Sie sind auf Seite 1von 6

3.

Uncertainties dened

Let us now dene the uncertainty precisely.


If we have a wave of amplitude A(x), then average of function F (x) over the associated probability
distribution A2 (x) (i.e. the intensity) is given by
F x =

dx F (x) A2 (x)

dx A2 (x)

(15)

and the root mean square (RMS) uncertainty is given by


F =
3.1.1

F 2 x F 2x

(16)

Gaussian wavepacket

Let us take a concrete example, of a Gaussian wave


packet where

2
(k) = ( a )1/4 eak /2 with the prefactor so that
dk (k) = 1.
Here k0 = 0, (k) is real, so x0 = 0 as well.

1 a 1/4
2
A(x) = ( )
dk eak /2 eikx

2
Doing this integral, we see that
A(x) = (

1 1/4 x2 /2a
) e
a

Here notice an interesting feature, if we put a = 1/a, then A(x) = ( a )


gaussian functions. Now let us calculate

1/4 a x2 /2

, i.e. duality of


dk k 2 (k)
2
1/2
kk =
=
(a/)
dk k eak = 0
2
dk (k)


a 1/2
1
2
2 2
1/2
2 ak2
dk k e
=( )
2=
k k = dk k (k) = (a/)
3/2

4a
2a

Thus k =

1 .
2a

Similarly xx = 0, x2 x = 1/(2a ) = a/2 and x =


and kx = 1/2.

a
2

A mathematical theorem (not proved here) gives that this is the minimum product of uncertainties
possible, so in general kx 1/2

Sourabh Dube - PHY202

The Schrodinger Wave Equation

Let us start with the DeBroglie relations:


p = h/ for photons Apply to electrons
Thus, p = h/ for electrons, giving the de Broglie wavelength for electrons as = h/p.
So how do we now treat an electron? We dene a wavefunction (x, t) which has certain properties. This wavefunction is the solution to the Schrodinger equation, which is the equation of
motion of our wave-particle. Usually, we might also say that the particle is the state (x, t).

4.1

The Schrodinger equation

We have seen that a wave can be written as (a plane wave exhibiting interference)

Let us use E = h
, thus = E/
h, and

p = h
k, thus k = p/h

(x, t) ei(kxt)

(x, t) e h (pxEt)
i

(Remember that we can make different wave functions by superimposing different waves of different p and E).
Let us differentiate w.r.t time t and position x.
i

= E
t
h

2
2 = p
= i p and thus

h
2

2 side by 2m and equating the


For a non-relativistic particle, E = p2 /(2m), So dividing the
two RHS

h2 2
(x, t) = ih (x, t)
(17)
2m
t
This is the free-particle Schrodinger equation.
If the particle is moving in some potential V (x), then classically we have E = p2 /2m + V (x) and
thus we can write

h2 2
+ V (x) (x, t) = ih (x, t)
(18)
2m
t
This Schrodinger equation cannot be derived. It is a basic postulate of quantum mechanics that
the wave function of a particle satises this equation. Although we motivated this equation using
plane waves, the solutions of this can be completely different from plane wave. Our goal will be
Sourabh Dube - PHY202

Figure 4: Figure shows a wavefunction squared (in red) and the area between two points is interpreted as the probability of nding particle between those points at given time.
to study this equation for various potentials, and try and nd explicit solutions to it, thus giving us
the wavefunctions for particles in that particular potential.

4.2

The wavefunction

Let us come back to look at what this (x, t) is. There is no physical meaning to (x, t). Instead
the accepted interpretation is that 2 describes the probability distribution for the particle, and (see
Fig: 4)
b
a

|(x, t)|2 dx = Probability of nding particle between a and b, at time t

(19)

If this is the interpretation, and we know that the particle exists, then the probability of nding it
somewhere must be one.

|(x, t)|2 dx = 1

But wait a minute, now we are starting to impose general conditions on the solution of the Schrodinger
equation. How do we ensure that a solution also obeys this probability business? Well it turns out
that since Schrodinger eqn is linear, so any multiple of is also a solution. So we take an unnormalized solution un and turn it into a normalized solution.
norm (
x, t) = N un (
x, t)
and

|N |2 d3 x|un (
x, t)|2 = 1
Thus,
|N | =

d3 x|un (
x, t)|2

(20)

In other words, once we have correctly determined N , we can then use the normalized wavefunc
tion norm and we are assured that the probability condition is met, i.e.
|(x, t)|2 dx = 1.

In the tutorial we shall see that as time progresses, the wavefunction does not lose its normalization.

d

|(x, t)|2 dx = 0
|(x, t)|2 dx =
dt
t

Sourabh Dube - PHY202

(21)
10

We shall see d/dt of the probability is zero: if is normalized at some time, it stays normalized at
some other time.

4.3

The average of x,p

Since |(x, t)|2 is the probability distribution, we can write down the average,

x =

x|(x, t)|2 dx

(22)

Is this the average position of the particle? NO.


Rather it is the average of all measurements (of position) made on many many identically prepared
systems.
We write p = mv = m dx
, so we can evaluate m dx
.
dt
dt

d
dx
x|(x, t)|2 dx =
x | 2 |dx
=
dt
dt
t

(23)

In the tutorial we shall see that we get


dx
ih
=
dx

dt
m
x

(24)

(25)

that is
p = ih
Customarily, this is how we write things
x =

dx
x

(x, t) x (x, t)dx

(x, t) ih
(x, t)dx
p =
x

4.4

(26)

(27)

Position and momentum bases

Recall that we learned earlier, for any wavefunction (x, t) we can calculate its Fourier transform,
which is another wavefunction (p, t) as follows
(p, t) =

1
(2)

3
2

d3 x (x, t) e h px
i

(28)

It turns out that if (x, t) normalized, then (p, t) is also normalized. Thus it is natural to interpret
|(p, t)|2 as the probability of the particle having momentum p at time t. In analogy to x, we can
write p as follows

p =

Sourabh Dube - PHY202

d3 p p |(p, t)|2

(29)
11

which is based on our generalised statement


g(p) =

dp g(p) |(p, t)|2

(30)

We have already written


(x, t) i
h
(x, t)dx
p =
x

(31)

which we arrived at by considering m dx


. Now let us consider the other denition
dt
p =
which in one dimension can be written as
p =

d3 p p |(p, t)|2
2

dp p |(p, t)| =

dp (p, t) p (p, t)

(32)

(33)

Examining these two equations, and since for a given system p is the same, we can see the
connection between two representations of p. And moreover (x, t) and (p, t) are related in a
special way (by a Fourier transform). Let us introduce some terminology to describe this. We refer
to (x, t) as the wavefunction in the x-basis and (p, t) as the wavefunction in the p-basis.
Then, we may say that in the x-basis, the momentum p is represented by the differential operator

(i.e. in one dimension it is i


ih
h x
). In fact, there is also a converse statement that one can
p . Here
p is
prove. In the p-basis, the position x is represented by the differential operator ih
the gradient with respect to the momentum:
(p )i =

, i = 1, 2, 3
pi

(34)

We can choose the basis to work in at our own convenience. We can work permanently in the x At the general level, both x
basis, in which the position is just x, but momentum is given by ih.
and p are to be thought of as operators. In different bases, their representations are different. Thus
in the x-basis, it just happens that position is just a multiplication and momentum a differentiation.
In QM, all observables are operators. And the converse is also true, all Hermitian operators are
observables (We shall dene Hermiticity later).

4.5

The Hamiltonian Operator

Let us now look at a very important operator, which is the QM-analogue of energy. We call it the
energy operator or for historical reasons, the Hamiltonian operator. It is dened as follows
2
= p + V (x)
H
2m

2,
and thus p2 = h2
In the x-basis, we have p i
h,
2

2
=h
+ V (x)
H
2m

Sourabh Dube - PHY202

12

and thus the Schrodinger equation looks like


= ih
H
t
The above discussion allows us to determine the expectation value of the energy of the particle in
state . It is just

H =
(x, t) H (x, t) d3 x

Sourabh Dube - PHY202

13

Das könnte Ihnen auch gefallen