You are on page 1of 39

PAGE 1

STRUCTURAL INVESTIGATION
OF TITANIUM SAPPHIRE
LASERS AND THEIR
APPLICATIONS IN
SPECTROSCOPY

NAME: Aayushi Paul


ROLL NUMBER: 186
COLLEGE: St. Xavier’s College, Kolkata
SUPERVISOR: Dr. Indranath Chaudhuri
PAGE 2

DECLARATION
I affirm that I have identified all my sources and that no part of my
dissertation paper uses unacknowledged materials.
PAGE 3

CONTENTS
1. INTRODUCTION.................................................................page 4

2. LASERS.................................................................................page 6

2.1 EINSTEIN’S COEFFICIENTS.....................................................page 7

2.2 LIGHT AMPLIFICATION.........................................................page 9

2.3 THRESHOLD CONDITION......................................................page 10

3. LASER RATE EQUATIONS.............................................page 13

3.1 TWO LEVEL SYSTEM.............................................................page 13

3.2 THREE LEVEL SYSTEM...........................................................page 15

3.3 FOUR LEVEL SYSTEM............................................................page18

4. TUNABLE LASER..............................................................page 21

5. TITANIUM SAPPHIRE LASER......................................page 22

5.1 LASER STRUCTURE..............................................................page 23

5.2 CRYSTAL STRUCTURE..........................................................page 23

5.3 PROPERTIES........................................................................page 25

5.4 ADVANTAGES OVER OTHER LASERS....................................page 27

6. FRANCK CONDON PRINCIPLE....................................page 28

6.1 BASIC MECHANISM.............................................................page 29

6.2 QUANTUM MECHANICAL FORMULATION...........................page 30

7. JAHN TELLER EFFECT..................................................page32

8. APPLICATION IN SPECTROSCOPIC STUDIES......page 34

9. CONCLUSION....................................................................page 36

10. REFERENCE....................................................................page 37

11. ACKNOWLEDGEMENT................................................page 38
PAGE 4

1. INTRODUCTION
Radiation is the emission or transmission of energy in the form of waves or particles
through space or material medium. This includes electromagnetic radiations, particle
radiations, neutron radiations, acoustic radiation and gravitational radiations.

Emission is the process by which a higher energy quantum mechanical state of a particle
becomes converted to a lower one by the emission of a photon resulting in the production
of light. The frequency of light emitted is a function of energy of the transition. Since the
energy must be conserved, the difference between the two states equals the energy carried
by the photon. The energy states of the transitions can lead to emissions over a very large
range of frequencies. The emittance of an object quantifies how much light is emitted by it.
This may be related to other properties of the object through the Stefan-Boltzmann law. For
most substances, the amount of emission varies with the temperature and spectroscopic
composition of the object, leading to the appearance of color temperature and emission
lines. Precise measurements at many wavelengths allow the identification of a substance
via emission spectroscopy.

Spontaneous emission is the process in which a quantum mechanical system such as


an atom or a molecule transitions from an excited energy state to a lower energy state and
emits a quantum in the form of a photon. Spontaneous emission cannot be explained on the
basis of classical electromagnetic theory and id fundamentally a quantum process. It
generally invokes the zero point energy of the quantum electromagnetic field.

Stimulated emission is the process by which an incoming photon of a specific


frequency can interact with an excited atomic electron causing it to drop to a lower energy
level. The liberated energy transfers to the electromagnetic field m to atomic creating a new
photon with a phase, polarisation, frequency and direction of travel that are identical to the
photons of the incident wave. This is in contrast to spontaneous emission, which occurs at
random intervals without regard to the ambient electromagnetic field. The process is
identical in form to atomic absorption in which the energy of an absorbed photon causes an
identical but opposite atomic transition: from the lower level to a higher energy level. In
normal media, at thermal equilibrium, absorption exceeds stimulated emission as there are
a greater number of electrons in the lower energy states than in the higher energy states.
However when a population inversion is present, the rate of stimulated emission exceeds
PAGE 5

that of the absorption and a net optical amplification can be obtained. Such a gain medium
along with an optical resonator is at the heart of a laser. Stimulated emission can provide
mechanism for optical amplification. If an external source of energy stimulates more than
50% of the atoms in the ground state to transition into the excited state, then what is called
a population inversion is created. When light of appropriate frequency passes through the
inverted medium, the photons are either absorbed by the atoms that remain in the ground
state or the photons stimulate the excited atoms to emit additional photons of the same
frequency, phase and direction. Since more atoms are in the excited state than in the
ground state, then an amplification of the input intensity results.

The population inversion, in units of atoms per cubic meter is,


𝑔(2)
∆N21 = N2– 𝑔(1) N1

Where g(1) and g(2)are the degeneracies of levels 1 and 2.

FIGURE 1: THE BASIC MECHANISM OF ABSORPTION, SPONTANEOUS EMISSION AND


STIMULATED EMISSION. [a]
PAGE 6

2.LASERS LASER is an acronym for Light Amplification by Stimulated

Emission of Radiation. As the name implies, the process of stimulated emission is


used for amplification of light waves. The fact that stimulated emission can be used
in the construction of coherent optical sources was the driving principle behind the
construction of lasers. Laser action can be demonstrated with gases, solids, liquids,
free electrons, semiconductors, etc.
The three main components of any laser are:-
1. The amplifying medium
2. The pump
3. The optical resonator

The amplifying medium is the source of optical gain within the laser. The medium
consists of a collection of atoms, molecules or ions which act as an amplifier for light waves.
Under normal conditions, the number of atoms in the lower energy state is always larger
than the number in the excited energy state; as such, a light wave passing through such a
collection of atoms would cause more absorption than emissions and therefore, the wave
will be attenuated. Thus in order to have an amplification, it is necessary to have population
inversion between two energy states in which there is a larger number of atoms in the lower
state than in the higher state. When a wave passes through a collection of atoms which are
in a state of population inversion, the wave will induce more emissions and will then be
amplified. The pump is the source of energy which maintains the medium in this population
inverted state. Pumping is basically the transfer of energy from an external source to the
lasing medium of the laser. The pump power must be higher than the lasing threshold of the
laser. The energy is usually provided in the form of light or electric current. The optical
resonator which consists of a pair of mirrors facing each other provides optical feedback to
the amplifier so that it can act as a source of radiation. The optical cavities surround the
lasing medium and the light confined in the cavity reflects multiple times producing standing
waves of certain resonant frequencies.

FIGURE 2: BASIC COMPONENTS OF A LASER. [b]


PAGE 7

2.1 EINSTEIN’S COEFFICIENTS [1]

We consider two levels of an atomic system and let N1 and N2 be the number of atoms per
unit volume present in the energy levels E1 and E2 respectively. If radiation at a frequency
corresponding to the energy difference (E1- E2) falls on the atomic system, it can interact in
three distinct ways:

(a) An atom in the lower energy level E1 can absorb the incident radiation and can be excited
to E2. This excitation process requires the presence of radiation. The rate at which
absorption takes place from level 1 to level 2 will be proportional to the number of atoms
present in the level E1 and also to the energy density of the radiation at the frequency ω =
(E2-E1)/ħ. Thus if u(ω)𝑑𝜔 represents the radiation energy per unit volume between ω and
ω+ 𝑑𝜔 then we may write the number of atoms undergoing absorption per unit per unit
volume from level 1 to level 2 as

ᴦ12 = B12 u (ω) N1

where B12 is a constant of proportionality and depends on the energy levels E 1 and E2.u(ω)
has the units of energy density per frequency interval.

(b) For the reverse process, namely the deexcitation of the atom from E2 to E1, Einstein
postulated that an atom can make a transition from E2 to E1 through two distinct processes
namely stimulated emission and spontaneous emission. In the case of stimulated emission,
the radiation which is incident on the atom stimulates it to emit radiation and the rate of
transition to the lower energy level is proportional to the energy density of radiation at the
frequency ω. Thus, the number of stimulated emissions per unit time per unit volume will
be,

ᴦ21 = B21u(ω)N2

Where B21 is the coefficient of proportionality and depends on the energy levels.

(c)An atom which is in the upper energy level E2 can also make a spontaneous emission; this
rate will be proportional to N2 only and thus we have for the number of atoms, making
spontaneous emissions per unit time per unit volume

U21 = A21N2

At thermal equilibrium between the atomic systems and the radiation fields, the number of
upward transitions must be equal to the number of downward transitions. Hence, at
thermal equilibrium,

N1B12u(ω) = N2A21 + N2B21u(ω)


PAGE 8

Or, u(ω) = A21/[(N1/N2)B12 – B21]

Using Boltzmann’s law, the ratio of the equilibrium populations of level 1 and 2 at
temperature T is

N2/N1 = 𝑒 ħ𝜔/𝑘𝑇

Where k is the Boltzmann’s constant. Hence,

u(ω) = A21/[B12𝑒 ħ𝜔/𝑘𝑇 − 𝐵 21] (1)

According to Planck’s law, the radiation energy density per unit frequency interval is given
by
1
u(ω) = ħ𝜔3n03/𝜋2c3 ħ𝜔 (2)
𝑒 𝑘𝑇 −1

Where c is the velocity of light in free space and n0 is the refractive index of the medium

Comparing equations (1) and (2) we obtain,

B12 = B21 = B and,

A21/B21 = ħ𝜔3n03/𝜋2c3 (3)

Thus the stimulated emission rate per atom, is the same as the absorption rate per atom
and the ratio of spontaneous to the stimulated emission coefficients is given by (3).

If we observe the spectrum of the radiation due to the spontaneous emission from a
collection of atoms, one finds that the radiation is not strictly monochromatic but is spread
over a certain frequency range. Similarly, if one measures the absorption by a collection of
atoms as a function of frequency, one again finds that the atoms are capable of absorbing
not just a single frequency but radiation over a band of frequencies. This implies that energy
levels have widths and the atoms can interact with radiation over a range of frequencies but
the strength of interaction is a function of frequency. This function is called the lineshape
function and is represented by g(ω). The function is normalised according to

∫ 𝑔(𝜔)𝑑𝜔 = 1 (4)

Therefore out of the total N2 and N1 atoms per unit volume, only N2g(ω)𝑑𝜔 and
N1g(ω)𝑑𝜔 atoms per unit volume will be capable of interacting with radiation of frequency
lying between ω and ω +𝑑𝜔. Hence the number of stimulated emissions per unit time per
unit volume will be given by

ᴦ21 = ∫ 𝑢(𝜔)B21 N2g(ω)𝑑𝜔

= N2𝜋2c3/ħn03tsp∫ 𝑢(𝜔)𝑔(𝜔)/ω3𝑑𝜔 (5)


PAGE 9

If the atoms are interacting with radiation whose spectrum is very broad compared to that
of g(ω), then one may essentially assume that over the region of integration where g(ω) is
appreciable, u(ω)/ω3 is essentially constant and thus may be taken out of the integral in (5).
Using the normalisation integral, (5) becomes

Γ21 =( N2𝜋2c3/ħω3n03tsp ) u(ω) (6)

Where ω represents transition frequency. Thus (6) represents the rate of stimulated
emission per unit volume when the atom interacts with broadband radiation.

The next consideration is of an atom interacting with near-monochromatic radiation. If the


frequency of the incident radiation is ω’, then the u(ω) curve will be extremely sharply
peaked at ω = ω’ as compared to g(ω) and thus g(ω)/ω3 can be taken outside the integral to
obtain

Γ21= N2( 𝜋2c3/ħω3n03tsp) g(ω’)∫ 𝑢(𝜔)𝑑𝜔

= N2 (𝜋2c3/ħω’3 no3tsp)g(ω’)u (7)

Where u = ∫ 𝑢(𝜔)𝑑𝜔 is the energy density of the incident near monochromatic radiation
having the dimensions of energy per unit volume unlike u(ω) which has the dimensions of
energy per unit volume per unit frequency interval. Thus when the atom described by the
lineshape function g(ω) interacts with near monochromatic radiation at frequency ω’, the
stimulated emission rate is given by (7)

And in a similar fashion, the number of stimulated absorptions per unit time per unit
volume will be

Γ12 = ( N1𝜋2c3/ħtspn03ω’3)g(ω’)u

2.2 LIGHT AMPLIFICATION


Considering a collection of atoms, we let a near-monochromatic radiation of energy density
u at frequency ω’ pass through it. we wish to obtain the rate of change of intensity of
radiation as it passes through the medium.

Let us consider two planes P1 and P2 of area S situated at z and z+𝑑𝑧; z being the direction of
propagation of the radiation. If I(z) and I(z+𝑑𝑧) represent the intensity of the radiation at z
and z+𝑑𝑧 respectively, then the net amount of energy entering the volume S𝑑𝑧 between P1
and P2 will be,
PAGE 10

𝑑𝐼
[I(z) – I(z+𝑑𝑧)]S = -(𝑑𝑧)S𝑑𝑧

This must be equal to the net energy absorbed by the atoms in the volume S𝑑𝑧. The energy
absorbed by the atoms in going from level 1 to level 2 will be Γ12S𝑑𝑧ħω’. Similarly, the
energy released through stimulated emission from level 2 to level 1 will beΓ21 S𝑑𝑧ħω’. We
shall neglect the energy arising from spontaneous emission since it appears over a broad
frequency range and is also emitted in all directions. Thus the fraction of the spontaneous
emission which would be at the radiation frequency ω’ and which would be travelling along
the z-direction will be very small. Thus the net energy absorbed per unit time in the volume
S𝑑𝑧 will be

(Γ12 - Γ21)ħω’S𝑑𝑧 = (𝜋2c3/ħω’3n03 tsp) ug(ω’)(N1 – N2) ħω’S𝑑𝑧

= (𝜋2c3/ω’2n03 tsp)ug(ω’)(N1 – N2)S𝑑𝑧

Now, the energy density u and the intensity of radiation I are related through the equation

I = 𝜐𝑢 = (c/n0 )𝑢
𝑑𝐼
Thus 𝑑𝑧 = -𝛼𝐼

Where 𝛼 = (𝜋2c3/ω2n03 tsp)g(ω’)(N1 – N2)

And the prime on ω removed with the understanding that ω represents the frequency of
the incident radiation. Hence, if N1>N2, 𝛼 is positive and the intensity decreases with z
leading to an attenuation of the beam. On the other hand, if N1<N2, then 𝛼 is negative and
the beam is amplified with z. this condition N2 > N1 is called population inversion and it is
under the condition that one can obtain optical amplification.

2.3 THRESHOLD CONDITION


A medium in the state of population inversion can amplify the incident radiation. The
medium will then act as an amplifier for those frequencies will fall within the linewidth. In
order to generate radiation this amplifying medium is placed in an optical resonator which
consists of a pair of mirrors facing each other. Radiations which bounce back and forth
between the mirrors is amplified by the amplifying medium and also suffers losses due to
the finite reflectivity of the mirrors and other scattering and diffraction losses. If the
oscillations have to be sustained in the cavity then the losses must be exactly compensated
PAGE 11

by the gain. Thus a minimum population inversion density is required to overcome the
losses and this is called the threshold population inversion.

Let us consider d to be the length of the resonator and R1 and R2 represent the reflectivities
of the mirrors. 𝛼 1 represents the average loss per unit length due to all loss mechanisms
such as scattering, diffraction due to finite mirror sizes. Let us consider a radiation with an
intensity I leaving mirror 1. As it propagates through the medium and reaches the second
mirror, it is amplified by 𝑒 −𝛼𝑑 and also suffers a loss of 𝑒 −𝛼1d . The intensity of the reflected
beam at the second mirror will be IR2𝑒 −(𝛼1+α)d . A second passage through the resonator and
a reflection at the first mirror leads to an intensity for the begin,

R1R2𝑒 −2(𝛼1+𝛼)𝑑 ≥ 1 (8)

The equality sign giving the threshold value for . Indeed, when the laser is oscillating in the
steady state with a continuous wave oscillation, then the equality sign in (8) must be
satisfied. If the inversion is increased, then the LHS becomes greater than unity; this implies
that the round trip gain is greater than the round trip loss. This would result in an increasing
intensity inside the laser till saturation effect takes over, which would result in a decrease in
the inversion. Thus the gain is brought back to the threshold value.

Equation (8) can be written as,

-𝛼 ≥ 𝛼1 – (1/2d)lnR1R2(9)

The RHS of the equation (9) depends on the passive cavity parameters only. This can be
related to the passive cavity lifetime tc which is the time in which the energy in the cavity
drops by a factor of 1/e. in the absence of the amplifying medium, the intensity at a point
reduces by a factor R1R2𝑒 −2𝛼1𝑑 = 𝑒 −(2𝛼1𝑑−𝑙𝑛𝑅1𝑅2) in a time corresponding to one round trip
time. One round trip time corresponds to t = 2d/(c/n0). Hence if the intensity reduces as
𝑒 −𝑡/𝑡𝑐 , then in a time t = 2dn0/c,

𝑒 −(2𝛼1𝑑−𝑙𝑛𝑅1𝑅2) = 𝑒 −2𝑑𝑛𝑜/𝑐𝑡𝑐

Or, 1/tc = (c/2dn0)(2𝛼 1d – lnR1R2)

Thus,

(N2 – N1) ≥ (4𝜐2n03tsp/c3tc)1/g(ω) (10)

Corresponding to the equality sign, we have the threshold population inversion density
required for the oscillation of the laser.

Conditions for threshold value being:

(a) The value of tc should be large as the cavity losses must be small.
PAGE 12

(b) Since g(ω) is normalised, the peak value will be inversely proportional to the width
∆𝜔 of the g(ω) function. The smaller widths give larger values of g(ω) which implies
smaller threshold values of (N2 – N1). Also since, the largest value of g(ω) appears on
the line centre, the resonator mode which lies closest to the line centre will reach
threshold first and begin to oscillate.
(c) Smaller values of tsp also lead to smaller values of threshold inversion. At the same
time for relaxation times, larger pumping power will be required to maintain a given
population inversion. In general, population inversion is more easily obtained on
transitions which have longer relaxation times.
(d) The value of g(ω) at the centre of the line is inversely proportional to ∆𝜔. Thus the
threshold population inversion increases approximately in proportion to ω 3. Thus it
is much easier to obtain laser action in infrared wavelengths than in the ultraviolet
region.
PAGE 13

3. LASER RATE EQUATIONS


3.1 TWO LEVEL SYSTEM
Considering a two level system consisting of energy levels E1 and E2 with N1 and N2natoms
per unit volume respectively. Let radiation at frequency ω with energy density u is incident
on the system. The number of atoms per unit volume which absorb the radiation and are
excited to the upper level is,

Γ12 = (𝜋2c3/ħω3n03tsp)ug(ω)N1 = W12N1 (11)

Where W12 = (𝜋2c3/ħω3n03tsp)ug(ω) (12)

The number of atoms undergoing stimulated emission from E2 to E1mper unit volume per
unit time will be

Γ21 W21N2 = W12N2 (13)

Where we considered the fact that absorption probability is the same as the stimulated
emission probability. In addition to the above two transitions, atoms in levelE 2 would also
undergo spontaneous transitions from E2 to E1. If A21 and S21 represent the radiative and
nonradiative transition rates from E2 to E1, then the number of atoms undergoing
spontaneous transitions from E2 to E1 will be T21N2 where,

T21 = A21 + S21 (14)

Thus we may write the rate of change of population energy levels E 2 and E1 as

dN1/dt = - W12(N1 – N2) + T21N2 (15)

dN2/dt = W12(N1 – N2) – T21N2 (16)

from (15) and (16),

(d/dt) (N1 + N2) = 0

i.e., N1 + N2 = Constant = N (17)

which is nothing but the fact that the total number of atoms per unit volume is constant. At
steady state,

dN1/dt = dN2/dt = 0 (18)

thus,
PAGE 14

N2/N1 = W12/(W12 +T21) (19)

Since bothW12 and T21 are positive, we can never obtain steady state population inversion by
optical pumping between just two levels.

Considering the population difference between two levels, from (19) we have,

(N2 – N1)/(N2 +N1) = - T21/(2W12 + T21)

∆𝑁/𝑁 = - 1/(1 + 2W12/T21) (20)

In order to put (20) in a different form, we consider that transitions from 2 to 1 is mostly
radiative, i.e., A21≫S21 and T21≈ A21. We also introduce a lineshape function ḡ(ω) which is
normalised to have unit value at ω = ω0, the centre of the line, i.e.,

ḡ(ω) = g(ω)/g(ω0)

since g(ω)≤ g(ω0) for all ω, 0< ḡ(ω) <1. Substituting the value of W12 in terms of u and
observing that u = n0I/c where I is the intensity of the incident radiation at ω, we have

W12/T21 = (𝜋2c3/ħω3n03tsp)I ḡ(ω)g(ω0)n0/ A21c

= (𝜋2c2/ħω3n02)I ḡ(ω)g(ω0) (21)

Hence (20) becomes,

∆𝑁/𝑁 = - 1/[1 + (I/IS) ḡ(ω)]

Where Is = ħω3n02/2c2𝜋2g(ω) is called the saturation intensity.


PAGE 15

3.2 THREE LEVEL SYSTEM [1]

FIGURE 3: ENERGY LEVELS OF A THREE LEVEL LASER. [c]

In order to produce a steady state population inversion, one makes use of either a three
level or a four level system.

We consider a three level system consisting of energy levels E1, E2 and E3 all of which are
assumed to be non degenerate. Let N1, N2 and N3 represent the population densities of the
three levels. The pump is assumed to lift atoms, from level 1 to 3 from which they decay
rapidly to level 2 through nonradiative process. Thus the pump effectively transfers atoms
from ground level 1 to excited level 2 which is now the upper level laser; the lower laser
level being the ground state 1. If the relaxation from level 3 to 2 is very fast, then the atoms
will relax down to level 2 rather than to level 1. Since the upper level 3 is not a laser level, it
can be a broad level so that a broadband light source may be effectively used as a pump
source.

If it is assumed that transitions take place between these three levels then,

N = N1 + N2 + N3 (22)

Where N represents the total number of atoms per unit volume.

The rate equations describing the rate of changing of N1, N2 and N3 are,

dN3/dt = Wp(N1 – N2) – T32N3 (23)


PAGE 16

where WP is the rate of pumping per atom from level 1 to 3 which depends on the pump
intensity. The first term in (23) represents stimulated transitions between levels 1 and 3 and
T32N2 represents the spontaneous transition rates from level 3 to level 2;

T32 = A32 + S32 (24)

A32 and S32 correspond respectively to the radiative and non radiative transition rates
between levels 3 and 2. In writing (23), T31N3 which corresponds to spontaneous transitions
between levels 3 and 1 since most atoms raised to level 3 make transitions to level 2 rather
than to level 1.

Also,

dN2/dt = W1(N1 – N2) + N3T32 – N2T21 (25)

and, dN1/dt = Wp(N3 – N1) + W1(N2 – N1) +N2T21 (26)

Where, W1 = ( 𝜋2c2/ħω3n02)A21g1(ω)I1 (27)

Represent the stimulated transition rate per atom between levels 1 and 2, I1 is the intensity
of the radiation in the 2 to 1 transitions and g1(ω) represents the line shape function
describing the transitions between levels 1 and 2. Also,

T21 = A21 +S21 (28)

With A21 and S21 represents the radiative and non radiative relaxation rates between levels 1
and 2. For good lasers since the transition must be mostly radiative, we assume, A 21≫ S21.

At steady state we must have,

dN1/dt = 0 = dN2/dt = dN3/dt (29)

from (23) we get,

N3 = WPN1/(WP + T32) (30)

From (25), (26) and (30),

N2 = [{W1(T32 + WP) + WPT32}N1]/(WP + T32)(W1 + T21) (31)

Thus from (22), (30) and (31),

(N2 – N1)/N = [WP(T32- T21) – T32T21]/[ 3WPW1 + 2WPT21 +2T32W1 + T32WP + T32T21] (32)

In order to get a steady population inversion, the sufficient condition is

T32> T21 (33)


PAGE 17

Since the lifetimes of level 3 and 2 are inversely proportional to the relaxation rates
according to (32), the lifetime of level 3 must be smaller than that of level 2 for attainment
of population inversion between levels 1 and 2. If the condition is satisfied, then according
to (32), there is a minimum pumping rate required to achieve population inversion which is
given by

Wpt = T32T21/(T32 – T21) (34)

If, T32 ≫ T21 and under the same approximations,

(N2 – N1)/N = ( Wp – T21)(WP + T21)W1/[1 + (3WP +2T32)/T32(WP + T21)] (35)

Below the threshold for laser oscillation, W1 is very small and hence we may write

(N2 – N1)/N =( Wp – T21)/(Wp + T21) (36)

Thus when W1 is small, i.e. when the intensity of the radiation corresponding to laser
transition is small, then the population inversion independent of I1 and thus there is an
exponential amplification of the beam. As the laser starts oscillating W1 becomes large and
from (35) there is a reduction if the inversion which in turn reduces the amplification.

If T32 is very large then, there will be very few atoms residing in level 3. Consequently we can
write,

N = N1 + N2 +N3≈ N1 + N2 (37)

Substituting in (36),

(N2 – N1)/(N2 + N1)= ( Wp –T21)/(Wp + T21)

Or, WPN1 = T21N2 (38)

The LHS of the above equation represents the number of atoms being lifted per unit volume
per unit time from level 1 to level 2 via level 3 and the RHS corresponds to the spontaneous
emission rate per unit volume from level 2 to level 1.
PAGE 18

3.3 FOUR LEVEL SYSTEM [1]

FIGURE 4: BASIC FOUR LEVEL LASER. [d]

In case of a three level system, one has to lift more than 50% of the atoms from the ground
level in order to obtain population inversion. By using another level of the atomic system
and having a lower level as an excited level instead of the ground level, lesser number of
atoms needs to be lifted. Level 1 is the ground level and levels 2, 3 and 4 are the excited
levels. Atoms from level 1 are pumped to level 4 from level 1 from where they a fast non
radiative relaxation to level 3. Level 3 serves as the upper laser level and is usually a
metastable level having a long lifetime. The transition from level 3 to 2 corresponds to a
laser transition. In order that the atoms do not accumulate in level 2 and hence destroy the
population inversion between the levels 3 and 2, level 2 must have a shorter lifetime so that
the atoms from level 2 are quickly removed to level 1 ready for pumping to level 4. If the
relaxation rate of atoms from level 2 to level 1 is faster than the rate of arrival of atoms to
level 2, then one can obtain population inversion between levels 3 and 2 even for low pump
powers. Level 4 can be a collection of large number of levels or a broad level. In such a case
an optical pump source emitting over a broad range of frequencies can be used to pump
atoms from level 1 to level 4 effectively. In addition, level 2 is required to be sufficiently
above the ground level so that, at ordinary temperatures, level 2 is mostly unpopulated. The
population of level 2 can also be reduced by lowering the temperature of the system.

Let us consider N1, N2, N3 and N4 are the population densities of the four levels. The rate of
change of N4 can be written as,

dN4/dt = Wp(N1 – N4) – T43N4 (39)


PAGE 19

where WpN1 is the number of atoms being pumped per unit time per unit volume. WpN4 is
the stimulated emission rate per unit volume

T43 = A43 + S43

Is the relaxation rate from level 4 to 3 and is the sum of radiative and nonradiative rates. In
writing (39), T42 and T41 are neglected in comparison to T43 assuming atoms relax from level
4 to level 3 and not to level 2 and 1. The rate equation for level 3 can be written as,

dN3/dt= W1(N2 – N3) +T43N4 – T32N3 (40)

where, W1 =( 𝜋2c2/ħ𝜔3n02)A32g(ω)I1 represents the stimulated transition rate per atom


between levels 3 and 2 and the subscript 1 stands for laser transitions.

T32 = A32 + S32 is the spontaneous relaxation rate level 3 to level 2 and consists of the
radiative and nonradiative contributions.

Also,

dN2/dt = W1(N3 – N2) + T32N3 - T21N2 (41)

dN1/dt = Wp(N4 – N1) + T21N2 (42)

Under steady state conditions,

dN1/dt = dN2/dt = dN3/dt = dN4/dt = 0

N = N1 + N2 +N3 + N4 (43)

From (39) we get,

WpN1 = (WP +T43)N4

Or, N4/N1 = WP/(WP + T43) (44)

From (40) we get,

T43N4 +W1N2 = (W1 + T32)N3 (45)

From (41) we get,

(W1 +T32)N3 = (W1 + T21)N2 (46)

N3/N2 = (W1+ T21)/(W1 + T32) (47)

From (39),

WP (N4 – N1) + T21N2 = 0

Substituting (44) we get,


PAGE 20

{WP2/(WP +T43) – WP}N1 + T21N2 = 0

N2/N1 = WPT43/T21(WP + T43) (48)

Substituting N2 and N4 in (46),

(W1 + T32)N3 = [WPT43(T21 + W1)/T21(WP + T43)]N1 (49)

Since the nonradiative decay is very fast, T43≫ WP

Using this approximation in, (47), (48) and (40),

N4/N1 =( WP/T43)/(WP/T43 +1 ) = 0 (50)

N2/N1 = WP/T21 (51)

From (50),

N3/N1 = (WPT43T21 + WPT43W1)/(T21WP + T21T43)(W1+ T32)

= (WPT21 + WPW1)/[T21T32 +(WPW1T21/T43) +W1T21 + (T21WPT32/T43)]

N3/N1 = WP(T21 + W1)/T21(W1 + T32) (52)

The total number of atoms,

N = N1 + N2 + N3 + N4

Substituting (50), (51) and (52) we get,

N = N1[{T21(W1 + T32) WP(W1 + T32) + WP(T21 + W1)}/T21(W1+ T32)]

Or, N/N1 = (T21W1 + T21T32 + WPW1 + WPT32 + WPT21 + WPW1)/T21(W1 + T32) (54)

Now,( N3 – N2)/N1 = WP/T21[(T21 – T32)/(W1+ T32) (55)

Dividing (55) by (54),

(N3 – N2)/N = (WPT21 – WPT32)/(T21W1 + T21T32 + 2WPW1 + WPT32 + WPT21) (56)

Using the approximation, T21≫ T32,

(N3 – N2)/N1 = [{WP – WP(T32/T21)}/{W1 +T32 + 2WP(W1/T21) + WP(T32/T21) +WP)

= WP/(T32 + WP)[1 + (T21W1 + 2WPW1)/T21(T32 + WP)

Just below the threshold for laser oscillations, W1≈ 0

Thus,

(N3 – N2)/N = WP/(WP + T32)


PAGE 21

4. TUNABLE LASER
A tunable laser is a tunable laser is a laser whose wavelength of operation can be altered in
a controlled manner. While all laser gain media allow small shifts in output wavelength, only
a few types of lasers allow continuous tuning over a significant wavelength range.

There are many types and categories of tunable lasers. They exist in the gas, liquid, and solid
state. Among the types of tunable lasers are excimer lasers, gas lasers (such as CO2 and He-
Ne lasers), dye lasers (liquid and solid state), transition metal solid-state lasers,
semiconductor crystal and diode lasers, and free electron lasers. Tunable lasers find
applications in spectroscopy, photochemistry, atomic vapor laser isotope separation, and
optical communications. Of operation can be altered in a controlled manner. While all laser
gain media allow small shifts in output wavelength, only a few types of lasers allow
continuous tuning over a significant wavelength range.

A few solid-state bulk lasers, in particular titanium–sapphire lasers and Cr:ZnS lasers allow
tuning over hundreds of nanometres in the near- and mid-infrared spectral region. (In
general, transition-metal doped gain media offer larger tuning ranges than rare-earth-doped
gain media, since the electrons involved in such media interact more strongly with the host
lattice; see the article on vibronic lasers.) Output powers can be hundreds or even
thousands of milliwatts.

In order to make a tunable laser the first element needed is a laser gain medium with a
broad gain profile. Once broadband laser emission is achieved in a simple mirror-mirror
cavity, lasing must be restricted to a single-transverse mode (TEM00) using a suitable
intracavity aperture. Once lasing is restricted to a TEM00 mode then one can proceed to limit
emission to a single longitudinal mode (SLM). This is done via the introduction of suitable
dispersive elements, such as prisms, gratings, or a combination of these. The dispersive
resonator assemblies can provide a very narrow window for intracavity transmission in the
frequency domain. A well designed multiple-prism grating assembly can restrict oscillation
to a SLM. Wavelength tuning is generally achieved by rotation of either a grating or a mirror.
PAGE 22

5.TITANIUM SAPPHIRE LASER [2]

Ti:sapphire lasers (also known as Ti:Al2O3 lasers, titanium-sapphire lasers, or Ti:sapphs) are
tunable lasers which emit red and near-infrared light in the range from 650 to 1100
nanometres. These lasers are mainly used in scientific research because of their tunability
and their ability to generate ultra short pulses. Lasers based on Ti:sapphire were first
constructed and invented in June 1982 by Peter Moulton at the MIT Lincoln Laboratory.

Titanium-sapphire refers to the lasing medium, a crystal of sapphire (Al2O3) that is doped
with titanium ions. A Ti:sapphire laser is usually pumped with another laser with a
wavelength of 514 to 532 nm, for which argon-ion lasers (514.5 nm) and frequency-doubled
Nd:YAG, Nd:YLF, and Nd:YVO lasers (527-532 nm) are used. Ti:sapphire lasers operate most
efficiently at wavelengths near 800 nm.

Titanium-doped sapphire (Ti3+:sapphire) is a widely used transition-metal-doped gain


medium for tunable lasers and femtosecond solid-state lasers. It was introduced in 1986,
and thereafter Ti:sapphire lasers quickly replaced most dye lasers, which had previously
dominated the fields of ultra short pulse generation and widely wavelength-tunable lasers.
Ti:sapphire lasers are also very convenient e.g. for pumping test setups of new solid-state
lasers (e.g. based on neodymium- or ytterbium-doped gain media), since they can easily be
tuned to the required pump wavelength and allow one to work with very high pump
brightness due to their good beam quality and high output power of typically several watts.

Sapphire is an ideal host crystal in both the ruby and the Ti:Al2O3 laser. It is transparent from
the ultraviolet to the infrared; also, it is non hygroscopic and very hard(it has a hardness of 9
on the Mohs scale, compared to10 for diamond), which is necessary for producing good
optical-quality surfaces that are not easily scratched. The thermal conductivity of sapphire,
which is one-tenth that of copper at room temperature and comparable to that of copper at
80 K, is high compared to other laser hosts. The excellent mechanical, thermal, and optical
properties ofTi:Al2O3 allow laser designs to be scaled to high average powers. For Ti:Al2O3
the overall power conversion efficiency can exceed 50%.
PAGE 23

FIGURE 5: CONFIGURATION DIAGRAM OF TITANIUM SAPPHIRE. [e]

5.1 LASER STRUCTURE


The laser uses an X-cavity design where the crystal typically ranges from 2 to 10mm in
length depending on the dopant level which is arranged with the output faces of the crystal
at Brewster’s angle. Longer length crystals with lower doping concentrations are used with
higher pumping flux intensities in order to obtain higher power output. Generally a CW
Argon ion laser or an Nd: YAG laser is used as a pumping source. The pump bema enters the
cavity at Brewster’s angle, can be rotated for wavelength tuning. A modified version of the
X- cavity is used to produce mode locked pulses which basically use two prisms for
intracavity dispersion compensation and also use Kerr lens mode locking technique. The
necessary aperture within the crystal to produce Kerr lens mode locking is provided by the
aperturing effect associated with the small diameter of the pump beam. Precise alignment
of the mirrors and cavity dimensions are essential to maintain a stable mode locked output
for this laser.

5.2 CRYSTAL STRUCTURE [3]

In titanium-doped sapphire the titanium ions substitute for the aluminium ions and (when
grown properly) exist in only the 3+ charge state. The energy levels of the titanium ions are
particularly simple to analyse because only a single d electron is in the outermost shell while
the remaining 18 electrons have the filled-shell configuration of a neutral argon atom. When
the titanium ions are placed in a host crystal, the electrostatic field of neighbouring atoms,
PAGE 24

or the crystal field, removes the fivefold angular momentum degeneracy of the single d
electron. In Ti:Al2O3 the 3d electron electrostatically interacts with the electronic charges of
six surrounding oxygen ions that are positioned at the corners of an octahedron. In three of
the five angular momentum states of the 3d electron (designated as the triplet T), the
orbitals do not point directly at the neighbouring oxygen atoms; these states have lower
energy than the two states in which the orbitals point directly at the oxygen atoms (the
doublet designated as E). This difference in energy corresponds to the energy of a green
photon (approximately 500 nm or 19,000cm-1), and absorption of green light causes
transitions from the ground state T to the excited state E. The same process occurs in
octahedrally coordinated [Ti(H2O)6]3+
complexes, which also absorb in the green.

The electronic energy levels of the Ti3+ ions


in Ti:Al2O3 are further perturbed by the
sapphire host lattice. When the Ti3+ ion is in
the excited state, the overall energy of the
system can be lowered if the position of the
Ti3+ ion displaces itself with respect to the
surrounding oxygen atoms (the Jahn-Teller effect). This displacement removes the
degeneracy of the two excited angular momentum states, which leads to a splitting of the
green absorption band. Also, as the Ti3+ion moves to its new equilibrium position, it kicks
the surrounding lattice and excites vibrations (or phonons); this action is why the Ti:Al2O3
laser is called a vibronic laser. The coupling of the electronic energy levels of theTi3+ ions
with the vibrational energy levels of the surrounding sapphire lattice is essential for Ti:Al2O3
to operate as a laser.

FIGURE 6: ENERGY LEVEL DIAGRAM OF TITANIUM. [f]


PAGE 25

5.3 PROPERTIES:-
Special properties of the Ti:sapphire gain medium are:

1. Sapphire (mono crystalline Al2O3) has an excellent thermal conductivity, alleviating


thermal effects even for high laser powers and intensities.
2. The Ti3+ ion has a very large gain bandwidth (much larger than that of rare-earth-
doped gain media), allowing the generation of very short pulses and also wide
wavelength tunability (typically using a birefringent tuner). The maximum gain and
laser efficiency are obtained around 800 nm. The possible tuning range is ≈ 650 nm
to 1100 nm, but different mirror sets are normally required for covering this huge
range, and exchanging mirror sets is a tedious task. (The number of mirror sets
required can be reduced by using ultra broadband chirped mirrors.)
3. There is also a wide range of possible pump wavelengths, which however are
located in the green spectral region (with the absorption peak at ≈490 nm), where
powerful laser diodes are not available. In most cases, several watts of pump
power are used, sometimes even 20 W. Originally, Ti:sapphire lasers were in most
cases pumped with 514-nm argon ion lasers, which are powerful, but very
inefficient, expensive to operate, and bulky. Other kinds of green lasers are now
available, and frequency-doubled solid-state lasers based on neodymium-doped
gain media are widely used. The pump wavelength is then typically 532 nm, with a
slightly reduced pump absorption efficiency compared with 514 nm. Direct diode
pumping at shorter wavelengths, e.g. at 455 nm with GaN-based laser diodes, is
also possible, but here one does not only have substantially reduced pump
absorption but also a detrimental induced loss which substantially further
degrades the performance.
4. The Ti3+ doping concentration has to be kept fairly low (e.g. 0.15% or 0.25%)
because otherwise no good crystal quality is possible. The therefore limited pump
absorption usually enforces the use of a crystal length of several millimetres,
which in combination with the small pump spot size (for high pump intensity)
means that rather high pump brightness is required.
PAGE 26

FIGURE 7: TRANSITION LEVELS IN THE CRYSTAL. [g]

5. The upper-state lifetime of Ti:sapphire is short (3.2 μs), and the saturation power
is very high. This means that the pump intensity needs to be high, so that a
strongly focused pump beam and thus a pump source with high beam quality are
required.
6. Despite the huge emission bandwidth, Ti:sapphire has relatively high laser cross
sections, which reduces the tendency of Ti:sapphire lasers for Q-switching
instabilities.

FIGURE 8: ABSORPTION AN EMISSION BANDS OF THE TITANIUM SAPPHIRE LASER.


[h]
PAGE 27

5.4 ADVANTAGES OVER OTHER LASERS [4]

1. ADVANTAGE OVER DYE LASERS: Though dye lasers possess a large gain bandwidth
which allows broad wavelength tunability and also ultra short pulse generation with
passive mode locking, the extremely short upper laser level lifetime of a few
nanoseconds make the dye lasers unsuitable for Q switching and long pulse pumping
operations. On the other hand, the longer upper laser level lifetime of about a few
microseconds and a larger laser cross section makes the titanium sapphire lasers
much resistant to Q switching instabilities. The lasing medium being a liquid makes
the dye lasers difficult to handle while the solid state crystals of titanium sapphire
are less hazardous and easier to maintain.
2. ADVANTAGE OVER EXCIMER LASERS: Operation of an excimer laser is extremely
limited in the ultraviolet region of the spectrum (150 to 350nm). Thus the laser does
not have a wide range of operation and cannot be tuned as compared to a titanium
sapphire laser that has a much larger range of operation (600 to 1100nm). Dealing
with poisonous gases makes the excimer lasers difficult to handle. The longer
lifetime of a titanium sapphire laser makes it more favourable than an excimer laser
as the gases which constitute the lasing medium are susceptible to contamination
due to their corrosive nature.
3. ADVANTAGE OVER LASER DIODES: As the working of semiconductor materials
depend on temperature, laser diode operation is highly temperature dependent. The
emission wavelength of laser diodes increase at a rate of 0.3nm per 1K rise in
temperature as a result of which extra care must be taken to keep the junction
temperature constant. The inconstancy of the junction temperature in turn affects
the tunability of the laser. Titanium sapphire lasers replace such lasers due to the
ease with which they can be tuned.
PAGE 28

6.FRANCK CONDON PRINCIPLE


Very often we can profit from a classical or semi classical picture of the interaction of light
with molecules. It is easy to imagine a hetero nuclear diatomic molecule as two partially
charged masses connected by a spring, oscillating and absorbing energy due to a resonantly
oscillating electric field. This picture enhances the understanding of infrared vibrational
transitions.

In electrical absorption and emission spectra, we also have a classical picture of sorts,
namely the Franck Condon principle of a vertical transition in which the sluggish nuclei
retain their position and momentum while the electrons make a quick transition.

The Franck–Condon principle is a rule in spectroscopy and quantum chemistry that explains the
intensity of vibronic transitions. Vibronic transitions are the simultaneous changes in electronic and
vibrational energy levels of a molecule due to the absorption or emission of a photon of the
appropriate energy. The principle states that during an electronic transition, a change from one
vibrational energy level to another will be more likely to happen if the two vibrational wave
functions overlap more significantly.

Classically, the Franck–Condon principle is the approximation that an electronic transition is


most likely to occur without changes in the positions of the nuclei in the molecular entity
and its environment. The resulting state is called a Franck–Condon state, and the transition
involved a vertical transition. The quantum mechanical formulation of this principle is that
the intensity of a vibronic transition is proportional to the square of the overlap integral
between the vibrational wavefunctions of the two states that are involved in the transition.
The Franck–Condon principle is applied equally to absorption and to fluorescence.
PAGE 29

FIGURE 9: TRANSITIONS BETWEEN TWO VIBRATION MODES OF A CRYSTAL. [i]

6.1 BASIC MECHANISM


The presumed mechanism was the excitation of a molecule by a photon, followed by a
collision with another molecule during the short period of excitation. The question was
whether it was possible for a molecule to break into photoproducts in a single step, the
absorption of a photon, and without a collision. In order for a molecule to break apart, it
must acquire from the photon a vibrational energy exceeding the dissociation energy, that
is, the energy to break a chemical bond. However, as was known at the time, molecules will
only absorb energy corresponding to allowed quantum transitions, and there are no
vibrational levels above the dissociation energy level of the potential well. High-energy
photon absorption leads to a transition to a higher electronic state instead of dissociation. In
examining how much vibrational energy a molecule could acquire when it is excited to a
higher electronic level, and whether this vibrational energy could be enough to immediately
break apart the molecule, he drew three diagrams representing the possible changes in
binding energy between the lowest electronic state and higher electronic states
PAGE 30

6.2 QUANTUM MECHANICAL


FORMULATION
Consider an electrical dipole transition from the initial vibrational state (𝜈) of the ground
electronic level (𝜖)│𝜖𝜈⟩, 𝑡𝑜 𝑠𝑜𝑚e vibrational level (ν’) of an excited electronic state (𝜖′) ,
|𝜖′𝜈′⟩. The molecular dipole operator 𝝁 is determined by the charge (-e) and location (ri) of
the electrons as well as the charges (Zje) and location (Rj) of the nuclei:

𝜇 = μe +μN

= -e∑ 𝑟I + e∑ 𝑍𝑗Rj

The probability amplitude P for the transition between these two states is given by,

P = ⟨𝜓′|𝝁|𝜓⟩ = ∫ 𝜓′ *μ𝜓𝑑𝜏

Where 𝜓′ and 𝜓 are, respectively, the overall wavefunctions of the initial and final state.
The overall wavefunctions are the product of the individual vibrational (depending on
spatial coordinates of the nuclei) and electronic space and spin wavefunctions:

𝜓 = 𝜓e𝜓v𝜓s

This separation of the electronic and vibrational wavefunctions is an expression of the Born–
Oppenheimer approximation and is the fundamental assumption of the Franck–Condon
principle. Combining these equations leads to an expression for the probability amplitude in
terms of separate electronic space, spin and vibrational contributions:

P = ⟨ 𝜓𝑒′𝜓𝑣′𝜓𝑠′|𝝁|𝜓𝑒𝜓𝑣𝜓𝑠⟩

= ∫ 𝜓e*’𝜓v’*𝜓s’*(μe+ μN) 𝜓e𝜓v𝜓s𝑑𝜏

=∫ 𝜓v’*𝜓v𝑑𝜏𝑛 ∫ 𝜓e’*μ𝜓e𝑑𝜏e∫ 𝜓s’*𝜓s𝑑𝜏s + ∫ 𝜓e’*𝜓e𝑑𝜏e∫ 𝜓v’*μN𝜓v𝑑𝜏v∫ 𝜓s’*𝜓s𝑑𝜏s

Where the first integral in the first term is the Franck Condon factor. The second integral in
the first term is the orbital selection rule and the third term is the spin selection rule.

The spin-independent part of the initial integral is here approximated as a product of two
integrals:

∬ 𝜓v’*𝜓e’*μ𝜓e𝜓v𝑑𝜏e𝑑𝜏n = ∫ 𝜓v’*𝜓v𝑑𝜏n∫ 𝜓e’*μ𝜓e𝑑𝜏e

This factorisation will be exact if the integral ∫ 𝜓e’*μ𝜓e𝑑𝜏eover the spatial coordinates of
the electrons would not depend on the nuclear coordinates. However, in the Born–
PAGE 31

Oppenheimer approximation 𝜓e and 𝜓e’do depend (parametrically) on the nuclear


coordinates, so that the integral (a so-called transition dipole surface) is a function of
nuclear coordinates. Since the dependence is usually rather smooth it is neglected (i.e., the
assumption that the transition dipole surface is independent of nuclear coordinates, called
the Condon approximation is often allowed).

The first integral after the plus sign is equal to zero because electronic wavefunctions of
different states are orthogonal. Remaining is the product of three integrals. The first integral
is the vibrational overlap integral, also called the Franck–Condon factor. The remaining two
integrals contributing to the probability amplitude determine the electronic spatial and spin
selection rules.

The Franck–Condon principle is a statement on allowed vibrational transitions between two


different electronic states; other quantum mechanical selection rules may lower the
probability of a transition or prohibit it altogether. Rotational selection rules have been
neglected in the above derivation. Rotational contributions can be observed in the spectra
of gases but are strongly suppressed in liquids and solids.

The quantum mechanical formulation of the Franck–Condon principle is the result of a


series of approximations, principally the electrical dipole transition assumption and the
Born–Oppenheimer approximation. For any given transition, the value of P is determined
by all of the selection rules; however spin selection is the largest contributor, followed by
electronic selection rules.
PAGE 32

7. JAHN TELLER EFFECT [6]

The Jahn Teller theorem essentially states that "A nonlinear polyatomic system in a spatially
degenerate electronic state distorts spontaneously in such a way that the degeneracy is
lifted and a new equilibrium structure of lower symmetry is attained."

The Jahn–Teller effect (JT effect or JTE) is an important mechanism of spontaneous


symmetry breaking in molecular and solid-state systems which has far-reaching
consequences for different fields, and it is related to a variety of applications in
spectroscopy, stereochemistry and crystal chemistry, molecular and solid-state physics, and
materials science. The Jahn–Teller effect, sometimes also known as Jahn–Teller distortion,
describes the geometrical distortion of molecules and ions that is associated with certain
electron configurations. The Jahn–Teller theorem essentially states that any nonlinear
molecule with a spatially degenerate electronic ground state will undergo a geometrical
distortion that removes that degeneracy, because the distortion lowers the overall energy
of the species. For a description of another type of geometrical distortion that occurs in
crystals with substitutional impurities.

The phenomenon is very common in six-coordinate copper(II) complexes. The d9 electronic


configuration of this ion gives three electrons in the two degenerate Eg orbitals, leading to a
doubly degenerate electronic ground state. Such complexes distort along one of the
molecular fourfold axes (always labeled the z axis), which has the effect of removing the
orbital and electronic degeneracies and lowering the overall energy. The distortion normally
takes the form of elongating the bonds to the ligands lying along the z axis, but occasionally
occurs as a shortening of these bonds instead (the Jahn–Teller theorem does not predict the
direction of the distortion, only the presence of an unstable geometry). When such an
elongation occurs, the effect is to lower the electrostatic repulsion between the electron-
pair on the Lewis basic ligand and any electrons in orbitals with a z component, thus
lowering the energy of the complex. Inversion centre is preserved after the distortion.

In octahedral complexes, the Jahn–Teller effect is most pronounced when an odd number of
electrons occupy the Eg orbitals. This situation arises in complexes with the configurations
d9, low-spin d7 or high-spin d4 complexes, all of which have doubly degenerate ground
states. In such compounds the Eg orbitals involved in the degeneracy point directly at the
ligands, so distortion can result in a large energetic stabilisation. Strictly speaking, the effect
also occurs when there is degeneracy due to the electrons in the T2g orbitals (i.e.
PAGE 33

configurations such as d1 or d2, both of which are triply degenerate). In such cases, however,
the effect is much less noticeable, because there is a much smaller lowering of repulsion on
taking ligands further away from the T2g orbitals, which do not point directly at the ligands
(see the table below). The same is true in tetrahedral complexes (e.g. manganate: distortion
is very subtle because there is less stabilisation to be gained because the ligands are not
pointing directly at the orbitals.

Each Titanium atom bonds to six neighboring Oxygen atoms, so the Titanium atoms
themselves exist effectively as Ti3+ions. The electronic structure is thus an inert argon shell
with a single 3d-orbital electron. In total there are five 3d-orbital arrangements, and they
are then split into two upper and lower energy manifolds via Stark shift due to the fields of
surrounding Oxygen atoms and spin-orbit coupling. The cubic field from the surrounding
Oxygen atoms splits the 5 free levels into an excited doublet level and ground state triplet
level, denoted as E and T respectively. The further fine splitting via the trigonal field and
spin-orbit coupling we can neglect. We can gain further insight by plotting the energies of E
and T as a function of the Ti3+ ion’s displacement from its equilibrium lattice position. As a
result of the Jahn-Teller effect, the energies are distorted in the displacement domain. The
continuum of energies formed by the upper and lower levels is sometimes called vibronic
levels1, since the electron can relax between them via emission of phonons and
displacement of the whole Ti3+ ion. At equilibrium, the 3d electron in the ground state T can
be excited into an upper vibronic metastable continuum in E by absorption in the
green/blue. The electron will then decay between the vibronic levels very quickly into the
upper lasing level. The distorted band gap is now smaller, allowing for emission in the red.
Another phonon emission process then allows for the Ti3+ to return to its equilibrium
position. We can therefore approximate the entire system as an effective four-level system,
with the lasing and metastable levels corresponding to different lattice configurations of the
Ti3+ ion.

FIGURE 10: TRANSITION METALS SHOWING WEAK JAHN TELLER DISTORTION. [j]
PAGE 34

8. APPLICATIONS IN
SPECTROSCOPIC STUDIES [7]

1. INTRACAVITY ABSORPTION SPECTROSCOPY


Intracavity laser absorption spectroscopy is a technique for highly sensitive spectroscopic
measurements. The basic principle is that the substance to be evaluated (e.g. some gas
sample) is placed within the resonator of a laser, which is preferably based on a gain
medium with broad gain bandwidth and a resonator with low losses. When the laser is
turned on, it starts to oscillate on many resonator modes simultaneously; only after many
resonator round trips will the optical spectrum of the generated light strongly concentrate
to the spectral region with highest gain. During this evolution, weak absorption features of
the substance under consideration can imprint signatures on the spectrum, because they
can influence the spectrum during many round trips. A measurement of the spectrum is
done some time after switching on the laser; this time should be long enough to allow for
strong spectral features to develop, but also short enough to prevent too strong narrowing
of the spectrum caused by the finite gain bandwidth.

2. TWO PHOTON SPECTROSCOPY


The passively mode-locked titanium: sapphire lasers help to acquire two-photon spectral
data in the near-infrared, a region not commonly accessible to synchronously pumped dye
lasers. This source generates pulses with peak powers near 100 kW at average powers over
1 W and is capable of yielding two-photon signals roughly two orders of magnitude larger
than is possible with synchronously pumped dye lasers. However, the multimode output of
this laser exhibits significant temporal and spectral pulse variations as the laser wavelength
is tuned. As a consequence, peak powers of the titanium:sapphire laser can vary
independently from average power across the tuning range. This wavelength dependence,
coupled with the quadratic dependence of the two-photon signal upon the instantaneous
power of the laser, precludes simple average power correction of nonlinear spectral band
shapes. Here, we investigate the key properties of the titanium:sapphire laser as an
excitation source for two-photon spectroscopy. We also identify a chemical reference
suitable for obtaining source-corrected excitation spectra in the near-infrared using a
double-beam, ratio metric approach; this is based on a source-in-dependent two-photon
excitation spectrum for the laser dye coumarin-480 that has been obtained with single-
frequency titanium sapphire laser.
PAGE 35

FIGURE 11: ENERGY LEVEL DIAGRAM OF TWO PHOTON SPECTROSCOPY. [k]


PAGE 36

9. CONCLUSION
Since the beginning of the 1990s, titanium sapphire has become the crystal of choice for the
development of ultra short laser systems producing very short and powerful pulses using
the Chirped Pulse Amplification technique and is considered to be the most successful solid
state laser material, in the near infrared wavelength range due to its high saturation energy,
large stimulated emission cross-section and very broad absorption gain. Owing to the given
characteristics, these lasers find huge applications in the high-energy physics, spectroscopy,
tunable optical parametric oscillators for continuous-wave(CW) operation, ultra short pulse
generations and power amplifications.

The only pronounced disadvantage of Ti:Sapphire is that its pumping requirements are
somewhat inconvenient in terms of pump wavelength and beam quality. Due to the
relatively weak absorption peak in the blue-green wavelength range, the successful
operation of such a laser requires high power blue-green pump sources. Usually, Titanium
sapphire lasers are pumped with multi-watt Argon ion lasers, frequency-doubled solid state
green lasers or Nd:YAG lasers which result in a fairly bulky, expensive and complicated
setups. So, further improvement can be brought about if new laser crystals doped with
Chromium or Ytterbium are employed in order to reach wavelength ranges directly and to overcome
the need to develop CW or pulsed green laser to pump the titanium sapphire crystal. On the other
hand optically-pumped semiconductor lasers in the green wavelength region can be used as the
principle to pump continuous wave titanium sapphire lasers.

Ultrashort pulses can have enormous peak powers even if the pulse energy stays moderate.
This kind of peak power applied to the typically quite small mode areas leads to enormous
optical intensities, which can give rise to a number of nonlinear effects or even optical
damage.

In addition to nonlinear effects, the large optical bandwidth of ultrashort pulses makes them
sensitive to chromatic dispersion. There are a substantial number of methods for keeping
such effects under control. Typically, they involve the introduction of additional dispersive
elements.
PAGE 37

10. REFERENCES
1. Ghatak and Thyagarajan; Optical electronics
2. P.F. Moulton; Spectroscopic and laser characteristics of Ti:Al2O3, paper
published in 1986.
3. K. F. Wall and Sanchez; Titanium Sapphire lasers
4. RP Photonics.
5. Aditya Vardhan Vutturi; Jahn Teller Distortion.
6. Research Gate; intracavity absorption spectroscopy, two photon
spectroscopy.

IMAGES:

1. Google images; [a], [b}, [c], [d], [g], [h], [i], [j], [k]
2. K. F. Wall and Sanchez; Titanium Sapphire lasers ; [e], [f]
PAGE 38

11. ACKNOWLEDGEMENT
I wish to express my deep sense of gratitude and respect to Dr. Indranath Chaudhuri,
professor of physics, St. Xavier’s College, Kolkata for the invigorating discussions and
constant encouragement during the course of this dissertation.
PAGE 39

You might also like