Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 14

1

Chapter II: Interaction of Electromagnetic


Radiation with Atoms and Molecules

Electromagnetic radiation
Electromagnetic radiation includes what we call light ‘light’, but also
radiation of much longer and much shorter wavelengths (Figure II.1).
These waves have both an electric and a magnetic component, best
illustrated by considering plane-polarized (also known as linearly polarized)
radiation (Figure II.2). The electric component is an oscillating electric eld
of strength E and the magnetic component, an oscillating magnetic eld of
strength H. These oscillating elds are at right angles to each other, as
shown, and, if the directions of the vectors E and H are y and z
respectively, then

Ey = A sin(2πvt — kx)
{Hz = A sin(2πvt — (2.1)
kx)
where A is the amplitude. K is the wave vector. Thus, the elds oscillate
sinusoidally with a frequency of 2πv and, because k is the same for each
component, they are in-phase.
Plane of polarization: plane containing the direction of E and that of
propagation (Figure II.2). The reason for this choice is that interaction of
electromagnetic radiation with matter is more commonly through the
electric component.
2

Absorption and emission of radiation


In Figure II.3, states m and n of an atom or molecule are stationary
states, so-called because they are time-independent. This pair of states
may be, for example, electronic, vibrational or rotational. We consider
the three processes that may occur when such a two-state system is
subjected to radiation of frequency v, or wavenumber v˜, corresponding to
the energy separation ∆E where
∆E = En — Em = hv = hcv˜ (2.2)
These processes are:
1. Induced absorption, the molecule (or atom) M absorbs a quantum of
radiation and is excited from m to n:
M + hcv˜ → M* (2.3)

2. Spontaneous emission, in which M * (in state n) spontaneously emits


a quantum of radiation:
M* → M + hcv˜ (2.4)

3. Induced, or stimulated, emission: A quantum of radiation of wavenumber


v˜ given by Equation (2.2) is required to induce, or stimulate, M* to go
from n to m. The process is represented by
M* + hcv˜ → M + 2hcv˜ (2.5)
Therefore induced emission requires the presence of radiation of the
correct wavenumber for it to occur. The reason why the absorption process
is strictly
referred to as induced absorption may now be appreciated since, of course,
it requires the presence of radiation of wavenumber ~ in order to occur.
3

Rate of change of population Nn of state n by induced absorption:


dnn
dt
= —Nn Bmn ρ(v˜) (2.6)

where Bmn is the B Einstein coef cient. ρ(v˜) is the spectral radiation density:
8πhcఔ˜3
ρ(v˜) =exp(hcv˜)–1 (2.7)
kT

Similarly, induced emission changes the population Nn by


dnn
= —Nn Bnm ρ(v˜) (2.8)
dt
where Bnm = Bmn . For spontaneous emission, we have
dnn
dt
= —NnAnm (2.9)

where Anm is the A Einstein coef cient. The lack of ρ(v˜)


indicates
a spontaneous process.
In presence of radiation of wavenumber all three processes occur and, at
v˜, equilibrium:
d nn
= (Nm — Nn )Bnm ρ(v˜) — Nn Anm = 0 (2.10)
dt

Also at equilibrium Nn and Nm, according to the Boltzmann distribution law:


nn
= gn
exp (— ΔE) = exp (— ΔE) (2.11)
nm gm kT kT

Assuming that the degrees of degeneracy gn =gm. Combining this relationship


and ρ(v˜) with Equ. (2.10) gives:
Anm = 8πhcv˜ 3 Bnm (2.12)
Very important: spontaneous emission increases rapidly relative to induced
emission as v˜ increases. Since lasers operate entirely by induced emission the
equation is particularly relevant to laser design.
Einstein coef cients related to wave functions ƒm and ƒn via transition moment:
Rnm = ƒ ƒ*μƒ
dr (2.13)
n m
for interaction with the electric component of the radiation. The quantity  is the
electric dipole moment operator:
μ = Σi qiri (2.14)
qi and ri = charge and position vector of ith particle (electron or nucleus). The
transition moment can be thought of as the oscillating electric dipole moment due to
the transition. Figure 2.4 shows the π and π* molecular orbitals of ethylene and,
if an
electron is promoted from π to π* in an electronic transition, there is a
corresponding non-zero transition moment. This example illustrates the important
point that a transition dipole moment may be non-zero even though the permanent
electric dipole moment is zero in both the states m and n.
Transition probability= |Rnm|2 and

Bnm
8π3 |Rnm|2 (2.15)
= (4πs0)3h2

How do we determine Bnm experimentally in an absorption experiment?


In Figure 2.5(a), incident radiation intensity is I0 on absorption cell of length l
containing absorbing material of concentration c ( for example, in the liquid
phase). The radiation emerges with intensity I, and scanning the radiation over v˜1 to
v˜2 , and measuring I0/I produces an absorption spectrum, as absorbance A. Beer–
Lambert law :

A = log10 I0I = s(v˜)cP (2.16)

s = molar absorption coefficient or molar absorptivity (also molar extinction


coef cient). A is dimensionless, s is in (concentration * length)–1 or
mol–1 dm3 cm–1 (Figure 2.5b). smax = maximum of A (also used as a measure of the
total absorption intensity). In Figure 2.5c, band has same smax but lower integrated
intensity, so it is dangerous to just use smax. One should integrate the area under
the curve; then, provided Nn « Nm so that decay of state n by induced emission is
negligible,
ఔ˜2
ƒ G(v˜)dv˜
˜nmBnm = (2.17)
nAhఔ
ఔ˜1 ln 10

v˜nm =average wavenumber of the absorption; NA=Avogadro’s number.


If absorption due to electronic transition then fnm, oscillator strength is often used to
quantify the intensity:
4G0 me c 2 ln 10 ఔ ˜2
f
G(v˜=
)dv˜ ƒ (2.18)
nm nAe2 ఔ˜1

fnm is dimensionless and is the ratio of the strength of the transition to that of an
electric dipole transition between two states of an electron oscillating in three
dimensions in a simple harmonic way, and its maximum value is usually 1.
Transition probability |Rnm|2 related to selection rules in spectroscopy (here
mainly electric dipoles ones).
|Rnm|2 = 0 for forbidden transitions
|Rnm|2 G 0 for allowed transitions
Electric dipole moment operator μ has components along the cartesian axes:
μx = Σi qixi
; μy = Σi qiyi ; μz = Σi qizi (2.19)

where qi and xi are, respectively, the charge and x coordinate of the ith particle
and so on.
Transition moment becomes:
Rnm
ƒ = ƒ ƒ*μ
dx
ƒ ; Rnm = ƒ ƒ*μ dx
ƒ ; Rnm = ƒ ƒ*μ dx (2.20)
x n x m y n y m z n z m

Transition probability is:


2
|Rnm|2 = (Rnm)2 + (Rnm) + (Rnm)2 (2.21)
x y z
Line width
Figure 2.6 shows, for a sample in the gas phase, a typical absorption line with a
HWHM (half-width at half-maximum) of ∆v and a characteristic line shape. The line is
not in nitely narrow even if we assume that the instrument used for observation has
not imposed any broadening of its own (called the instrumental line width). Three
important factors may contribute to line width and shape:

Natural line broadening


If state n in Figure 2.3a is populated by absorption, the species M* in this state will
decay to the lower state until the Boltzmann population is regained. Decay is rst-
order, i.e.

— dnn = kN n (2.22)
dt

where k is ( rst-order) rate constant and


1
=r (2.23)
k

Here, τ = time taken for Nn to fall to 1/e of its initial value = lifetime of state n. If
spontaneous emission is the only process by which M* decays, comparison with
Equation (2.9) gives:
k = Anm (2.24)
Time-energy Heisenberg uncertainty principle:
r∆E ≥ ԰ (2.25)
Illustrates the point that state n has an exactly de ned energy only if τ is in nite.
Never the case, so all energy levels are smeared out to some extent, with resulting line
broadening.
Combining Equations (2.12) and (2.15) gives:

Anm
64π4 ఔ 3 |Rnm|2 (2.26)
= (4πs0)3hc3
so that, from Equation (2.25),

32π3 ఔ 3 |Rnm|2 (2.27)


∆v ≥ (4πs0)3hc3

Dependence of ∆v on v3: much larger value for an excited electronic state, typically
30 MHz, than for an excited rotational state, typically 10–4 — 10–5Hz, because of
the much greater v for an excited electronic state.
Equation (2.27) illustrates what is called the natural line broadening. Since each atom
or molecule behaves identically in this respect it is an homogeneous line broadening,
which results in a characteristic Lorentzian line shape. Natural line broadening <<
other causes of broadening, which we now examine.

Doppler broadening
In a gas, frequency of absorbed or emitted radiation depends on the velocity of the
atom or molecule relative to the detector. This is known as the Doppler effect.
If an atom or molecule is travelling towards the detector with velocity va , then
frequency va at which a transition is observed to occur is related to the actual
transition frequency v of a stationary atom or molecule by
ఔ a –1
va = v (1 — ) (2.28)
c

where c is the speed of light. Because of the usual Maxwell velocity distribution there
is a spread of values of va and a characteristic line broadening given by
ఔ 2kT ln 2
∆v = ( 1/2
) (2.29)
c m
where m is the mass of the atom or molecule; ∆v >> natural line width. The
broadening is inhomogeneous, since not all atoms or molecules in a particular
sample behave in the same way, and results in a line shape known as gaussian.

Pressure broadening
When collisions occur between gas phase atoms or molecules there is an exchange
of energy, which leads effectively to a broadening of energy levels. If r is the
mean time between collisions and each collision results in a transition between two
states there is a line broadening ∆v of the transition, where
∆v = (2πr)–1 (2.30)
derived from the uncertainty principle of Equation (1.16). This broadening is, like
natural line broadening, homogeneous and usually produces a Lorentzian line
shape except for transitions at low frequencies, when an unsymmetrical line shape
results.
However, there is an additional contribution to pressure broadening which arises
from the interaction between the system under study and collisional partners.
Depending on the shapes of the ground and excited state potential curves between
the system and its collisional partners, and on the average distance between them, an
additional inhomogeneous broadening arises. Pushed to its extreme in the case of
atoms or molecules embedded in a crystal lattice, this inhomogeneous broadening is
caused by the slight structural differences between the different trapping sites.

Power, or saturation, broadening


The equilibrium populations of energy levels, given by the Boltzmann distribution
law of Equation 2.11, are maintained through exchange of energy by collisions.
However, when the ground and excited energy levels m and n, respectively, are
close together the population ratio Nn /Nm at equilibrium is close to 1. As a
result, the equilibrium may be seriously disturbed when, in obtaining an absorption
spectrum, the molecular sample is subjected to radiation of intensity I. As I is
increased the limit of Nn /Nm = 1 is approached, the Beer-Lambert law
breaks
down, s becomes dependent on I, and saturation is said to occur. Clearly, this occurs
more readily the closer the energy levels n and m. Consequently, saturation is more
common in microwave and millimeter wave spectroscopy, although the extremely
high power of a laser source may result in saturation effects in higher energy regions
of the spectrum.
One effect of saturation, and the dependence of s on I, is to decrease the
maximum absorption intensity of a spectral line. The central part of the line is
attened and the intensity of the wings is increased. The result is that the line is
broadened, and the effect is known as power, or saturation, broadening. Typically,
microwave power of the order of 1 mWcm–2 may produce such broadening.
Minimizing the power of the source and reducing the absorption path length λ
can limit the effects of power broadening.

Removal of line broadening


Of the four types of broadening that have been discussed, that due to the natural
line width is, under normal conditions, much the smallest and it is the removal, or the
decrease, of the effects of only Doppler, pressure and power broadening that can be
achieved.
Except at very low frequencies, pressure broadening may be removed simply by
working at a suf ciently low pressure. Doppler broadening may be reduced or
removed by two general methods, which will be discussed brie y below.
Effusive atomic or molecular beams
An effusive beam of atoms or molecules is produced by pumping them through a
narrow slit, with a pressure of a few mbars on the source side of the slit. The beam
may be further collimated by suitable apertures along it.
Such beams have many uses, including some important applications in
spectroscopy. In particular, pressure broadening of spectral lines is removed in an
effusive beam and, if observations are made perpendicular to the direction of the
beam, Doppler broadening is considerably reduced because the velocity component
in the direction of observation is very small.
Figure 2.1: range of electromagnetic radiation

Figure II.2 : Plane-polarized electromagnetic radiation travelling along the x


axis; Ey is the electric component; Hz is the magnetic component
Figure II.3: Absorption and emission processes between states m
and n.

Figure 2.4 (a) A  and (b) a * molecular orbital of ethylene


Figure 2.5 (a) An absorption experiment. (b) A broad and (c) a narrow
absorption band with the same smax; c is the concentration of the
absorbing material in the liquid phase
Figure 2.6: Typical (gaussian) absorption line showing a HWHM (half
width at half maximum) of n.

You might also like