Vous êtes sur la page 1sur 10

Statistical Thermodynamics: Molecules to Machines

Venkat Viswanathan
May 20, 2015

Module 1: Classical and Quantum Mechanics


Learning Objectives:
The formulation of classical mechanics in the Lagrangian form as a
preliminary setup for quantum mechanics
Introduction to basic concepts in quantum mechanics, key
differences from the classical concepts.
Example problems to highlight key features of classical and quantum
mechanics, which will also be be exploited further in the statistical
thermodynamics part of this course

Key Concepts:
Lagrangian formulation of classical (Newtonian) mechanics, path of minimal action, quantum mechanical amplitude, path integration,
Schrdinger equation, quantum mechanical modes.

statistical thermodynamics: molecules to machines

Classical Mechanics
Classical mechanics, also called Newtonian mechanics, is based
Newtons laws of motion which govern the motion of macroscopic
objects. It allows a continuous spectrum of energies and a continuous
spatial distribution of matter. Newtons laws of motion are:
1. First Law When viewed in an inertial reference frame, an
object at rest tends to stay at rest and that an object in uniform
motion tends to stay in uniform motion unless acted upon by a net
external force.
2. Second law An applied force F on an object equals the time rate
of change of its momentum p, leading directly to the equation F =
ma, where m is the mass of the object (independent of time), and a
is the acceleration.
3. Third law For every action there is an equal and opposite reaction
Various mathematical formulations exist for describing motion of
ob- jects in classical mechanics, which are useful in understanding
quantum mechanics. We begin with the Lagrangian formalism, which
is based on the principle of stationary action. The lagrangian,
L, of a particle is defined as the difference between its kinetic energy,
T , and potential energy, V , using generalized coordinates for space, q
= (qx, qy , qz , ....), and time, t, for describing the motion as:
L= TV= m

V (q, t)

(1)

2
q
The action, S, is defined as the integral of the lagrangian between
dq
two given instants of time (where q =d ) as:

S=

t2

L(q, q, t) dt

(2)

t1

Now, the principle of stationary (or least) action states that


the path taken by the system between times t1 and t2, as shown in Fig.
2, is the one for which the action is stationary (no change) to first
order. Mathematically, for indicating a small change, this principle
states:
S = S[q + q] S[q] = 0

(3)

As the end points are fixed at q1 and q2, the perturbation has the
condition q1 = q2 = 0. Using the definition of S as in Eq. (2), we

Figure 1: Sir Issac Newton: "I do not


know what I may appear to the world,
but to myself I seem to have been only
like a boy playing on the seashore, and
diverting myself in now and then finding a smoother pebble or a prettier shell
than ordinary, whilst the great ocean of
truth lay all undiscovered before me."

have:

S[q + q] =

L(q + q, q + q, t) dt

t2

L(q, q, t) + q
L dt

t1

t2
t1

= S[q] +

+ q

t2
t1

(4)

q
+ q
q
q

dt
Figure 2: Motion of a particle from
q1, at time t1 to q2, at time t2 in the
ex- ternal potential V (x, t). Among
the several possible paths in q and t
that the particle can traverse, the one
denoted in red is the classical path that

Therefore, using integration by parts, the variation S can be


as: written

S =

t2

L
q

t2

L
+
q

dt =
q

t2

dt (5)
q dt
q
q
q
t1
t1
q. 1
q
The first term in Eq. 5 is zero as q1t = q2 = 0. Therefore, regardless
of q, the path with the minimum action will satisfy the condition:
d L
dt q

L
= 0

(6)

Finally, using the definition of L, we have the equation of motion as:


dm2
qdt2 +
V

=0

(7)

V
Defining the force due to the external potential to be F
=
Now,
q
we have V = ma, which is Newtons second law of motion.

we considerqexample problems to draw some conclusions about classical


mechanics.
Example 1: A free particle
Consider a free particle with 1-D motion along the x axis and the external potential V (x, t) = 0. Therefore, its equation of motion will be:
m

d2 x

(8)
dt2 = 0 x(t) = C1t + C2
Where C1, C2 are constants determined using the initial conditions.
Considering x = 0 at t = 0 and x = v = 0 at t = 0, we get x = vt as
the equation describing the particles motion. This result is in
agreement
with Newtons law of motion.
Example 2: A particle in a harmonic potential field
Consider the same particle as in Exmple 1, but with the external
k 2
poten- tial V (x,
2 t) = x . This will result in an equation of motion as:

dm2
xdt2 + kx = 0

d2x dt2

the particle chooses to traverse along,


the other path curves (in purple) are
not taken by the particle.

+ x=0
(9)

.
Where =

is called the characteristic frequency. Considering

x = 0 at t = 0 and x = v = v0 at t = 0, we get the equations


describing
the particles motion as:
x(t) = A sin(t) + B cos(t) =

v0

sin(t)

(10)

dx

= v0 cos(t)
(11)
dt
Like the Lagrangian formulation, the Hamiltonian formulation of
classical mechanics describes the the equations of motion, albeit using a
different quantity, H, called the hamiltonian, which is defined as the
sum of the kinetic and potential energies as:
v(t) =

H=T+V =

q + V (q, t)

2
The hamiltonian of the particle in Example 2, would hence be:
1
1 2 2
1
2
2
2

(12)

H = mv0 cos (t) + 2 v0 sin (t) = mv0


(13)

2
2
Which is independent
of time. 2
From these two simple examples we infer some key conclusions.
Clas- sical mechanics predicts particle motion to be deterministic, i.e.
the con- ditions of a particle at a given time will chart out its future
trajectory. The Lagrangian formulation teaches us that particle
traverses along a path that action S to be an extremum. A particle
that is free from the influence of any external potential (and thus
forces) will maintain a constant velocity, as proposed by Newtons
first law of motion. Finally, the motion of a particle in a stationary or
time independent potential will be governed by the constraint of
maintaining constant total energy H = T + V , as described by the
Hamiltonian formulation.

Quantum Mechanics
Although classical mechanics is successful when applied for macroscopic
objects, several experimental observations demonstrate the inadequacy
of classical mechanics in treating microscopic phenomena. For example:
1.The Rayleigh-Jeans formula for spectral intensity of black body radiation, which was based on laws of mechanics, electromagnetic theory
and statistical thermodynamics failed for short wavelengths in what
was called as the Ultraviolet Catastrophe. Max Planck later postulated that the oscillating atoms of a black body radiate energy only in
discrete, i.e. quantized amounts which was found to be in agreement
with experimental observations (Fig. 3).

Figure 3: Plancks law (colored curves)


accurately describes black body radiation and resolved the Ultraviolet
Catastrophe (black curve)

2.The interference patterns that arise from light impinging on a


double- slit experiment, originally done by Young, brought into
forefront the fact that light and matter can display
characteristics of both classically defined waves and particles.
Young showed by means of a diffraction experiment that light
behaved as waves. He also pro- posed that different colors were
caused by different wavelengths of light (Fig. 4).
3.The photoelectric effect, explained by Albert Einstein, which is
the phenomenon of emission of electrons from a metallic surface that
is subjected to electomagnetic radiation. In case light was only a
wave, the energy contained in one of those waves would depend only
on its amplitude, i.e. on the intensity of the light. Other factors, like
the frequency, should make no difference. However, electron emission
was found to occur at a threshold frequency (not intensity) and the
maximum kinetic energy of the emitted electrons was found to depend
on the frequency of the incident light (Fig. 5).

Figure 4: Two-slit diffraction pattern


due to interference of plane waves.

Quantum mechanics shows, that physical processes are not predetermined in a mathematically exact sense. The particle motion is
not restricted to a single path determined by the principle of least
action; instead all the paths, as shown in Fig. 2, have a probability of occurring. We define the probability P (2, 1) of going
from 2 = (q2, t2) to 1 = (q1, t1) in terms of a total amplitude K(2,
1), such
that P (2, 1) = |K(2, 21)| . Using the previously defined quantity,
action S of a particular path, the total amplitude can be considered as
a sum of contributions [q(t)] from each and every path
connecting 1 to 2, such that:
K(2, 1) =

[q(t)]

(14)

all paths

Where the contribution of each path can be determined in terms of


its action as:
2i S[q(t)])
(15)
[q(t)] = C.
h
exp(
Where h = 6.626 1034 J s is Plancks constant, and the constant
C is chosen such that K(2, 1) can be normalized. We saw earlier that
q was one of the several paths chosen by the particle to go from 1
to 2, however, the overall amplitude K(2, 1) includes contributions
from each path, however improbable. Here we introduce the concept
of path
integrals 1 that formally defines the summation over all possible paths
going from 1 to 2 as:
K[2, 1] =
C

12
allpaths

exp(

2i S[q(t)]) d[q(t)]
h

(16)

Figure 5: The maximum kinetic energy


as a function of the frequency of light,
as observed in the photoelectric effect

All objects are quantum mechanical in nature, i.e. they traverse along
paths with probabilities dictated by the action S of each path. Macroscopic objects that have comparably large masses have actions which
are large when compared to the quanta of action which is h. Therefore,
macroscopic objects posses only one dominant path which determines
their behavior; this path corresponds to the classical path q as
deter- mined by S = 0. While such a formulation smoothly merges into
New- tonian mechanics for macroscopic physical processes, it has far
reaching implications on the interpretation of microscopic physical
processes.
As discussed before, the amplitude K(2, 1) is related to the
probabil- ity of going from 1 to 2. To find the probability of locating a
particle at a location q at time t, we define the wave-packet (q, t) to
give the
2
time-dependent probability distribution P (q, t) = |(q, t)
| . Using the
condition that the probability must be Markovian, we can write:

[q2, t2]
=

K(2, 1)[q1, t1] dq1

(17)

This property is used to find a diffusion equation for the wave-packet,


. further details can be found elsewhere 2. The governing equation for
the wave-packet is :

h (q,
t)
2i

h2 2(q, t)

=
t

q 2

+ V (q, t)(q, t)

(18)

82
m
This equation is the famous Schrdinger equation that forms the
basis of most of quantum mechanical calculations. Using r as the
position vector, the same equation can be expressed in 3 dimensions
as:

h
2i (r, t)

+ V (r,
8 2 m
t)

.
(r, t)

(19)

= t
In order to predict the expectation value of energy, we note that the
Hamiltonian operator is:
2

H= 2 + V
8 mof energy, E, as:
Which gives the expectation value
H = E

(20)

(21)

This is also known as the time independent Schrdinger equation.


Next, we consider some cases where we consider the primary molecular
behavior of a particle in equilibrium using this equation.
Example 3: Particle in a box
Consider a particle with 1-D motion along the x axis in a box of

length L from x = 0 to x = L.
The external potential is assigned
as V (x, t) = 0

Figure 6: The potential barriers out- side the 1-D box are infinitely large, while the
interior of the box has a con- stant, zero potential.

for 0 < x < L and V (x, t) for x L and x 0 as shown in Fig. 6.


The governing equation for the particle inside the box is:
h2

d2

2
= E
(22)
8 m dx2
This equation has the boundary conditions = 0 at x = 0 and
at x = L. The equation can be written in the same form as that of
a harmonic oscillator as:
d2

+G =0

Where G2 =

8mE2
h

dx2
. The solution for this system is given

= C1 sin(Gx) + C2 cos(Gx)

(23)
as:
(24)

To satisfy the boundary conditions, we must have C2 = 0. The


remaining solution has infinite possibilities because sin(n) = 0 for n
= 1, 2, 3, 4....... The condition for the solution thus results in:
.
8mEn
GnL
(25)
h2
=
This implies:
h2

2
(26)
En =
n
2 8mL2
Using the condition that || has to be normalized, we have C1 =
,
th
quantum
( 2). Hence, the solution for the wave-packet for the
L
n
state of the particle is:

nx

)
(27)
L
L
One can easily extend this to 3-dimensions, which instead of n would
n(x) =

2sin(

result in nx, ny , nz . However, what is more important here is to understand the quantization of the energy levels in terms of n. For varying n,
we get different solutions of the Schrdinger equation in 1-dimension, as
shown in Fig. 7.
Next, we look at the quantum-mechanical analogue of the particle
in a harmonic potential field.
Example 4: The quantum mechanical harmonic oscillator
The vibrational modes of a diatomic molecule can be determined by
con- sidering a single particle in a harmonic potential. Consider a
diatomic molecule with atomic masses m1 and m2. The covalent bond
between the two atoms can be modeled as a harmonic spring with
spring con-

Figure 7: Solution for the wave-packets


for the first four states, n = 1, 2, 3, 4, in
a one-dimensional particle in a box

stant k. If we define x to be the distance of separation between the two


atoms, we have the governing equation for the wave-packet as:
h2 d2
+

Where

m m
= 1 2.
m1
2

kx = E

(28)
82 dx2
2
Upon solving this equation, similar to the case in

Example 3, we have an infinite number of solutions with discreet energy


levels. In general, the nth wave-packet can be described by:
n =

.
n!

2na
Where a4 =

h2

.1/2

.
.
2
. x
exp x 2
2a
H
.
n

(29)

and the function Hn (u) represents the Hermite


42k
polynomials as H0(u) = 1, H1(u) = 2u, H2(u) = 4u2 2, H3(u) =
8u3 12u for the first four states. Following this procedure, the
of the nth state can be described as:
.
.
1
h
E
=
n
.
n+
2
Where =
k

energy

(30)

2
and n = 0, 1, 2, 3, 4..... are the quantum mechanical
modes of motion. As in the case of classical mechanics, the
characteristic frequency plays an important role in determining the
solutions using the quantum mechanical solutions, as shown in Fig. 8.

Figure 8: Wave-packet representations


for the eigenstates, n = 0 to 7 for the
harmonic oscillator.The horizontal
axis
shows the position x

Vous aimerez peut-être aussi