Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

82 Chapter 2

Lorentz, H. A., Theory of Electrons, Second Edition, Dover, New York


(1952).
Lorrain, P. and Corson, D., Electromagnetic Fields and Waves, Second
Edition, W. H. Freeman, New York (1962).
Maxwell, J. C., Treatise on Electricity and Magnetism, Third Edition,
Dover, New York (1954).
Mie, G., “Beitrage zur Optik trüber Medien speziell kolloidaler
Metallösungen,” Ann. Phys. 25, 377–445 (1908).
Palik, E. D., Handbook of Optical Constants of Solids, Academic Press,
Orlando, FL (1985).
Pust, N. J. and Shaw, J. A., “Digital all-sky polarization imaging of partly
cloudy skies,” Appl. Optics 47, 190–198 (2008).
Rayleigh, Lord, “On the light from the sky, its polarization and colour,”
Philos. Mag. J. Sci. 41, 107–112 (1871).
Schott, J. R., Remote Sensing: The Image Chain Approach, Second
Edition, Oxford University Press, Oxford, New York (2007).
Smith, F. G. (Ed.), The Infrared and Electro-Optical Systems
Handbook: Volume 2, Atmospheric Propagation of Radiation, SPIE Press,
Bellingham, WA (1993).

Downloaded From: http://ebooks.spiedigitallibrary.org/ on 04/22/2013 Terms of Use: http://spiedl.org/terms


Chapter 3
Atomic and Molecular
Spectroscopy
A quantum mechanical model is necessary to accurately understand the
spectroscopy of atoms and molecules. While the classical mechanical
model provides some insight into the spectral characteristics of materials
and a useful basis for semi-empirical modeling of amorphous solids and
liquids, it fails to capture a variety of observed spectral features such as
rotational lines in gaseous spectra, band structures of crystalline solids,
and quantized emission spectra of atoms.
The quantum mechanical treatment accurately models these effects by
characterizing particles such as electrons in an atom or molecule through
wave functions. Such wave functions exist only at discrete energy levels,
and interaction with optical radiation involves the allowed transitions
between such energy levels. In this way, the spectral characteristics of
emission and absorption depend on the specific energy levels of the atom
or molecule, the dipole moment associated with the energy level between
which the system transitions, and the state populations.
According to the Born–Oppenheimer approximation, a complex
molecule can be analyzed by decomposing the wave function and energy
levels into atomic, vibrational, and rotational components. Therefore,
the mathematical treatment in this chapter separately addresses these
individual components to elucidate characteristic features, and then
discusses the more complicated case of aggregate spectral properties.
While this chapter only scratches the surface of the field of atomic and
molecular spectroscopy, the insights gained are of great importance in
understanding the observed spectral properties of real materials, as well as
the effects of the atmosphere on hyperspectral measurements in a remote
setting.

3.1 Quantum Mechanics


From a quantum mechanical perspective, the state of a molecular system
is described in terms of wave function Ψ(x, y, z, t). This wave function
83
Downloaded From: http://ebooks.spiedigitallibrary.org/ on 04/22/2013 Terms of Use: http://spiedl.org/terms
84 Chapter 3

represents either a particle or system of particles (atoms and electrons),


while the effect of the local environment (bonding forces within atoms
and molecules) and external fields is characterized by a potential function
where the particles reside. The squared magnitude of the wave function,

Ψ∗ (x, y, z, t)Ψ(x, y, z, t), (3.1)

is proportional to the probability that the particle resides at specified time


t and location (x, y, z). As this quantity is a probability density function, it
can be normalized such that

Ψ∗ (x, y, z, t)Ψ(x, y, z, t)dxdydz = 1. (3.2)

All of the characteristics of the system state are defined by the wave
function. Because the system state is defined probabilistically as opposed
to deterministically, system characteristics can only be defined as expected
values, as opposed to deterministic quantities. Three such expected values
of particular interest relate to the expected location in space,

x̂(t) = Ψ∗ (x, y, z, t)x Ψ(x, y, z, t)dxdydz

ŷ(t) = Ψ∗ (x, y, z, t)y Ψ(x, y, z, t)dxdydz (3.3)

ẑ(t) = Ψ∗ (x, y, z, t)z Ψ(x, y, z, t)dxdydz,

the expected momentum,


  
∗ ∂
p̂ x (t) = Ψ (x, y, z, t) −i Ψ(x, y, z, t) dxdydz
∂x
  
∗ ∂
p̂y (t) = Ψ (x, y, z, t) −i Ψ(x, y, z, t) dxdydz (3.4)
∂y
  
∗ ∂
p̂z (t) = Ψ (x, y, z, t) −i Ψ(x, y, z, t) dxdydz,
∂z

and the expected energy level,


  
∗ ∂
Ê(t) = Ψ (x, y, z, t) i Ψ(x, y, z, t) dxdydz. (3.5)
∂t

Downloaded From: http://ebooks.spiedigitallibrary.org/ on 04/22/2013 Terms of Use: http://spiedl.org/terms


Atomic and Molecular Spectroscopy 85

In Eq. (3.5),  = h/2π, where h is Planck’s constant (h = 6.626 ×


10−34 J s). If the expected energy level is constant in time, the wave
function represents a stationary state of the system.

3.1.1 Stationary states of a quantum mechanical system

The wave function must be a solution to the Schrödinger wave equation


(Schrödinger, 1926),
 
2 2 ∂
− ∇ + V(x, y, z, t) Ψ(x, y, z, t) = i Ψ(x, y, z, t), (3.6)
2m ∂t

where m is the particle mass and V(x, y, z, t) is the potential function. The
left-hand side of Eq. (3.6) is known as the Hamiltonian of the system, while
the right-hand side is recognized as the energy operator within Eq. (3.5).
Stationary states of a system can be represented by the time-harmonic
solution,

Ψ(x, y, z, t) = ψ(x, y, z) e−iEt/ , (3.7)

or superpositions thereof, where the constant E represents the energy level


of the stationary state. The time-independent wave function ψ(x, y, z) is a
solution to the time-independent Schrödinger wave equation

2 2
− ∇ ψ(x, y, z) + V(x, y, z) ψ(x, y, z) = Eψ(x, y, z). (3.8)
2m

When Eq. (3.8) is solved for the stationary states of a particular


quantum mechanical system defined by V(x, y, z), there are normally
multiple solutions {ψi (x, y, z); i = 1, 2, . . . N} that correspond to different
energy levels {Ei }. Mathematically, the solutions {ψi (x, y, z)} are the
eigenfunctions of the system, and the corresponding energy levels Ei
are the eigenvalues. While any linear combination of solutions is also a
solution of the Schrödinger wave equation, one typically considers the
orthonormal set of eigenfunctions that exhibit the property

ψ∗i (x, y, z, t)ψ j (x, y, z, t)dxdydz = δi j , (3.9)

where δi j is the Kronecker delta function: δi j = 1 if and only if i = j. As


an example, consider the triple quantum well in Fig. 3.1, defined by the

Downloaded From: http://ebooks.spiedigitallibrary.org/ on 04/22/2013 Terms of Use: http://spiedl.org/terms


86 Chapter 3

Figure 3.1 Potential function for a triple quantum well.

potential function produced by epitaxial crystal growth in the x direction:





⎪ V0 x<0


⎪ ≤ x≤a



0 0


⎪ V a < x < a+b

⎨ 0
V(x, y, z) = ⎪
⎪ 0 a + b ≤ x ≤ 2a + b . (3.10)




⎪ V 2a + b < x < 2a + 2b



0

⎪ 0 2a + 2b ≤ x ≤ 3a + 2b

⎩V
0 3a + 2b < x

The first four wave-function solutions in terms of increasing energy level


for V0 = 0.1 eV, a = 100 Å, and b = 50 Å are illustrated in Fig. 3.2.
A few typical characteristics of wave functions are evident from these
examples. First, wave functions exhibit higher-spatial-frequency content
as the energy level increases. Second, they tend to alternate between
even and odd symmetry. Finally, wave functions are nonzero even in the
barrier regions, meaning that particles can exist within the barriers and
move through the barriers between the wells. This characteristic, known
as tunneling, can only be explained through a quantum mechanical model.

3.1.2 Interaction with electromagnetic radiation


In the presence of electromagnetic radiation of angular frequency ω,

V(x, y, z, t) = V0 (x, y, z) + e(E f · r) cos(ωt), (3.11)

where V0 (x, y, z) is the stationary field due to the local environment, E f is


the vector electric field amplitude, and r is a spatial vector. Equation (3.11)
simply introduces the Lorentz force from the electromagnetic radiation
into the potential function. Solutions to Eq. (3.6) with the potential

Downloaded From: http://ebooks.spiedigitallibrary.org/ on 04/22/2013 Terms of Use: http://spiedl.org/terms


Atomic and Molecular Spectroscopy 87

9000

8000

7000
Wavefunction (m-0.5)

6000

5000

4000

3000

2000

1000

0
-300 -200 -100 0 100 200 300 400 500 600 700
x (angstroms)
(a)
10000

8000

6000
Wavefunction (m-0.5)

4000

2000

-2000

-4000

-6000

-8000

-10000
-300 -200 -100 0 100 200 300 400 500 600 700
x (angstroms)
(b)

Figure 3.2 Wave functions corresponding to the lowest four energy levels of the
triple quantum well: (a) first level, (b) second level, (c) third level, and (d) fourth
level.

Downloaded From: http://ebooks.spiedigitallibrary.org/ on 04/22/2013 Terms of Use: http://spiedl.org/terms


88 Chapter 3

8000

6000

4000
Wavefunction (m-0.5)

2000

-2000

-4000

-6000

-8000

-10000
-300 -200 -100 0 100 200 300 400 500 600 700
x (angstroms)
(c)

8000

6000

4000
Wavefunction (m-0.5)

2000

-2000

-4000

-6000

-8000
-300 -200 -100 0 100 200 300 400 500 600 700
x (angstroms)
(d)

Figure 3.2 (continued)

Downloaded From: http://ebooks.spiedigitallibrary.org/ on 04/22/2013 Terms of Use: http://spiedl.org/terms


Atomic and Molecular Spectroscopy 89

function in Eq. (3.11) are of the form

Ψ(x, y, z, t) = ci (t) ψi (x, y, z) e−iEi t/ + c j (t) ψ j (x, y, z) e−iE j t/ , (3.12)

which can be interpreted as follows. The energy absorbed from the


electromagnetic radiation causes a transition from a wave function ψi with
energy Ei to a wave function ψ j with energy E j , according to coefficients
ci (t) and c j (t). These coefficients are bound by the initial conditions
ci (0) = 1 and c j (0) = 0, meaning that the particle starts in the i’th state.
The probability that a state transition to ψ j occurs at time t is given by the
quantity c∗j (t)c j (t). By inserting Eq. (3.12) into Eq. (3.6), this is shown to
be

| E f · Mi j |2 sin2 [(E j − Ei − ω) t/2]


c∗j (t) c j (t) = , (3.13)
 [(E j − Ei − ω)/2] 2

where Mi j is the transition-moment integral defined by



Mi j = ψ∗i (x, y, z)μ ψ j (x, y, z) dxdydz, (3.14)

and μ is the electric dipole moment of the atomic or molecular system. The
dipole moment is typically defined as

μ= qn rn , (3.15)
n

where qn and rn represent the charges and positions of all charged particles
in the system. In the quantum mechanical model, the dipole moment can
be represented by

μ= ψ∗n (x, y, z) qn r ψn (x, y, z) dxdydz, (3.16)
n

where ψn represents the wave functions for all the charged particles.
A few important observations should be drawn from Eqs. (3.13) through
(3.16). First, transitions can only occur between states for which the
transition-moment integral is nonzero for the polarization direction of
incident electromagnetic radiation. Note that this integral is identically
zero for transitions between states where both wave functions are either
symmetric or antisymmetric in the polarization direction because the
integrand in Eq. (3.14) is an odd function in these cases, and, therefore,
the integral is identically zero. Thus, transitions can occur between the

Downloaded From: http://ebooks.spiedigitallibrary.org/ on 04/22/2013 Terms of Use: http://spiedl.org/terms


90 Chapter 3

first and second levels of the triple quantum well depicted in Figs. 3.1
and 3.2, but not between the first and third levels. Such transitions are
called forbidden transitions, while those for which the integral is nonzero
are called allowed transitions. For allowed transitions, energy is only
substantially absorbed (causing a substantial state transition probability) if
the radiation frequency is such that the second term in Eq. (3.13) is large.
This occurs near the transition frequency or wavelength, such that

hc
ωi j = = E j − Ei , (3.17)
λi j

which is equivalent to the resonance condition of the classical model. For


forbidden transitions, energy is not absorbed irrespective of the radiation
frequency.

3.1.3 Born–Oppenheimer approximation


The Hamiltonian of an atomic or molecular system represents the sum of
kinetic and potential energy of the system, that is, the first and second terms
of the Schrödinger wave equation. In a molecule, kinetic energy includes
the motions of both the negatively charged electrons and the positively
charged nuclei. Rigorously, one needs to find solutions to Eq. (3.8) for this
complete system; these solutions will be very complex, even for simple
atoms and molecules. Born and Oppenheimer (1927) proposed a useful
approximation based on the recognition that vibrating nuclei move much
slower than electrons, such that electronic states can be computed based on
fixed nuclei. Mathematically, the Born–Oppenheimer approximation holds
that

ψ(x, y, z) = ψe (x , y , z ) ψn (x, y, z), (3.18)

where (x , y , z ) are electron coordinates relative to fixed nuclei, and


(x, y, z) are nuclear coordinates. It follows that

E = Ee + En , (3.19)

where Ee and En are the energy levels corresponding to the electronic and
nuclear wave functions ψe and ψn , respectively.
The Born–Oppenheimer approximation can similarly be extended by
factoring the nuclear wave functions into components related to the rigid
rotation of the molecule as a whole, designated by ψr , and the vibrational
motion of atoms within a molecule, designated by ψv , such that

ψ(x, y, z) = ψe (x , y , z ) ψv (x , y , z ) ψr (x, y, z) (3.20)

Downloaded From: http://ebooks.spiedigitallibrary.org/ on 04/22/2013 Terms of Use: http://spiedl.org/terms


Atomic and Molecular Spectroscopy 91

and

E = Ee + Ev + Er . (3.21)

In Eq. (3.20), the double-primed coordinates correspond to fixed nuclei,


and the primed coordinates correspond to nonrotating nuclei. Through
this approximation, the electronic, vibrational, and rotational spectral
characteristics can be addressed separately and then aggregated. The
treatment in this chapter begins by considering the simpler cases of
electronic spectroscopy of single atoms and rotational spectroscopy of
rigid molecules, and eventually aggregates the simpler models to address
the complex case where electronic, vibrational, and rotational transitions
can simultaneously occur.

3.2 Electromagnetic Absorption and Emission

In the quantum mechanical model, absorption and emission of radiation


occurs during transitions between electronic states, such that the energy
of the radiation equals the change in kinetic energy corresponding to the
electron transition. This means that the photon energy

hc
E p = ω = = hcσ (3.22)
λ

equals the difference in energy levels E p = E j − Ei , which is just


a restatement of Eq. (3.17). It is common in spectroscopy to define a
wavenumber σ that is the reciprocal of the wavelength. Typically, σ is
measured in cm−1 units, such that:

10, 000
σ (cm−1 ) = . (3.23)
λ (μm)

When a charged particle transitions between states, one of the


three processes depicted in Fig. 3.3 can occur: stimulated absorption,
spontaneous emission, or stimulated emission. In stimulated absorption,
an atom or molecule absorbs a quantum of radiation, or photon, and is
excited from a lower state with energy Ei to a higher state with energy
E j . In spontaneous emission, a photon is emitted as the atom or molecule
relaxes from a higher state E j to a lower state Ei . In stimulated emission,
the relaxation from a higher to lower state is induced by a photon, and
another photon is emitted that is coherently in phase with the incident
photon. This latter process is the basis of laser radiation.

Downloaded From: http://ebooks.spiedigitallibrary.org/ on 04/22/2013 Terms of Use: http://spiedl.org/terms

You might also like