0% found this document useful (0 votes)
114 views20 pages

Quantum Physics: 1 Some Definitions

1. In the early 20th century, experiments produced results that could not be explained by classical physics, necessitating new laws of quantum physics. 2. Blackbody radiation played a major role in the development of quantum physics. Experiments showed that the wavelength of blackbody radiation's peak intensity shifts to lower values as temperature increases. 3. Neither Wien's distribution law nor Rayleigh-Jeans law could fully explain the spectral distribution of blackbody radiation.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
114 views20 pages

Quantum Physics: 1 Some Definitions

1. In the early 20th century, experiments produced results that could not be explained by classical physics, necessitating new laws of quantum physics. 2. Blackbody radiation played a major role in the development of quantum physics. Experiments showed that the wavelength of blackbody radiation's peak intensity shifts to lower values as temperature increases. 3. Neither Wien's distribution law nor Rayleigh-Jeans law could fully explain the spectral distribution of blackbody radiation.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 20

Quantum Physics

In the beginning of 20th century, scientists felt that the known laws of physics are inadequate to
explain the events that occurs in nature. It was common belief that the nature consists of particles
and radiation. The particles obey Newton’s laws of motion while radiation follows Maxwell’s laws of
electromagnetism. These well established laws at that time are known as ‘Classical Physics’.
The results of some experiments performed at that time could not be explained by the classical
physics. Some of these experiments are,

• Black body radiation

• Photo electric effect

• Stability of atom

• Compton effect etc.

In order to explain these effects, new laws of physics are to be developed. These new laws, which
explain above mentioned phenomenon are known as ‘Quantum Physics’. The black body radiation
has a major role in development of quantum physics. In order to have a better understanding of
quantum physics, we have to explain some definitions.

1 Some definitions

1.1 Radiation
Energy emitted or absorbed by the moving particle or wave is called radiation. In case of electromag-
netic waves, the radiation is electromagnetic and velocity of the radiation is given by,
1
v = √ (1.1)
µ

1.2 Black body


The ability of a body to radiate is closely related to its ability to absorb radiation. This is to be
expected, since a body at a constant temperature is in thermal equilibrium with its surroundings and
must absorb energy from them at the same rate as it emits energy. It is convenient to consider an
ideal body that absorbs all radiation incident upon it, regardless of frequency. Such a body is called
a black body.
The point of introducing the idealized black body in a discussion of thermal radiation is that we
can now disregard the precise nature of whatever is radiating, since all blackbodies behave identically.
In the laboratory a blackbody can be approximated by a hollow object with a very small hole leading
to its interior. Any radiation striking the hole enters the cavity, where it is trapped by reflection back
and forth until it is absorbed. The cavity walls are constantly emitting and absorbing radiation, and
it is in the properties of this radiation (black body radiation) that we are interested.

1.3 Black body radiation


When a black body is heated and kept at fixed temperature em radiations are emitted from the inner
walls of the cavity (black body). The intensity and spectrum of black body radiation depends only on
the temperature of the black body or cavity. The fixed temperature indicates the thermal equilibrium
of the system. The thermal equilibrium means that the radiation emitted per second by the walls is
equal to the radiation absorbed.

1.4 Energy density


At a particular temperature T , the energy in the per unit volume of the cavity is called energy density.

1
Figure 2: The spectral radiancy of a black
body radiator as a function of the frequency of
radiation. Note that the frequency at which
the maximum radiancy occurs (dashed line)
Figure 1: The spectral radiancy curves for
increases linearly with increasing temperature,
cavity radiation at four selected temperatures as
and that the total power emitted per square
function of wavelength. Note that as the tem-
meter of the radiator (area under curve) in-
perature increases, the wavelength of maximum
creases very rapidly with temperature.
spectral radiancy shifts to lower values.

1.5 Spectral energy density


At certain temperature the average energy density between wavelength λ and λ+dλ is given by uλ dλ.
Thus the total energy density is defined as,
Z ∞
U = uλ dλ (1.2)
0

where uλ is spectral energy density.

1.6 Spectral emissive power and total emissive power


The spectral emissive power R(T ) at a certain temperature T is defined as,
Z ∞
R(T ) = R(T, λ) dλ (1.3)
0

where R(λ, T ) is known as spectral emissive power or spectral radiancy. The spectral emissive power
is power radiated per unit area per unit time. Spectral emissive power is related to average energy
density by the relation,
c
R(T ) = U (T ) (1.4)
4

2 Experimental studies of black body radiation

First attempt to study the black body radiation was made by Lummer and Pringsheim. The plot
of spectral radiancy and wavelength are plotted at various temperatures as shown in figure 1. In
figure 2, spectral radiancy, which now can be considered as a function of frequency and temperature
R(ν, T ).

2.1 Characteristic of black body radiation


the black body radiation is described by following properties,

1. As it is clear from the figure 1 that the graph is continuous, which means that at every temper-
ature radiation for all wavelengths are emitted but the spectral radiancy is different for different
wavelengths.

2
2. Spectral radiancy for each λ increases with temperature.

3. At a particular temperature, first spectral radiancy increases with λ but after reaching a certain
highest value, it goes on decreasing. The value of wavelength at which highest value of spectral
radiancy occurs is λm .

4. The Wien’s displacement law: We can see from the spectral radiancy curves of Figure
1 that λm , the wavelength at which the spectral radiancy is a maximum, decreases as the
temperature increases. Wilhelm Wien was able to show from classical theory that, in fact, the
product λm T (= w) is a universal constant; its experimental value proves to be

w = λm T = 2898 µm.K (2.1)

The above equation is known as the Wien’s displacement law and w as the Wien’s constant.

5. The Stefan- Boltzmann law: The total radiated output per unit area of the cavity aperture,
called its radiancy R(T ), is given by

R(T ) = σ T 4 (2.2)

in which σ (= 5.670x10−8 W/m2 K 4 ) is a universal constant called the Stefan-Boltzmann con-


stant.
6. × 10-9
3 Failure of classical physics

There are many attempts to explain the nature of


black body radiation described in previous sec-
tion. Now we are going to study Wien’s and 4. × 10-9
Rayleigh-Jeans distribution law and to see that
these distributions are unable to explain the ob-
ℜ(ν)

served phenomenon of blackbody radiation. The


comparison between Rayleigh-Jeans’ law, Wien’s
distribution law and Planck’s distribution law is 2. × 10-9
shown in figure 3. Planck
Rayleigh-Jeans
3.1 Wien’s distribution law
Wien

Wien’s distribution law involved a conjecture


that the spectral energy density function can be 0
14 14 14
0 1 × 10 2 × 10 3 × 10
related to the Maxwell speed distribution func- ν
tion for the molecules of an ideal gas. It is given
by,
ρ(ν, T ) = a ν 3 e−bν/T (3.1) Figure 3: Three theoretical predictions for the spec-
tral energy density of cavity radiation at 2000 K. The
where a and b are constants and ρ(ν, T ) is spec- arbitrary constants in the Planck formula were chosen
tral energy density function. to give the best fit to the experimental data. These
same constants were used in plotting the Wien curve.
3.2 Calculation of frequency modes The Rayleigh-Jeans curve(R-J) requires no adjustment
of constants.
Consider a large (meaning side length L >> λ,
where λ is the longest wavelength of interest) cubical cavity filled with radiation in thermodynamic
equilibrium. The purpose of the cavity is to confine the radiation long enough for it to reach equilib-
rium. The walls of a cavity capable of confining radiation must have non-zero conductivity because
walls of zero conductivity would be transparent; they would generate no currents in response to the
electric field of the incoming radiation, so the radiation would pass through unaffected. For walls
having non-zero conductivity, the equilibrium electric field strength at the walls is E = 0 because
modes with E 6= 0 at the walls are lossy. Only those standing waves with E = 0 at the walls will
persist after some time t >> L/c . We can enumerate all possible standing-wave modes in the cavity.

3
Figure 4: Standing waves corre-
sponding to jx λx /2 = L for jx =
1; 2; 3. Abscissa:x-axis of the cav-
ity, bounded by walls at x = 0 Figure 5: This two-dimensional figure illustrates
and x = L . Ordinate: Electric standing waves propagating in a cavity with wave nor-
field strengths of the longest stand- mals at angles α and β from the x and y axes, respec-
ing waves satisfying the boundary tively. Examples of wave nodes, where |E| = 0 , are
condition E = 0 at the walls. indicated by dashed lines for the case jx = 3, jy = 2.

For example, consider all standing waves whose wave normals point in the x direction. The boundary
conditions E = 0 at x = 0 and at x = L mean that only the waves having the discrete wavelengths
λx 2λx 3λx
= L, = L, = L, . . . (3.2)
2 2 2
can have nonzero amplitudes. Thus,
jx λx
= L (3.3)
2
where jx = 1, 2, 3, . . .. Similarly,
jy λy jz λz
= = L (3.4)
2 2
The above expressions are for waves normal in the x, y, z directions.
What about a wave whose normal is in some arbitrary direction? Let α, β, γ be the angles
between the wave normal and the x; y; z axes, respectively. From the figure 5 it is clear that

λ = λx cos α (3.5)

where λ is the wavelength measured in the direction of the wave normal and λx ≥ λ is the spacing
between the wave nodes measured along the x-axis. Thus
λ
λx = (3.6)
cos α
Similarly,
λ λ
λy = , λz = (3.7)
cos β cos γ
Now our boundary conditions,
2L 2L 2L
jx = , jy = , jz = (3.8)
λx λy λz

become,
2L cos α 2L cos β 2L cos γ
jx = , jy = , jz = , (3.9)
λ λ λ
Squaring and summing the last three equations gives,
 2
2L
jx2 jy2 jz2 cos2 α + cos2 β + cos2 γ .

+ + = (3.10)
λ

Using directional cosine rule, the last term in the parentheses with various cosine squared terms is

4
equal to one, thus our formula reduces to,
 2
2L
jx2 + jy2 + jz2 = (3.11)
λ

Since the allowed frequencies for a particular


wavelength follows the relation ν = c/λ, we can
write,
 2
2L
jx2 + jy2 + jz2 = ν2 (3.12)
c

The permitted standing-wave modes can be rep-


resented as points in the positive octant of the
space whose axes are jx , jy and jz . Each point in
the positive octant represents one possible mode
of electromagnetic radiation. The space density
of points in this space is unity, so the number of
points in any volume is equal to the volume of
positive octant.
Let j be the radius of coordinates jx , jy and jz ,
then
 2
Figure 6: The (x, y) plane in the imaginary space + 2 2 2 2L
jx jy + jz = j = ν 2. (3.13)
whose axes are (jx , jy , jz ) . Permitted standing waves c
in the (jx , jy ) plane are indicated by dots at positive
integer values of these axes. The number g(ν)dν of modes having frequencies
in the range ν to ν + dν is the volume of the
spherical octant shell between j and j + dj , multiplied by 2 (two polarizations):
1
g(ν)dν = × 2 × 4π j 2 dj
8
= π j 2 dj (3.14)

Now putting value of j from 3.13, we get number of frequency modes as,

8πL3 2
g(ν) dν = ν dν (3.15)
c3
Thus the density of frequency of standing waves is,
8π 2
G(ν) dν = ν dν, (3.16)
c3
which is independent of size of the cavity.

3.3 Rayleigh-Jeans distribution law


The Rayleigh-Jeans law treats the distribution of energy as classical, i.e. classical Boltzmann distri-
bution law. The Boltzmann distribution function is given by,

P (E) = e−E/kT (3.17)

Using the Boltzmann distribution function, the average energy is given by the formula,
R∞
E P (E) dE
¯ = 0R ∞ (3.18)
0 P (E) dE

5
Putting value of distribution function P (E), we get the following integrals,
Z ∞
E e−E/kT dE = (k T )2
0
Z ∞
e−E/kT dE = k T (3.19)
0

Substituting the values of different integrals in the equation 3.18, we get average energy as,

¯ = k T (3.20)

Using this formula for average energy, we can write Rayleigh-Jeans distribution as,

ρ(ν, T ) dν = ¯ G(ν) dν
8πkT 2
= ν dν. (3.21)
c3
This expression shows the energy density of standing waves in lying in the region between ν and
ν + dν.

3.4 Planck’s quantum hypothesis


Shortly after the derivation of the Rayleigh-Jeans formula, Planck found a simple way to explain the
experimental behaviour, but in doing so he contradicted the wave theory of light, which had been so
carefully developed over the previous hundred years. Planck realized that the ultraviolet catastrophe
could be eliminated by assuming that a mode of frequency ν could only take up energy in well defined
discrete portions (small packets or quanta) each having the energy,
 
h
E = h ν = ~ ω, ~= , ω = 2πν (3.22)

where the constant h is adjusted to fit the experimentally observed R(λ, T ). Contrast between the
wave and Planck’s hypothesis is that in the classical case the mode energy can lie at any position
between 0 and ∞ of the energy line, whereas in the quantum case the mode energy can only take on
discrete (point) values. The assumption of the discrete energy distribution required a modification of
the equipartition theorem. Planck introduced “discrete portions” so that he might apply Boltzmann’s
statistical ideas to calculate the energy density distribution of the black-body radiation.

3.5 Planck’s distribution law


The difference between Planck’s law and Rayleigh-Jeans law arose due to the consideration of quan-
tized oscillators and Maxwell’s distribution function in former rather than consideration of classical
distribution of energy in the later. Planck considered quantized oscillator which have energy,

E = nhν (3.23)

where n is an integer. This equation shows that the energy absorbed in walls of the cavity or emitted
must be discretized. This discretized energy must be considered in the evaluation of average energy.
The average energy of quantized oscillator is given as,
P∞ −nhν/kT
n=0 (nhν)e
¯ = P∞ −nhν/kT (3.24)
n=0 e

Let x = hν/kT , the formula of average energy with new variables becomes,
P∞ −nx
n=0 nxe
¯ = kT P ∞ −nx
. (3.25)
n=0 e

6
Since

d −n x X 1
(e ) = − n e−n x and e−nx = , (3.26)
dx 1 − e−x
n=0
one can write,
d X −nx  e−x
e = − (3.27)
dx (1 − e−x )2
and equation 3.25 as,
P∞ −nx ) d P∞
d
n=0 x dx (e e−nx
¯ = − kT P∞ −nx = − x kT P∞n=0 −nx .
dx
(3.28)
n=0 e n=0 e

Substituting values of numerator and denominator from 3.26 and 3.27, we get,

e−x e−x
   
1 1
¯ = x kT / = x kT = x kT x . (3.29)
(1 − e−x )2 1 − e−x (1 − e−x ) (e − 1)

Substituting the value of x, one get the desired average energy of quantized oscillator as,

¯ = (3.30)
ehν/kT −1
Thus according to Planck, the energy density of standing waves in the range ν to ν + dν is given by,

8π h ν 3
ρ(ν, T ) dν = dν Planck radiation formula (3.31)
c3 (ehν/kT − 1)

4 Some calculations based on Planck’s formula

4.1 Stefen-Boltzmann law


Integrating Planck’s radiation formula 3.31 in the frequency range 0 to ∞1 ,
Z ∞
c
R(T ) = ρ(ν, T )dν
0 4
Z ∞
c 8π h ν 3
= dν. (4.1)
0 4 c3 (ehν/kT − 1)

Let hν/kT = x ⇒ h dν/kT = dx, this substitution leads to

2π k 4 T 4 ∞ x3
Z
R(T ) = dx. (4.2)
c2 h3 0 ex − 1

Substituting the value of integral,



x3 π4
Z
dx = ,
0 ex − 1 15
one gets the total radiated energy,

2 k4 π5
 
R(T ) = T 4. (4.3)
15c2 h3
2 k4 π 5
The term in the parentheses is constant ant let it be σ(= 15c2 h3
), the total radiated energy becomes,

R(T ) = σ T 4 (4.4)

which is Stefen-Boltzmann law of radiation.


1
formula relating spectra radiancy and energy density is R(ν, T ) = 4c ρ(ν, T ) which on integrating gives relation
between energy radiated per unit area per second and total energy density.

7
4.2 Wien’s displacement law
To obtain Wien’s formula from Planck’s distribution, one have to express Planck’s energy spectrum
in terms of wavelength rather than frequency. The quantity ρ(λ, T ) is defined from the equality
ρ(λ, T ) dλ = −ρ(ν, T ) dν. The minus sign indicates that, though ρ(λ, T ) and ρ(ν, T ) are both positive,
dλ and dν have opposite signs. (An increase in frequency gives rise to a corresponding decrease in
wavelength.) From relation ν = c/λ, one have dν/dλ = −(c/λ2 ), so that

dν c
ρ(λ, T ) = − ρ(ν, T ) = ρ(ν, T ) 2 (4.5)
dλ λ
Now substituting value of ρ(ν, T ) from 3.31, one obtains

8π h c dλ
ρ(λ, T ) dλ = . (4.6)
λ5 ehc/λkT − 1
Differentiating distribution function and equating it to zero for maximum value,

d 8π h c h c ehc /λkT − 5λkT ehc/λkT − 1
ρ(ν, T ) = = 0 (4.7)

2
dλ λm kλ7 T ehc /λkT − 1
λm

one gets,
h c hc /λm kT  
e − 5 ehc/λm kT − 1 = 0 (4.8)
λm kT
Let define x = hc /λm kT , above equation becomes,
5
x ex − 5(ex − 1) = 0 ⇒ ex = , (4.9)
5−x
which is non algebraic. The solution of this equation leads to the value x ≈ 4.965. Thus
hc hc
= 4.965 ⇒ λm T = 4.965 (4.10)
λm kT k
Since k (Boltzmann constant), h (Planck constant) and c (speed of light) are constant, the right hand
of above equation is constant, i.e. product of the wavelength at which the spectral radiancy is a
maximum corresponding temperature is constant (Wien’s displacement law).

5 Wave particle duality and de Boglie matter waves

In 1924, a French graduate student Louis de Broglie, proposed in his dissertation thesis that the
behaviour of matter possesses dual nature i.e. matter also behaves as waves. His speculation was
particular for electrons, but it is characteristic of matter. The originality of his idea was paid attention
in the physics world but lack of experimental evidence, de-Broglie’s ideas were not considered to have
any physical reality. Later, Einstein and Paul Langevin recognized the importance of de Broglie’s
ideas and draw the attention of other physicists. After some years, C. J. Davisson and L. H. Germer
and G. P. Thomson verified the de Broglie hypothesis directly by observing interference patterns, a
characteristic of waves, with electron beams.
Since our universe is entirely composed of matter and radiation, de-Broglie’s views are consistent
with the symmetries of spacetime. de Broglie proposed that the wave aspects of matter were related
quantitatively to the particle aspects in exactly the same way that we found for radiation, namely,
E = hν and p = h/λ. The first equation,
E = hν (5.1)
relates the total relativistic energy of the entity to the frequency ν of the wave associated with its
motion, while the second equation,
p = h/λ (5.2)

8
shows the momentum p of the entity associated with the wavelength λ. Above two equations show
that how two quantities E and p with particle concept are related to the wave concepts frequency ν
and wavelength λ. Radiation corresponds to particles of zero rest mass (moving at speed c), whereas
matter corresponds to particles of finite rest mass (moving at speeds less than c), but the same general
transformation properties apply to both entities. The de Broglie wavelength sis defined as
h h
λ = = , (5.3)
p mv
where m is relativistic mass of the particle. In low speed classical limit, the relativistic mass can be
replaced by rest mass m0 , i.e. λ = h/m0 v.

5.1 Different forms of de Broglie’s formula


1) Kinetic energy of material particle,

1 p2
E = m v2 = (5.4)
2 2m
leads to wavelength of material particle as,
h
λ = √ (5.5)
2mE
2) Kinetic theory of gases gives the formula for average energy as
3 1
E = kT = m v 2 (5.6)
2 2
This formula leads to,
h
λ = √ (5.7)
3mkT
3) If we accelerate electrons through a potential difference V , then the kinetic energy of electrons is
given by,
1
e V = m v2 (5.8)
2

which leads to momentum as p = 2mev. Now the formula for wavelength is given as,
h
λ = √ (5.9)
2meV
5.2 How fast do de Broglie waves travel?
Since de Broglie waves are associated with moving bodies, one assume that it has same velocity as
that of the body, but it is not true. Let us see how this happen. If we call the de Broglie wave velocity
vp , we can apply the usual formula,
vp = ν λ (5.10)
to evaluate vp . The wavelength λ = h/γmv is de Broglie wavelength. The quantum expression for
energy E = hν and relativistic expression for energy E = mc2 are equated and one gets,

h ν = γ m c2
γ m c2
ν = . (5.11)
h
Thus substituting values of de Broglie wavelength and ν from above equation in 5.10, one gets,

γ m c2 c2
  
h
vp = = (5.12)
h γ mv v

9
Because the particle velocity v must be less than the velocity of light c, the de Broglie waves always
travel faster than light! In order to understand this unexpected result, we must look into the dis-
tinction between phase velocity and group velocity. (Phase velocity is what we have been calling
wave velocity.)

5.3 Mathematical review of waves


In this subsection we review the mathematical description of waves. A wave can be written as a
periodic function of time, i.e.
y = A cos 2π ν t. (5.13)
Now, if the wave is propagating with velocity vp , we can write the variation of y as,
 
νx
y = A cos 2π ν t − . (5.14)
vp

At x = 0, above equation gives the equation 5.13.


Since speed of the wave is related with frequency
and wavelength by relation vp = νλ, we can write
above equation as,
 x
y = A cos 2π ν t − . (5.15)
λ
In terms of angular frequency ω(= 2πν) and
wave number k(= 2π/λ = ω/vp ), we can write
wave equation as,

y = A cos(ω t − k x) (5.16)

5.4 Phase velocity and group velocity


Amplitude of de Broglie waves ⇒ probability of
finding a moving body at a particular place and
Figure 7: Two waves of slightly different wavelength
time. de Broglie wave are not simple waves but
and frequency produce beats. (a) Shows at a given
are wave packets or wave group related to moving instant for each of the two y(x) waves. (b) The sum
body. Wave group is superposition of individ- of these waves. The spatial extent of the group ∆x is
ual waves of different wavelengths which cause in inversely proportional to the difference in wave numbers
the variation of amplitudes. ∆k. Identical figures are obtained if y is plotted versus
time t at a fixed point x. In that case the extent in time
Phase velocity: velocity of individ- ∆t is inversely proportional to the frequency difference
ual waves of wave group. ∆ω.
Group velocity: velocity of wave
packet or group which is consist
of de Broglie waves.

Let us suppose wave group arises from combination of two waves,

y1 = A cos (ω t − k x) (5.17)
y2 = A cos ((ω + ∆ω)t − (k + ∆k) x) . (5.18)

The resultant wave is compose of above two waves and given as,

y = y1 + y2 . (5.19)

Using the identity cos C + cos D = 2 cos(C + D)/2 cos(C − D)/2, we get
1 1
y = 2A cos ((2ω + ∆ω)t − (2k + ∆k) x) cos (∆ω t − ∆k x) . (5.20)
2 2

10
Since ∆ω and ∆k are small compared to ω and k respectively, we can neglect them in the first term
of the above equation. Now with this approximation above equation reduces to,
1
y = 2A cos (ω t − k x) cos (∆ω t − ∆k x) . (5.21)
2
Equation 5.21 represents a wave of angular frequency ω and wave number k that has been superim-
posed with a modulation wave of angular frequency ∆ω/2 and of wave number ∆k/2. The modulation
of is to produce successive wave groups. The phase velocity is,
ω
vp = (5.22)
k
and group velocity,
∆ω dω
vg = in the continuum limit vg = . (5.23)
∆k dk
Now for de Broglie waves, we have,

c2 2 π m c2
ω = 2πν = 2πγ m = p (5.24)
h h 1 − v 2 /c2

and wave number,


2π v 2πmv
k = = 2πγ m = p . (5.25)
λ h h 1 − v 2 /c2
Definition of group velocity
dω dω/dv
vg = =
dk dk/dv
leads to
c2
vg = v and vp = .
v
The derivatives are calculated as,
dω 2 πm v
=
dv h(1 − v 2 /c2 )3/2
dk 2 πm
=
dv h(1 − v 2 /c2 )3/2

5.5 Relation between group and phase velocity


vg = and angular frequency ω = vp k
dk
Thus
d dvp
vg = (vp k) = vp + k .
dk dk
Since k = 2π/λ ⇒ dk = −(2π/λ2 ) dλ, we get above eqaution as,
vp
vg = vp − λ (5.26)

5.6 Davisson-Germer experiment
In 1927 Clinton Davisson and Lester Germer in the United States
and G. P. Thomson in England independently confirmed de Broglie’s
hypothesis by demonstrating that electron beams are diffracted when
they are scattered by the regular atomic arrays of crystals. (All three Figure 8: The schematic dia-
received Nobel Prizes for their work). We are going to look at the gram of Davisson-Germer exper-
experiment performed by Davisson and Germer as its interpretation iment.
is more direct. Davisson and Germer were studying the scattering of
11
Figure 9: Results of the Davisson-Germer experiment, showing how the number of scattered electrons varied
with the angle between the incoming beam and the crystal surface. The Bragg planes of atoms in the crystal
were not parallel to the crystal surface, so the angles of incidence and scattering relative to one family of these
planes were both 65o .

electrons from a solid using an apparatus like that sketched in Fig 8. The energy of the electrons
in the primary beam, the angle at which they reach the target, and the position of the detector
could all be varied. Classical physics predicts that the scattered electrons will emerge in all directions
with only a moderate dependence of their intensity on scattering angle and even less on the energy
of the primary electrons. Using a block of nickel as the target, Davisson and Germer verified these
predictions. In the midst of their work an accident occurred that allowed air to enter their apparatus
and oxidize the metal surface. To reduce the oxide to pure nickel, the target was baked in a hot
oven. After this treatment, the target was returned to the apparatus and the measurements resumed.
Now the results were very different. Instead of a continuous variation of scattered electron intensity
with angle, distinct maxima and minima were observed whose positions depended upon the electron
energy! Typical polar graphs of electron intensity after the accident are shown in Fig. 9. The
method of plotting is such that the intensity at any angle is proportional to the distance of the
curve at that angle from the point of scattering. If the intensity were the same at all scattering
angles, the curves would be circles centered on the point of scattering. Two questions come to
mind immediately: What is the reason for this new effect? Why did it not appear until after the
nickel target was baked? De Broglie’s hypothesis suggested that electron waves were being diffracted
by the target, much as x-rays are diffracted by planes of atoms in a crystal. This idea received
support when it was realized that heating a block of nickel at high
temperature causes the many small individual crystals of which it is
normally composed to form into a single large crystal, all of whose
atoms are arranged in a regular lattice.
Now let us analyze the result of Davisson-Germer
experiment. For 54 eV beam electron beam, a sharp
peak is observed at an angle 50◦ from the incident
beam. This shows that the beam form a diffraction
pattern from crystal planes as shown in the fig 10. The
Bragg equation for maxima in the diffraction pattern
is,
n λ = 2d sin θ. (5.27)
The inter planar spacing d = 0.091nm and n = 1 for
principle maxima. From fig. 10, we can see,

2θ + φ = 180
2θ + 50 = 180
⇒ θ = 65 Figure 10: Scattering of electrons by a crystal.
Electron waves are strongly scattered if the Bragg
Thus, the experimental value of the de Broglie wave- condition is met.

12
length λ of the diffracted electrons is,

λ = 2 × 0.091 × sin 65 ⇒ λ = 0.165 nm (5.28)

Now we use de Broglie’s formula λ = h/γmv to find the expected wavelength of the electrons. The
kinetic energy of electron is 54 eV, which is small as compared to its rest
√ mass energy 0.51 MeV, so
we can take γ = 1. Since kinetic energy K = mv 2 /2, we get p = mv = 2mK. Use of this relation
in de Broglie’s fromula, we get,
λ = 0.166 nm. (5.29)
These observations shows that there is an excellent agreement of electron wavelength between exper-
imental value (calculated by Davisson-Germer) and theoretical value (by de Broglie’s hypothesis).

6 Bohr’s postulates

Study of atomic absorption and emission spectra and many other predictions are explained by Bohr.
The basic postulates of Bohr’s are,

1. Coulomb interacion is the cause of the circular orbits of the electrons around the nucleas. This
Coulomb interactio provides the necessary centripetal force (classical one) for the circular orbits.

2. Accordng to classical mechanics, infinite number of circular orbits ae possible, but electrons
are bound to move in orbits which have angular momentum L as integral multiple of Planck’s
constant h divided by 2π.

3. Electrons moving in these fixed orbits do not radiate although they are accelerated, i.e. there
total energy is fixed.

4. If an electron jumps form lower to higher energy states or from higher to lower energy states,
the energy absorbed or emitted is given by the relation,

E = hν = Ef − Ei (6.1)

where Ei and Ef are initial and final energy states and ν is frequency of emitted or absorbed
radiation.

6.1 Bohr’s model


Let us consider a simple model of atom with nuclear charge Ze and mass M and a electron of charge
−e. We assume that the electron revolves in a circular orbit about the nucleus. Initially we suppose
that the mass of the electron is neglible as compared to that of the nucleas and the nucleus remains
fixed in space. The condition of mechanical stability of the electron is,

1 Ze2 v2
2
= m (6.2)
4π0 r r
where v is the speed of the electron in its orbit, and r is the radius of the orbit. The left side of
this equation is the Coulomb force acting on the electron, and the right side is ma, where a is the
centripetal acceleration keeping the electron in its circular orbit. Now, th eorbital angular momentum
must be constant and given by the relation,

L = mvr = n~ n = 1, 2, 3, . . . (6.3)

Substitutig the solution of v from above equation in 6.2, we get,

n2 ~2
 
2 2 n~
Ze = 4π0 mv r = 4π0 mr = 4π0 (6.4)
mr mr

13
which leads to
n2 ~2
⇒ r = 4π0 (6.5)
mZe2
n~ 1 Ze2
and v = = (6.6)
mr 4π0 n~
The kinetic energy and potential energy of the atom is written as,

1 1 Ze2
K = mv 2 = (6.7)
2 8π0 r
1 Ze2
V = − (6.8)
4π0 r
Thus the total energy of the atom is,
2
1 Ze2 1 m Ze2
E = K +V = − = − (6.9)
8π0 r n2 (4π0 )2 2~2

The equations 6.5, 6.6 and 6.9 shows the dependence of various quabtities on n. This dependence
of r, v and E on n is due to quantization of angular momentum. From equation 6.5, we se that the
radius is diectly propotional to n2 and from 6.9, the total energy is inversely propotional n2 . This
model of atoms explain correctly hydrogen like atoms but fails for more complex atoms.

7 Heisenberg’s uncertainty principle

According to classical physics, given the initial conditions and the forces acting on a system, the
future behavior (unique path) of this physical system can be determined exactly. That is, if the
initial coordinates r~0 , velocity v~0 , and all the forces acting on the particle are known, the position
~r(t) and velocity ~v (t) are uniquely determined by means of Newton’s second law. Classical physics is
thus completely deterministic.
But this determnistic behaviour of claasical mechanics does not holds at microscopic level. The
reason behind this is the concept of wave particle duality. At quantum level, this duality plays an
important role as particles are represented by means of a wave function corresponding to the parti-
cleâĂŹs wave, and since wave functions cannot be localized, then a microscopic particle is somewhat
spread over space (unlike classical particle which is localized). The classical concepts of exact position,
exact momentum, and unique path of a particle therefore make no sense at the microscopic scale.
This is the essence of Heisenberg’s uncertainty principle.

If the x-component of the momentum of a particle is measured with an uncer-


tainty ∆px , then its x-position cannot, at the same time, be measured more
accurately than ∆x = ~/(2∆px ).

Other forms of uncertainty principle


8 Wave function

The idea that the electron’s orbits in atoms correspond to standing matter waves was taken by
Schrödinger in 1926 to formulate wave mechanics. The basic quantity in wave mechanics is the wave
function Ψ(~r, t), which measures the wave disturbance of matter waves at time t and at a point ~r.

8.1 Physical meaning of the wave function


The wave function of a particle describes the probability distribution of a particle in space, just as
the wave function of an electromagnetic field describes the distribution of the EM field in space.
Study of intereference and diffraction pattern tells us that the intensity of the interference pattern
is proportional to the square of the field amplitude (the probability that the waves interfere positively

14
or negatively at some points). Maing same analogy, Born suggested that the quantity |Ψ(~r, t)|2 at
any point ~r is a measure of the probability density that the particle will be found near that point.
Since |Ψ(~r, t)|2 dV is interpreted as the probability, it is normalized to one as,
Z Z
|Ψ(~r, t)|2 dV = Ψ∗ (~r, t)Ψ(~r, t)dV = 1 (8.1)
V V

which shows that the particle will be found anywhere in the volume considered. The one dimensional
form of the above equation can be written as,
Z +∞ Z +∞
2
|Ψ(x, t)| dx = Ψ∗ (x, t)Ψ(x, t)dx = 1 (8.2)
−∞ −∞

with the probability density of fnding the particle between x and x + dx is |Ψ(x, t)|.
Wave functions are usually complex with both real and imaginary parts. A proba- bility, however,
must be a positive real quantity. Generally one can write a wave function as

Ψ = A + iB and conjugate Ψ∗ = A − i B (8.3)

which leads to probability amplitude Ψ∗ Ψ = A2 + B 2 is real.


Wave function Ψ(~r, t) of a particle is mathematical construct only. Only |Ψ(~r, t)|
has physical meaning - probability density, and |Ψ(~r, t)|2 dV is the probability
of finding the particle in the volume dV .

8.2 Properties of wave function


A well behaved wave function represents a real body. The properties of well behaved wave functions
are summarized as,
• Ψ must be continuous and single-valued everywhere.
• Derivatives (∂Ψ/∂x, ∂Ψ/∂y ∂Ψ/∂z) must be continuous and single-valued everywhere.
• Ψ must be normalizable, which means that Ψ must go to 0 as x ± ∞, y ± ∞, z ± ∞ in order
that |Ψ(~r, t)|2 dV over all space be a finite constant.

9 Expectation value
The average position x̄ of a number of identical particles distributed along the x axis in such a way
that there are N1 particles at x1 , N2 particles at x2 , and so on? The average position in this case is
the same as the center of mass of the distribution, and so,
P
N1 x1 + N2 x2 + N3 x3 + . . . Ni xi
x̄ = = P (9.1)
N1 + N2 + N3 + . . . Ni

When we are observing the behaviour of a single particle, we must replace the number Ni of particles
at xi by the probability Pi that the particle be found in an interval dx at xi . The probability is,

Pi = |Ψi |2 dx = Ψi ∗ Ψi dx (9.2)

where Ψi is wave function at position xi . Making this substitution and changing the summations to
integrals, we see that the expectation value of the position of the single particle is,
R +∞ 2
R +∞ ∗
−∞ x|Ψ| dx Ψ xΨ dx
hxi = R +∞ = R−∞ +∞ . (9.3)
2 ∗
−∞ |Ψ| dx −∞ Ψ Ψ dx

If our wave function is properly normalized, we can write,


Z +∞ Z +∞
hxi = 2
x|Ψ| dx = Ψ∗ xΨ dx (9.4)
−∞ −∞
15
10 Operator

An operator tells us what operation to carry out on the quantity that follows it. The operator d/dx
tells us that the we have to take the first derivative of the quantity that it follows. The energy and
momentum operator in quantum mechanics are defined as,
∂ ~ ∂
Ê → i~ and p̂ → (10.1)
∂t i ∂x
which applied on wave function in the manner,
∂ ~ ∂
ÊΨ = i~ Ψ and p̂Ψ = Ψ (10.2)
∂t i ∂x
10.1 Operators and Expectation Values
Since p and E can be replaced by their corresponding operators in an equation, we can use these
operators to obtain expectation values for p and E. Thus the expectation value for p is
Z +∞ Z +∞
~ +∞ ∗ ∂
 
∗ ~ ∂
Z
2
hpi = p̂|Ψ| dx = Ψ Ψ dx = Ψ Ψ dx (10.3)
−∞ −∞ i ∂x i −∞ ∂x

and the expectation value for E is


Z +∞ +∞   Z +∞
∂ ∂
Z

hEi = 2
Ê|Ψ| dx = Ψ i~ Ψ dx = i~ Ψ∗ Ψ dx (10.4)
−∞ −∞ ∂t −∞ ∂t

In the case of algebraic quantities such as x and V (x), the order of factors in the integrand is unim-
portant, but when differential operators are involved, the correct order of factors must be observed.
We can write the expectation value of any general operator Ĝ as,
Z +∞ Z +∞
hGi = Ĝ|Ψ|2 dx = Ψ∗ ĜΨ dx (10.5)
−∞ −∞

11 Schrödinger equation

Schrödinger’s equation, which is the fundamental equation of quantum mechanics in the same sense
that the second law of motion is the fundamental equation of Newtonian mechanics, is a wave equation
in the variable. The one dimentional wave equation is written as,

∂2 1 ∂2
Ψ(x, t) = Ψ(x, t) (11.1)
∂x2 v 2 ∂t2
which has general solution of form,

Ψ(x, t) = A e−iω(t−x/v) . (11.2)

Using values E = hν and equation of de Broglie wave p = h/p, we can write the solution of wave
equation as,
h
Ψ(x, t) = A e−i/~(Et−px) where ~ = (11.3)

11.1 Time dependent
Taking the time derivative of 11.3, we get,
 
∂ iE
Ψ(x, t) = − A e−i/~(Et−px)
∂t ~
 
iE ∂
= − Ψ(x, t) ⇒ EΨ(x, t) = i~ Ψ(x, t) (11.4)
~ ∂t

16
which shows that the energy operator E is i~∂/∂t.
Again differentiating the 11.3 w.r.t. s twice, we get,
2
∂2

ip
Ψ(x, t) = A e−i/~(Et−px)
∂x2 ~
p2 ∂2
= Ψ(x, t) ⇒ p2 Ψ(x, t) = − ~2 Ψ(x, t) (11.5)
~2 ∂x2
At speeds small compared with that of light, the total energy E of a particle is the sum of its kinetic
energy p2 /2m and its potential energy U , where U is in general a function of position x and time t:

p2
E = + U (x, t) (11.6)
2m
Multiplying both sides of above equation by the wave function Ψ(x, t) gives,

p2
EΨ(x, t) = Ψ(x, t) + U Ψ(x, t) (11.7)
2m
Substituting values from 11.4 and 11.5, we get,

∂ ~2 ∂ 2
i~ Ψ(x, t) = − 2
Ψ(x, t) + U Ψ(x, t)
∂t  2m2∂x 2 
~ ∂
= − + U Ψ(x, t) (11.8)
2m ∂x2

which is time dependent Schrödinger equaton. The three dimensional form is given as,

~2
 2
∂2 ∂2
  
∂ ∂
i~ Ψ(x, t) = − + + + U Ψ(x, t) (11.9)
∂t 2m ∂x2 ∂y 2 ∂z 2

11.2 Time independent


In many situations the potential energy of a particle does not depend on time explicitly i.e. it
varis with the position of the particle only. When this is true, SchrÃűdingerâĂŹs equation may be
simplified by removing all dependencis on time t.
We begin by noting that the one-dimensional wave function Ψ(x, t) of an unrestricted particle
may be written as

Ψ(x, t) = A e−i/~(Et−px) = A e(i/~)px e(−i/~)Et = ψ(x)e−(i/~)Et . (11.10)

It is evident fro above equation that Ψ is product of two functions, one is space (position) dependent
part ψ and second one is time dependent part e−i/~Et . Substituting 11.10 in 11.8,

~2 ∂ 2
 
∂  −(i/~)Et

i~ ψ(x)e = − + U ψ(x)e−(i/~)Et
∂t 2m ∂x2
~2 ∂ 2
 
−(i/~)Et
⇒ Eψ(x)e = − + U ψ(x)e−(i/~)Et (11.11)
2m ∂x2

cancelling same terms on the both sides, we get time independent SchrÃűdingerâĂŹs equation for
ψ(x) as,
~2 ∂ 2
 
Eψ(x) = − + U ψ(x). (11.12)
2m ∂x2
After some simple rearrangement of terms, one can write above equation as,

∂2 2m
2
ψ(x) + 2 (E − U )ψ(x) = 0 (11.13)
∂x ~

17
For a free particle, potential energy U is zero, the equation further reduces to,

∂2 2m
2
ψ(x) + 2 Eψ(x) = 0 (11.14)
∂x ~
11.3 Particle in box
The simplest quantum-mechanical problem is that of a par- ∞
ticle trapped in a box with infinitely hard walls. The in-
finitely hard walls here repesents that the probabillity of
U
finding particle outside the box is zero. We may specify the
particle’s motion by saying that it is restricted to traveling
along the x axis between x = 0 and x + L by infintely hard
walls. A particle does not lose energy when it collides with
such walls, so that its total energy stays constant. The
restrictions imposed by the boundary conditions are,
0 L X
U = 0 for 0 < x < L
U = ∞ for x ≤ 0 and x ≥ L.
Since particle can not exist outside the box, wave function Figure 11: A square potential well with
infinitely high barriers at each end corre-
ψ is zero outside the box. Within the box Schrödinger’s
sponds to a box with infinitely hard walls.
equation for free particle becomes

d2 2m
ψ(x) + 2 Eψ(x) = 0. (11.15)
dx2 ~
or r
d2 2 2mE
ψ(x) + K ψ(x) = 0 where K = (11.16)
dx2 ~
We have replaced partial derivatives by total derivatives as our wave function ψ is a function only of
x in this problem. The solution of equation 11.16 is,

ψ(x) = A sin (Kx) + B cos (Kx) (11.17)

which can be verified easily.


The particle is strictly confined in the boundaries of the box, therefore we have boundary condi-
tions,

ψ(x) = 0 at x = 0 (i)
ψ(x) = 0 at x = L (ii)

The first boundary codition leads us to B = 0 as sin 0 = 0. The second condition gives,

sin(KL) = 0 ⇒ KL = nπ. (11.18)

Substituting the value of k, we get


r
2mEn n2 π 2 ~2
L = nπ ⇒ En = . (11.19)
~ 2mL2
In above equation En denotes the energy corresponding to each value of n.
The wave functions of a particle in a box whose energies are En are, from Eq. 11.16 with B = 0.
r
2mE
ψn (x) = A sin(Kx) with K = (11.20)
~
substiuting value of K from Eq. 11.18, we get

ψn (x) = A sin x (11.21)
L
18
(a) for wave function (b) for probability density

Figure 12: Wave functions and probability densities of a particle confined to a box with rigid walls.

It is easy to verify that these eigenfunctions meet all the requirements discussed for well behaved
wave function. Furthermore, the integral of |ψn |2 over all space is finite, as we can see by integrating
|ψn |2 dx from x = 0to x = L (since the particle is confined within these limits). Using trigonometric
identity sin2 θ = 1−cos
2
θ
, we get
Z +∞ Z L
2
|ψn | dx = |ψn |2 dx
−∞ 0
L
nπx L2
Z
= A2 sin2 dx = A2 . (11.22)
0 L 2

To normalize the wave function, we must assign a value to A such that the probability |ψn |2 dx of
finding the particle in the beween 0 to L is finite and must be equal to one, therefore,
Z +∞ Z L r
2
2 2 2L 2
|ψn | dx = |ψn | dx = A = 1 ⇒ A = (11.23)
−∞ 0 2 L

Thus, our final normalized wave function can be written as,


r
2 nπ
ψ(x) = sin x where n = 1, 2, 3, . . . (11.24)
L L

The normalized wave functions ψ1 , ψ2 and ψ3 together with the probability densities |ψ1 |2 , |ψ2 |2 and
|ψ3 |2 are plotted in Figure 12. Although n may be negative as well as positive, |ψn |2 is never negative
and, since ψn is normalized, its value at a given x is equal to the probability density of finding the
particle there. In every case |ψn |2 = 0 at x = 0 and x = L, the boundaries of the box.

12 Eigen values and Eigen functions

A operator Ô transforms a function in way,

Ôf (x) = λf (x) (12.1)

19
the function f (x) is said to be eigen function and the constant value λ is eigen value.

References
[1] Basic Concepts in Relativity and Early Quantum Theory, Second Edition; Robert Resnick and David
Halliday; John Wiley and Sons.
[2] QUANTUM PHYSICS of Atoms, Molecules, Solids, Nuclei, and Particles, Second Edition; Robert
Eisberg; John Wiley and Sons.
[3] Concepts of Modern Physics, Sixth Edition; Arthur Beiser; The McGraw-Hill Company.
[4] Lecture notes of various authors available freely on web.

20

You might also like