You are on page 1of 13

188 Atomic Radiation Chap.

orthogonal to the source, and the field of view (FOV) is limited to match the incoming beam.
Thus the output consists of the input intensity plus that added by the following processes:

1. Stimulated emission: the amount of radiation stimulated by the incoming wave. Since
this is stimulated emission, the frequency, phase, and direction of the added signal
are the same as the incoming wave (this is indicated by process 1 on the energy-level
diagram)
minus:
2. Absorption: the amount of radiation absorbed by the atoms in state 1
plus:
3. Spontaneous emission: the amount of radiation emitted spontaneously by the atoms in
state 2 in the direction ofthe input wave and in the same frequency within bandwidth
~v of our detector (this is indicated by the wavy line going from state 2 to state 1).

Now the easiest way to do the bookkeeping on these processes is to arrange the
physical factors into the vertical columns. We recognize that each transition caused by the
foregoing processes contributes (or subtracts) a package of energy hv (column 1), which,
when multiplied by the rate per atom, is the power contributed by each atom (column 2), but
we also recognize that this contribution (or interaction) scales according to the line shape
(column 3). Column 4 is the probability that the photon involved has the proper polarization,
and column 5 is the probability that it is within the solid angle specified by the experimental
setup (which is heavily weighted favoring the first two processes).
However, some of the spontaneous power reaches the detector. The radiation from the
slab is spread uniformly into 4:rr steradians, or equivalently into the various electromagnetic
modes defined by the beam size and the length Az. Hence, the detector, with an acceptance
cone of dQ, will only collect a fraction, dQ/4:rr, ofthe power radiated by these atoms, and
only one half of that fraction has the proper polarization. Finally, column 6 is the number
of atoms involved in the interaction. The arrangement is given in (7.5.1).

2 3 4 5 6
t,
~Iv = +hv X B2 1 - - X g(v) X X x N2~Z
(c/n g )
t; (7.5.1a)
-hv x B 12 - - x g(v) X x X NI~Z
(c/n g )
1 dQ
+hv x A21~V x g(v) X - X X N2~Z
2 4:rr
Further manipulation yields

~I" d I;
-- ---? - [ -hv- (B 21N2 - B 12N I)g(v) ] Iv
~z dz (c/n g )

+ -1 [ hvA2IN2g(V)~V-
dQ] (7.5.1b)
2 4:rr
Sec. 7.5 Amplification by an Atomic System 189

We can readily appreciate that this last term can be called "noise," since this signal is present
at the detector without any input I; to the slab. Even though it is essential to the laser (to
initiate the oscillation), its presence is not required at this time and we neglect it for now.
It is convenient to use some of the relations found in Sec. 7.3 to change the appearance
of (7.5.1b). Recall that B I 2I B21 = g21gl and A21! B21 = 8Jrn 2nghv 3 le 3 • Thus, the basic
equation becomes

(7.5.2)

The intensity increases with distance if all terms on the right-hand side of (7.5.2) are positive
and thus the coefficient y (v) is called the gain coefficient and (7.5.2) is the central equation
of laser theory and thus treated with due respect. (Memorize it!). If the intensity does not
change the population density of states 2 and I, then it is a "small signal," and a subscript
is attached to y(v) ---* Yo (v) to indicate that fact. All lasers start with a small signal value,
but eventually the intensity does affect the populations, and thus N2 and N I are implicit
functions of L; That situation will be addressed in the next chapter.
All terms on the right are always positive other than the difference in populations,
and thus

for gain (7.5.3)

This is an inversion of the normal state of affairs such as that given by the Boltzmann
relationship* and points out a fundamental requirement for amplification. If the population is
"normal" and N 2 < (g21gl)N I , then dIvldz is negative and (7.5.2) predicts the absorption
coefficient. t Much of our immediate surroundings are in a quasinonnal state of affairs, and
thus absorption is the rule. Thus, (7.5.2) applies to all cases in which photons are added to
or subtracted from any electromagnetic wave at any frequency,"
The first bracket in the brace is the stimulated emission cross section, as defined by
(7.4.7) and is a collection of atomic constants describing the atom (through A and A 21)
and its environment through g(v). It is a most convenient specification of a fundamental
parameter of the system, since one merely multiplies this "area" by the effective inversion
density N2 - (g21gl)N I to obtain the gain coefficient. If, for example, the peak value of
a (vo) were 10- 16 em? (typical), then it requires a net inversion of 10 14 cm- 3 to obtain a gain
coefficient of 0.01 cm' (or l%/cm). Sometimes the absorption cross section is specified,

*Ifwe use the Boltzmann factor to describe N,j N 1 = (g2/g1) exp[ -hv/ kT], then the temperature T must
be negative to obtain gain.
t (In the older literature, the absorption coefficient is called the "extinction" coefficient and is abbreviated
by k(v). A formula such as (7.5.2) is in books printed in 1934. See Ref. [1).)
'The case of the "scattering" photons out of a beam does not quite fit the details of (7.5.2). However, the
generic equation still applies; namely, dt id; = (Na sc,,) . I where ascot would be the scattering cross section.)
190 Atomic Radiation Chap. 7

which is related to the stimulated emission value by


aabs(V) = (g2/gJ)a stim(v) (7.5.4)
Unless the subscript on a is indicated explicitly, we will always specify the stimulated
value.
Having found the differential form of the gain, we can integrate (7.5.2) to obtain
Iv(z) = Iv (0) exp(yo(v)z]
= Go(v)/v(O) (7.5.5)
Go = exp[yo(v)d]
where Go is the small signal (power) gain of an amplifier of length d.
It is most important to remember that the gain coefficient is frequency dependent;
consequently, the gain Go is even more so, since Yo(v) appears in the exponent. Thus we
have a narrow bandpass amplifier. In Chapter 8 we put this narrow band amplifier into a
feedback loop to obtain oscillation.
There is another logic path that can be used to arrive at (7.5.2). It requires us to
translate the differential equation definition of the gain coefficient into simple words. The
gain coefficient is defined by

y(v) =L. II (ddIZv) (7.5.6)


v

which translates to
L. (dlv/dz] = net power emitted per unit of volume
. that vo Iume
y(v) = 1v = power per unit. 0 f area traversing

Note the net power emitted per unit of volume is equal to h v times the number of transitions,
per unit of volume, per unit of time by stimulated emission minus that by absorption. The
last tenn in (7.4.6) is that difference, hence

v
y(v) = 2-
~
. hv { dN21
dt
= a(v)/ [N2 - (g2/gJ)N1J}
hv
(stim-abs)

=a(V)[N2- ~~Nl]
which is identical to (7.5.2). There is not any real difference in the two derivations, but it
is convenient to have the different pictures in mind. In the first, we should have the mental
picture provided by Fig. 7.7 where the optical beam is highlighted. In the second approach
denoted by (7.5.6), we are observing the atoms' response to the optical beam. In either case,
we are merely conserving total energy, the sum of that carried by the beam and that stored
in the atoms.
Sec. 7.6 Broadening of Spectral Lines 191

7.6 BROADENING OF SPECTRAL LINES

There are many factors that affect the shape and width of a line and thus affect the gain
coefficient. The purpose of this section is to identify the various classifications of broadening
and, where possible, provide an analytic expression for g (v). Sometimes, however, we must
wave the white flag and give up on an analytic approach, because of incomplete knowledge
or because of the excessive complexity of the mathematics. Then we use a measurement of
the spectral shape of the spontaneous emission, absorption, or gain coefficient and apply
the normalization condition (7.4.2) to evaluate its peak value.
All transitions have a finite width if, for no other reason, to ensure compliance with
the uncertainty principle. There are two generic classifications of processes that contribute
to the width of a spectral line. They are

1. Homogeneous broadening: a mechanism that applies to all atoms.


2. Inhomogeneous broadening: one that is caused by some identifiable difference
between groups of atoms.

In any practical case, we usually have a mixture of both types and the combination is almost
always intractable or not worth the mathematical effort for the minimal gain in precision.
In such cases, we choose the dominant one and proceed.

7.6.1 Homogeneous broadening mechanisms

Lifetime broadening. We have already argued in Sec. 7.4 that a perfectly


sharp energy level is impossible, and thus there is a distribution of transitions from (oc-
cupied) states near E2 to (empty) states" near EI and, thus, there is a spectral width
associated with the state 2 ---? 1 "line" emission. Let us consider an important special case
of symmetrical level broadening that illustrates the procedure to be followed in any case.
Let P2(E) dE be the probability of a state existing in the energy interval E to E + dE
near the energy E2 and thus N2P2(E) dE represents the number of atoms that have an energy
between E and E + dE. A similar definition is used for the states around E I. To make
the arithmetic tractable, we assume a Lorentzian distribution of energy levels with different
widths /)"E2.1.

P = (/)"E2,1) (7.6.1)
2 1(E)
, 2n [(E - E2,1)2 + (/)"E 2,J/2)2]
where /)"E2,1 = FWHM of the energy spread around E2,1 assuming that /)"E2,1 « (E2 -
EI) as implied in Fig. 7.6.
The spontaneous emission of a photon of energy hv requires that there be a (filled)
band dE near E2 and an (empty) band dE at exactly hv below the first. For any given value
of hv ; we must sum over the various possibilities that yield the same value of hv, and thus
'For an atomic system, the words occupied or empty are superfluous since the atom is either in states 2, I,
or elsewhere. However, those words will play an important role in transitions in semiconductors (see Sec. 11.4).
192 Atomic Radiation Chap. 7

the line shape reflects the "joint probability" of the parts of the two bands with intervals dE
being separated by hv. Let

for band 1: E = x + E, and a = !:i.Ed2 (7.6.2a)

for band 2: E = x + E1 + hv and (7.6.2b)

and define 8 = hv - (E2 - E,) (7.6.3)

Then

(7.6.4)

where x is the energy integration variable and the two braces represent the joint probability
that the two dE exist. This integral can be evaluated by the following steps:

1. Expand the integrand into partial fractions:


Ax + B C(X + 8) + D
---::---:::- +
x 2 + a2 (x + 8)2 + b2
to find

2. The tenus involving A and C are antisymmetric and vanish in the integration process.
3. The integration of B and D tenus yield Btt/ a and Dst / b, respectively. After
combining these tenus, the factor 82 + (b - a)2 cancels yielding

1 !:i.E2 + !:i.E,
g(hv) = - . 2 2 (7.6.5)
2n [hv - (E2 - E,)] + [(!:i.E2 + !:i.E,)/2]
Equation (7.6.5) demonstrates the fact that the broadening of both levels contributes to the
line shape function. In tenus of frequency units, we have
!:i.v
g (v) = -~---,-------:-.". (7.6.6)
2n [(VQ - v)2 + (!:i.v/2)2]
where !:i.v = _1 {~+ ~} (7.6.7)
2n T2 T,
and the energy widths of the levels are related to the lifetimes by (!:i.E1,2) . (T',2) = 1,
which is consistent with assuming a Lorentzian probability distribution for the broadened
levels.
There are many causes of the finite lifetime of the various states. For instance, state 2
can decay to lower states and can transfer its stored energy to a different type of atom or
Sec. 7.6 Broadening of Spectral Lines 193

can be lost by "quenching" collisions*

with a similar equation applying to state 1. The sum of the A coefficients is the inverse of
the radiative lifetime of state 2:

(7.6.8a)

The branching ratio is defined by


rate of decay from states 2 to 1
¢21= (7.6.8b)
sum of all decay rates from state 2
We can never do any better than eliminate processes that quench the upper state, and hence
the radiative branching ratio is usually specified with k2 = O.
Since the radiative processes are always present, there is a minimum width of any
spectral line caused by the radiative rates ofthe two states, and that width has the unfortunate
name of the natural width (implying by default that there is something un-natural about
other widths). Equation (7.6.7) becomes'

(7.6.9)

Most practical lasers use environments leading to transitions whose width is many times
that specified by (7.6.9), and, hence, natural broadening is seldom important.

Collision broadening (also known as pressure broadening in


a gas). Probably the most important homogeneous broadening mechanism is that
described by the following reaction equation:

[N2,d + [M] ---? [N2,d + [M] (7.6.10)

Yes, it does appear that nothing is happening since the left-hand side is the same as the
right, and, no, it is not a misprint! Equation (7.6.10) describes an "elastic" collision of
the atomic states (2,1) with something whose density or concentration is [M] . During the
"collision," the energy levels £2,1 can be considered as functions of time with some of the
potential energy being converted to a kinetic variety due to the attractive or repulsive forces
between the states and [M]. But since it is an elastic collision, the atom emerges with a
final energy equal to initial value. Now the time scale for this collision is extremely short,
roughly (diameter of the atoms) ~ 1.0A-;.- [thermal velocity of the atoms ~ 3 x 104 ern/sec]
"" 3 x 10- 13 sec-which is much shorter than the radiative lifetime. This will be discussed
in greater detail in Chapter 14. But for now, a simple picture is that such collisions interrupt

*Quenching expresses the known fact of the loss of population that proceeds at the rate k[N], but the details
as to where the atom goes or why may be unknown.
194 Atomic Radiation Chap. 7

E(t)

Elastic collisions

E(t)

FIGURE 7.8. Classical picture of the


effect of (b) elastic collisions on the
emission radiation shown in (a).

the phase of the wavefunctions that, in turn affect the classical field being radiated by the
group of atoms in the manner shown in Fig. 7.8.
A classical picture of this process is that shown in Fig. 7.8 (with great exaggeration).
In Fig. 7.8(a), the classical field associated with the radiation from a large number of
noninteracting atoms is shown. Fig. 7.8(b) is for atoms undergoing elastic collisions. The
field decays at the same rate, indicative of the finite lifetime of the state; however, the phase
of the radiation takes discontinuous jumps. This, in turn, shows up as a broadening in the
frequency domain.
It is a rather unpleasant task in mathematics to account for the random nature of
the collision rate around some mean rate Veol. and then to Fourier-analyze its effect on the
spectrum, but the result is somewhat transparent to a physical interpretation. Inasmuch as
there are two quantum states involved in the transition 2 ---? 1, each atomic wave function is
interrupted Veol times per second; hence, the additional broadening is 2Veol. The full width
of the transition is given by
1
L1v = -
2n
[(A2 + k2) + (AI + kl ) + 2Veol] (7.6.11)

where the grouping (A + k) indicates the decay rates of the states.


Whether the factor of 2 belongs in (7.6.11) is somewhat academic, since we must
depend on experimental measurements to infer the pressure-broadened line width of a
transition. As an example, E. T. Gerry and D. A. Leonard presented the data of Fig. 7.9
on the absorption coefficient of 10.6 J.Lm radiation by C02 at various pressures. Without
becoming unnecessarily complex about the energy levels of C02, it should be clear to all that
the number of absorbing molecules increases as the pressure is increased. If we followed our
Sec. 7.6 Broadening of Spectral Lines 195

intuition, we would therefore expect the absorption to increase linearly with the number of
molecules. The absorption does increase linearly with pressure at low pressures, but beyond
about 10 to 20 torr becomes independent of pressure.
As shown by (7.5.2), the absorption coefficient is proportional to the number of ab-
sorbers (as one would expect) and the line-shape function g(vo). For a pressure-broadened
line, g(vo) is inversely proportional to the collision rate. Hence, the product becomes
independent of pressure, as shown.
The prediction of the broadening from first principles is a significant chore and beyond
the scope of this book, but we should be able to use experimental results such as that given
in Fig. 7.9. Usually, this contribution is specified in the form of "MHz of broadening per unit
of pressure or density of gas," where a typical value might be 20 MHz/torr. Alternatively, a
collision cross section might be given that expresses the size (area) of one atom, the target,
to the other, the projectile (M). The collision rate is proportional to the density of projectiles,
N m , the collision cross section, and the relative velocity between the active atom and the
gas broadening the transitions.
For a Maxwellian distribution specified by a temperature T, with a cross section
independent of velocity, the collision rate becomes

Vcol = Nmu [8kT


.lr
(_1
u;
+ _1 )]1/2
Mn
(7.6.12)

where the M are the masses of the colliding atoms of type m and n.

10-2 r - - - - - - - . - - - - - - - , - - - - - - - - , - - - - - - - - - - ,

~lIoowl'" = ~Vcolli'iOl1 at 5.2 torr


where u; = 5.33 X 10 7 S-I
~

_\~~
'"
r-
N 10-3
'OJ
/
/
'5 /
/
~ /
;::
ll)
/

'u Combined Dopper and


""
4-<
ll)
0
U 10-4
collision contour for
ao= 5.7 x 10- 15 ern?
I':
T,=4.7s
'i
E
o 298K
-< o 353K
c: 404K

10-5

0.1 1.0 10 100 1000


Gas pressure (torr)

FIGURE 7.9. Absorption coefficient in CO 2 at 10.6 Ilm as a function of CO 2 pressure.


(After E. T. Gerry and D. A. Leonard, Appl. Phys. Lett. 8,227, 1966.)
196 Atomic Radiation Chap. 7

Note that every atom has been treated on equal footing in this subsection. We have
assumed that every atom is more or less the same as the other one, or that there was no dis-
tinguishing feature about anyone group. This is characteristic of homogeneous broadening,
and thus we add the subscript h to ~V.
In all the cases studied, the line shape g(v) has the same functional form: the
Lorentzian.

(7.6.6)

where
1
~Vh = 2rr [(A 2 + k2 ) + (AI + k\) + 2vcol] (7.6.11)

In most practical cases, the last term of (7.6.11) dominates, and the width of the
homogeneously broadened line becomes
A ~VcOI
ilVh ~ - - (7.6.12)
n rrT2
where T2 is the mean time between phase interrupting collisions. If we can distinguish
between different groups of atoms under special circumstances, a different functional form
of the line shape results.

7.6.2 Inhomogeneous Broadening

Although all atoms are broadened by the homogeneous process just indicated, there are
"shifts" in the central frequency of same type of atoms (ie, neon) but with different charac-
teristics (i.e., mass) that contribute to the overall line shape. These shifts are usually caused
by some feature that distinguishes one group of atoms from another, such as: isotope effect,
atoms of the same type but with different masses; Doppler effect, atoms of the same type
and mass but different velocities; nuclear spin, coupling to the angular momentum (hyper-
fine splitting); Zeeman splitting, owing to magnetic fields; and Stark splitting of the lines,
owing to the local crystalline electric field in a solid. More often than not, more than one
mechanism is operative in addition to the homogeneous processes for a given transition. A
very common example is the 2537 Atransition in Hg whose radiation excites the phosphor
in the common fluorescent lamp: there are 7 isotopes, which lead to 10 lines owing to the
combination of isotope effect and nuclear spin, and each of those are Doppler broadened
by the random velocity of the atoms and also broadened by the homogeneous processes of
the previous section. With all of these complicated processes, the 2537 Atransition is still
only r - 15 GHz wide (or 0.032 A).
Probably the simplest practical case of importance for the gas laser that emphasizes
the fact that these shifts lead to an effective broadening is illustrated in Fig. 7.10. Naturally
occurring neon has two isotopes with any significant abundance: 80% will have a mass of
20 AMU (atomic mass units = a multiple of 1.67252· 10- 27 kg) and most of the remainder
(20%) has a mass of 22. Because of the isotope effect, the center of the common "red" He:Ne
laser at 6328 A (3s 2 ~ 2p4) in neon changes slightly depending upon which isotope is
Sec. 7.6 Broadening of Spectral Lines 197

considered. Each contribution to the line shape scales according to the relative abundance
and are symmetrical about the respective center frequencies. The composite line shape, the
quantity of interest to us, is the sum of the two and is asymmetrical with its peak shifted
from either center. It is obvious that the line width /";. v of the composite is larger than that
for either one and illustrates the fact that some line shapes might be asymmetrical and a
terrible arithmetic mess to describe. Nevertheless, we must learn to tolerate such a situation
since the effective line-shape function appears in the laser gain equation. *
A case of considerable practical importance is due to the Doppler effect from the
thermal velocity of the atoms in a gas. If a select group of atoms emit a frequency Va in
their rest frame which is moving towards an observer in the laboratory with a velocity vz ,
then the observer would measure the Doppler-shifted frequency

z
va, = Va (
1+ ~
V )

Thus the homogeneous line width radiated by that particular group identified by their
velocity is

(7.6.13)

Composite
1 - - - - + - - - - lineshape .::>......."f-7"!"'"-~....--1--------+-------i

v(20) v( 2)

FIGURE 7.10. Inhomogeneous broadening in neon owing to the isotope effect. Each com-
ponent line is symmetrically broadened owing to the Doppler effect and other homogeneous
causes, but the composite line shape is slightly asymmetrical.

"(Since the gain coefficient at line center is proportional to 1-:- (line width), we would obtain the highest
gain with a single isotope.)
198 Atomic Radiation Chap. 7

Now we must multiply (7.6.13) by the probability of occurrence of a group of atoms traveling
away from us at a velocity of V z (within dv z ). This is given by the Maxwell-Boltzmann
distribution function.
fraction of the atoms with
velocity in the z direction
with velocity between V z
-dN = ( M
- - )1/2 exp (MV;)
- - - dv (7.6.14)
N 2rrkT 2kT z
and V z + du.:
By multiplying (7.6.13) by (7.6.14) and integrating over all velocities, we obtain the more
realistic line-shape function for these atoms.

g(v) = ( -M- ) 1/2


2rrkT
£+00 {
-00 2rr [(v - vo - Vovde)2
bVh
+ (~vh/2)2]
I
x exp MV2)
- __ z dv; (7.6.15)
( 2kT

This function has been tabulated and is called the Voigt function. We usually resort to the
approximation of replacing the Lorentzian, the quantity in the braces, by a delta function.
(We will see later when this approximation is valid.) That is,
~x
L(x - x )
I
= 2rr[(x - x ' )2 + (~x/2)2J
~ 8(x - x )
I

Thus (7.5.15) can be integrated easily:

g(v) = (2rr~T ) 1/2 [:00 8 (v - Vo -


z
VOeV ) exp (- ~1) d (V~o) :

or

(7.6.16)

This is the "pure" Doppler line shape. The peak value occurs at Vo (naturally) and falls to
one half of that at frequencies given by
2
Me (V+,_ - Vo )2 = ln2
--
2kT Vo
Thus the full width of the transition is given by
8kT In 2 ) 1/2
(V+ - V_) = ~VD = ( Me 2 Vo (7.6.17)

and (7.5.16) can be reexpressed in the following manner:

4 In-
g(v) = ( - 2) 1/2 -1- exp [ -4ln2 ( - Vo ) 2]
v -- (7.6.18)
rr ~VD ~VD
Sec. 7.6 Broadening of Spectral Lines 199

We can now evaluate the validity of our approximation of replacing the Lorentzian by a
delta function. If the Doppler width (7.6.17) is much larger than the homogeneous width.
(7.6.18) is valid. If the reverse is true, we should replace the exponential term in (7.6.15)
by a delta function centered at V z = 0 and thus recover the homogeneous line shape.
An electronic (i.e., visible) transition in a low-pressure gas tends to be dominated by
inhomogeneous or Doppler broadening. If, however, the pressure is high enough, or the
center frequency Vo is low enough, homogeneous or pressure broadening will dominate.
The Stark effect is very important for solid state lasers in which the active atoms,
say a rare earth atom such as neodymium (Nd) or erbium (Er), is doped into a solid, say
crystalline YAG or silica (Si02 = glass). Usually, the local electric field experienced by
the active atoms is so strong that the levels split into distinct and separate parts. For instance,
the ground state of Nd in YAG has the name 419/2 and, due to the crystalline field, splits
into five distinct levels with the group called the 419/2 manifold. * This is shown in Fig. 7.11.
This major effect is not a broadening; rather, it is a (Stark) splitting. We will assume that
some one has measured this shift, and thus we have the correct value for the center of each
level.
Because of the lattice vibrations (phonons) and/or the slight inhomogeneities in the
composition of the host, the local electric field experienced by each atom may not be the
same from site to site, and thus the center of each level will reflect the statistical distribution
of the local electric field. Since the phonons affect all atoms, it is a homogeneous broadening
mechanism, whereas the differences between different sites is an inhomogeneous one.
In most cases, we must resort to experiment to ascertain the line width and, more often
than not, use a Lorentzian to describe g ( v) even though the dominant broadening process
may be inhomogeneous.

800

%/ 2 - 0
:
II . t •
311
200

132 - - - - -
o

FIGURE 7.11. The Stark splitting of the 4/9/2 level of neodymium in YAG (After Kamin-
ski [17]). The numerical values refer to the energies of the levels in em:". More detail will
be discussed in Chapter 10.

•A short primer on the meaning of the spectroscopic names, such as 4/9/ 2, is given in Chapter 15, with
more detail on the Nd system in Chapter 10. For now, these are names and additional information is not needed.
200 Atomic Radiation Chap. 7

7.6.3 General Comments on the Line Shape

The reader should be cautioned against assuming a direct relationship between the amount
of mathematics expended here on a broadening mechanism and its relative importance.
Although Doppler and pressure-broadening mechanisms are important, they do not over-
whelm all other types (indeed, they do not even apply in a solid). In fact, only the central
portion of some transitions in a gas is adequately described by the theory presented here.
However, the idea of a line shape is most important, quite general, and independent
of the maze of mathematics surrounding its development. The line-shape function, g (v) d v,
is the relative probability that

1. A photon emitted by a spontaneous transition will appear between v and v + dv.


2. Radiation in the frequency interval v to v + d v can be absorbed by atoms in state 1.
3. Radiation in this interval will stimulate atoms in state 2 to give up their internal energy.

Obviously, the first applies to spontaneous emission, the second to absorption, and
the third to stimulated emission. However, the same line-shape function applies to all three
processes.
Many of the real-life line-shape functions are asymmetric and mathematically in-
tractable. However, the atoms have no knowledge of and no trouble with our arithmetic.
In response, we must be prepared to tolerate and use a real-life line-shape function about
which we have imperfect information.

7.7 REVIEW

Before we proceed to laser oscillation, it is appropriate to stand back and review the material
in this chapter. Although a lot of arithmetic was used, most of it was used to convert one
equation into another form. Consequently, definitions are an important part of this chapter.
If we remember the physical meaning of the quantities defined, the few lines of arithmetic
follow most naturally.
For instance, the Einstein A and B coefficients were introduced to describe the rate
at which the atoms respond to the electromagnetic field. In covering this, we are naturally
led to the physical interpretation of A and B and then to the idea of line shape. This was
the probability of a group of atoms interacting with a wave by stimulated emission or
absorption, or in producing the wave by spontaneous emission. We also observed that there
are different types of mechanism leading to different types of line shapes, or broadened
transitions. If there is no distinguishing feature about the atoms, we have homogeneous
broadening, whereas if two (or more) groups can be identified, we have inhomogeneous
broadening.
All of these concepts will playa major role in laser oscillation.

You might also like