You are on page 1of 14

CH 2: Time-independent Schrödinger equation

2.4 The free particle: 𝑽(𝒙) = 𝟎 everywhere

This should have been the simplest case of all. Classically, this would just mean motion at
constant velocity, but in QM the problem is surprisingly subtle and tricky. The TISE reads

ℏ 𝑑 𝜓
− + 𝑉(𝑥) 𝜓 = 𝐸𝜓
2𝑚 𝑑𝑥
ℏ 𝑑 𝜓
− = 𝐸𝜓
2𝑚 𝑑𝑥
𝑑 𝜓 2𝑚𝐸
=− 𝜓
𝑑𝑥 ℏ

We introduce 𝑘 ≡ such that 𝐸 > 0, then above equation becomes

𝑑 𝜓
= −𝑘 𝜓
𝑑𝑥
This is the same as inside of infinite square well where potential is also zero. However, here for
some reasons we prefer to write the general solution in exponential form instead of sines or
cosines. The solution in exponential form is

𝜓(𝑥) = 𝐴𝑒 + 𝐵𝑒

Unlike the infinite well, there are no boundary conditions to restrict the possible values of 𝑘 and
hence the energy; and the free particle can have any positive energy. Tacking on the standard
time dependence
⁄ℏ
𝜙(𝑡) = 𝑒

Ψ(𝑥, 𝑡) = 𝜓(𝑥)𝜙(𝑡)
⁄ℏ
Ψ(𝑥, 𝑡) = 𝐴𝑒 + 𝐵𝑒 𝑒
⁄ℏ ⁄ℏ
Ψ(𝑥, 𝑡) = 𝐴𝑒 𝑒 + 𝐵𝑒 𝑒
⁄ℏ ⁄ℏ
Ψ(𝑥, 𝑡) = 𝐴𝑒 + 𝐵𝑒
( ⁄ ℏ) ( ⁄ ℏ)
Ψ(𝑥, 𝑡) = 𝐴𝑒 + 𝐵𝑒

√2𝑚𝐸 2𝑚𝐸 𝑘 ℏ
𝑘≡ → 𝑘 = then 𝐸 =
ℏ ℏ 2𝑚

39
CH 2: Time-independent Schrödinger equation

ℏ ℏ
Ψ(𝑥, 𝑡) = 𝐴𝑒 ℏ + 𝐵𝑒 ℏ

ℏ ℏ
Ψ(𝑥, 𝑡) = 𝐴𝑒 + 𝐵𝑒

Now, any function of 𝑥 and 𝑡 that depends on these variables in special combination 𝑥 ± 𝑣𝑡 for
some constant 𝑣 represents a wave of fixed profile, traveling in the ±𝑥 direction at speed 𝑣.

A fixed point on the waveform, for example a maximum or a minimum, corresponds to a fixed
value of the argument and hence to 𝑥 and 𝑡 are such that

𝑥 ± 𝑣𝑡 = constant , or 𝑥 ∓ 𝑣𝑡 = constant

Since every point on the waveform is moving along with the same velocity, its shape doesn’t
change as it propagates. Thus the first term in Ψ(𝑥, 𝑡) represents wave traveling to the right,
and the second represents a wave of the same energy, going to the left.

Since, they only differ by the sign in front of 𝑘, we might as well write


Ψ (𝑥, 𝑡) = 𝐴𝑒

ℏ𝑘 ℏ𝑘 𝜔
𝜔= & 𝑣= & 𝜔 = 𝑣𝑘 or 𝑣 =
2𝑚 2𝑚 𝑘
and let 𝑘 run negative to cover the case of waves traveling to the left

√ 𝑘 > 0 traveling to the right


𝑘=± , with
ℏ 𝑘 < 0 traveling to the left

These are "stationary states" of the free particle are propagating waves with wavelength

2𝜋
𝜆=
|𝑘|

According to de Broglie formula, they carry momentum

ℎ 2𝜋ℏ
𝑝= = → 𝑝 = ℏ𝑘
𝜆 𝜆
The corresponding speed of such waves is

ℏ|𝑘| ℏ √2𝑚𝐸 𝐸
𝑣 = = =
2𝑚 2𝑚 ℏ 2𝑚

40
CH 2: Time-independent Schrödinger equation

On the other hand, the classical speed of the free particle with energy 𝐸 = 𝑚𝑣 (pure

kinetic since 𝑉 = 0), is 𝑣 = , so

2𝐸
𝑣 𝑚
= =2
𝑣 𝐸
2𝑚

𝑣 = 2𝑣

Therefore, it appears that the quantum mechanical wavefunction travels at only half the speed
of the particle that it is supposed to represent!

The other problem is that the resulting wave function is not normalizable. For

Ψ∗ Ψ 𝑑𝑥 = |𝐴| 𝑑𝑥 = |𝐴| (∞)

What does it mean? It means that for the free particle, the separable solutions do not represent
physically realizable states. A free particle cannot exist in a stationary state, or to put it another
way, there is no such thing as a free particle with definite energy.

But that doesn’t mean the separable solutions are of no use to us since they play a mathematical
role that is entirely independent of their physical interpretation.

The general solution to the time dependent Schrödinger equation Ψ(𝑥, 𝑡) is still a linear
combination of separable solutions.

Since 𝑘 is continuous, the resulting expression is an integral instead of sum over the discrete
index 𝑛


Ψ (𝑥, 𝑡) = 𝐴𝑒

1 ℏ
Ψ(𝑥, 𝑡) = 𝜙(𝑘)𝑒 𝑑𝑘
√2𝜋

The constant 1⁄√2𝜋 is putted for convenience, and 1⁄√2𝜋 𝜙(𝑘)𝑑𝑘 plays rule for coefficient 𝑐
in the equation

Ψ (𝑥, 𝑡) = 𝑐 Ψ (𝑥, 𝑡)

41
CH 2: Time-independent Schrödinger equation

For appropriate 𝜙(𝑘), the last wavefunction

1 ℏ
Ψ(𝑥, 𝑡) = 𝜙(𝑘)𝑒 𝑑𝑘
√2𝜋

can be normalized. But it is necessarily carries a range of 𝑘′s and hence a range of energies and
speeds. We call it a wave packet.

If we know the initial wavefunction at time 𝑡 = 0, then the solution is as follow

1 ℏ
Ψ(𝑥, 𝑡) = 𝜙(𝑘)𝑒 𝑑𝑘
√2𝜋
1 ℏ
Ψ(𝑥, 𝑡) = 𝜙(𝑘)𝑒 𝑒 𝑑𝑘
√2𝜋

1
Ψ(𝑥, 0) = 𝜙(𝑘)𝑒 𝑒 𝑑𝑘
√2𝜋
1
Ψ(𝑥, 0) = 𝜙(𝑘)𝑒 𝑑𝑘
√2𝜋

We can determine the function 𝜙(𝑘) [and, therefore Ψ(𝑥, 𝑡)], by

1
𝜙(𝑘) = Ψ(𝑥, 0)𝑒 𝑑𝑥
√2𝜋

This is a classical problem in Fourier analysis; the answer is provided by Plancherel's theorem:

1 1
𝑓(𝑥) = 𝐹(𝑘)𝑒 𝑑𝑘 ⟺ 𝐹(𝑘) = 𝑓(𝑥)𝑒 𝑑𝑥
√2𝜋 √2𝜋

where 𝑓(𝑥) is inverse Fourier transform of 𝐹(𝑘), and 𝐹(𝑘) is Fourier transform 𝑓(𝑥). Note that:
the only difference is the sign of exponent, and the integrals have to exist.

So the solution of generic quantum problem, for the free particle, is the equation

1 ℏ
Ψ(𝑥, 𝑡) = 𝜙(𝑘)𝑒 𝑑𝑘
√2𝜋

with

1
𝜙(𝑘) = Ψ(𝑥, 0)𝑒 𝑑𝑥
√2𝜋

42
CH 2: Time-independent Schrödinger equation

Example 2.6 A free particle that is initially localized in the range −𝑎 < 𝑥 < 𝑎 is released at
time 𝑡 = 0.

𝐴, if − 𝑎 < 𝑥 < 𝑎
Ψ(𝑥, 0) =
0, otherwise

where 𝐴 and 𝑎 are positive real constants. Find Ψ(𝑥, 𝑡)?

Solution: First normalize Ψ(𝑥, 0)

1= |Ψ(𝑥, 0)| 𝑑𝑥

𝑎
= |𝐴| 𝑑𝑥 = |𝐴| 𝑥| = |𝐴| [𝑎 − (−𝑎)]
−𝑎
1
= 2𝑎|𝐴| → 𝐴 =
√2𝑎

Next, calculate 𝜙(𝑘), using

1
𝜙(𝑘) = Ψ(𝑥, 0)𝑒 𝑑𝑥
√2𝜋
1 1 1
𝜙(𝑘) = 𝐴𝑒 𝑑𝑥 = 𝑒 𝑑𝑥
√2𝜋 √2𝜋 √2𝑎

1 1 1 𝑒 𝑎
𝜙(𝑘) = 𝑒 𝑑𝑥 =
√2𝜋 √2𝑎 √4𝜋𝑎 −𝑖𝑘 −𝑎

( )
1 𝑒 −𝑒
𝜙(𝑘) =
2√𝜋𝑎 −𝑖𝑘

1 𝑒 −𝑒
𝜙(𝑘) =
2√𝜋𝑎 −𝑖𝑘

1 1 𝑒 −𝑒
𝜙(𝑘) =
√𝜋𝑎 𝑘 2𝑖

𝑒 −𝑒
sin(𝑥) =
2𝑖
1 sin(𝑘𝑎)
𝜙(𝑘) =
√𝜋𝑎 𝑘

Finally, we put this in the equation

43
CH 2: Time-independent Schrödinger equation

1 ℏ
Ψ(𝑥, 𝑡) = 𝜙(𝑘)𝑒 𝑑𝑘
√2𝜋

1 1 sin(𝑘𝑎) ℏ
Ψ(𝑥, 𝑡) = 𝑒 𝑑𝑘
√2𝜋 √𝜋𝑎 𝑘

This integral need to be solved numerically since it cannot be solved in terms of elementary
functions expect for few invaluable cases. [END]

Problem 2.21 A free particle has the initial wavefunction


| |
Ψ(𝑥, 0) = 𝐴𝑒

where 𝐴 and 𝑎 are positive real constants.

(a) Normalize Ψ(𝑥, 0)


(b) Find 𝜙(𝑘)
(c) Construct Ψ(𝑥, 𝑡), in the form of an integral.

Solution: (a)

| |
𝑒 ∞
|Ψ(𝑥, 0)| 𝑑𝑥 = |𝐴| 𝑒 𝑑𝑥 = |𝐴| (2) 𝑒 𝑑𝑥 = |𝐴| (2)
−2𝑎 0

𝑒 𝑒 1 1 1
|𝐴| (2) − = |𝐴| (2) 0 − = |𝐴| (2) = |𝐴| =1
−2𝑎 −2𝑎 −2𝑎 2𝑎 𝑎

|𝐴| = 𝑎 → 𝐴 = √𝑎
| | | |
∴ Ψ(𝑥, 0) = 𝐴𝑒 → Ψ(𝑥, 0) = √𝑎𝑒

(b) To find 𝜙(𝑘)

1
𝜙(𝑘) = Ψ(𝑥, 0)𝑒 𝑑𝑥
√2𝜋
1 | |
𝜙(𝑘) = √𝑎𝑒 𝑒 𝑑𝑥
√2𝜋

√𝑎 | | | |
𝜙(𝑘) = 𝑒 𝑑𝑥 + 𝑒 𝑑𝑥
√2𝜋

|𝑥| = 𝑥 if 𝑥 > 0
−𝑥 if 𝑥 < 0

44
CH 2: Time-independent Schrödinger equation

𝑎 ( )
𝜙(𝑘) = 𝑒 𝑑𝑥 + 𝑒 𝑑𝑥
2𝜋

𝑎
𝜙(𝑘) = 𝑒( )
𝑑𝑥 + 𝑒( )
𝑑𝑥
2𝜋

𝑎 𝑒( )
0 𝑒( )

𝜙(𝑘) = +
2𝜋 𝑎 − 𝑖𝑘 −∞ −𝑎 − 𝑖𝑘 0

𝑎 𝑒 𝑒 𝑒 𝑒
𝜙(𝑘) = − + −
2𝜋 𝑎 − 𝑖𝑘 𝑎 − 𝑖𝑘 (−𝑎 − 𝑖𝑘) −𝑎 − 𝑖𝑘

𝑎 1 1
𝜙(𝑘) = −0 + 0−
2𝜋 𝑎 − 𝑖𝑘 −𝑎 − 𝑖𝑘

𝑎 1 1
𝜙(𝑘) = −0 + 0+
2𝜋 𝑎 − 𝑖𝑘 𝑎 + 𝑖𝑘

𝑎 (𝑎 + 𝑖𝑘) (𝑎 − 𝑖𝑘)
𝜙(𝑘) = +
2𝜋 (𝑎 − 𝑖𝑘)(𝑎 + 𝑖𝑘) (𝑎 + 𝑖𝑘)(𝑎 − 𝑖𝑘)

𝑎 (𝑎 + 𝑖𝑘) (𝑎 − 𝑖𝑘)
𝜙(𝑘) = +
2𝜋 𝑎 + 𝑘 𝑎 +𝑘

𝑎 𝑎 + 𝑖𝑘 + 𝑎 − 𝑖𝑘
𝜙(𝑘) =
2𝜋 𝑎 +𝑘

𝑎 2𝑎
𝜙(𝑘) =
2𝜋 𝑎 + 𝑘

1 ℏ
Ψ(𝑥, 𝑡) = 𝜙(𝑘)𝑒 𝑑𝑘
√2𝜋

1 𝑎 2𝑎 ℏ
Ψ(𝑥, 𝑡) = 𝑒 𝑑𝑘
√2𝜋 2𝜋 𝑎 + 𝑘

2𝑎 𝑎 1 ℏ
Ψ(𝑥, 𝑡) = 𝑒 𝑑𝑘
√2𝜋 2𝜋 𝑎 +𝑘

𝑎 1 ℏ
Ψ(𝑥, 𝑡) = 𝑒 𝑑𝑘
𝜋 𝑎 +𝑘

45
CH 2: Time-independent Schrödinger equation

Now we return to the issue of velocity. First, there is really no problem because separable
solutions are not physically realizable. The wave packet is a superposition of sinusoidal functions
with their amplitudes modulated by 𝜙.

It can be visualized as "ripples" inside an "envelope". The speed of


envelope (group velocity) corresponds to the particle velocity.
What we found earlier was a speed of individual ripples (phase
velocity).

In our case, for the wave function of a free particle in QM, the
group velocity 𝑣 is twice the phase velocity 𝑣 . Since

ℏ𝑘
𝜔=
2𝑚
then

𝜔 ℏ𝑘 ℏ𝑘
𝑣 = = =
𝑘 2𝑚𝑘 2𝑚
and

𝑑𝜔 𝑑 ℏ𝑘 2ℏ𝑘 ℏ𝑘
𝑣 = = = =
𝑑𝑘 𝑑𝑘 2𝑚 2𝑚 𝑚

𝑣 =𝑣 = 2𝑣

2.5 The Delta – Function Potential

2.5.1 Bound States and Scattering States

We have encountered two very different kinds of solutions to TISE:

(1) Normalizable and labeled by discrete index 𝑛 (infinite well, and harmonic oscillator).

(2) Non-normalizable and labeled by continuous index 𝑘 (free particle).

The first one represents physically realizable states, while the second do not. However, in both
cases the general solution to the time TISE is a linear combination of stationary states .The first
type takes the form of a sum over 𝑛, whereas the second type is an integral over 𝑘.

Now, what is the physical significance of this distinction?

46
CH 2: Time-independent Schrödinger equation

In classical mechanics, a 1-D time independent


potential can give rise to two rather different kinds of
motion.

If the potential 𝑉(𝑥) rises higher than the particle’s


total energy 𝐸 on the either side (figure a), then the
particle is stuck in the potential well. It rocks back and
forth between the turning points, but it cannot escape
unless one provides it with a source of extra energy. We call this bound state.

If, on the other hand, the total


energy 𝐸 exceeds the potential
𝑉(𝑥) on one side or both, then the
particle comes in form infinity , slow
down or speeds up under the
influence of potential, and returns
to infinity (figure b). It cannot get
trapped in the potential unless there is some mechanism, such as friction, to dissipate energy.
We call this a scattering state.

Some potential admits only bound states (for instance, harmonic oscillator). Some allow only
scattering states (a potential hill with no dips in it, for
example). Some permit both kinds depending on the
energy of the particle.

The two kinds of solutions to the Schrödinger equation


correspond precisely to bound and scattering states.

The distinction is even cleaner in quantum domain,


because the phenomenon of tunneling that allows the particle to leak through any finite
potential barrier, so the only thing that matters is the potential at infinity [see figure c]

𝐸 < [𝑉(−∞) and 𝑉(+∞)] ⟹ bound state


𝐸 > [𝑉(−∞) or 𝑉(+∞)] ⟹ scattering state

In “real life” most of potentials go to zero at infinity, in which case the criterion simplifies even
further to

𝐸<0⟹ bound state


𝐸 > 0 ⟹ scattering state

47
CH 2: Time-independent Schrödinger equation

Because the infinite square well and harmonic oscillator the potentials go to infinity as 𝑥 → ±∞,
they admit bound states only, and since the free particle potential is zero everywhere, it only
allows scattering states.

2.5.2 The delta – function potential

The Dirac delta function is an infinitely high, infinitesimally


narrow spike at the origin, whose area is 1 (see figure)

0, if 𝑥 ≠ 0
𝛿(𝑥) ≡ , with 𝛿(𝑥)𝑑𝑥 = 1
∞, if 𝑥 = 0

Technically, it is not a function at all, since it is not finite


at 𝑥 = 0. The mathematicians call it a generalized function, or distribution, and it is very useful
construct in theoretical physics.

Notice that 𝛿(𝑥 − 𝑎) would be a spike of area 1 at point 𝑎 away from the origin in +𝑥 axis. If we
multiply 𝛿(𝑥 − 𝑎) by an ordinary function 𝑓(𝑥) , it’s the same as multiplying by 𝑓(𝑎)

𝑓(𝑥)𝛿(𝑥 − 𝑎) = 𝑓(𝑎)𝛿(𝑥 − 𝑎)

because the product 𝑓(𝑥)𝛿(𝑥 − 𝑎) is zero anyway except 𝑎. Particularly, the most important
property of the delta – function is

𝑓(𝑥)𝛿(𝑥 − 𝑎)𝑑𝑥 = 𝑓(𝑎) 𝛿(𝑥 − 𝑎)𝑑𝑥 = 𝑓(𝑎)

Under the integral sign it serves to “pick out” the value of 𝑓(𝑥) at the point 𝑎. (Of course, the
integral needs not to go from −∞ to +∞; all that matters is that the domain of integration
includes the point 𝑎, so 𝑎 − 𝜖 to 𝑎 + 𝜖 would do, for any 𝜖 > 0)

Let’s consider a potential of form

𝑉(𝑥) = −𝛼𝛿(𝑥)

such that 𝛼 is some positive constant. The Schrödinger equation


for delta-function reads

ℏ 𝑑 𝜓
− + 𝑉𝜓 = 𝐸𝜓
2𝑚 𝑑𝑥
ℏ 𝑑 𝜓
− − 𝛼𝛿(𝑥)𝜓 = 𝐸𝜓
2𝑚 𝑑𝑥
It yields both bound states (𝐸 < 0) and scattering states (𝐸 > 0).

48
CH 2: Time-independent Schrödinger equation

Bound state solutions to the delta function potential

We’ll look First at the bound states, in the region where 𝑥 < 0, 𝑉(𝑥) = 0, so

ℏ 𝑑 𝜓
− − 𝑉(𝑥) 𝜓 = 𝐸𝜓
2𝑚 𝑑𝑥
𝑑 𝜓 2𝑚𝐸
=− 𝜓
𝑑𝑥 ℏ

𝑑 𝜓 √−2𝑚𝐸
∴ =𝜅 𝜓 , where 𝜅 ≡
𝑑𝑥 ℏ
In which 𝐸 is negative by assumption, so 𝜅 is real and positive. The general solution to the last
equation = 𝜅 𝜓 is given as

𝜓(𝑥) = 𝐴𝑒 + 𝐵𝑒

The first term 𝐴𝑒 blows up 𝜓 → ∞ as 𝑥 → −∞, so we must choose 𝐴 = 0, then

∴ 𝜓(𝑥) = 𝐵𝑒 (𝑥 < 0)

In the region where 𝑥 > 0, 𝑉(𝑥) is again zero, and the general solution is of the form

𝜓(𝑥) = 𝐹𝑒 + 𝐺𝑒

this time it’s the second term 𝐺𝑒 blows up 𝜓 → ∞ as 𝑥 → ∞, so

𝜓(𝑥) = 𝐹𝑒 (𝑥 > 0)

It remains only to stitch these two functions together, using the appropriate boundary conditions
at 𝑥 = 0

1. 𝜓 is always continuous
2. 𝑑𝜓⁄𝑑𝑥 is continuous except at points where the potential is infinite

The first boundary condition yields

lim 𝜓(𝑥) = lim 𝜓(𝑥)


→ →

lim 𝐹𝑒 = lim 𝐵𝑒
→ →

𝐹𝑒 = 𝐵𝑒

∴𝐹=𝐵

49
CH 2: Time-independent Schrödinger equation

𝐵𝑒 (𝑥 ≤ 0)
∴ 𝜓(𝑥) =
𝐵𝑒 (𝑥 ≥ 0)

The second condition 𝑑𝜓⁄𝑑𝑥 tells us nothing;


this is (like the infinite square well) the
exceptional case where 𝑉 is infinite at the join,
and it’s clear from the graph that this function
has a kink at 𝑥 = 0.

Up to this point the delta function has not


come into story at all.

Evidently, the discontinuity in the derivative of 𝜓, at 𝑥 = 0, must be determined by the delta


function. We'll now show how it works, and as a result, we'll see why 𝑑𝜓⁄𝑑𝑥 is typically
continuous.

The idea is to integrate the Schrödinger equation from −𝜖 to +𝜖, and take the limit at 𝜖 → 0

ℏ 𝑑 𝜓
− − 𝑉(𝑥)𝜓 = 𝐸𝜓
2𝑚 𝑑𝑥
ℏ 𝑑 𝜓
− 𝑑𝑥 − 𝑉(𝑥)𝜓𝑑𝑥 = 𝐸𝜓(𝑥)𝑑𝑥
2𝑚 𝑑𝑥

ℏ 𝑑𝜓 𝜖
− − 𝑉(𝑥)𝜓𝑑𝑥 = 𝐸𝜓(𝑥)𝑑𝑥
2𝑚 𝑑𝑥 −𝜖

The second integral

𝐸𝜓(𝑥)𝑑𝑥 → lim 𝐸𝜓(𝑥)𝑑𝑥 = 0


since it's the area has vanishing width and finite height, thus

ℏ 𝑑𝜓 𝜖
− − 𝑉(𝑥)𝜓𝑑𝑥 = 0
2𝑚 𝑑𝑥 −𝜖

Now,

𝑑𝜓 𝑑𝜓 +𝜖 2𝑚
Δ ≡ lim = lim 𝑉(𝑥)𝜓(𝑥)𝑑𝑥
𝑑𝑥 → 𝑑𝑥 −𝜖 ℏ →

50
CH 2: Time-independent Schrödinger equation

𝑑𝜓 𝑑𝜓 2𝑚
lim − = lim 𝑉(𝑥)𝜓(𝑥)𝑑𝑥
→ 𝑑𝑥 𝑑𝑥 ℏ →

Typically, the limit on the right is again zero, and that’s why the derivative 𝑑𝜓⁄𝑑𝑥 is ordinarily
continuous. But when the potential 𝑉(𝑥) is infinite at the boundary this argument failed. In
particular, if

𝑉(𝑥) = −𝛼𝛿(𝑥)

the equation

𝑑𝜓 𝑑𝜓 2𝑚
lim − = lim 𝑉(𝑥)𝜓(𝑥)𝑑𝑥
→ 𝑑𝑥 𝑑𝑥 ℏ →

𝑑𝜓 𝑑𝜓 2𝑚
lim − = lim −𝛼𝛿(𝑥)𝜓(𝑥)𝑑𝑥
→ 𝑑𝑥 𝑑𝑥 ℏ →

𝑑𝜓 𝑑𝜓 2𝑚𝛼
lim − =− lim 𝛿(𝑥)𝜓(𝑥)𝑑𝑥
→ 𝑑𝑥 𝑑𝑥 ℏ →

∵ 𝑓(𝑥)𝛿(𝑥 − 𝑎)𝑑𝑥 = 𝑓(𝑎) 𝛿(𝑥 − 𝑎)𝑑𝑥 = 𝑓(𝑎)

𝑑𝜓 𝑑𝜓 𝑑𝜓 2𝑚𝛼
∴Δ ≡ lim − =− 𝜓(0)
𝑑𝑥 → 𝑑𝑥 𝑑𝑥 ℏ

Now, we are going to determine Δ . For this case, the equation

𝐵𝑒 (𝑥 ≤ 0)
𝜓(𝑥) =
𝐵𝑒 (𝑥 ≥ 0)

gives

𝑑𝜓⁄𝑑𝑥 = −𝐵𝜅𝑒 for (𝑥 > 0) so, 𝑑𝜓⁄𝑑𝑥 | = −𝐵𝜅


𝑑𝜓⁄𝑑𝑥 = 𝐵𝜅𝑒 for (𝑥 < 0) so, 𝑑𝜓⁄𝑑𝑥 | = +𝐵𝜅

Hence, the expression

𝑑𝜓 𝑑𝜓 𝑑𝜓 2𝑚𝛼
Δ ≡ lim − =− 𝜓(0)
𝑑𝑥 → 𝑑𝑥 𝑑𝑥 ℏ

2𝑚𝛼
[−𝐵𝜅 − 𝐵𝜅] = − 𝜓(0)

( ) ( )
𝜓(0) = 𝐵𝑒 = 𝐵𝑒 =𝐵

51
CH 2: Time-independent Schrödinger equation

2𝑚𝛼
−2𝐵𝜅 = − 𝐵

𝑚𝛼
𝜅=

Since

√−2𝑚𝐸 ℏ 𝜅
𝜅≡ → 𝐸=−
ℏ 2𝑚
Then, the allowed energy is

𝑚𝛼 𝑚 𝛼
ℏ 𝜅 ℏ ℏ
𝐸=− =− ℏ =− ℏ = − 𝑚𝛼
2𝑚 2𝑚 2𝑚 2ℏ
Finally the normalization gives

1= |𝜓(𝑥)| 𝑑𝑥 = |𝐵𝑒 + 𝐵𝑒 | 𝑑𝑥 = |𝐵| 𝑒 + 𝑒 𝑑𝑥

( ) ( )
𝑒 ∞ 𝑒 𝑒
= 2|𝐵| 𝑒 𝑑𝑥 = 2|𝐵| = 2|𝐵| −
−2𝜅 0 −2𝜅 −2𝜅

1 1 |𝐵|
= 2|𝐵| 0− = 2|𝐵| =
−2𝜅 2𝜅 𝜅
|𝐵|
=1 → 𝐵 = √𝜅
𝜅
𝑚𝛼
𝜅= is real and positive

𝑚𝛼 √𝑚𝛼
𝐵 = √𝜅 = =
ℏ ℏ

Evidently the delta-function well, regardless of its strength 𝛼, has exactly one bound state.
| |
𝜓(𝑥) = 𝐵𝑒

√𝑚𝛼 | |⁄ℏ
𝑚𝛼
𝜓(𝑥) = 𝑒 , such that 𝐸 = −
ℏ 2ℏ

52

You might also like