𝑔⟩ + ⟨𝐵̂ 𝑓|𝑔⟩ = ⟨(𝐴̂ + 𝐵̂)𝑓|𝑔⟩
+
(𝐴̂ + 𝐵̂) = (𝐴̂ + 𝐵̂)
(c) Consider two Hermitian operators 𝐴̂ and 𝐵̂. Under what condition their product
𝐴̂𝐵̂ is Hermitian?
𝐴̂+ = 𝐴̂ => ⟨𝑓|𝐴̂𝑔⟩ = ⟨𝐴̂† 𝑓|𝑔⟩ = ⟨𝐴̂𝑓|𝑔⟩
𝐵̂ + = 𝐵̂ => ⟨𝑓|𝐵̂𝑔⟩ = ⟨𝐵̂† 𝑓|𝑔⟩ = ⟨𝐵̂𝑓|𝑔⟩
If their product is Hermitian we have
!
(𝐴̂𝐵̂) = (𝐴̂𝐵̂)† => ⟨𝑓|(𝐴̂𝐵̂)𝑔⟩ =
⏞ ⟨(𝐴̂𝐵̂)𝑓|𝑔⟩
!
⟨𝑓|(𝐴̂𝐵̂ )𝑔⟩ = ⟨𝐴̂𝑓|𝐵̂𝑔⟩ = ⟨𝐵̂𝐴̂𝑓|𝑔⟩ =
⏞ ⟨(𝐴̂𝐵̂)𝑓|𝑔⟩ =>
Therefore, if
𝐵̂𝐴̂ = 𝐴̂𝐵̂ => [𝐴̂, 𝐵̂] = 0
Then 𝐴̂𝐵̂ is Hermitian.
̂ corresponding to energy 𝐸 with
(d) Energy operator Hamiltonian operator 𝐻
̂ 𝜓ℓ = 𝐸ℓ 𝜓ℓ
𝐻
Thus the Hamiltonian operator should be a Hermitian operator
! 𝑝̂ 2
̂†
𝐻 = ̂ with
⏞ 𝐻 ̂=
𝐻 + 𝑉(𝑥)
2𝑚
From EX(3) we have
(𝑝̂ 2 )† = 𝑝̂ † 𝑝̂ † = 𝑝̂ 2
For V(x), which is a real function 𝑉̂ † =?
⟨𝑉̂ † 𝜓ℓ |𝜓𝑛 ⟩ = ⟨𝜓ℓ |𝑉̂ 𝜓𝑛 ⟩
∞ ∞
Determinate States
In each of the experiments, the particles are described by the same function
Ψ(𝑥, 0) at some initial time t=0.
There are two possible cases for the choice of Ψ(𝑥, 0):
1. Wave function Ψ(𝑥, 0) may be a general superposition of the possible
solutions of the Sch.eq:
∞
Ψ(𝑥, 0) = ∑ 𝑐𝑛 𝜓𝑛 (𝑥 )
𝑛=1
2. In each of the experiments, the particles may be described by the function given
by
Ψ(𝑥, 0) = 𝜓𝑢 (𝑥 ) (∗)
such that every measurement of the observable 𝑄 is certain to return the same
value, say 𝑞𝑢 . Such a state is called a determinate state.
2
2
𝑄 〉) 〉 = 〈(𝑄 − 𝑞𝑢 )2 〉 = ⟨𝜓𝑢 |(𝑄̂ − 𝑞𝑢 ) 𝜓𝑢 ⟩
𝝈𝟐 = 𝟎 => 〈(𝑄 − 〈⏟
=𝑞𝑢
̂.
Determinate states are eigenfunctions of 𝑸
22
For example, determinate states of the total energy are eigenfunctions of the
Hamiltonian operator
̂ 𝜓𝑛 = 𝐸𝑛 𝜓𝑛
𝐻
Continuous Spectra
So far, we have excluded continuous spectra from the discussion. From a physical
point of view, a continuous spectrum (e.g. the energy of a free quantum object) makes
perfect sense. But we have the problem that the corresponding eigenfunctions are not
square integrable, and therefore we cannot properly define a scalar product.
Since these unphysical states are not square integrable, the
previously-developed probability concept of quantum mechanics cannot work with
them (or at least not readily)—that is the essential difficulty.
Example 3.2
Find the eigenfunctions and eigenvalues of the momentum operator.
Let 𝜙𝑝 (𝑥 ) be the eigenfunction and 𝑝 the eigenvalue:
𝑑
𝑝̂𝑥 𝜙𝑝 (𝑥 ) = 𝑝 𝜙𝑝 (𝑥 ) ⟹ −𝑖ℏ 𝜙𝑝 (𝑥 ) = 𝑝 𝜙𝑝 (𝑥 )
𝑑𝑥
OR
𝑑𝜙𝑝 (𝑥 ) 𝑝
− 𝑖 𝜙𝑝 (𝑥 ) = 0
𝑑𝑥 ℏ
The general solution is
𝑝𝑥
𝑖
𝜙𝑝 (𝑥 ) = 𝐴 𝑒 ℏ 𝑜𝑟 with 𝑝 = 𝑘ℏ
This is not square integrable,
∞ ∞
Ψ(𝑥, 0) = ∑ 𝑐𝑛 𝑓𝑛 (𝑥 ) (∗∗∗)
𝑛
The eq.(*) is similar to (**) but the indices (𝑝, 𝑝′) are now continuous variables, and
the Kronecker delta has become a Dirac Delta , but otherwise it looks just the same.
Note that the eigenfunctions {𝝓𝒑 (𝒙)} are complete and since the indices are now
continuous variables, the sum in (***) is replaced by an integral to write any
function 𝑔(𝑥 )
∞
∫ 𝑔(𝑥) 𝜙𝑝∗ ′ (𝑥 )𝑑𝑥 = ∫ 𝑐(𝑝) 𝑑𝑝 ∫ 𝜙𝑝∗ ′ (𝑥 )𝜙𝑝 (𝑥 )𝑑𝑥 = ∫ 𝑐(𝑝) 𝛿 (𝑝 − 𝑝′ )𝑑𝑝 = 𝑐(𝑝′ )
−∞ −∞ ⏟
−∞ −∞
⟨𝜙𝑝′ |𝜙𝑝 ⟩
𝑥 𝜓𝑦 (𝑥 ) = 𝑦 𝜓𝑦 (𝑥)
𝑓(𝑥 ) = ∫ 𝑐 (𝑦′) 𝜓𝑦′ (𝑥 ) 𝑑𝑦′ => 𝑐 (𝑦) = ⟨𝜓𝑦 |𝑓⟩ = ∫ 𝑓(𝑥 ) 𝜓𝑦∗ (𝑥 ) 𝑑𝑥
−∞ −∞
∞
𝑐 (𝑦) = ∫−∞ 𝑓 (𝑥 ) 𝛿 (𝑥 − 𝑦 ) 𝑑𝑥 => 𝑐 (𝑦) = 𝑓(𝑦) (For its interpretation see page 29)
26
In Ch2, the time dependent Schrödinger equation with a time independent Hamiltonian
𝜕Ψ(𝑥, 𝑡)
𝑖ℏ = 𝐻Ψ(𝑥, 𝑡)
𝜕𝑡
has the most general solution given by
∞ ∞
𝑖𝐸𝑛 𝑡
Ψ(𝑥, 𝑡) = ∑ 𝑐𝑛 Ψ𝑛 (𝑥, 𝑡) = ∑ 𝑐𝑛 𝜓𝑛 (𝑥 ) 𝑒 − ℏ
𝑛=1 𝑛=1
Here
𝑖𝐸𝑛 𝑡
Ψ𝑛 (𝑥, 𝑡) = 𝜓𝑛 (𝑥 ) 𝑒 − ℏ
With
𝐻𝜓𝑛 (𝑥 ) = 𝐸𝑛 𝜓𝑛 (𝑥 )
are stationary state solutions, in the sense that all probabilities and expectation
values are independent of time.
Possible outcomes of an energy measurement in state Ψ(𝑥, 𝑡) will be 𝐸𝑛 with the
probability |𝑐𝑛 |2 .
27
𝑄̂ (𝑥, 𝑝) is hermitian.
{𝑓𝑛 (𝑥 )} : orthogonal => ⟨𝑓𝑚 |𝑓𝑛 ⟩ = 𝛿𝑛𝑚
and square integrable =>
∞
∫ 𝑓𝑛∗ (𝑥 ) 𝑓𝑛 (𝑥 )𝑑𝑥 < ∞
−∞
=>eigenfunctions are normalizable.
Eigenvalues 𝑞𝑛 : real.
If {𝑓𝑛 (𝑥 )} is also complete, any wave function 𝑔(𝑥 ) can be written uniquely and
exactly as a linear combination of them:
Then,
|𝑔⟩ = ∑ 𝑐𝑛 |𝑓𝑛 ⟩ => |𝑔⟩ = ∑|𝑓𝑛 ⟩⟨𝑓𝑛 | 𝑔⟩ => ∑|𝑓𝑛 ⟩⟨𝑓𝑛 | = 𝕀 ∶ 𝑡ℎ𝑒 𝑢𝑛𝑖𝑡 𝑜𝑝𝑒𝑟𝑎𝑡𝑜𝑟
𝑛 𝑛 𝑛
Now, if you measure the observable 𝑄 on a state 𝜓(𝑥) you are certain to get one of
the eigenvalues:
28
!
∑|𝑐𝑛 =
⏞1 |2
𝑛
∗
1 = ⟨𝑔|𝑔⟩ = ⟨(∑𝑚 𝑐𝑚 𝑓𝑚 (𝑥 ))|(∑𝑛 𝑐𝑛 𝑓𝑛 (𝑥 ))⟩ = ∑ ∑ 𝑐𝑚 ⟨𝑓𝑚 |𝑓𝑛 ⟩
𝑐𝑛 ⏟
𝑛 𝑚 𝛿𝑛𝑚
1 = ∑|𝑐𝑛 |2
𝑛
Similarly,
𝑞𝑛 𝑓𝑛
⏞
〈𝑄 〉 = ⟨𝑔|𝑄̂ 𝑔⟩ = ⟨(∑𝑚 𝑐𝑚 𝑓𝑚 (𝑥 ))| (∑𝑛 𝑐𝑛 𝑄
̂ 𝑓𝑛 (𝑥 ))⟩ = ∑ ∑ 𝑐𝑚
∗
𝑐𝑛 𝑞𝑛 ⟨⏟𝑓𝑚 |𝑓𝑛 ⟩
𝑛 𝑚 𝛿𝑛𝑚
= ∑|𝑐𝑛 |2 𝑞𝑛 =>
𝑛
〈𝑄 〉 = ∑|𝑐𝑛 |2 𝑞𝑛
𝑛
29
𝑥 𝜓𝑦 (𝑥 ) = 𝑦 𝜓𝑦 (𝑥)
where
𝜓𝑦 (𝑥 ) = 𝛿 (𝑥 − 𝑦 )
𝑔(𝑥 ) = ∫ 𝑐(𝑦′) 𝜓𝑦′ (𝑥 ) 𝑑𝑦′ => 𝑐 (𝑦) = ⟨𝜓𝑦 |𝑔⟩ = ∫ 𝜓𝑦∗ (𝑥 )𝑔(𝑥 ) 𝑑𝑥
−∞ −∞
∞
𝑄̂ 𝑓𝑛 (𝑥 ) = 𝑞𝑛 𝑓𝑛 (𝑥 ) 𝑥 𝜓𝑦 (𝑥 ) = 𝑦 𝜓𝑦 (𝑥)
1 𝑖
𝑝𝑥
𝜙𝑝 (𝑥 ) = 𝑒 ℏ
√2𝜋ℏ
If we restrict ourselves to real eigenvalues, we do recover a kind of “orthonormality”.
⟨𝜙𝑝′ |𝜙𝑝 ⟩ = 𝛿 (𝑝 − 𝑝′ ) : Dirac orthogonality
Note that the eigenfunctions {𝜙𝑝 (𝑥)} are complete so that any function 𝜓(𝑥) can be
expanded as
∞ ∞
1 𝑝𝑥
𝜓(𝑥 ) = ∫ 𝑐(𝑝) 𝜙𝑝 (𝑥 )𝑑𝑝 => 𝑐 (𝑝) = ⟨𝜙𝑝 |𝜓⟩ = ∫ 𝑒 −𝑖 ℏ 𝜓(𝑥 )𝑑𝑥
√2𝜋ℏ
−∞ −∞
∞
1 −𝑖
𝑝𝑥
𝑐 (𝑝) → Φ(𝑝) = ∫𝑒 ℏ 𝜓(𝑥 )𝑑𝑥 ∶ momentum space wave function
√2𝜋ℏ
−∞
∞
1 𝑝𝑥
𝜓(𝑥 ) = ∫ 𝑒 𝑖 ℏ Φ(𝑥 )𝑑𝑝 ∶ position space wave function
√2𝜋ℏ
−∞
|Φ(𝑝)|2 𝑑𝑝
∶ probability of finding the particle in state 𝜓(𝑥 ) at the momentum p in the range dp
(𝑝, 𝑝 + 𝑑𝑝)
32
𝛼 1/4 −𝛼𝑥 2 /2
Ψ(𝑥, 0) = ( ) 𝑒
𝜋
What is the probability that its momentum is in the range (𝑝, 𝑝 + 𝑑𝑝)?
1/4
1
( p ) dx e x /2 e ipx /
2
2
1
1/4 1/2
dxe ( xip / /2
)2
e p
2 2
2
1/4
1 p2 /2
2
2
e
From this we find that the probability the momentum is in the range (p, p + dp) is
1/2
1 p2 /
| ( p) |2 dp
2
2
e
33
1
𝑚𝜔 4 −𝑚𝜔𝑥 2
𝜓0 (𝑥 ) = ( ) 𝑒 2ℏ
𝜋ℏ
1
𝑚𝜔 4 −𝑚𝜔𝑥 2 −𝑖𝐸0 𝑡
Ψ0 (𝑥, 𝑡) = ( ) 𝑒 2ℏ 𝑒 2 𝑤𝑖𝑡ℎ 𝐸0 = ℏ𝜔/2
𝜋ℏ
∞
1 −𝑖
𝑝𝑥
Φ(𝑝, 𝑡) = ∫𝑒 ℏ Ψ(𝑥, 𝑡)𝑑𝑥 ∶ momentum space wave function
√2𝜋ℏ
−∞
1 ∞
1𝑚𝜔 4 −𝑖𝐸0 𝑡 −𝑖
𝑝𝑥 𝑚𝜔 2
Φ(𝑝, 𝑡) = ( ) 𝑒 2 ∫𝑒 ℏ 𝑒 2ℏ 𝑥 𝑑𝑥
−
√2𝜋ℏ 𝜋ℏ
−∞
𝑝𝑥 𝑚𝜔 2 𝑚𝜔 𝑝𝑥
𝑒𝑥𝑝 [−𝑖 − 𝑥 ] = 𝑒𝑥𝑝 [− (2𝑖 + 𝑥 2 )]
ℏ 2ℏ 2ℏ 𝑚𝜔
𝑚𝜔 𝑝 2 𝑝2
= 𝑒𝑥𝑝 [− (𝑥 + 𝑖 ) ] 𝑒𝑥𝑝 [− ]
2ℏ 𝑚𝜔 2ℏ𝑚𝜔
1 ∞
1𝑚𝜔 4 −𝑖𝐸0 𝑡 − 𝑝2 𝑚𝜔 𝑝 2
Φ(𝑝, 𝑡) = ( ) 𝑒 2 𝑒 2ℏ𝑚𝜔 ∫ 𝑒𝑥𝑝 [− (𝑥 + 𝑖 ) ] 𝑑𝑥
√2𝜋ℏ 𝜋ℏ ⏟
2ℏ 𝑚𝜔
−∞
𝜋
√𝑚𝜔/2ℏ
1
1 4 − 𝑝2 𝑖𝐸0 𝑡
Φ(𝑝, 𝑡) = ( ) 𝑒 2ℏ𝑚𝜔 𝑒 − 2
𝜋ℏ𝑚𝜔
34
Classical momentum is
𝑝2
𝐸= => 𝑝 = ±√2𝑚𝐸.
2𝑚
In the ground state:𝐸 = ℏ𝜔/2. Then,
𝑝 = ±√𝑚ℏ𝜔
Thus the classical range of the momentum is
−√𝑚ℏ𝜔 ≤ 𝑝 ≤ √𝑚ℏ𝜔
The normalization=>
∞
1 = ∫ |Φ(𝑝, 𝑡)|2 𝑑𝑝 =
−∞
𝑜𝑢𝑡𝑠𝑖𝑑𝑒 𝑡ℎ𝑒 𝑐𝑙𝑎𝑠𝑠𝑖𝑐𝑎𝑙 𝑟𝑎𝑛𝑔𝑒
𝑜𝑢𝑡𝑠𝑖𝑑𝑒 𝑡ℎ𝑒 𝑐𝑙𝑎𝑠𝑠𝑖𝑐𝑎𝑙 𝑟𝑎𝑛𝑔𝑒
⏞
−√𝑚ℏ𝜔 0 √𝑚ℏ𝜔 ⏞ ∞
Thus, the probability that a measurement of p on a particle in the ground state would yield a value
outside the classical range is given by
𝑜𝑢𝑡𝑠𝑖𝑑𝑒 𝑡ℎ𝑒 𝑐𝑙𝑎𝑠𝑠𝑖𝑐𝑎𝑙 𝑟𝑎𝑛𝑔𝑒
𝑜𝑢𝑡𝑠𝑖𝑑𝑒 𝑡ℎ𝑒 𝑐𝑙𝑎𝑠𝑠𝑖𝑐𝑎𝑙 𝑟𝑎𝑛𝑔𝑒
⏞
−√𝑚ℏ𝜔 ⏞ ∞
0 √𝑚ℏ𝜔
Now,
1 2
√𝑚ℏ𝜔 √𝑚ℏ𝜔
1 4 𝑝2 𝑖𝐸 𝑡
− − 0
∫ |Φ(𝑝, 𝑡)|2 𝑑𝑝 = ∫ |( ) 𝑒 2ℏ𝑚𝜔 𝑒 2 | 𝑑𝑝
𝜋ℏ𝑚𝜔
0 0
1 √𝑚ℏ𝜔
1 2 𝑝2
−
=( ) ∫ 𝑒 ℏ𝑚𝜔 𝑑𝑝
𝜋ℏ𝑚𝜔
0
Chance of variable:
2 ℏ𝑚𝜔
𝑧≡√ 𝑝 => 𝑑𝑝 = √ 𝑑𝑧
ℏ𝑚𝜔 2
√𝑚ℏ𝜔 1 √2
1 ℏ𝑚𝜔 𝑧22
−
2
∫ |Φ(𝑝, 𝑡)| 𝑑𝑝 = ( ) √ ∫ 𝑒 2 𝑑𝑧
⏟𝜋ℏ𝑚𝜔 2
0 0
√1
2𝜋
√2
𝑧2 𝜋 𝜋
−
∫𝑒 2 𝑑𝑧 = √ 𝐸𝑟𝑓[1] = √ (0.843) = 1.056
2 2
0
√𝑚ℏ𝜔
1 𝜋 (0.843)
∫ |Φ(𝑝, 𝑡)|2 𝑑𝑝 = √ ∗ √ (0.843) =
2𝜋 2 2
0
√𝑚ℏ𝜔
(0.843)
𝑃 = 1 − 2 ∫ |Φ(𝑝, 𝑡)|2 𝑑𝑝 = 1 − 2 = 0.157 ≅ 0.16
2
0
1
ℏ
𝜎𝑥 𝜎𝑝 ≥
2
The more precise a wave’s position is, the less precise is its wavelength
(=momentum) , and vice versa.
In Ch1 and we have checked it several times, in the problems.
For any two state vectors |𝜓⟩ and |𝜙⟩ of the Hilbert space, we can show that
|⟨𝜓|𝜙⟩|2 ≤ ⟨𝜓|𝜓⟩⟨𝜙|𝜙⟩
2
𝜎𝐴2 = 〈(𝐴 − 〈𝐴〉)2 〉 = ⟨𝜓|(𝐴̂ − 〈𝐴〉) 𝜓⟩ = ⟨𝜓|(𝐴̂ − 〈𝐴〉)(𝐴̂ − 〈𝐴〉)𝜓⟩
= ⟨(𝐴̂ − 〈𝐴〉)𝜓|(𝐴̂ − 〈𝐴〉)𝜓⟩
Likewise
2
𝜎𝐵2 = 〈(𝐵 − 〈𝐵〉)2 〉 = ⟨𝜓|(𝐵̂ − 〈𝐵〉) 𝜓⟩ = ⟨𝜓|(𝐵̂ − 〈𝐵〉)(𝐵̂ − 〈𝐵〉)𝜓⟩
where
where
|⟨𝑓|𝑔⟩|2 = ⟨𝑓|𝑔⟩⟨𝑓|𝑔⟩∗
2
1
|𝑧 |2 = [𝑅𝑒(𝑧) ]2 + [𝐼𝑚(𝑧) ]2 ≥ [𝐼𝑚(𝑧 )]2 ∗
= [ (𝑧 − 𝑧 )]
2𝑖
Letting
2 2
1 1
⟨𝑓|𝑔⟩ ≡ 𝑧 => |⟨𝒇|𝒈⟩|𝟐 = |𝑧|2 = [ (𝑧 − 𝑧 ∗ )] = [ (⟨𝑓|𝑔⟩ − ⟨𝑔|𝑓⟩]
2𝑖 2𝑖
2
1
𝜎𝐴2 𝜎𝐵2 ≥ | (⟨𝑓|𝑔⟩ − ⟨𝑔|𝑓⟩| (∗∗)
2𝑖
Now,
Similarly,
So that
where
2
1
𝜎𝐴2 𝜎𝐵2 ≥ | 〈[𝐴̂, 𝐵̂]〉| (3 ∗)
2𝑖
𝑑
EX: Suppose 𝐴̂ = 𝑥̂ and 𝐵̂ = 𝑝̂ = −𝑖ℏ . Their commutator is
𝑑𝑥
[𝑥̂, 𝑝̂ ] = 𝑖ℏ
So,
1 2
ℏ2
𝜎𝑥2 𝜎𝑝2 ≥ | 𝑖ℏ| = [ ]
2𝑖 2
ℏ
𝜎𝑥 𝜎𝑝 ≥
2
-------------------------------------------------------------------------------------------------
Note that the uncertainty principle is not an extra assumption in quantum
theory, but rather a consequence of the statistical interpretation.
When you measure the position of the particle the act of the
measurement collapses the wave function to a narrow spike.
(An eigenstate of the position operator!) To produce a matter
wave that was spatially localized that much, we put several
matter waves of particle with different momenta together.
This wave function therefore necessarily carries a broad
range of wavelengths (and hence momenta) in its Fourier
decomposition.
Poisiton =precise
Momentum=unknown (∞′ 𝑙𝑦 𝑚𝑎𝑛𝑦)
-------------------------------------------------------------------------------------------
6
Suppose that 𝑃̂ and 𝑄̂ are two operators that have a complete set of common
eigenfunctions {𝑓𝑛 }:
𝑃̂𝑓𝑛 = 𝜆𝑛 𝑓𝑛
𝑄̂ 𝑓𝑛 = 𝑞𝑛 𝑓𝑛
That is 𝑓𝑛 (𝑥) is an eigenfunction of both of 𝑃̂ and 𝑄̂ . Since the set {𝑓𝑛 } is complete
=> any function 𝑔(𝑥) in Hilbert space can be expressed as a linear combination
𝑔(𝑥 ) = ∑ 𝑐𝑛 𝑓𝑛
𝑛
Now,
[𝑃̂, 𝑄̂ ]𝑔(𝑥 ) = (𝑃̂𝑄̂ − 𝑄̂ 𝑃̂)𝑔(𝑥 ) = (𝑃̂𝑄̂ − 𝑄̂ 𝑃̂) ∑ 𝑐𝑛 𝑓𝑛
𝑛
= 𝑃̂ ∑ 𝑐𝑛 𝑄̂ 𝑓𝑛 − 𝑄̂ ∑ 𝑐𝑛 𝑃̂𝑓𝑛 = ∑ 𝑐𝑛 𝑞𝑛 𝑃̂𝑓𝑛 − ∑ 𝑐𝑛 𝜆𝑛 𝑄̂ 𝑓𝑛
𝑛 𝑛 𝑛 𝑛
= ∑ 𝑐𝑛 𝑞𝑛 𝜆𝑛 𝑓𝑛 − ∑ 𝑐𝑛 𝜆𝑛 𝑞𝑛 𝑓𝑛 = 0
𝑛 𝑛
Since this is true for any function g=>[𝑃̂, 𝑄̂ ] = 0
̂ , 𝐿̂2 , 𝐿̂𝑧 } are mutually compatible observables=>
For example, in the H-atom, {𝐻
̂ , 𝐿̂2 ] = [𝐻
[𝐻 ̂ , 𝐿̂𝑧 ] = [𝐿̂2 , 𝐿̂𝑧 ] = 0 =>
2
1
𝜎𝐴2 𝜎𝐵2 ≥ | 〈[𝐴̂, 𝐵̂]〉|
2𝑖
𝑝̂2
𝐴̂ = 𝑥̂ and 𝐵̂ = 𝐻 = + 𝑉(𝑥)
2𝑚
2
1
𝜎𝑥2 𝜎𝐻2 ̂ ] 〉|
≥ | 〈[𝑥̂, 𝐻
2𝑖
𝑝̂ 2 1
̂ ] = [𝑥̂,
[𝑥̂, 𝐻 + 𝑉(𝑥)] = [𝑥̂, 𝑝̂ 2 ] + [⏟𝑉(𝑥), 𝑥̂ ] =
2𝑚 2𝑚
=0
1 1 𝑖ℏ𝑝̂
([⏟
𝑥̂, 𝑝̂ ]𝑝̂ + 𝑝̂ [𝑥̂, 𝑝̂ ]) = ̂] =
(2𝑖ℏ𝑝̂ ) => [𝑥̂, 𝐻
2𝑚 2𝑚 𝑚
𝑖ℏ
𝑑〈𝑥〉
Also 〈𝑝〉 = 0 =>Ehrenfest’s theorem:〈𝑝〉 = 𝑚 . (Expectation values obey
𝑑𝑡
classical laws)
𝑑〈𝑥〉
In a stationary state, expectations do not change in time, and so .
𝑑𝑡
8
𝜕𝛹 𝜕𝛹 1
𝑖ℏ = 𝐻Ψ => = 𝐻Ψ
𝜕𝑡 𝜕𝑡 𝑖ℏ
𝑑 1 𝜕𝑄̂ 1
〈𝑄 〉 = ⟨ 𝐻Ψ|𝑄̂ 𝛹⟩ + ⟨𝛹| 𝛹⟩ + ⟨𝛹|𝑄̂ 𝐻Ψ⟩
𝑑𝑡 𝑖ℏ 𝜕𝑡 𝑖ℏ
1 1 𝜕𝑄̂
=− ⟨𝐻Ψ|𝑄̂ 𝛹⟩ + ⟨𝛹|𝑄̂ 𝐻Ψ⟩ + ⟨𝛹| 𝛹⟩
𝑖ℏ 𝑖ℏ 𝜕𝑡
1
̂
1
̂ 𝜕𝑄̂
= − ⟨Ψ|𝐻𝑄 𝛹⟩ + ⟨𝛹|𝑄 𝐻Ψ⟩ + ⟨𝛹| 𝛹⟩
𝑖ℏ 𝑖ℏ 𝜕𝑡
1 𝜕𝑄̂
=− ⟨Ψ|(𝐻𝑄̂ − 𝑄̂ 𝐻)𝛹⟩ + ⟨𝛹| 𝛹⟩
𝑖ℏ 𝜕𝑡
𝑑 𝑖 𝜕𝑄̂
〈𝑄 〉 = 〈[𝐻, 𝑄̂ ]〉 + 〈 〉
𝑑𝑡 ℏ 𝜕𝑡
𝜕𝑄̂ 𝑑 𝑖
= 0 => 〈𝑄 〉 = 〈[𝐻, 𝑄̂ ]〉
𝜕𝑡 𝑑𝑡 ℏ
𝑑 𝑖 𝜕𝑄̂
̂
〈𝑄 〉 = 〈[𝐻, 𝑄 ]〉 + 〈 〉
𝑑𝑡 ℏ 𝜕𝑡
a) 𝑄 = 1 . [𝐻, 1] = 0
𝑑 𝑑
〈𝑄 〉 = 0 = ⟨𝛹 |𝛹 ⟩
𝑑𝑡 𝑑𝑡
(this is the conservation of normalization, which we originally proved in Eq. 1.27).
b)
𝑄̂ = 𝐻
𝜕𝐻 𝑑 𝑑
= 0 , [𝐻, 𝐻 ] = 0 => 〈𝐻 〉 = 〈𝐸 〉 = 0
𝜕𝑡 𝑑𝑡 𝑑𝑡
=> 〈𝐸 〉 𝑖𝑠 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 𝑜𝑓 𝑚𝑜𝑡𝑖𝑜𝑛
𝑝̂ 2 1
[𝐻, 𝑥̂ ] = [ + 𝑉(𝑥), 𝑥̂] = [𝑝̂ 2 , 𝑥̂ ] + [⏟𝑉(𝑥), 𝑥̂ ] =
2𝑚 2𝑚
=0
1 1 𝑖ℏ𝑝̂
[𝐻, 𝑥̂ ] = (𝑝̂ [⏟𝑝̂ , 𝑥̂ ] + [𝑝̂ , 𝑥̂ ]𝑝̂ ) = (−2𝑖ℏ𝑝̂ ) => [𝐻, 𝑥̂ ] = −
2𝑚 2𝑚 𝑚
−𝑖ℏ
𝑑 𝑖 𝑖ℏ𝑝̂ 𝑑 〈𝑝̂ 〉
〈𝑥 〉 = 〈− 〉 => 〈𝑥 〉 =
𝑑𝑡 ℏ 𝑚 𝑑𝑡 𝑚
10
d) 𝑄 = 𝑝
𝑑 𝑖 𝜕𝑝̂
〈𝑝〉 = 〈[𝐻, 𝑝̂ ]〉 + 〈 〉
𝑑𝑡 ℏ 𝜕𝑡
𝑝̂ 2 1
[𝐻, 𝑝̂ ] = [ + 𝑉(𝑥), 𝑝̂ ] = [𝑝̂ 2 , 𝑝̂ ] + [𝑉(𝑥), 𝑝̂ ]
2𝑚 2𝑚
𝑑𝑓 𝑑 𝑑𝑉
[𝑉 (𝑥 ), 𝑝̂ ]𝑓 (𝑥 ) = −𝑖ℏ (𝑉(𝑥 ) − (𝑉𝑓)) = 𝑖ℏ 𝑓
𝑑𝑥 𝑑𝑥 𝑑𝑥
𝑑𝑉
[𝐻, 𝑝̂ ] = 𝑖ℏ
𝑑𝑥
𝑑 𝑖 𝑑𝑉 𝑑 𝑑𝑉
〈𝑝〉 = 〈𝑖ℏ 〉 => 〈𝑝 〉 = − 〈 〉
𝑑𝑡 ℏ 𝑑𝑥 𝑑𝑡 𝑑𝑥
This is in accord with the Ehrenfest’s theorem.
11
𝑑 𝑖
〈𝑄 〉 = 〈[𝐻, 𝑄̂ ]〉
𝑑𝑡 ℏ
Then,
2
2 2
ℏ2 𝑑〈𝑄 〉 ℏ 𝑑 〈𝑄 〉
𝜎𝐻 𝜎𝑄 ≥ | | 𝑂𝑅 , 𝜎
⏟𝐻 𝜎𝑄 ≥ | |
4 𝑑𝑡 2 𝑑𝑡
∆𝐸
OR
𝜎𝑄 ℏ
∆𝐸 ≥
𝑑 〈𝑄 〉 2
| |
𝑑𝑡
where 𝜎𝐻 ≡ ∆𝐸. We now identify
𝜎𝑄
∆𝑡 ≡
𝑑 〈𝑄 〉
| |
𝑑𝑡
Then
ℏ
Δ𝐸Δ𝑡 ≥
2
It remains to interpret the quantity 𝚫𝒕.
𝑑 〈𝑄 〉
| | : 𝑀𝑒𝑎𝑠𝑢𝑟𝑒𝑠 ℎ𝑜𝑤 𝑓𝑎𝑠𝑡 〈𝑄 〉 𝑐ℎ𝑎𝑛𝑔𝑒𝑠 𝑜𝑣𝑒𝑟 𝑡𝑖𝑚𝑒
𝑑𝑡
∆𝑡 then tells you the approximate amount of time it takes for the expectation value
of an observable to change by a standard deviation.
𝜎𝑄 ← 𝑎𝑚𝑜𝑢𝑛𝑡 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒
∆𝑡 ≡
𝑑 〈𝑄 〉 ← 𝑣𝑒𝑙𝑜𝑐𝑖𝑡𝑦 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒
| |
𝑑𝑡
13
Problem 3.18 Test the energy-time uncertainty principle for the wave function of
𝑑〈𝑥〉
Problem 2.5 and observable x, by calculating 𝜎𝐻 , 𝜎𝑥 and exactly.
𝑑𝑡
In a stationary state for which the energy is uniquely determined all expectation
values are constant in time:
𝑑 〈𝑄 〉
=0
𝑑𝑡
And ∆𝐸 = 0 => ∆𝑡 = ∞. To make something happen you must take a time
dependent linear combination state.
1
Ψ(𝑥, 𝑡) = (𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )
√2
Energy-time uncertainty relation says
ℏ 𝜎𝑄
Δ𝐸Δ𝑡 ≥ , 𝜎𝐻 ≡ ∆𝐸 𝑤𝑖𝑡ℎ ∆𝑡 ≡
2 𝑑 〈𝑄 〉
| |
𝑑𝑡
For the position observable 𝑄 = 𝑥
𝜎𝑥
∆𝑡 ≡
𝑑 〈𝑥 〉
| |
𝑑𝑡
Now
2
ℏ 2 2
ℏ2 𝑑〈𝑥 〉
Δ𝐸Δ𝑡 ≥ => 𝜎𝐻 𝜎𝑥 ≥ [ ]
2 4 𝑑𝑡
1 1
(Δ𝐸)2 ≡ 𝜎𝐻2 = 〈𝐻 2 〉 − 〈𝐻 〉2 = (𝐸1 − 𝐸2 )2 = (3ℏ𝜔)2
4 4
𝑎2 1 5 32 2
𝜎𝑥2 = 〈𝑥 2〉
− 〈𝑥 〉2 = [ − 2
− ( 2 ) cos2 3𝜔]
4 3 4(𝜋) 9𝜋
14
𝑑 〈𝑥 〉 𝑑 𝑎 32 𝑎 32 𝑑 〈𝑥 〉 8ℏ
= [1 − 2 cos 3𝜔] = 3𝜔 sin 3𝜔 => = sin 3𝜔
𝑑𝑡 𝑑𝑡 2 9𝜋 2 9𝜋 2 𝑑𝑡 3𝑚𝑎
2
2 2
ℏ2 𝑑 〈𝑥 〉
𝜎𝐻 𝜎𝑥 ≥ [ ]
4 𝑑𝑡
2 2
1 𝑎 1 5 32
𝐿𝐻𝑆 ≡ 𝜎𝐻2 𝜎𝑥2 = (3ℏ𝜔)2 [ − 2
− ( 2 ) cos 2 3𝜔]
4 4 3 4(𝜋) 9𝜋
3 2 1 5 32 2
= ( ) (𝑎ℏ𝜔)2 [ − 2
− ( 2 ) cos2 3𝜔]
4 3 4(𝜋) 9𝜋
2 2
ℏ2 𝑑 〈𝑥 〉 ℏ2 8ℏ ℏ 8ℏ 2 2
𝑅𝐻𝑆 ≡ [ ] = [ sin 3𝜔] = ( ) sin 3𝜔
4 𝑑𝑡 4 3𝑚𝑎 2 3𝑚𝑎
2
ℏ2 𝑑〈𝑥 〉 8 2
𝑅𝐻𝑆 ≡ [ ] = ( 2 ) (ℏ𝜔𝑎)2 sin2 3𝜔
4 𝑑𝑡 3𝜋
?
3 2 1 5 32 2
8 2
2
( ) (𝑎ℏ𝜔) [ − 2 ⏞(
− ( 2 ) cos 3𝜔] ≥ ) (ℏ𝜔𝑎 )2 sin2 3𝜔
4 3 4(𝜋) 2 9𝜋 3𝜋 2
?
1 5 32 2 4 × 8 2
[ − ⏞(
− ( 2 ) cos2 3𝜔] ≥ ) sin2 3𝜔
3 4(𝜋) 2 9𝜋 3 × 3𝜋 2
?
1 5 4×8 2 2 32 2 32 2
[ − ⏞(
]≥ 2
) sin 3𝜔 + ( 2 ) cos 3𝜔 = ( 2 )
3 4(𝜋)2 3 × 3𝜋 2 9𝜋 9𝜋
𝑌𝐸𝑆
1 5 32 2
[ − ⏞ (
] ≥ )
3 4 (𝜋 )2 ⏟9𝜋 2
⏟
0.20668 0.12978
15
Δ𝐸 ≡ 𝜎𝐻 = √〈𝐻 2 〉 − 〈𝐻 〉2
〈𝐻 〉 = ⟨Ψ|𝐻 |Ψ⟩
1 1
=⟨ (𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )|𝐻 (𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )⟩
√2 √2
1
= ⟨(𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )|(𝐸1 𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝐸2 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )⟩
2
1
〈𝐻 〉 = (𝐸 + 𝐸2 )
2 1
Note that all the cross terms vanish due to orthogonality of energy eigenfunctions:
〈𝐻 2 〉
1 1
= ⟨ (𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )|𝐻 2 (𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )⟩
√2 √2
1
= ⟨(𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )|𝐻(𝐸1 𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝐸2 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )⟩
2
1
= ⟨(𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )|(𝐸12 𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝐸22 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )⟩
2
Here
1 2 1 1
(𝐸1 + 𝐸22 ) − (𝐸12 + 𝐸22 ) + 2𝐸1 𝐸2 => 𝜎𝐻2 = (𝐸12 + 𝐸22 − 2𝐸1 𝐸2 )
2 4 4
1
= (𝐸1 − 𝐸2 )2
4
1
(Δ𝐸)2 ≡ 𝜎𝐻2 = (𝐸1 − 𝐸2 )2
4
2 2 2 2 2 2
ℏ 𝜋 ℏ 𝜋
(𝐸1 − 𝐸2 )2 = [ 2
(1 − 4 )] = [− 2
3] ≡ (3ℏ𝜔)2
2𝑚𝑎 2𝑚𝑎
1 1
(Δ𝐸)2 ≡ 𝜎𝐻2 = (𝐸1 − 𝐸2 )2 = (3ℏ𝜔)2 (𝑎)
4 4
∞ ∞
∗
〈𝑥 〉 = ∫ Ψ 𝑥̂ Ψ 𝑑𝑥 = ∫ 𝑥̂ |Ψ(x, t)|2 𝑑𝑥
−∞ −∞
From problem 2.5(c)
17
𝑎 32
〈𝑥 〉 = [1 − 2 cos 3𝜔]
2 9𝜋
𝑑〈𝑥〉 𝑑 𝑎 32 𝑎 32 𝑑〈𝑥〉 8ℏ
= [1 − cos 3𝜔] = 3𝜔 sin 3𝜔 => = sin 3𝜔 (b)
𝑑𝑡 𝑑𝑡 2 9𝜋2 2 9𝜋2 𝑑𝑡 3𝑚𝑎
〈𝑥 2 〉
1 1
= ⟨ (𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )|𝑥 2 (𝜓1 (𝑥)𝑒 −𝑖𝐸1 𝑡/ℏ + 𝜓2 (𝑥)𝑒 −𝑖𝐸2 𝑡/ℏ )⟩
√2 √2
1
= [⟨𝜓1 |𝑥 2 |𝜓1 ⟩ + ⟨𝜓2 |𝑥 2 |𝜓2 ⟩ + 𝑒 𝑖(𝐸1 −𝐸2 )𝑡/ℏ ⟨𝜓1 |𝑥 2 |𝜓2 ⟩
2
+ 𝑒 −𝑖(𝐸1 −𝐸2 )𝑡/ℏ ⟨𝜓2 |𝑥 2 |𝜓1 ⟩]
18
𝑎
2 𝑛𝜋𝑥 1 1
⟨𝜓𝑛 |𝑥 2 |𝜓𝑛 ⟩ = ∫ 𝑥 2 sin2 ( ) 𝑑𝑥 = 𝑎2 [ − ]
𝑎 𝑎 3 2(𝑛𝜋)2
0
1 1 1 1
⟨𝜓1 |𝑥 2 |𝜓1 ⟩ = 𝑎2 [ − ] , ⟨ 𝜓 2 | 𝑥 2|
𝜓 2 ⟩ = 𝑎 2
[ − ]
3 2(𝜋)2 3 2(2𝜋)2
𝑎
2 𝑛𝜋𝑥 𝑚𝜋𝑥
⟨𝜓𝑛 |𝑥 2 |𝜓𝑚 ⟩ = ∫ 𝑥 2 sin ( ) sin ( ) 𝑑𝑥 =
𝑎 𝑎 𝑎
0
𝑎
1 (𝑛 − 𝑚)𝜋𝑥 (𝑛 + 𝑚)𝜋𝑥
= ∫ 𝑥 2 [cos ( ) − cos ( )] 𝑑𝑥
𝑎 𝑎 𝑎
0
𝑎 𝑎
2
𝑢 2𝑎2 𝑥 𝑢 𝑎 3 𝑢𝜋𝑥 2 𝑢
∫ 𝑥 [cos ( 𝜋𝑥)] 𝑑𝑥 = { 2 2 cos ( 𝜋𝑥) + ( ) [( ) − 2] sin ( 𝜋𝑥)}
𝑎 𝑢 𝜋 𝑎 𝑢𝜋 𝑎 𝑎 0
0
𝑎
(𝑛 − 𝑚)𝜋𝑥
∫ 𝑥 2 [cos ( )]
𝑎
0
𝑎
2𝑎2 𝑥 𝑢 𝑎 3 𝑢𝜋𝑥 2 𝑢
= { 2 2 cos ( 𝜋𝑥) + ( ) [( ) − 2] sin ( 𝜋𝑥)}
𝑢 𝜋 𝑎 𝑢𝜋 𝑎 𝑎 0
19
2𝑎3 𝑎 3
= { 2 2 cos(𝑢𝜋) + ( ) [(𝑢𝜋)2 − 2] ⏟
sin(𝑢𝜋)} , 𝑢 = 𝑛 − 𝑚
𝑢 𝜋 𝑢𝜋 =0
𝑎
2
(𝑛 − 𝑚)𝜋𝑥 2𝑎3
∫ 𝑥 [cos ( )] = 2 2
(−1)𝑛−𝑚
𝑎 (𝑛 − 𝑚) 𝜋
0
𝑎
2
(𝑛 + 𝑚)𝜋𝑥 2𝑎3
∫ 𝑥 [cos ( )] = 2 2
(−1)𝑛+𝑚
𝑎 (𝑛 + 𝑚) 𝜋
0
2|
1 2𝑎3 (−1)𝑛−𝑚 (−1)𝑛+𝑚
⟨𝜓𝑛 |𝑥 𝜓𝑚 ⟩ = [ − ]
𝑎 𝜋 2 (𝑛 − 𝑚)2 (𝑛 + 𝑚)2
1 2𝑎3 (𝑛2 + 𝑚2 + 2𝑛𝑚)(−1)𝑛−𝑚 (𝑛2 + 𝑚2 − 2𝑛𝑚)(−1)𝑛+𝑚
= [ − ]
𝑎 𝜋2 (𝑛 − 𝑚)2 (𝑛 + 𝑚)2
=0 =2(−1)𝑛+𝑚
1 2𝑎3 (𝑛2 + 𝑚2 ) ⏞
[(−1)𝑛−𝑚 − (−1)𝑛+𝑚 ] + (2𝑛𝑚) ⏞
[(−1)𝑛−𝑚 + (−1)𝑛+𝑚 ]
=
𝑎 𝜋2 (𝑛 − 𝑚)2 (𝑛 + 𝑚)2
[ ]
2|
2𝑎2 (4𝑛𝑚)(−1)𝑛+𝑚
⟨𝜓𝑛 |𝑥 𝜓𝑚 ⟩ = 2 [ ]
𝜋 (𝑛2 − 𝑚2 )2
2| 2|
2𝑎2 (8)(−1)3 16𝑎2
⟨𝜓1 |𝑥 𝜓2 ⟩ = ⟨𝜓2 |𝑥 𝜓1 ⟩ = 2 [ ]=−
𝜋 (1 − 4)2 9𝜋 2
Thus,
1 2 1 1 1 1
〈𝑥 2 〉 = [𝑎 [ − ] + 𝑎 2
[ − ]
2 3 2(𝜋)2 3 2(2𝜋)2
(𝐸1 − 𝐸2 ) ℏ2 𝜋 2 ℏ𝜋 2
= (1 − 4) = −3 ≡ −3𝜔
ℏ 2𝑚𝑎2 ℏ 2𝑚𝑎2
2〉
𝑎2 2 5 32
〈𝑥 = [ − − cos 3𝜔]
2 3 8(𝜋)2 9𝜋 2
Therefore
2
𝑎2 2 5 32 𝑎 32
𝜎𝑥2 = 〈𝑥 2〉
− 〈𝑥 〉2 = [ − − cos 3𝜔] − ( [1 − 2 cos 3𝜔])
2 3 8(𝜋)2 9𝜋 2 2 9𝜋
𝑎2 2 5 32 𝑎2 32 2 2
32
= [ − − cos 3𝜔] − (1 + ( ) cos 3𝜔 − 2 cos 3𝜔)
2 3 8(𝜋)2 9𝜋 2 4 9𝜋 2 9𝜋 2
𝑎2 1 5 32 2
𝜎𝑥2 = [ − 2
− ( 2 ) cos2 3𝜔] (𝑐)
4 3 4(𝜋) 9𝜋
2 2
1 𝑎 1 5 32
𝐿𝐻𝑆 ≡ 𝜎𝐻2 𝜎𝑥2 = (3ℏ𝜔)2 [ − 2
− ( 2 ) cos 2 3𝜔]
4 4 3 4(𝜋) 9𝜋
3 2 2
1 5 32 2
= ( ) (𝑎ℏ𝜔) [ − 2
− ( 2 ) cos2 3𝜔]
4 3 4(𝜋) 9𝜋
22
ℏ2 𝑑 〈𝑥 〉 ℏ2 8ℏ ℏ 8ℏ 2 2
𝑅𝐻𝑆 ≡ [ ] = [ sin 3𝜔] = ( ) sin 3𝜔
4 𝑑𝑡 4 3𝑚𝑎 2 3𝑚𝑎
2
ℏ2 𝑑〈𝑥 〉 8 2
𝑅𝐻𝑆 ≡ [ ] = ( 2 ) (ℏ𝜔𝑎)2 sin2 3𝜔
4 𝑑𝑡 3𝜋
?
3 2 1 5 32 2
8 2
( ) (𝑎ℏ𝜔)2 [ − ⏞(
− ( 2 ) cos 2 3𝜔] ≥ ) (ℏ𝜔𝑎 )2 sin2 3𝜔
4 3 4(𝜋) 2 9𝜋 3𝜋 2
21
?
1 5 32 2 4×8 2 2
[ − 2 ⏞(
− ( 2 ) cos 3𝜔] ≥ ) sin 3𝜔
3 4(𝜋)2 9𝜋 3 × 3𝜋 2
?
1 5 4×8 2 2 32 2 32 2
[ − ⏞(
]≥ 2
) sin 3𝜔 + ( 2 ) cos 3𝜔 = ( 2 )
3 4(𝜋)2 3 × 3𝜋 2 9𝜋 9𝜋
𝑌𝐸𝑆
1 5 32 2
[ − ⏞ (
] ≥ )
⏟3 4(𝜋)2 ⏟9𝜋 2
0.20668 0.12978
1
At the end of 1926, there was two different mathematical descriptions of quantum
mechanics. It was subsequently shown that the two different mathematical models
are equivalent because they may be transformed into one another.
Here the expansion coefficient is determined from the initial linear combination state
∞
Ψ(𝑥, 0) = ∑ 𝑐𝑛 𝜑𝑛 (𝑥 ) (∗)
𝑛=1
given by
∞
We can understand this situation a little differently: We take into account the fact that
the eigenfunctions {𝜑𝑛 } represent an orthonormal basis of the vector space ℋ (a
Hilbert space), like the 3 unit vectors {𝑒̂𝑖 } in the space of 3 dimensional vectors (ℝ3 ).
3
⃗ = ∑ 𝑽𝒊 𝑒̂𝑖
𝑉 Ψ = ∑ 𝒄𝒊 𝜑𝑖 (𝑥)
𝑖=𝑥,𝑦,𝑧 𝑖
Projection
∞
⃗ ∙ 𝑒̂𝑖
𝑽𝒊 = 𝑉 𝒄𝒊 = ∫ 𝜑𝑖∗ (𝑥 )Ψ(𝑥 )𝑑𝑥
−∞
𝒄𝒊 = ⟨𝜑𝑖 |Ψ⟩
4
𝑽𝒛 = ⃗𝑽 ∙ 𝑒̂𝑧
⃗ or 𝑽𝒙 , 𝑽𝒚 , 𝑽𝒛 , we can calculate 𝑉
it makes no difference whether we specify 𝑽 ⃗
uniquely from 𝑉𝑥 , 𝑉𝑦 , 𝑉𝑧 and vice versa if the unit vectors are known.
The situation described in (*) and (**) is quite analogous by simple replacements
⃗ → 𝜳(𝒙)
𝑉
{𝑒̂𝑖 } → {𝜑𝑛 }
𝑉𝑖 → 𝒄𝒊
For example, when {𝜑𝑖 } is known, 𝒄𝒊 can be determined uniquely, if 𝜳(𝒙) is given,
and vice versa:
∞
⟨𝜓|𝜙⟩ = ∑ 𝑐𝑖∗ 𝑑𝑖 (3 ∗)
𝑖,𝑗
We now show that this is exactly the same result when we take the scalar
product of the two coordinate vectors 𝒄 and 𝒅.
However when we try to multiply two vectors when they are both viewed as column
matrices
𝑐1 𝑑1
𝒄 ∙ 𝒅 = (𝑐2 ) ∙ (𝑑2 ) (4 ∗)
⋮ ⋮
𝒄 ∙ 𝒅 = (𝑛 × 1) ∙ (𝑛 × 1): 𝑁𝑜𝑡 𝑑𝑒𝑓𝑖𝑛𝑒𝑑
we cannot multiply them.
To rectify this problem, we can take the transpose of the first vector, turning it into
a 1 × 𝑛 row matrix. With this change, the product is well defined. Therefore, instead
of (4*), we now have
𝑑1
†
𝑐 ∙𝑑 = (𝑐1∗ 𝑐2∗ …) ∙ (𝑑2 ) = 𝑐1∗ 𝑑1 + 𝑐2∗ 𝑑2 + ⋯ = ∑ 𝑐𝑖∗ 𝑑𝑖
⏟ ⋮ 𝑖,𝑗
(1×𝑛)∙(𝑛×1)
𝑐 † ∙ 𝑑 = ∑ 𝑐𝑖∗ 𝑑𝑖
𝑖,𝑗
we see that
∞
The expression ⟨𝝍|𝝓⟩ = ∫−∞ 𝝍∗ (𝒙)𝝓(𝒙)𝒅𝒙 is clearly a scalar product.
6
We conclude that
Any function Ψ in ℋ
Ψ = ∑ 𝒄𝒊 𝜑𝑖 (𝑥)
𝑖
The first step is to expand the wave function in terms of the eigenfunctions {𝜑𝑖 } of
the Hamiltonian (or any other basis system in ℋ):
∗ ( )
Multiply by 𝜑𝑚 𝑥 ≡ ⟨𝜑𝑚 |
̂ |𝝋𝒏 ⟩ = 𝑞 ∑ 𝑐𝑛 ⏟
∑ 𝑐𝑛 ⟨𝝋𝒎 |𝑸 ⟨𝜑𝑚 |𝜑𝑛 ⟩
𝑛 𝑛 𝛿𝑛𝑚
Thus
This is a matrix equation: LHS is the product of a matrix ℚ with a column vector 𝒄
and RHS is also a column vector 𝑞𝒄.
Now, this is a matrix representation, in the basis {𝜑𝑛 }, of the eigenvalue equation we
started with
𝑄̂ |𝜓⟩ = 𝑞|𝜓⟩ →
⏟ ℚ𝒄=𝑞𝒄
𝑖𝑠
𝑟𝑒𝑝𝑟𝑒𝑠𝑒𝑛𝑡𝑒𝑑
𝑏𝑦
⟨𝜑1 |𝜓⟩
⟨𝜑 |𝜓⟩
|𝜓⟩ ⟶
⏟ 𝒄=( 2 )
𝑖𝑠 ⟨𝜑3 |𝜓⟩
𝑟𝑒𝑝𝑟𝑒𝑠𝑒𝑛𝑡𝑒𝑑 ⋮
𝑏𝑦
8
• The set of numbers 𝑐𝑛 = ⟨𝜑𝑛 |𝜓⟩ constitute a complete description of the state |𝜓⟩
• The set of numbers 𝑄𝑚𝑛 ≡ ⟨𝜑𝑚 |𝑄̂ |𝜑𝑛 ⟩ constitute a complete description of
Now consider a Hermitian operator 𝐺̂ and suppose that {𝜑𝑛 } is the eigenfunctions of 𝐺̂
with the eigenvalues 𝑔𝑛 :
𝐺̂ |𝜑𝑛 ⟩ = 𝑔𝑛 |𝜑𝑛 ⟩
In basis {𝜑𝑛 }, the matrix representation of 𝐺̂ is obtained by
𝐺𝑚𝑛 ≡ ⟨𝜑𝑚 |𝐺̂ |𝜑𝑛 ⟩ = 𝑔𝑛 ⟨𝜑𝑚 |𝜑𝑛 ⟩ = 𝑔𝑛 𝛿𝑚𝑛
Then
𝐺11 𝐺12 𝐺13 ⋯ 𝑔1 0 0 ⋯
𝐺23 ⋯ 0 𝑔2
𝐺̂ → 𝔾 = (𝐺21 𝐺22 )=( 0 ⋯)
𝐺31 𝐺32 𝐺33 ⋯ 0 0 𝑔3 ⋯
⋮ ⋮ ⋮ ⋱ ⋮ ⋮ ⋮ ⋱
Thus, the matrix of a Hermitian operator in a basis of the eigenfunctions of that operator
is diagonal.
How to represent eigenfunctions of a diagonal matrix?
In basis {𝜑𝑛 }; the column vector representation of the eigenfunctions are
(𝑛)
|𝜑𝑛 ⟩ = 𝕀|𝜑𝑛 ⟩ = ∑ |𝜑𝑚 ⟩ ⏟
⟨𝜑𝑚 |𝜑𝑛 ⟩ => |𝜑𝑛 ⟩ = ∑ 𝑐𝑚 |𝜑𝑚 ⟩
𝑚 (𝑛) 𝑚
𝑐𝑚
where
(𝑛)
𝑐𝑚 ≡ ⟨𝜑𝑚 |𝜑𝑛 ⟩ = 𝛿𝑚𝑛 ≠ 0 𝑖𝑓 𝑚 = 𝑛
9
(1) (2)
𝑐1 ⟨𝜑1 |𝜑1 ⟩ 1 𝑐1 ⟨𝜑1 |𝜑2 ⟩ 0
(1) (2)
⟨𝜑 |𝜑 ⟩ ⟨𝜑 |𝜑 ⟩
|𝜑1 ⟩ ⟶ 𝑐2 ≡ ( 2 1 ) = (0) , |𝜑2 ⟩ ⟶ 𝑐2 ≡ ( 2 2 ) = (1) ,
(1) ⟨𝜑3 |𝜑1 ⟩ 0 (2) ⟨𝜑3 |𝜑2 ⟩ 0
𝑐3 ⋮ 𝑐3 ⋮
⋮ ⋮
( ⋮ ) ( ⋮ )
𝑒𝑡𝑐
The matrix representation of the eigenvector |𝜑𝑛 ⟩ is a column vector with a single
nonzero unit entry in the n.th slot.
The Energy Representation
We learned that the matrix representing an operator is always diagonal in its own
In general,
𝐻|𝜓𝑛 〉 = 𝐸𝑛 |𝜓𝑛 〉
𝐸1 0 0 ⋯ 1 0
̂ →ℍ=(0
𝐻
𝐸2 0 ⋯ ) , |𝜓 〉 → (0) , |𝜓 〉 → (1) , 𝑒𝑡𝑐.
1 2
0 0 𝐸3 ⋯ 0 0
⋮ ⋮ ⋮ ⋱ ⋮ ⋮
10
If 𝑈 is unitary , 𝑈 † 𝑈 = 𝑈 𝑈 † = 𝐼
( 𝑈† 1𝑈)𝑛𝑞 = 𝛿𝑛𝑞 = ⟨𝜑𝑛 | 𝑈 † 1𝑈|𝜑𝑞 ⟩ = ∑⟨𝜑𝑛 | 𝑈 † |𝜑𝑝 ⟩⟨𝜑𝑝 |𝑈|𝜑𝑞 ⟩
𝑝
𝐸1 0 0 ⋯ 1/2 0 0 ⋯
̂ → ℍ = ( 0 𝐸2 0 ⋯ ) = ℏ𝜔 ( 0
𝐻
3/2 0 ⋯ )
0 0 𝐸3 ⋯ 0 0 5/2 ⋯
⋮ ⋮ ⋮ ⋱ ⋮ ⋮ ⋮ ⋱
To simplify the notation we instead write
𝐻|𝑛〉 = 𝐸𝑛 |𝑛〉
Then, the matrix elements of the position operator are
1 1
𝑥𝑛𝑘 = 〈𝑛|𝑥̂|𝑘〉 = 〈𝑛|(𝑎̂ − + 𝑎̂ + )|𝑘〉 = [〈𝑛|𝑎̂ − |𝑘〉 + 〈𝑛|𝑎̂+ |𝑘〉]
√2𝛽 √2𝛽
1 − + ]
𝑥𝑛𝑘 = [𝑎̂𝑛𝑘 + 𝑎𝑛𝑘
√2𝛽
1 − + ]
𝑥𝑛𝑘 = [𝑎̂𝑛𝑘 + 𝑎𝑛𝑘
√2𝛽
This gives the matrix
0 √1
0 0 0 ⋯ 0 0 0 0 0 ⋯
1 0 0
√2 0 0 ⋯ √1 0 0 0 0 ⋯
𝑥̂ ⟶ 0 0 √3 0 ⋯ + 0
0 √2 0 0 0 ⋯
√2𝛽
0 0
0 0 √4 ⋱ 0 0 √3 0 0 ⋱
[( ⋮ ⋮ ⋮ ) ( ⋮ ⋮ ⋮ )]
0 √1 0 0 0 ⋯
√1 0 √2 0 0 ⋯
1
𝑥̂ ⟶ 0 √2 0 √3 0
√2𝛽 0 0 √3 0 ⋯
√4
0 0 0 √4 0 ⋱
( ⋮ ⋮ ⋮ )
Similarly, for the momentum operator,
1
𝑝̂ = (𝑎̂ − − 𝑎̂+ ) ,
𝑖√2𝛽
1 1
𝑝𝑛𝑘 = 〈𝑛|𝑝̂ |𝑘〉 = 〈𝑛|(𝑎̂ − − 𝑎̂+ )|𝑘〉 = [〈𝑛|𝑎̂|𝑘〉 − 〈𝑛|𝑎̂† |𝑘〉]
𝑖√2𝛽 𝑖√2𝛽
1 − + ]
𝑝𝑛𝑘 = [𝑎̂𝑛𝑘 − 𝑎𝑛𝑘
𝑖√2𝛽
14
0 √1 0 0 0 ⋯ 0 0 0 0 0 ⋯
1 0 0 √2 0 0 ⋯ √1 0 0 0 0 ⋯
𝑝̂ ⟶ 0 √3 0 ⋯ − 0 0 0 0 ⋯
𝑖√2𝛽 0 0 √2
0 0 0 0 √4 ⋱ 0 0 √3 0 0 ⋱
[( ⋮ ⋮ ⋮ ) ( ⋮ ⋮ ⋮ )]
0 −𝑖√1 0 ⋯
0 0
𝑖√1 0 −𝑖√2 0 0 ⋯
1
𝑝̂ ⟶ 0 𝑖√2 0 −𝑖√3 0
√2𝛽 0 0 −𝑖√3 0 −𝑖√4 ⋯
0 0 0 −𝑖√4 0 ⋱
( ⋮ ⋮ ⋮ )
𝑝̂ 2 1
𝐻= + 𝑚𝜔2 𝑥 2
2𝑚 2
1
0 0 ⋯
2
3
= ℏ𝜔 0 2
0 ⋯ ≡ℍ
5 ⋯
0 0
2 ⋱
(⋮ ⋮
⋮ )
15
ℎ 𝑔 𝐸 0 ℎ−𝐸 𝑔
𝑑𝑒𝑡 [( )−( )] = 𝑑𝑒𝑡 ( ) = 0 =>
𝑔 ℎ 0 𝐸 𝑔 ℎ−𝐸
(ℎ − 𝐸 )2 − 𝑔2 = 0 => ℎ − 𝐸 = ±𝑔 => 𝐸± = ℎ ± 𝑔
16
ℎ𝛼 + 𝑔𝛽 = (ℎ + 𝑔)𝛼 => 𝛼 = 𝛽
𝑔𝛼 + ℎ𝛽 = (ℎ + 𝑔)𝛽 => 𝛼 = 𝛽
1
|ℊ+ ⟩ = 𝛼 ( )
1
Normalization=>
1
1 = ⟨ℊ+ |ℊ+ ⟩ = 𝛼 2 (1 1) ( ) = 2𝛼 2 => 𝛼 = 1/√2
1
1 1
|ℊ+ ⟩ = ( ) : Normalized eigenvector of 𝐻 with 𝐸 = 𝐸+ = ℎ + 𝑔
√2 1
𝐸 = 𝐸− = ℎ − 𝑔
ℎ𝛼 + 𝑔𝛽 = (ℎ − 𝑔)𝛼 => 𝛼 = −𝛽
𝑔𝛼 + ℎ𝛽 = (ℎ − 𝑔)𝛽 => 𝛼 = −𝛽
1
|ℊ− ⟩ = 𝛼 ( )
−1
Normalization=>
1
1 = ⟨ℊ− |ℊ− ⟩ = 𝛼 2 (1 −1) ( ) = 2𝛼 2 => 𝛼 = 1/√2
−1
1 1
|ℊ− ⟩ = ( ) : Normalized eigenvector of 𝐻 with 𝐸 = 𝐸− = ℎ − 𝑔
√2 −1
17
1 1 1 1
|ℊ+ ⟩ = ( ) , |ℊ− ⟩ = ( )
√2 1 √2 −1
Therefore, we expand the initial state |ℊ(0)⟩ as a combination of |ℊ+ ⟩ and |ℊ− ⟩
1
|ℊ(0)⟩ = ( ) = 𝑐+ |ℊ+ ⟩ + 𝑐− |ℊ− ⟩
0
1 1 1
𝑐+ = ⟨ℊ+ |ℊ(0)⟩ = (1 1) ( ) => 𝑐+ =
√2 0 √2
1 1 1
𝑐− = ⟨ℊ− |ℊ(0)⟩ = (1 −1) ( ) => 𝑐− =
√2 0 √2
1 1
|ℊ(0)⟩ = ( ) = (|ℊ+ ⟩ + |ℊ− ⟩)
0 √2
1
|𝒢(t)⟩ = ∑ 𝑐𝑛 |ℊ𝑛 (0)⟩𝑒 −𝑖𝐸𝑛 𝑡/ℏ = (|ℊ+ ⟩𝑒 −𝑖𝐸+𝑡/ℏ + |ℊ− ⟩𝑒 −𝑖𝐸−𝑡/ℏ )
𝑛=+,−
√2
1 1 1 −𝑖(ℎ+𝑔)𝑡/ℏ 1 1
= (
( )𝑒 + ( ) 𝑒 −𝑖(ℎ−𝑔)𝑡/ℏ )
√2 √2 1 √2 −1
1 1 −𝑖(ℎ+𝑔)𝑡/ℏ 1
= (( ) 𝑒 + ( ) 𝑒 −𝑖(ℎ−𝑔)𝑡/ℏ )
2 1 −1
1 −𝑖ℎ𝑡/ℏ 1 −𝑖𝑔𝑡/ℏ 1
= 𝑒 (( ) 𝑒 + ( ) 𝑒 𝑖𝑔𝑡/ℏ )
2 1 −1
ℎ cos(𝑔𝑡/ℏ) − 𝑖𝑔 sin(𝑔𝑡/ℏ) ⋎
−𝑖ℎ𝑡/ℏ ℎ 𝑔 −𝑖ℎ𝑡/ℏ cos(𝑔𝑡/ℏ)
𝑒 ( 𝑔𝑡 )=
⏞( )𝑒 ( )
−𝑖ℎ sin ( ) + 𝑔 cos(𝑔𝑡/ℏ) 𝑔 ℎ −𝑖 sin(𝑔𝑡/ℏ)
ℏ
19
Projection Operators
𝐴̂ 𝐵̂|𝜓⟩⟨𝜙|𝐶̂ |χ⟩
The operators 𝐵̂ and 𝐴̂ act successively on the vector |𝜓⟩ multiplied by the scalar
⟨𝜙|𝐶̂ |χ⟩
𝐴̂ 𝐵̂|𝜓⟩ ⟨𝜙|𝐶
⏟ ⏟ ̂ |χ⟩ = 𝑎 𝑣𝑒𝑐𝑡𝑜𝑟
𝑎 𝑣𝑒𝑐𝑡𝑜𝑟 𝑎 𝑠𝑐𝑎𝑙𝑎𝑟
⏟̂ 𝐵̂|𝜓⟩⟨𝜙|𝐶̂ )
(𝐴 ⏟⟩
|𝜒 = 𝑎 𝑣𝑒𝑐𝑡𝑜𝑟
𝑎𝑛 𝑜𝑝𝑒𝑟𝑎𝑡𝑜𝑟 𝑎 𝑣𝑒𝑐𝑡𝑜𝑟
In this case, we see that we can construct operators with the combination of a ket and
̂ ≡ |𝜓⟩⟨𝜙| ,
operator ∶ 𝐷 ̂ † ≡ |𝜙⟩⟨𝜓|
adjoint operator: 𝐷
𝑃̂ ≡ |𝜓⟩⟨𝜓|
20
With the property that it picks out the portion of any other vector |𝛽 ⟩ that lies along|𝜓⟩:
since if we let this operator act on any function |𝜓⟩ we recover the expansion of |𝜓⟩
=>
|𝝍⟩⟩ = ∑ 𝑐𝑛 |𝜑𝑛 ⟩
𝑛
21
𝑃̂ ≡ |𝜓⟩⟨𝜓|
Let it acts on the state |𝛽 ⟩
𝑃̂ |𝛽 ⟩ = |𝜓⟩⟨𝜓| 𝛽 ⟩
𝜆 = 0: =>
𝑃̂ |𝜂 ⟩0 = 0
Any vector orthogonal to |𝜂 ⟩1 is an eigenvector of 𝑃̂ with the eigenvalue 𝜆 = 0.
22
̂ is unitary ,
If 𝑈
̂†𝑈
𝑈 ̂=𝑈
̂𝑈̂ † = 𝕀̂
Matrix elements:
̂ † 𝕀̂𝑈
( 𝑈 † 𝑈)𝑛𝑚 = 𝛿𝑛𝑚 = ⟨𝜑𝑛 | 𝑈 ̂|𝜑𝑚 ⟩ = ∑⟨𝜑𝑛 | 𝑈
̂ † |𝜑𝑝 ⟩⟨𝜑𝑝 |𝑈
̂|𝜑𝑚 ⟩
𝑝
With unitary operators we can define unitary transformations of states and operators.
Common notations are
̂ |𝜓⟩ and
|𝜓⟩ → |𝜓 ′ ⟩ = 𝑈 𝐴̂ → 𝐴̂′ = 𝑈
̂ 𝐴̂ 𝑈
̂†
The interesting thing about the unitary transformations is that they leave important
properties and quantities unchanged. In this respect, a unitary transformation is an
analogue of the rotation in elementary vector calculus. One can obtain a new
orthogonal basis in 3-space through a rotation of axes about the origin. In this new
basis,
We already know that a given Hilbert space has many basis. The transformation from
one basis to another is a unitary transformation. It has all the properties listed above
that a rigid rotation in 3-space has.
24
Transformation of Basis:
Let {𝜑𝑛 } and {𝑓𝑛 } are two basis :
⟨𝜑𝑛 |𝜑𝑘 ⟩ = 𝛿𝑛𝑘 and ∑|𝜑𝑛 ⟩⟨𝜑𝑛 | = 𝕀̂
𝑛
and
⟨𝑓𝑛 |𝑓𝑘 ⟩ = 𝛿𝑛𝑘 and ∑|𝑓𝑛 ⟩⟨𝑓𝑛 | = 𝕀̂
𝑛
Now, write
|𝑓𝑛 ⟩ = 𝕀̂|𝑓𝑛 ⟩ = ∑|𝜑𝑝 ⟩ ⟨𝜑
⏟ 𝑝 |𝑓𝑛 ⟩
𝑝 ∗
⟨𝑓𝑛 |𝜑𝑝 ⟩
̂
with the matrix elements of a operator 𝑈
∗
𝑈𝑛𝑝 ≡ ⟨𝑓𝑛 |𝜑𝑝 ⟩ and then 𝑈𝑛𝑝 ≡ ⟨𝜑𝑝 |𝑓𝑛 ⟩
Then
∗
|𝑓𝑛 ⟩ = ∑ 𝑈𝑛𝑝 |𝜑𝑝 ⟩ ∶ Transformation of basis
𝑝
̂ is unitary, i.e., 𝑈
Let us prove that 𝑈 ̂†𝑈
̂=𝑈
̂𝑈̂ † = 𝕀̂
∗
⟨𝑓𝑖 |𝑓𝑗 ⟩ = 𝛿𝑖𝑗 ⟹ ⟨∑ 𝑈𝑖𝑘 𝜑𝑘 | ∑ 𝑈𝑗𝑙∗ 𝜑𝑙 ⟩ = 𝛿𝑖𝑗 ⟹ ∑ 𝑈𝑖𝑘 𝑈𝑗𝑙∗ ⟨𝜑𝑘 |𝜑𝑙 ⟩ = 𝛿𝑖𝑗
𝑘 𝑙 𝑘
∗
̃𝑘𝑗 ̂𝑈̂ † = 𝕀̂ . Therefore U is a unitary matrix.
⟹ ∑ 𝑈𝑖𝑘 𝑈 = 𝛿𝑖𝑗 𝑂𝑅 𝑈
𝑘
̂.
{𝝋𝒏 } and {𝒇𝒏 } are related through the unitary transformation 𝑼
25
∗
|𝑓𝑛 ⟩ = ∑ 𝑈𝑛𝑝 |𝜑𝑝 ⟩
𝑝
Then
𝜓𝑛′ = ∑ 𝑈𝑛𝑝 𝜓𝑝
𝑝
This is the matrix representation of the equation
̂ |𝜓⟩
|𝜓 ′ ⟩ = 𝑈
That is,
⏟ ̂ |𝜓⟩ ⟶
|𝜓 ′ ⟩ = 𝑈 𝜓𝑛′ = ∑ 𝑈𝑛𝑝 𝜓𝑝
𝑇ℎ𝑒 𝑒𝑞𝑢𝑎𝑡𝑖𝑜𝑛 ⏟ 𝑝
𝑖𝑡𝑠 𝑚𝑎𝑡𝑟𝑖𝑥 𝑟𝑒𝑝𝑟𝑒𝑠𝑒𝑛𝑡𝑎𝑡𝑖𝑜𝑛
with 𝑈𝑛𝑝 ≡ ⟨𝑓𝑛 |𝜑𝑝 ⟩
This eq. says how an arbitrary state vector transforms under a change of basis.
26
Problem 11.12 (Liboff): Show that the scalar product ⟨𝜓|𝜙⟩ is preserved under a
unitary transformation.
Let |𝜓⟩ and |𝜙⟩ be two arbitrary vectors in Hilbert space. Under the transformation
̂ of basis
𝑈
̂
𝑈
{𝜑𝑛 } ⟶
⏞ {𝑓𝑛 }
State vectors transform
̂ |𝜓⟩
|𝜓⟩ ⟶ |𝜓 ′ ⟩ = 𝑈
̂ |𝜙⟩
|𝜙⟩ ⟶ |𝜙 ′ ⟩ = 𝑈
How the scalar product ⟨𝜓|𝜙⟩ transform?
̂ 𝜓|𝑈
⟨𝜓|𝜙⟩ ⟶ ⟨𝜓 ′ |𝜙 ′ ⟩ = ⟨𝑈 ̂ 𝜙⟩ = ⟨𝜓| 𝑈
̂†𝑈
̂ 𝜙⟩ = ⟨𝜓|𝕀̂𝜙⟩ = ⟨𝜓|𝜙⟩
𝐹̂ ′ =?
From
|𝜓 ′ ⟩ = 𝐹̂ ′ |𝜙 ′ ⟩ ⟹ 𝑈
̂ |𝜓⟩ = 𝐹̂ ′ 𝑈
̂ |𝜙⟩
̂ −1 :
Multiply from the left by 𝑈
̂ −1 𝑈
⟹ ⏟
𝑈 ̂ |𝜓⟩ = 𝑈
̂ −1 𝐹̂ ′ 𝑈 |𝜙⟩ ̂ −1 𝐹̂ ′ 𝑈
⟹ |𝜓⟩ = 𝑈 ̂ |𝜙⟩
𝕀̂
𝐹̂ = 𝑈
̂ −1 𝐹̂ ′ 𝑈
̂ 𝑂𝑅 𝐹̂ ′ = 𝑈
̂ 𝐹̂ 𝑈
̂ −1 ∶ Unitary-similarity Transformation
The more general case of transformation:
𝐴̂ ⟶ 𝐴̂′ = 𝑆̂𝐴̂ 𝑆̂ −1 𝑠𝑢𝑐ℎ 𝑡ℎ𝑎𝑡 𝑆̂ −1 ≠ 𝑆̂ † .Then 𝑆̂ is similarity transformation
28
----------------------------------------------------------------------------------------
A Hermitian matrix is always diagonalizable by a unitary-similarity transformation.
Let 𝐴̂ be a Hermitian 𝑛 × 𝑛 matrix. Let the column vectors of the 𝑛 × 𝑛 matrix 𝑆̂ be
comprised of the orthonormalized eigenvectors of 𝐴̂. Then
𝑆̂ is unitary.
𝑆̂ −1 𝐴̂𝑆̂ is a diagonal matrix comprised of the eigenvalues of 𝐴̂.
EX: Consider the matrix
0 0 1
𝐴 = (0 0 0)
1 0 0
Find the eigenvalues and corresponding eigenvectors.
(𝐴 − 𝜆𝕀̂)𝜓 = 0 => 𝑑𝑒𝑡(𝐴 − 𝜆𝕀̂) = 0 => 𝜆1 = 0, 𝜆2 = 1, 𝜆3 = −1
𝑎
𝜆1 = 0: 𝜓1 = (𝑏 )
𝑐
0 0 1 𝑎
(0 0 0) (𝑏 ) = 0 => 𝑐 = 0, 𝑎 = 0
1 0 0 𝑐
0
𝜓1 = (1)
0
𝑎
𝜆2 = 1: 𝜓2 = (𝑏 )
𝑐
0 0 1 𝑎 𝑎
(0 0 0) (𝑏 ) = (𝑏 ) => 𝑐 = 𝑎, 𝑏 = 0
1 0 0 𝑐 𝑐
11
𝜓2 = ( 0)
√2 1
29
𝑎
𝜆3 = −1: 𝜓3 = (𝑏 )
𝑐
0 0 1 𝑎 𝑎
(0 0 0) (𝑏 ) = − (𝑏 ) => 𝑐 = −𝑎, 𝑏 = 0
1 0 0 𝑐 𝑐
−11
𝜓3 = (0)
√2 1
Then, construct the matrix whose columns are comprised of orthonormalized
eigenvectors of 𝐴̂:
0 1 0 0 1/√2 1/√2 0 0 0
= ( 1/√2 0 1/√2) (0 0 0 ) = (0 1 0)
−1/√2 0 1/√2 0 1/√2 −1/√2 0 0 −1
It is a diagonal matrix whose diagonal elements are eigenvalues (0,1,-1) of A. Thus, the
unitary matrix 𝕊 diagonalizes the matrix A.
30
Invariance of Eigenvalues
Since the eigenvalues of an operator corresponding to an observable are physically
measurable quantities, one doesn’t expect these values to be affected by a
transformation of basis in Hilbert space. For example, the eigenenergies of SHO are
𝐸𝑛 = ℏ𝜔(𝑛 + 1/2)
in all representations. In a similar vein, the eigenvalues of a Hermitian operator must
be real. It follows that
(1) The eigenvalues of a Hermitian operator are preserved under a unitary-
similarity transformation.
(2) The Hermiticity of an operator is preserved under a unitary-similarity
transformation.
Let
In {𝜑𝑛 } basis:
𝐴̂|𝜓⟩ = 𝑎 |𝜓⟩
In {𝑓𝑛 } basis:
𝐴̂† → 𝐴̂†′ = (𝑈 ̂ −1 )† = (𝑈
̂ 𝐴̂𝑈 ̂ −1 )† 𝐴̂† 𝑈
̂†
̂ is unitary, 𝑈
Since 𝑈 ̂ −1 = 𝑈 ̂ † )† = 𝑈
̂ † and ( 𝑈 ̂. Thus,
𝐴̂† → 𝐴̂†′ = 𝑈
̂ 𝐴̂𝑈
̂ −1 = 𝐴̂
32
matrix
0 0 0
ℍ = ℏ𝜔 (0 1 0)
0 0 −1
What values will we obtain when measuring the energy and with what probabilities?
------------------------------------------------------------------------------------------------
𝐸1 = 0, 𝐸2 = ℏ𝜔, 𝐸3 = −ℏ𝜔
Which of these possible results of the energy values will appear after the
Case1: Measure the energy in one of the eigenstates of H, |𝜓1 ⟩, |𝜓2 ⟩, |𝜓3 ⟩.
1 0 0 0 1 1
𝐸1 = 0 in state |𝜓1 ⟩ = (0) with 𝑃 = 100%, since ℏ𝜔 (0 1 0 ) (0) = 0 (0)
0 0 0 −1 0 0
0 0 0 0 0 0
𝐸2 = ℏ𝜔 in state |𝜓2 ⟩ = (1) 𝑤𝑖𝑡ℎ 𝑃 = 100%, since ℏ𝜔 (0 1 0 ) (1) = ℏ𝜔 (1)
0 0 0 −1 0 0
0 0 0 0 0 0
𝐸3 = −ℏ𝜔 in state |𝜓3 ⟩ = (0) 𝑤𝑖𝑡ℎ 𝑃 = 100%, since ℏ𝜔 (0 1 0 ) (0) = −ℏ𝜔 (0)
1 0 0 −1 1 1
33
1 1−𝑖
|𝜙⟩ = (1 − 𝑖 )
√5 1
Since the set {|𝜓1 ⟩, |𝜓2 ⟩, |𝜓3 ⟩} forms an orthonormal system, we write |𝜙⟩ in terms
of them,
1−𝑖 1 1−𝑖 0 1 0
|𝜙⟩ = 𝑐1 |𝜓1 ⟩ + 𝑐2 |𝜓2 ⟩ + 𝑐3 |𝜓3 ⟩ = (0 ) + ( 1) + (0 )
√5 0 √5 0 √5 1
2
1−𝑖 2
𝑃(𝐸 = 𝐸1 ) = |𝑐1 |2 = | | =
√5 5
2
1−𝑖 2
𝑃(𝐸 = 𝐸2 ) = |𝑐2 |2 =| | =
√5 5
2
1 1
𝑃(𝐸 = 𝐸3 ) = |𝑐3 |2 =| | =
√5 5
34
1 −1 0
ℍ = 𝜀 (−1 1 0)
0 0 −1
a) What values will we obtain when measuring the energy and with what
probabilities?
𝜀−𝐸 −𝜀 0
(ℍ − 𝐸𝕀̂)𝜓 = 0 => 𝑑𝑒𝑡(ℍ − 𝐸𝕀̂) = 0 => 𝑑𝑒𝑡 ( −𝜀 𝜀−𝐸 0 )=0
0 0 −(𝜀 + 𝐸)
−(𝜀 − 𝐸 )(𝜀 − 𝐸 )(𝜀 + 𝐸 ) + 𝜀 2 (𝜀 + 𝐸 ) = 0 => (𝜀 + 𝐸 )[𝜀 2 − (𝜀 − 𝐸 )2 ] = 0
𝐸1 = 0 , 𝐸2 = −𝜀 , 𝐸3 = 2𝜀
The corresponding energy eigenstates:
𝑎
𝐸1 = 0: 𝜓1 = (𝑏 )
𝑐
1 −1 0 𝑎
𝜀 (−1 1 0 ) (𝑏 ) = 0 => 𝑎 − 𝑏 = 0, 𝑏 − 𝑎 = 0, 𝑐 = 0
0 0 −1 𝑐
11
𝜓1 = (1 )
√2 0
𝑎
𝐸2 = −𝜀: 𝜓2 = (𝑏 )
𝑐
1 −1 0 𝑎 𝑎
𝜀 (−1 1 0 ) (𝑏 ) = −𝜀 (𝑏 ) => 𝑎 − 𝑏 = −𝑎, 𝑏 − 𝑎 = −𝑏, 𝑐 = 𝑐
0 0 −1 𝑐 𝑐
35
0
𝜓 2 = (0 )
1
𝑎
𝐸3 = 2𝜀: 𝜓3 = (𝑏 )
𝑐
1 −1 0 𝑎 𝑎
𝜀 (−1 1 0 ) (𝑏 ) = 2𝜀 (𝑏 ) => 𝑎 − 𝑏 = 2𝑎, 𝑏 − 𝑎 = 2𝑏, 𝑐 = −2𝜀𝑐
0 0 −1 𝑐 𝑐
1 −1
𝜓3 = (1)
√2 0
1 1 1 −1 0 1 1 1 1
𝐸1 = 0 in state |𝜓1 ⟩ = (1) with 𝑃 = 100%, since 𝜀 (−1 1 0) (1 ) = 0 (1)
√2 0 0 0 −1 √2 0 √2 0
0 1 −1 0 0 0
𝐸2 = −𝜀 in state |𝜓2 ⟩ = (0) with 𝑃 = 100%, since 𝜀 (−1 1 0 ) (0) = −𝜀 (0)
1 0 0 −1 1 1
1 −1 1 −1 0 1 −1 1 −1
𝐸3 = 2𝜀 in state |𝜓3 ⟩ = ( 1 ) with 𝑃 = 100%, since 𝜀 (−1 1 0) ( 1 ) = 2𝜀 (1)
√2 0 0 0 −1 √2 0 √2 0
b) Now consider another observable of the same system, 𝐴, represented by the matrix
0 4 0
𝔸 = 𝛼 (4 0 1)
0 1 0
Suppose that when we measure the energy, we obtain a value of −𝜀. Immediately
afterwards, we measure 𝐴. What values will we obtain for A and with what
probabilities?
36
𝐴̂𝜙 = 𝜆𝜙
−𝜆 4𝛼 0
(𝔸 − 𝜆𝕀̂)𝜙 = 0 => 𝑑𝑒𝑡(𝔸 − 𝜆𝕀̂) = 0 => 𝑑𝑒𝑡 ( 4𝛼 −𝜆 𝛼𝛼 ) = 0
0 𝛼 −𝜆
𝑎
𝜆1 = 0: 𝜙1 = (𝑏 )
𝑐
0 4 0 𝑎
𝛼 (4 0 1) (𝑏 ) = 0 => 4𝛼𝑏 = 0, 4𝛼𝑎 + 𝛼𝑐 = 0, 𝛼𝑏 = 0
0 1 0 𝑐
1 1
𝜙1 = (0)
√17 −4
𝑎
𝜆2 = √17𝛼: 𝜙2 = (𝑏 )
𝑐
0 4 0 𝑎 𝑎
𝛼 (4 0 1) (𝑏 ) = √17𝛼 (𝑏 ) => 4𝑏 = √17𝑎, 4𝑎 + 𝑐 = √17𝑏, 𝑏 = √17𝑐
0 1 0 𝑐 𝑐
1 4
𝜙2 = (√17)
√34 1
37
𝑎
𝜆3 = −√17𝛼: 𝜙3 = (𝑏 )
𝑐
0 4 0 𝑎 𝑎
𝛼 (4 0 1) (𝑏 ) = −√17𝛼 (𝑏 ) => 4𝑏 = −√17𝑎, 4𝑎 + 𝑐 = −√17𝑏, 𝑏 = −√17𝑐
0 1 0 𝑐 𝑐
1 4
𝜙3 = (−√17)
√34 1
The set {|𝜙1 ⟩, |𝜙2 ⟩, |𝜙3 ⟩} forms an orthonormal system. (Show this)They are the
determinate states of A.
If a measurement of the energy yields 𝐸2 = −𝜀, this means that the system is left in
the state 𝜓2 . The state 𝜓2 is not a determinate state for the observable A. However,
we can expand it in terms of the determinate states of A:
|𝜓2 ⟩ = 𝑐1 |𝜙1 ⟩ + 𝑐2 |𝜙2 ⟩ + 𝑐3 |𝜙3 ⟩
1 0 1 −4
𝑐1 = ⟨𝜙1 |𝜓2 ⟩ = (1 0 −4) (0) = (−4) => 𝑐1 =
√17 1 √17 √17
1 0 1 1
𝑐2 = ⟨𝜙2 |𝜓2 ⟩ = (4 √17 1) (0) = (1) => 𝑐2 =
√34 1 √34 √34
1 0 1 1
𝑐3 = ⟨𝜙3 |𝜓2 ⟩ = (4 −√17 1) (0) = (1) => 𝑐3 =
√34 1 √34 √34
−4 2 16
𝑃(𝐴 = 0 in state 𝜓2 ) = |𝑐1 |2 =| | =
√17 17
2
1 1
𝑃(𝐴 = √17𝛼 in state 𝜓2 ) = |𝑐2 |2 =| | =
√34 34
2
1 1
𝑃(𝐴 = −√17𝛼 in state 𝜓2 ) = |𝑐3 |2 =| | =
√34 34