26.09.2014 Views

Lecture 8: Laser amplifiers

Lecture 8: Laser amplifiers

Lecture 8: Laser amplifiers

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Lecture</strong> 8: <strong>Laser</strong> <strong>amplifiers</strong><br />

<br />

<br />

<br />

<br />

Optical transitions<br />

Optical absorption and amplification<br />

Population inversion<br />

Coherent optical <strong>amplifiers</strong>: gain,<br />

nonlinearity, noise<br />

References: This lecture follows the materials from Photonic Devices, Jia-Ming Liu,<br />

Chapter 10. Also from Fundamentals of Photonics, 2 nd ed., Saleh & Teich,<br />

Chapter 14.<br />

1


Intro<br />

<br />

<br />

<br />

<br />

<br />

<br />

The word laser is an acronym for light amplification by stimulated<br />

emission of radiation.<br />

However, the term laser generally refers to a laser oscillator, which<br />

generates laser light without an input light wave.<br />

A device that amplifies a laser beam by stimulated emission is called a<br />

laser amplifier.<br />

<strong>Laser</strong> light is generally highly collimated with a very small divergence<br />

and highly coherent in time and space. It also has a relatively narrow<br />

spectral linewidth and a high intensity in comparison with light generated<br />

from ordinary sources (e.g. light-emitting diodes)<br />

Due to the process of stimulated emission, an optical wave amplified by a<br />

laser amplifier preserves most of the characteristics including the<br />

frequency spectrum, the coherence, the polarization, the divergence and<br />

the direction of propagation of the input wave.<br />

Here, we discuss the characteristics of laser <strong>amplifiers</strong>. We will discuss<br />

laser oscillators in <strong>Lecture</strong> 9.<br />

2


Optical transitions<br />

3


Optical transitions<br />

<br />

<br />

<br />

Optical absorption and emission occur through the<br />

interaction of optical radiation with electrons in a material<br />

system that defines the energy levels of the electrons.<br />

Depending on the properties of a given material, electrons<br />

that interact with optical radiation can be either those<br />

bound to individual atoms or those residing in the energyband<br />

structures of a material such as a semiconductor.<br />

The absorption or emission of a photon by an electron is<br />

associated with a resonant transition of the electron<br />

between a lower energy level |1> of energy E 1 and an<br />

upper energy level |2> of energy E 2 .<br />

4


Photon-matter interaction processes<br />

There are three fundamental processes electrons make transitions<br />

between two energy levels upon a photon of energy.<br />

E = h 12 = E 2 – E 1<br />

• A two-level system is a model system that only contains<br />

two energy levels with which the photon interacts.<br />

5


Three basic photon-matter interaction processes<br />

<br />

Absorption – when the quantum energy h equals the energy<br />

difference between the two energy levels (a resonant<br />

condition); the atom gains a quantum of energy<br />

<br />

<br />

Stimulated emission – the emission of a photon is triggered<br />

by the arrival of another, resonant photon<br />

Spontaneous emission – when an atom emits a photon,<br />

losing a quantum of energy in the process<br />

<br />

<br />

Einstein in 1917 first pointed out that stimulated emission is<br />

essential in the overall balance between emission and<br />

absorption, about reaching thermal equilibrium for a system<br />

of atoms. (Einstein Relations)<br />

Stimulated emission was demonstrated in 1953 in the<br />

microwave frequency by Basov, Prokhorov and Townes<br />

(Nobel 1964)<br />

6


Three basic photon-matter interaction processes<br />

<br />

<br />

<br />

<br />

A photon emitted by stimulated emission has the same<br />

frequency, phase, polarization and propagation direction as<br />

the optical radiation that induces the process.<br />

Spontaneously emitted photons are random in phase and<br />

polarization and are emitted in all directions, though their<br />

frequencies are still dictated by the separation between the<br />

two energy levels, subject to a degree of uncertainty<br />

determined by the linewidth of the transition.<br />

Therefore, stimulated emission results in the amplification<br />

of an optical signal, whereas spontaneous emission adds<br />

noise to an optical signal.<br />

Absorption leads to the attenuation of an optical signal.<br />

7


Spontaneous emission<br />

An electron spontaneously falls from a higher energy level E 2<br />

to a lower one E 1 , the emitted photon has frequency<br />

= (E 2 –E 1 ) / h<br />

|2><br />

<br />

<br />

This photon is emitted in a random direction with arbitrary<br />

polarization.<br />

The probability of such a spontaneous jump is given<br />

quantitatively by the Einstein coefficient for spontaneous<br />

emission (known as the “Einstein A coefficient”) defined as<br />

|1><br />

A 21 = “probability” per second of a spontaneous jump from<br />

level |2> to level |1>.<br />

8<br />

8


Probability per second of a spontaneous emission<br />

<br />

<br />

For example, if there are N 2 population per unit volume in<br />

level |2> then N 2 A 21 per second make jumps to level |1>.<br />

The total rate at which jumps are made between the two<br />

levels is<br />

dN 2 /dt = -N 2 A 21<br />

A negative sign because the population<br />

of level 2 is decreasing<br />

<br />

Generally an electron can make jumps to more than one lower<br />

level, unless it is in the first (lowest) excited level.<br />

9<br />

9


Natural lifetime<br />

<br />

The population of level |2> falls exponentially with time as<br />

electrons leave by spontaneous emission.<br />

N 2 = N 20 exp(-A 21 t)<br />

<br />

The time in which the population falls to 1/e of its initial<br />

value is called the natural lifetime of level |2>,<br />

2 = 1/A 21<br />

<br />

The magnitude of this lifetime is determined by the actual<br />

probabilities of jumps from level |2> by spontaneous<br />

emission.<br />

10<br />

10


Spectral lineshape<br />

<br />

<br />

<br />

<br />

<br />

The spectral characteristic of a resonant transition is therefore<br />

never infinitely sharp.<br />

Any allowed resonant transition between two energy levels<br />

has a finite relaxation time constant because at least the<br />

upper level has a finite lifetime due to spontaneous emission.<br />

From Quantum Mechanics, the finite spectral width of a<br />

resonant transition is dictated by the uncertainty principle of<br />

quantum mechanics.<br />

Intuitively, any response that has a finite relaxation time in<br />

the time domain must have a finite spectral width in the<br />

frequency domain. (recall the impulse response discussed in<br />

<strong>Lecture</strong> 2)<br />

We will see that the rate of the induced transitions between<br />

two energy levels in a given system is directly proportional to<br />

the spontaneous emission rate from the upper to the lower<br />

level.<br />

11


Spectral lineshape<br />

<br />

<br />

For each particular resonant transition between two energy<br />

levels, there is a characteristic lineshape function g()<br />

^<br />

of<br />

finite linewidth that characterizes the optical processes<br />

associated with the transition.<br />

The lineshape function is generally normalized as<br />

<br />

ĝ() d ĝ() d 1<br />

0<br />

<br />

<br />

0<br />

ĝ() 2ĝ()<br />

Area = 1<br />

<br />

<br />

<br />

12<br />

12


Homogeneous broadening<br />

<br />

<br />

<br />

<br />

If all of the atoms in a material that participate in a resonant<br />

interaction associated with the energy levels |1> and |2> are<br />

indistinguishable, their responses to an electromagnetic field<br />

are characterized by the same resonance frequency 21 and<br />

the same relaxation constant 21 .<br />

In such a homogeneous system, the physical mechanisms that<br />

contribute to the linewidth of the transition affect all atoms<br />

equally.<br />

Spectral broadening due to such mechanisms is called<br />

homogeneous broadening.<br />

Previously (in Lect. 2), we discussed that such<br />

homogeneously broadened systems can be described as the<br />

damped response characterized by a single resonance<br />

13<br />

frequency and a single relaxation constant.


Homogeneous broadening<br />

<br />

<br />

<br />

In the interaction of a material with an optical field, the<br />

absorption and emission of optical energy are characterized<br />

by the imaginary part ” of the susceptibility of the material.<br />

Therefore, the spectral characteristics of optical absorption<br />

and emission due to a resonant transition in a homogeneously<br />

broadened medium are described by the Lorentzian lineshape<br />

function of ”(). (recall from Lect 2)<br />

Using the normalization condition, we find that the resonant<br />

transition between |1> and |2> has the following normalized<br />

Lorentzian lineshape function:<br />

ĝ() <br />

h<br />

2[( 21<br />

) 2 ( h<br />

/2) 2 ]<br />

where h is the FWHM of the lineshape<br />

14


Inhomogeneous broadening<br />

<br />

<br />

<br />

<br />

However, in many practical situations, the simple picture that<br />

gives rise to Lorentzian lineshape is not adequate.<br />

For example, because of the Doppler effect, gas atoms with<br />

different velocities have different effective resonance<br />

frequencies even if they are otherwise identical.<br />

In solids the slightly different environments in which the<br />

resonant atoms find themselves, such as random dislocations,<br />

impurities and strain fields, also give rise to different<br />

effective resonance frequencies for differently located but<br />

otherwise identical atoms.<br />

Thus, in many cases the actual emission line must be thought<br />

of as a superposition of a large number of Lorentzian lines,<br />

each with homogeneous width k and each with a distinct<br />

center frequency k .<br />

Ref: Optical resonance and two-level atoms, L. Allen and J. H. Eberly, pp. 7-10<br />

15


Spectral lineshape<br />

Spread in<br />

frequencies<br />

Gaussian lineshape<br />

<br />

<br />

<br />

The origin of inhomogeneous broadening. The individual Lorentzian<br />

emission lines associated with different atomic dipoles are oscillating at<br />

multiple distinct frequencies.<br />

If a dielectric material is made up of those atoms with such individual<br />

lines, its emission line will be the sum of the curves. When the individual<br />

lines are densely spaced over a frequency range large compared with their<br />

own individual widths, the total lineshape is termed inhomogeneously<br />

broadened.<br />

<br />

Ref: Optical resonance and two-level atoms, L. Allen and J. H. Eberly, pp. 7-10<br />

16


Transition rates<br />

<br />

<br />

<br />

<br />

<br />

<br />

The transition rate of a resonant optical process measures the<br />

probability per unit time for the process to occur.<br />

The transition rate of an induced process is a function of the<br />

spectral distribution of the optical radiation and the spectral<br />

characteristics of the resonant transition.<br />

The spectral distribution of an optical field is characterized by<br />

its spectral energy density u() – the energy density of the<br />

optical radiation per unit frequency interval at the optical<br />

frequency .<br />

The total energy density of the radiation<br />

The spectral intensity distribution I() = (c/n) u(), n is the<br />

refractive index of the medium<br />

The total intensity<br />

I <br />

<br />

<br />

0<br />

I()d<br />

u <br />

<br />

<br />

0<br />

u()d<br />

17


Spectral energy density<br />

<br />

<br />

The energy density of a radiation field u() (joules per unit<br />

volume per unit frequency interval) can be simply related to<br />

the intensity of a plane electromagnetic wave.<br />

If the intensity of the wave is I() (watts per unit area per<br />

frequency interval)<br />

u() c = I()<br />

where c is the velocity of light in free space (in the medium<br />

of refractive index n, u() c/n = I())<br />

V<br />

A<br />

Length c in a second<br />

18<br />

18


Transition rates<br />

<br />

For the upward transition from |1> to |2> associated with<br />

absorption in the frequency range between and +d is<br />

W 12<br />

()d B 12<br />

u()ĝ()d<br />

(s -1 )<br />

<br />

For the downward transition from |2> to |1> associated with<br />

stimulated emission<br />

W 21<br />

()d B 21<br />

u()ĝ()d<br />

(s -1 )<br />

<br />

The spontaneous emission rate is independent of the energy<br />

density of the radiation and is solely determined by the<br />

transition lineshape function<br />

W sp<br />

()d A 21 ĝ()d<br />

(s -1 )<br />

The A and B constants are the Einstein A and B coefficients.<br />

19


Radiative processes connecting two energy levels in<br />

thermal equilibrium<br />

Population N 2<br />

E 2<br />

Spontaneous<br />

emission<br />

Stimulated<br />

emission<br />

absorption<br />

h<br />

Population N 1<br />

<br />

Einstein (1917) demonstrated that the rates of the three<br />

transition processes of absorption (B 12 ), stimulated emission<br />

(B 21 ) and spontaneous emission (A 21 ) are related.<br />

E 1<br />

20<br />

20


Transition rates<br />

<br />

<br />

The total induced transition rates<br />

W 12<br />

<br />

W 21<br />

<br />

<br />

W 12<br />

()d B 12<br />

u()ĝ()d<br />

0<br />

<br />

<br />

<br />

0<br />

W 21<br />

()d B 21<br />

u()ĝ()d<br />

0<br />

<br />

<br />

0<br />

The total spontaneous emission rate is<br />

W sp<br />

<br />

<br />

<br />

0<br />

W sp<br />

()d A 21<br />

u()<br />

W 12<br />

= B 12<br />

u()<br />

N 2<br />

|2><br />

W 21<br />

= B 21<br />

u() W sp<br />

= A 21<br />

|1><br />

N 1<br />

21


Transition rates<br />

<br />

<br />

<br />

The induced and the spontaneous transition rates for a given<br />

system are directly proportional to one another.<br />

The relationship can be obtained by considering the<br />

interaction of blackbody radiation with an ensemble of<br />

identical atomic systems in thermal equilibrium.<br />

The spectral energy density of the blackbody radiation at a<br />

temperature T (known as thermal radiation or blackbody<br />

radiation) is given by Planck’s formula:<br />

u() 8n3 h 3<br />

c 3 1<br />

e h /k BT 1<br />

where k B is the Boltzmann constant, k B T is the thermal<br />

energy (k B T = 26 meV @ T = 300 K)<br />

22


Blackbody radiation<br />

<br />

<br />

<br />

A system under thermal equilibrium produces a radiation<br />

energy density u()(J Hz -1 m -3 ) which is identical to<br />

blackbody radiation.<br />

A blackbody absorbs 100% all the radiation falling on it,<br />

irrespective of the radiation frequency.<br />

If the inside of this body is in thermal equilibrium it must<br />

radiate as much energy as it absorbs and the emission from<br />

the body is therefore characteristic of the equilibrium<br />

temperature T inside the body<br />

=> this type of radiation is often called “thermal” radiation or<br />

blackbody radiation<br />

Thermal<br />

radiation<br />

23<br />

23


Planck’s law of blackbody radiation<br />

<br />

Planck showed that the radiation energy density for a<br />

blackbody radiating within a frequency range to +d is given<br />

by<br />

u = (8n h 3 /c 3 ) [exp(h/k B T) – 1] -1<br />

= (8n 2 /c 3 ) h [exp(h/k B T) – 1] -1<br />

Photon energy<br />

Photon density of states in a medium<br />

Of refractive index n<br />

(number of photon modes per<br />

volume per frequency interval)<br />

Photon probability of occupancy<br />

(average number of photons in each<br />

mode according to Bose‐Einstein<br />

distribution)<br />

24<br />

24


Planck’s law of blackbody radiation<br />

180<br />

Radiation energy density u() (JHz ‐1 m ‐3 )<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

1500 K<br />

1000 K<br />

0 5E+13 1E+14 1.5E+14 2E+14 2.5E+14 3E+14 3.5E+14<br />

Frequency (Hz)<br />

25<br />

25


Transition rates<br />

<br />

If N 2 and N 1 are the population densities per unit volume of<br />

the atoms in levels |2> and |1>, the number of atoms per unit<br />

volume making the downward transition per unit time<br />

accompanied by the emission of radiation in a frequency<br />

range from to +d<br />

N 2<br />

(W 21<br />

()W sp<br />

())d<br />

<br />

The number of atoms per unit volume making the upward<br />

transition per unit time<br />

N 1<br />

W 12<br />

()d<br />

26


Transition rates<br />

<br />

In thermal equilibrium, both the blackbody radiation spectral<br />

density and the atomic population density in each energy<br />

level should reach a steady state<br />

N 2<br />

(W 21<br />

()W sp<br />

()) N 1<br />

W 12<br />

()<br />

<br />

This is the principle of detailed balance in thermal<br />

equilibrium. The steady-state population distribution in<br />

thermal equilibrium:<br />

N 2<br />

N 1<br />

<br />

W 12<br />

()<br />

W 21<br />

()W sp<br />

() <br />

B 12<br />

u()<br />

B 21<br />

u() A 21<br />

27


Transition rates<br />

<br />

In thermal equilibrium at temperature T, the population ratio<br />

of the atoms in the upper and the lower levels follows the<br />

Boltzmann distribution.<br />

N 2<br />

N 1<br />

g 2<br />

g 1<br />

exp(h / k B<br />

T )<br />

where g 2 and g 1 are the degeneracy factors* of these energy<br />

levels, and the energy density<br />

u() <br />

A 21<br />

/ B 21<br />

g 1<br />

B 12<br />

<br />

exp(h / k B<br />

T)1<br />

g 2<br />

B 21 <br />

*In an atomic or molecular system, a given energy level<br />

usually consists of a number of degenerate quantummechanical<br />

states, which have the same energy.<br />

28


Boltzmann distribution<br />

Energy E<br />

E 2<br />

N 2


Transition rates<br />

Identify u() with Planck’s formula:<br />

A 21<br />

8n3 h 3<br />

B 21<br />

c 3<br />

g 1<br />

B 12<br />

g 2<br />

B 21<br />

The spontaneous radiative lifetime of the atoms in the level |2><br />

associated with the radiative spontaneous transition from |2> to<br />

|1> is<br />

sp<br />

1 1<br />

W sp<br />

A 21<br />

30


Transition rates<br />

<br />

<br />

Therefore, the spectral dependence of the spontaneous<br />

emission rate<br />

W sp<br />

() 1 ĝ()<br />

sp<br />

The transition rates of both of the induced processes of<br />

absorption and stimulated emission are directly proportional<br />

to the spontaneous emission rate.<br />

W 21<br />

() <br />

c 3<br />

8n 3 h 3 sp<br />

u()ĝ() <br />

c 2<br />

8n 2 h 3 sp<br />

I()ĝ()<br />

W 12<br />

() g 2<br />

g 1<br />

W 21<br />

()<br />

31


Transition cross section<br />

We often express the transition probability of an atom in its<br />

interaction with optical radiation at a frequency in terms of the<br />

transition cross section, () [m 2 , cm 2 ].<br />

W 21<br />

() I()<br />

h 21 ()<br />

W 12<br />

() I()<br />

h 12 ()<br />

hphoton-flux<br />

density)<br />

The emission cross section<br />

e<br />

() 21<br />

() <br />

c 2<br />

8n 2 2 sp<br />

ĝ()<br />

The absorption cross section<br />

a<br />

() 12<br />

() g 2<br />

g 1<br />

21<br />

() g 2<br />

g 1<br />

e<br />

()<br />

32


Transition cross section<br />

<br />

<br />

<br />

For the ideal Lorentzian lineshape in a homogeneously<br />

broadened medium<br />

ĝ() <br />

h<br />

2[( 21<br />

) 2 ( h<br />

/2) 2 ]<br />

The peak value of the lineshape occurs at the center of the<br />

spectrum and is a function of linewidth h only.<br />

2<br />

gˆ ( <br />

21)<br />

<br />

<br />

Thus, the peak value of the emission cross section at the<br />

center wavelength of the spectrum<br />

<br />

e<br />

h<br />

h<br />

2<br />

<br />

2<br />

4<br />

n <br />

<br />

2<br />

sp<br />

33


Characteristics of some laser materials<br />

Gain medium<br />

Wavelength<br />

(m)<br />

System<br />

Peak cross<br />

section e<br />

(m 2 )<br />

Spontaneous<br />

linewidth<br />

(gain<br />

bandwidth)<br />

<br />

sp<br />

2<br />

HeNe 0.6328 I, 4 3.0x10 -17 1.5 GHz 300 ns 30 ns<br />

Ruby (Cr 3+ :Al 2<br />

O 3<br />

) 0.6943 H, 3 1.25-2.5 x<br />

10 -24 330 GHz 3 ms 3 ms<br />

Nd:YAG 1.064 H, 4 2-10 x 10 -23 150 GHz 515 s 240 s<br />

Nd:glass 1.054 I, 4 4.0 x 10 -24 6 THz 330 s 330 s<br />

Er:fiber 1.53 H/I, 3 6.0 x 10 -25 5 THz 10 ms 10 ms<br />

Ti:sapphire 0.66-1.1 H,Q2 3.4x10 -23 100 THz 3.9 s 3.2 s<br />

Semiconductor 0.37-1.65 H/I, Q2 1-5 x 10 -20 10-20 THz ~1 ns ~1 ns<br />

H: homogeneously broadened; I: inhomogeneously broadened<br />

34<br />

34


Optical absorption and<br />

amplification<br />

35


Optical absorption and amplification<br />

For a monochromatic optical field at frequency and intensity<br />

I() = I(’-)<br />

W 21 = (I/h) e () and W 12 = (I/h) a ()<br />

<br />

The net power (time-averaged) that is transferred from the optical<br />

field to the material is the difference between that absorbed by the<br />

material and that emitted due to stimulated emission:<br />

W p = hW 12 N 1 –hW 21 N 2<br />

= [N 1 a () – N 2 e ()]I<br />

<br />

<br />

W p > 0 => net power absorption from the optical field<br />

W p < 0 => net power flows from the medium to the optical field<br />

36


Optical absorption and amplification<br />

absorption coefficient [m -1 , cm -1 ]<br />

() = N 1 a () – N 2 e () = (N 1 –(g 1 /g 2 )N 2 ) a ()<br />

gain coefficient [m -1 , cm -1 ]<br />

() = N 2 e () – N 1 a () = (N 2 –(g 2 /g 1 )N 1 ) e ()<br />

() > 0 and () < 0 if N 1 > (g 1 /g 2 ) N 2<br />

() > 0 and () < 0 if N 2 > (g 2 /g 1 ) N 1<br />

<br />

<br />

A material absorbs optical energy in its normal state of thermal<br />

equilibrium when the lower energy level is more populated than the upper<br />

energy level.<br />

A material must be in a nonequilibrium state of population inversion with<br />

the upper energy level more populated than the lower energy level in<br />

order to provide a net optical gain to the optical field. 37


Optical absorption and amplification<br />

For simplicity, in some later discussion we can assume the<br />

degeneracy of levels 1 and 2 are equal, i.e. g 1 = g 2<br />

absorption coefficient [m -1 , cm -1 ]<br />

() = N 1 a () – N 2 e () = (N 1 –N 2 ) ()<br />

gain coefficient [m -1 , cm -1 ]<br />

() = N 2 e () – N 1 a () = (N 2 –N 1 ) ()<br />

() > 0 and () < 0 if N 1 > N 2<br />

() > 0 and () < 0 if N 2 > N 1<br />

And e () = a () = ()<br />

38


Resonant optical susceptibility<br />

<br />

For resonant interaction of an isotropic medium with a<br />

monochromatic plane optical field at a frequency = 2,<br />

we have<br />

E(t) Ee it E *e it<br />

P r es<br />

(t) 0<br />

( res<br />

()Ee it * res()E *e it )<br />

<br />

where P res is the polarization contributed by the resonant<br />

transitions and res is the resonant susceptibility.<br />

The time-averaged power density absorbed by the medium is<br />

P<br />

2 <br />

W<br />

p<br />

E 2<br />

0"<br />

res<br />

( )<br />

| E | "<br />

res<br />

( )<br />

I<br />

t<br />

nc<br />

t<br />

39


Resonant optical susceptibility<br />

<br />

<br />

<br />

<br />

<br />

Relate the time-averaged power density absorbed by the<br />

medium to the population relation<br />

<br />

W<br />

p<br />

" res<br />

( )<br />

I [ N1<br />

a<br />

( )<br />

N<br />

2<br />

e<br />

( )]<br />

I<br />

nc<br />

The imaginary part of the susceptibility contributed by the<br />

resonant transitions between energy levels |1> and |2> is<br />

" res<br />

() nc<br />

[N 1 a<br />

() N 2<br />

e<br />

()]<br />

The real part ’ res () can then be found through the Kramers-<br />

Kronig relations (recall from Lect. 2)<br />

Recall from Lect. 2 that a medium has an optical loss if ”><br />

0, and it has an optical gain if ”< 0.<br />

It is also clear that there is a net power loss from the optical<br />

field to the medium if ” res > 0, but there is a net power gain<br />

for the optical field if ” res < 0.<br />

40


Resonant optical susceptibility<br />

<br />

The medium has an absorption coefficient given by<br />

() nc " () res<br />

<br />

( )<br />

"<br />

( )<br />

in the case of normal population distribution when ” res >0,<br />

whereas it has a gain coefficient given by<br />

<br />

nc<br />

In the case of population inversion when ” res


Resonant optical susceptibility<br />

<br />

When the phase information of the optical wave is of no<br />

interest, we can find the evolution of the intensity of the<br />

optical wave as it propagates through the medium.<br />

dI/dz = -I<br />

(-ve sign represents attenuation)<br />

in the case of optical attenuation when ” res > 0, and<br />

dI/dz = I<br />

in the case of optical amplification when ” res < 0<br />

42


Population inversion<br />

43


Population inversion and optical gain<br />

<br />

<br />

<br />

<br />

<br />

Population inversion is the basic condition for the presence of<br />

an optical gain.<br />

In the normal state of any system in thermal equilibrium, a<br />

low-energy state is always more populated than a high-energy<br />

state – no population inversion<br />

Population inversion in a system can only be accomplished<br />

through a process called pumping – actively exciting the<br />

atoms in a low-energy state to a high-energy state.<br />

Population inversion is a nonequilibrium state that cannot be<br />

sustained without active pumping. To maintain a constant<br />

optical gain we need continuous pumping to keep the<br />

population inversion at a constant level.<br />

Many different pumping techniques depending on the gain<br />

media: optical excitation, current injection, electric<br />

discharge, chemical reaction, and excitation with ion beams<br />

44


Population inversion<br />

<br />

A nonequilibrium distribution showing population inversion<br />

Energy E<br />

E 2<br />

E 1<br />

N 1<br />

N 2<br />

Population N<br />

45


Population inversion and optical gain<br />

<br />

<br />

<br />

<br />

The use of a particular pumping technique depends on the<br />

properties of the gain medium being pumped.<br />

The lasers and optical <strong>amplifiers</strong> are often made of either<br />

dielectric solid-state media doped with active ions, such as<br />

Nd:YAG and Er:glass fiber, or direct-gap semiconductors,<br />

such as GaAs and InP.<br />

For dielectric media, the most commonly used pumping<br />

technique is optical pumping either with incoherent light<br />

sources, such as flashlamps and light-emitting diodes, or<br />

with coherent light sources from other lasers.<br />

Semiconductor gain media can also be optically pumped,<br />

but they are usually pumped with electric current<br />

injection.<br />

46


Rate equations<br />

The net rate of change of population density in a given energy<br />

level is described by a rate equation.<br />

Here we only write the rate equations for the upper laser level<br />

|2> and the lower laser level |1>.<br />

In the presence of a monochromatic, coherent optical wave of<br />

intensity I at a frequency ,<br />

dN 2 /dt = R 2 –N 2 / 2 – (I/h) (N 2 e –N 1 a )<br />

dN 1 /dt = R 1 –N 1 / 1 + N 2 / 21 + (I/h) (N 2 e –N 1 a )<br />

where R 2 and R 1 are the total rates of pumping into energy<br />

levels |2> and |1>, and 2 and 1 are the fluorescence lifetimes<br />

(total lifetimes) of levels |2> and |1>. The rate of population<br />

decay, including radiative and nonradiative spontaneous<br />

relaxation from |2> to |1> is 1/ 21 .<br />

47


Rate equations<br />

<br />

Because it is possible for the population in level |2> to relax<br />

to other energy levels, the total population decay rate of level<br />

|2> is 1/ 2 1/ 21 .<br />

2 21 sp<br />

<br />

<br />

<br />

In an optical gain medium, level |2> is known as the upper<br />

laser level and level |1> is known as the lower laser level.<br />

The fluorescence lifetime 2 of the upper laser level is an<br />

important parameter that determines the effectiveness of a<br />

gain medium.<br />

In general, the upper laser level has to be a metastable state<br />

with a relatively large 2 for a gain medium to be useful.<br />

48


Population inversion<br />

<br />

Population inversion in a medium is generally defined as<br />

N 2 > (g 2 /g 1 ) N 1<br />

(N 2 > N 1 for g 1 = g 2 )<br />

<br />

<br />

However, this condition does not guarantee an optical gain at a particular<br />

optical frequency when the population in each level, |1> or |2>, is<br />

distributed unevenly among its sublevels.<br />

A better condition for population inversion to guarantee an optical gain at<br />

a given frequency <br />

N 2 e () – N 1 a () > 0<br />

<br />

The pumping requirement for the condition to be satisfied depends on the<br />

properties of a medium. For atomic and molecular media, there are three<br />

different basic systems. Each has a different pumping requirement to<br />

reach effective population inversion for an optical gain. The pumping<br />

requirement can be found by solving the coupled rate equations.<br />

49


Two-level systems<br />

pump<br />

h p<br />

h<br />

|2><br />

|1><br />

When the only energy levels involved in the pumping and the<br />

relaxation processes are the upper and the lower laser levels<br />

|2> and |1>, the system can be considered as a two-level<br />

system. (i.e. p = )<br />

Level |1> is the ground state with 1 = ∞, and level |2><br />

relaxes only to level |1> so that 21 = 2 .<br />

The total population density is N t = N 1 + N 2 .<br />

50


Two-level systems<br />

<br />

<br />

No matter how a true two-level system is pumped, it is not<br />

possible to achieve population inversion for an optical gain<br />

in the steady state.<br />

The optical pump for a two-level system has to be in<br />

resonance with the transition between the two levels –<br />

inducing both downward and upward transitions.<br />

|2><br />

pump<br />

h p<br />

h p<br />

<br />

While a pumping mechanism excites atoms from the lower<br />

energy level to the upper energy level, the same pump also<br />

stimulates atoms in the upper energy level to relax to the<br />

lower energy level.<br />

|1><br />

51


Two-level systems<br />

<br />

While a pumping mechanism excites atoms from the lower<br />

laser level to the upper laser level, the same pump also<br />

stimulates atoms in the upper laser level to relax to the lower<br />

laser level.<br />

R 2 = -R 1 = W 12p N 1 -W 21p N 2 ,<br />

where W 12p and W 21p are the pumping rates from 1 to 2 and<br />

from 2 to 1.<br />

<br />

<br />

Under these conditions, dN 2 /dt and dN 1 /dt are equivalent to<br />

each other (N 1 + N 2 = N t = constant).<br />

The upward (W 12p ) and downward (W 21p ) pumping rates are<br />

not independent of each other but are directly proportional to<br />

each other because both are associated with the interaction of<br />

the same pump source with a given set of energy levels.<br />

52


Two-level systems<br />

Take the upward pumping rate W 12p = W p and the downward<br />

pumping rate to be W 21P = p W p , where p is a constant that depends<br />

on the detailed characteristics of the two-level atomic system and<br />

the pump source.<br />

In the steady state when dN 2 /dt = dN 1 /dt = 0,<br />

N 2 e –N 1 a = [W p 2 ( e -p a )- a ]N t [1+(1+p)W p 2 + (I 2 /h)( e + a )] -1<br />

<br />

For optical pumping<br />

p = ep / ap = e ( p )/ a ( p ),<br />

where ap and ep are the absorption and emission cross sections at<br />

the pump wavelength.<br />

53


Two-level systems<br />

<br />

<br />

In a true two-level system, the energy levels |2> and |1> can<br />

each be degenerate with degeneracies g 2 and g 1 , but the<br />

population densities in both levels are evenly distributed<br />

among the respective degenerate states.<br />

In this situation, p = ep / ap = g 1 /g 2 = e / a<br />

N 2 e –N 1 a = - a N t [1+( e + a )(I/hW p / a ) 2 ] -1 < 0<br />

<br />

No matter how a true two-level system is pumped, it is clearly<br />

not possible to attain population inversion for an optical gain<br />

in the steady state.<br />

54


Two-level systems<br />

<br />

<br />

Intuitively, the pump for a two-level system has to be in<br />

resonance with the transition between the two levels, thus<br />

inducing downward transitions and upward transitions.<br />

In the steady state, the two-level system would reach thermal<br />

equilibrium with the pump at a finite temperature T, resulting<br />

in a Boltzmann population distribution<br />

N 2 /N 1 = (g 2 /g 1 ) exp(-h/k B T) without population inversion.<br />

55


Quasi-two-level systems<br />

|2><br />

pump<br />

h p<br />

h<br />

|1><br />

<br />

<br />

<br />

However, many laser gain media including laser dyes,<br />

semiconductor gain media, and some solid-state gain media,<br />

are often pumped as a quasi-two-level system.<br />

An energy level is split into a band of closely spaced, but not<br />

exactly degenerate, sublevels with its population density<br />

unevenly distributed among these sublevels.<br />

A system is a quasi-two-level system if either or both of the<br />

two levels involved are split in such a manner.<br />

56


Quasi-two-level systems<br />

<br />

<br />

<br />

By pumping such a quasi-two-level system properly, it is<br />

possible to reach the needed population inversion in the steady<br />

state for an optical gain at a particular laser frequency .<br />

Now the ratio p = ep / ap at the pump frequency p can be<br />

made different from the ratio e / a at the laser frequency due<br />

to the uneven population distribution among the sublevels<br />

within an energy level.<br />

The pumping requirements for a steady-state optical gain from<br />

a quasi-two-level system (see p.53)<br />

p = ep / ap < e / a ; W p > (1/ 2 ) a /( e –p a )<br />

<br />

Because the absorption spectrum is generally shifted to the<br />

short-wavelength side of the emission spectrum, these<br />

conditions can be satisfied by pumping sufficiently strongly at a<br />

higher transition energy than the photon energy corresponding<br />

to the peak of the emission spectrum.<br />

57


Three-level systems<br />

|3><br />

Nonradiative<br />

relaxation<br />

pump<br />

h p<br />

h<br />

|2><br />

|1><br />

<br />

<br />

Population inversion in steady state is possible for a threelevel<br />

system.<br />

The lower laser level |1> is the ground state (or is very close<br />

to the ground state, within an energy separation of above the upper<br />

laser level |2>.<br />

58


Population inversion in three-level systems<br />

Over a period the population in the metastable state N 2<br />

increases above those in the ground state N 1 .<br />

=>The population inversion is obtained between levels |2><br />

and |1>.<br />

<br />

Drawback: the three-level system generally requires very<br />

high pump powers because the terminal state of the<br />

stimulated transition is the ground state. More than half the<br />

ground state atoms must be pumped into the metastable state<br />

to attain population inversion.<br />

59


Three-level systems<br />

<br />

An effective three-level system satisfies the following<br />

conditions:<br />

• Population relaxation from level |3> to level |2> is very<br />

fast and efficient, ideally 2 >> 32 ≈ 3<br />

s. t. the atoms excited by the pump quickly end up in level<br />

|2><br />

• Level |3> lies sufficiently high above level |2> with E 32<br />

>> k B T s. t. the population in level |2> cannot be<br />

thermally excited back to level |3><br />

• The lower laser level |1> is the ground state, or its<br />

population relaxes very slowly if it is not the ground state.<br />

<br />

Under these conditions<br />

R 2 ≈ W p N 1 , R 1 ≈ -W p N 1 , and N 1 +N 2 ≈ N t<br />

1 ≈∞and 21 ≈ 2<br />

60


Three-level systems<br />

The parameter W P is the effective pumping rate for exciting an atom in the<br />

ground state to eventually reach the upper laser level. It is proportional to<br />

the power of the pump.<br />

In the steady state with a constant pump, W p is a constant and dN 2 /dt =<br />

dN 1 /dt = 0<br />

N 2 e –N 1 a = (W p 2 e - a )N t [1+W p 2 + (I 2 /h)( e + a )] -1<br />

<br />

The pumping condition for a constant optical gain under steady-state<br />

population inversion<br />

W p > a / 2 e<br />

<br />

<br />

This condition sets the minimum pumping requirement for effective<br />

population inversion to reach an optical gain in a three-level system.<br />

Note that almost all of the population initially resides in the lower laser<br />

level |1>. To attain population inversion, the pump has to be strong<br />

enough to depopulate sufficient population density from level |1>, while<br />

the system has to be able to keep it in level |2>. In the case when a = e<br />

(i.e. g 1 = g 2 ), no population inversion occurs before at least one-half of<br />

the total population is transferred from level |1> to level |2>.<br />

61


Erbium-doped silica fibers<br />

<br />

Er 3+ :silica fiber amplifier is a three-level system.<br />

3<br />

32<br />

pump<br />

2<br />

1.55 m<br />

• Pumping at 980 nm using InGaAs laser diodes; a mixture of<br />

homogeneous/inhomogeneous broadening; ~ 5.3 THz<br />

1<br />

• The laser transition can also be directly pumped at 1.48 m by<br />

light from InGaAsP laser diodes – like a quasi-two-level scheme<br />

62<br />

62


Four-level systems<br />

pump<br />

h p<br />

Nonradiative<br />

relaxation<br />

Nonradiative<br />

relaxation<br />

h<br />

|3><br />

|2><br />

|1><br />

|0><br />

<br />

<br />

A four-level system is more efficient than a three-level<br />

system.<br />

The lower laser level |1> lies sufficiently high above the<br />

ground level |0>, with E 10 >> k B T. Thus, in thermal<br />

equilibrium, the population in |1> is negligibly small<br />

compared with that in |0>. Pumping takes place from level<br />

|0> to level |3>.<br />

63


Four-level systems<br />

<br />

<br />

<br />

Levels |3> and |2> need to satisfy the same conditions as in a three-level<br />

system.<br />

The population in level |1> relaxes very quickly back to the ground level,<br />

ideally 1 ≈ 10 remains relatively unpopulated in<br />

comparison with level |2> when the system is pumped.<br />

Under these conditions,<br />

N 1 ≈ 0; R 2 ≈ W p (N t –N 2 )<br />

where the effective pumping rate W p is proportional to the pump power.<br />

In the steady state when W p is held constant, by taking dN 2 /dt = 0,<br />

(ignoring dN 1 /dt because N 1 ≈ 0)<br />

N 2 e –N 1 a ≈ N 2 e =(W p 2 e )N t [1 + W p 2 + (I 2 /h) e ] -1<br />

=> No minimum pumping requirement for an ideal four-level system because<br />

level |1> is initially empty. A practical four-level system is much more<br />

efficient than a three-level system.<br />

64


Neodymium-doped glass<br />

<br />

Nd 3+ :glass amplifier is a four-level system.<br />

3<br />

32<br />

pump<br />

0<br />

1<br />

2<br />

1.053 m<br />

1<br />

65<br />

65


Neodymium-doped glass<br />

<br />

Level 1 is 0.24 eV above the ground state. This is substantially<br />

larger than the thermal energy 0.026 eV at room temperature, so<br />

that the thermal population of the lower laser is negligible.<br />

Level 3 is a collection of four absorption bands, centered at 805,<br />

745, 585, and 520 nm.<br />

<br />

<br />

The excited ions decay rapidly from level 3 to level 2 and then<br />

remain in level 2 for a substantial time sp = 330 s. 1 is very<br />

short (~ 300 ps)<br />

The 2→1 transition is inhomogeneously broadened because of the<br />

amorphous nature of the glass, which presents a different<br />

environment at each ionic location. This material therefore has a<br />

large spontaneous linewidth (gain bandwidth) ≈ 6 THz<br />

66<br />

66


Coherent optical <strong>amplifiers</strong><br />

Gain, nonlinearity, noise


Coherent optical <strong>amplifiers</strong><br />

<br />

<br />

<br />

<br />

<br />

A coherent optical amplifier is a device that increases the<br />

amplitude of an optical field while maintaining its phase.<br />

If the optical field at the input to such an amplifier is<br />

monochromatic, the output will also be monochromatic with<br />

the same frequency.<br />

The output amplitude is increased relative to the input while<br />

the phase remains unchanged or is shifted by a fixed amount.<br />

In contrast, an incoherent optical amplifier increases the<br />

intensity of an optical wave without preserving its phase.<br />

Coherent optical <strong>amplifiers</strong> are important, for example, in the<br />

amplification of weak optical pulses that have traveled<br />

through a long length of optical fiber, and as a basis to<br />

understanding laser oscillators.<br />

68


Coherent light amplification<br />

<br />

<br />

<br />

As seen earlier, stimulated emission allows a photon in a<br />

given mode to induce an atom whose electron is in an upper<br />

energy level to undergo a transition to a lower energy level<br />

and, in the process, to emit a clone photon into the same<br />

mode as the initial photon. A clone photon has the same<br />

frequency, direction and polarization as the initial photon.<br />

These two photons in turn serve to stimulate the emission of<br />

two additional photons, and so on, while preserving these<br />

properties.<br />

The result is coherent light amplification. Because<br />

stimulated emission occurs only when the photon energy is<br />

nearly equal to the transition energy difference, the process is<br />

restricted to a band of frequencies determined by the<br />

transition linewidth.<br />

69


<strong>Laser</strong> amplification vs. electronic <strong>amplifiers</strong><br />

<br />

<br />

<br />

<strong>Laser</strong> amplification differs in a number of respects from<br />

electronic amplification.<br />

Electronic <strong>amplifiers</strong> rely on devices in which small changes<br />

in an injected electric current or applied voltage result in large<br />

changes in the rate of flow of charge carriers (electrons and<br />

holes in a semiconductor field-effect transistor). Tuned<br />

electronic <strong>amplifiers</strong> make use of resonant circuits (e.g. a<br />

capacitor and an inductor) to limit the gain of the amplifier to<br />

the band of frequencies of interest.<br />

In contrast, atomic, molecular, and solid-state laser <strong>amplifiers</strong><br />

rely on differences in their allowed energy levels to provide<br />

the principal frequency selection. These entities act as<br />

natural resonators that select the frequency of operation and<br />

bandwidth of the device.<br />

70


Population inversion<br />

<br />

<br />

<br />

<br />

Light transmitted through matter in thermal equilibrium is<br />

attenuated.<br />

This is because absorption by the large population of atoms in<br />

the lower energy level is more prevalent than stimulated<br />

emission by the smaller population of atoms in the upper<br />

level.<br />

An essential ingredient for attaining laser amplification is the<br />

presence of a greater number of atoms in the upper energy<br />

level than in the lower level. This is a nonequilibrium<br />

situation.<br />

Attaining such a population inversion requires a source of<br />

power to excite (pump) the atoms to the higher energy level.<br />

71


Ideal coherent <strong>amplifiers</strong><br />

<br />

<br />

<br />

<br />

An ideal coherent amplifier is a linear system that increases<br />

the amplitude of the input signal by a fixed factor, the<br />

amplifier gain.<br />

A sinusoidal input leads to a sinusoidal output at the same<br />

frequency, but with larger amplitude.<br />

The gain of the ideal amplifier is constant for all frequencies<br />

within the amplifier spectral bandwidth.<br />

The amplifier may impart to the input signal a phase shift that<br />

varies linearly with frequency (corresponding to a time delay<br />

at the output with respect to the input).<br />

phase<br />

gain<br />

<br />

<br />

Output amp.<br />

Input amplitude<br />

72


Real coherent <strong>amplifiers</strong><br />

<br />

<br />

<br />

<br />

<br />

Real coherent <strong>amplifiers</strong> deliver a gain and phase shift that are frequency<br />

dependent. The gain and phase shift determine the amplifier’s transfer<br />

function.<br />

For a sufficiently large input amplitude, real <strong>amplifiers</strong> generally exhibit<br />

saturation, a form of nonlinear behavior in which the output amplitude<br />

does not increase in proportion to the input amplitude.<br />

Saturation introduces harmonic components into the output, provided that<br />

the amplifier bandwidth is sufficiently broad to pass them.<br />

Real <strong>amplifiers</strong> also introduce noise, s.t. a random fluctuating component<br />

is present at the output, regardless of the input.<br />

An amplifier may therefore be characterized by the following features:<br />

• Gain<br />

• Bandwidth<br />

• Phase shift<br />

• Power source<br />

• Nonlinearity and gain saturation<br />

• Noise<br />

73


Theory of laser amplification<br />

<br />

<br />

<br />

A monochromatic optical plane wave traveling in the z<br />

direction with frequency , electric field<br />

Intensity<br />

E( z)<br />

Re[ E(<br />

z)exp(<br />

i2t<br />

)]<br />

I( z)<br />

|<br />

E(<br />

z)<br />

Photon-flux density (photons per second per unit area)<br />

2<br />

|<br />

<br />

<br />

( z ) I(<br />

z)<br />

/ h<br />

Consider the atomic medium (gain or active medium) with<br />

two relevant energy levels whose energy difference nearly<br />

matches the photon energy h.<br />

The numbers of atoms per unit volume in the lower and upper<br />

energy levels are denoted N 1 and N 2 . (assume g 1 = g 2 )<br />

74


Gain and bandwidth<br />

<br />

<br />

The wave is amplified with a gain coefficient () (per unit<br />

length) and undergoes a phase shift () (per unit length).<br />

() > 0 corresponds to amplification, () < 0 corresponds<br />

to attenuation.<br />

Recall that the probability density (s -1 ) that an unexcited<br />

atom absorbs a single photon is<br />

<br />

W i<br />

( )<br />

where the transition cross section at the frequency <br />

2<br />

c<br />

( ) <br />

2 2<br />

8n<br />

<br />

gˆ(<br />

)<br />

Here we assume the probability density for stimulated<br />

emission is the same as that for absorption. ( a () = e () =<br />

())<br />

sp<br />

75


Gain coefficient<br />

<br />

<br />

<br />

<br />

The average density of absorbed photons (number of photons<br />

per unit time per unit volume) is N 1 W i .<br />

Similarly, the average density of clone photons generated as a<br />

result of stimulated emission is N 2 W i .<br />

The net number of photons gained per second per unit<br />

volume is therefore NW i , where N = N 2 –N 1 is the<br />

population density difference.<br />

N is referred to as the population difference.<br />

• If N > 0, a population inversion exists, in which case the<br />

medium acts as an amplifier and the photon-flux density<br />

can increase.<br />

• If N < 0, the medium acts as an attenuator and the photonflux<br />

density decreases.<br />

• If N = 0, the medium is transparent.<br />

76


Gain coefficient<br />

<br />

<br />

<br />

As the incident photons travel in the z direction, the<br />

stimulated-emission photons also travel in this<br />

direction.<br />

An external pump providing a population inversion<br />

N > 0 then causes the photon-flux density (z) to<br />

increase with z.<br />

Because emitted photons stimulate further emissions,<br />

the growth at any position z is proportional to the<br />

population at that position. (z) thus increases<br />

exponentially.<br />

77


Gain coefficient<br />

<br />

<br />

Consider the incremental number of photons per unit area per<br />

unit time, d(z), is the number of photons gained per unit<br />

time per unit volume, NW i , multiplied by the thickness dz<br />

d<br />

<br />

NW dz<br />

In the form of a differential equation<br />

d<br />

dz<br />

( ) (<br />

z)<br />

i<br />

)<br />

where the gain coefficient<br />

2<br />

c<br />

( ) N<br />

( ) N<br />

2 2<br />

8n<br />

<br />

sp<br />

gˆ(<br />

<br />

78


Gain coefficient<br />

<br />

<br />

The coefficient () represents the net gain in the photon-flux<br />

density per unit length of the medium.<br />

The photon-flux density therefore is given as<br />

( z)<br />

(0)exp[<br />

( ) z]<br />

<br />

The optical intensity I(z) = h(z)<br />

I( z)<br />

I(0)exp[<br />

( ) z]<br />

<br />

Thus, () also represents the gain in the intensity per unit<br />

length of the medium.<br />

79


Gain<br />

<br />

For an interaction region of total length d, the overall gain of<br />

the laser amplifier G() is defined as the ratio of the photonflux<br />

density at the output to the photon-flux density at the<br />

input,<br />

G( ) (<br />

d) / (0)<br />

<br />

G( ) exp[ ( ) d]<br />

<br />

Note that in the absence of a population inversion, N is<br />

negative (N 2 < N 1 ) and so is the gain coefficient. The<br />

medium will then attenuate light traveling in the z direction.<br />

A medium in thermal equilibrium cannot provide laser<br />

amplification.<br />

80


Gain bandwidth<br />

<br />

<br />

<br />

<br />

<br />

The dependence of the gain coefficient () on the frequency<br />

of the incident light is contained in its proportionality to the<br />

lineshape function g().<br />

The latter is a function of width centered about the atomic<br />

resonance frequency 0 = (E 2 –E 1 )/h.<br />

The laser amplifier is therefore a resonant device, with a<br />

resonance frequency and bandwidth determined by the<br />

lineshape function of the atomic transition.<br />

This is because stimulated emission and absorption are<br />

governed by the atomic transition.<br />

The linewidth in frequency (Hz) and in wavelength<br />

(nm) are related by<br />

= |(c/)| = (c/ 2 ) = ( 2 /c)<br />

81


Gain bandwidth<br />

<br />

If the lineshape function is Lorentzian, the gain coefficient is<br />

then also Lorentzian with the same width<br />

( )<br />

2<br />

( <br />

/ 2)<br />

( <br />

0)<br />

2<br />

( <br />

) ( <br />

/ 2)<br />

0<br />

2<br />

where the peak gain coefficient at the central frequency 0<br />

( ) <br />

0<br />

2<br />

c<br />

2 2<br />

4<br />

n <br />

N 2<br />

sp<br />

<br />

82


Phase‐shift coefficient<br />

<br />

<br />

The laser amplification process also introduces a phase shift.<br />

When the lineshape is Lorentzian with linewidth <br />

g() = (/2) / [( – 0 ) 2 +(/2) 2 ]<br />

<br />

The amplifier phase shift per unit length turns out to be<br />

() = [( – 0 )/] ()<br />

<br />

This phase shift is in addition to that introduced by the medium<br />

hosting the laser atoms.<br />

83


Gain coefficient and phase-shift coefficient for a laser<br />

amplifier with a Lorentzian lineshape function<br />

gain<br />

coefficient<br />

()<br />

<br />

<br />

<br />

Phase-shift<br />

coefficient<br />

()<br />

<br />

<br />

(Compare these with the ideal coherent amplifier gain and phase responses on p. 72)<br />

84


Rate equations<br />

2<br />

R 2<br />

1/ 21 = 1/ sp + 1/ nr<br />

1<br />

sp<br />

nr<br />

1/ 2 = 1/ 21 + 1/ 20<br />

R 1<br />

1<br />

20<br />

1/ nr<br />

: non-radiative decay rate<br />

<br />

<br />

Steady-state populations of levels 1 and 2 can be maintained only if<br />

the energy levels above level 2 are continuously excited by<br />

pumping and ultimately populate level 2.<br />

Pumping serves to populate level 2 at rate R 2 and depopulate level<br />

1 at rate R 1 (per unit volume per second)<br />

=> levels 1 and 2 can attain non-zero steady-state populations.<br />

85<br />

85


Rate equations in the absence of amplifier radiation<br />

The rates of increase of the population densities of levels 2<br />

and 1 arising from pumping and decay are<br />

dN 2 /dt = R 2 –N 2 / 2<br />

dN 1 /dt = -R 1 –N 1 / 1 + N 2 / 21<br />

<br />

Steady-state population difference in the absence of amplifier<br />

radiation (dN 1 /dt = dN 2 /dt = 0)<br />

N 0 = N 2 –N 1 = R 2 2 (1- 1 / 21 ) + R 1 1<br />

<br />

A large gain coefficient requires a large population difference<br />

( 0 () = N 0 e ())<br />

86<br />

86


Rate equations in the absence of amplifier radiation<br />

To increase population difference N 0<br />

• Increase pumping and de-pumping rate (R 2 and R 1 )<br />

• Long 2 , but sp must be sufficiently short so as to make<br />

the radiative transition rate large<br />

• Short 1 (if R 1 < ( 2 / 21 )R 2 )<br />

<br />

The physical picture:<br />

• the upper level should be pumped strongly and decay<br />

slowly so that it retains its population.<br />

• The lower level should be de-pumped strongly so that it<br />

quickly disposes of its population.<br />

<br />

Ideally, 21 ≈ sp


Rate equations in the presence of amplifier radiation<br />

<br />

<br />

The presence of radiation near the resonance frequency 0 enables<br />

transitions between levels 2 and 1 to occur via stimulated emission and<br />

absorption.<br />

These processes are characterized by the probability density<br />

W i = ()<br />

where = I/h (assuming g 1 = g 2 and thus e () = a ())<br />

R 2<br />

2<br />

W<br />

-1<br />

i<br />

1<br />

R 1<br />

sp<br />

nr<br />

1<br />

20<br />

88<br />

88


Rate equations in the presence of amplifier radiation<br />

dN 2 /dt = R 2 –N 2 / 2 –N 2 W i + N 1 W i<br />

dN 1 /dt = -R 1 –N 1 / 1 + N 2 / 21 + N 2 W i -N 1 W i<br />

<br />

<br />

The population density of level 2 is decreased by stimulated<br />

emission from level 2 to level 1 and increased by absorption<br />

from level 1 to level 2.<br />

Under steady-state conditions (dN 1 /dt = dN 2 /dt = 0), the<br />

population difference in the presence of amplifier radiation<br />

(assuming g 1 = g 2 )<br />

N = N 2 –N 1 = N 0 /(1 + s W i )<br />

<br />

The characteristic time s (saturation time constant) is always<br />

positive ( 2 ≤ 21 ) is given by<br />

s = 2 + 1 (1 – 2 / 21 )<br />

89<br />

89


Depletion of the steady-state population difference<br />

Population difference N<br />

N 0<br />

N 0<br />

/2<br />

0<br />

s<br />

W i<br />

0.01 0.1 1 10<br />

If the radiation is sufficiently weak so that s W i


Four-level pumping<br />

Pump R<br />

<strong>Laser</strong> W i<br />

-1<br />

<br />

Rapid<br />

decay<br />

<br />

Rapid<br />

<br />

decay<br />

<br />

Short-lived |3><br />

Long-lived |2><br />

Short-lived |1><br />

Ground state |0><br />

<br />

Here we assume that the rate of pumping into level 3, and out<br />

of level 0, are the same.<br />

91<br />

91


Four-level pumping<br />

<br />

<br />

<br />

<br />

An external source of energy pumps atoms from level 0 to<br />

level 3 at a rate R.<br />

If the decay from level 3 to level 2 is sufficiently fast, it may<br />

be taken to be instantaneous, in which case pumping to level<br />

3 is equivalent to pumping level 2 at the rate R 2 = R.<br />

However, in this case, atoms are neither pumped into nor out<br />

of level 1, s.t. R 1 = 0.<br />

Thus, in the absence of amplifier radiation (W i = = 0), the<br />

steady-state population difference is given by (see p.86)<br />

<br />

N 0<br />

R 2<br />

1 1<br />

<br />

<br />

21<br />

<br />

<br />

<br />

92


Four-level pumping<br />

<br />

In most four-level systems, the nonradiative decay component<br />

in the 2 to 1 transition is negligible ( sp > sp<br />

>> 1 , s.t.<br />

N 0<br />

R sp<br />

s<br />

sp<br />

And therefore<br />

N <br />

R sp<br />

1 sp<br />

W i<br />

<br />

We have assumed that the pumping rate R is independent of<br />

the population difference N = N 2 –N 1 .<br />

93<br />

93


Four-level pumping<br />

<br />

This is not always the case because the population densities<br />

of the ground state and level 3, N g and N 3 , are related to N 1<br />

and N 2 by<br />

N g<br />

N 1<br />

N 2<br />

N 3<br />

N a<br />

<br />

where the total atomic density in the system, N a , is a constant.<br />

If the pumping involves a transition between the ground state<br />

and level 3 with transition probability W, then<br />

R (N g<br />

N 3<br />

)W<br />

<br />

If levels 1 and 3 are short-lived, then N 1 ≈ N 3 ≈ 0, N g + N 2 ≈<br />

N a s.t. N g ≈ N a -N 2 ≈ N a - N<br />

94<br />

94


Four-level pumping<br />

<br />

<br />

<br />

Under these conditions, the pumping rate can be<br />

approximated as<br />

R ( Na N)<br />

W<br />

which reveals that the pumping rate is a linearly decreasing<br />

function of the population difference N and is thus not<br />

independent of it.<br />

This arises because the population inversion established<br />

between levels 2 and 1 reduces the number of atoms available<br />

to be pumped.<br />

We obtain<br />

N <br />

sp<br />

N a<br />

W<br />

1 sp<br />

W sp<br />

W i<br />

95<br />

95


Four-level pumping<br />

<br />

The population difference can be written in the general form<br />

N<br />

N0<br />

1 <br />

s W i<br />

N 0<br />

spN a<br />

W<br />

1 sp<br />

W<br />

s<br />

<br />

sp<br />

1 sp<br />

W<br />

96<br />

96


Three-level pumping<br />

<br />

Rapid<br />

decay<br />

Short-lived |3><br />

Long-lived |2><br />

Pump R<br />

<strong>Laser</strong> W i<br />

-1<br />

<br />

Ground state |1><br />

Here we assume that the rate of pumping into level 3<br />

is the same as the rate of pumping out of level 1.<br />

97<br />

97


Three-level pumping<br />

<br />

Under rapid 3 to 2 decay, the three-level system (assumed R<br />

is independent of N)<br />

R 1<br />

R 2<br />

R 1<br />

2<br />

21<br />

<br />

In steady state, both the rate equations provide the same result<br />

0 R N 2<br />

21<br />

N 2<br />

W i<br />

N 1<br />

W i<br />

<br />

As 32 is very short, level 3 retains a negligible steady-state<br />

population. All of the atoms that are raised to it immediately<br />

decay to level 2.<br />

N 1<br />

N 2<br />

N a<br />

98<br />

98


Three-level pumping<br />

<br />

The population difference N can be cast in the form:<br />

<br />

Where<br />

N <br />

N 0<br />

1 s<br />

W i<br />

N 0<br />

2R 21<br />

N a<br />

s<br />

2 21<br />

<br />

When nonradiative decay from level 2 to level 1 is negligible<br />

( sp


Three-level pumping<br />

<br />

<br />

<br />

Attaining a population inversion N 0 > 0 in the three-level<br />

system requires a pumping rate R > N a /2 sp . A substantial<br />

pump power density given by E 3 N a /2 sp .<br />

The large population in the ground state (which is the lowest<br />

laser level) is an inherent obstacle to attaining a population<br />

inversion in a three-level system that is avoided in a fourlevel<br />

system (in which level 1 is normally empty as 1 is<br />

short).<br />

The saturation time constant s ≈ sp for the four-level<br />

pumping scheme is half that for the three-level scheme.<br />

100<br />

100


Three-level pumping<br />

<br />

The dependence of the pumping rate R on the population<br />

difference N can be included in the analysis of the three-level<br />

system by writing<br />

R (N 1<br />

N 3<br />

)W<br />

<br />

N 3 ≈ 0, N 1 = (N a -N)/2,<br />

R 1 2 (N a<br />

N)W<br />

<br />

Substituting this in the principal equation<br />

N 2R sp N a<br />

1 2 sp<br />

W i<br />

101<br />

101


Three-level pumping<br />

<br />

We can write the population difference in the usual form<br />

N <br />

N 0<br />

1 s<br />

W i<br />

N 0<br />

N a ( sp W 1)<br />

1 sp<br />

W<br />

s<br />

<br />

2 sp<br />

1 sp<br />

W<br />

As in the four-level scheme, N 0 and s saturates as the<br />

pumping transition probability W increases.<br />

102<br />

102


Saturated gain in homogeneously broadened media<br />

The gain coefficient () of a laser medium depends on the<br />

population difference N.<br />

N is governed by the pumping level.<br />

N depends on the transition rate W i .<br />

W i depends on the photon-flux density .<br />

=> the gain coefficient () of a laser medium is dependent on<br />

the photon-flux density that is to be amplified. This is the<br />

origin of gain saturation and laser amplifier nonlinearity.<br />

103


Saturation photon-flux density<br />

<br />

Substituting<br />

W i<br />

()<br />

<br />

Into steady-state population difference (in the presence of<br />

amplifier radiation)<br />

N <br />

N 0<br />

1 s<br />

W i<br />

=><br />

N <br />

N 0<br />

1 / s<br />

()<br />

where<br />

1<br />

s<br />

() s () <br />

c 2<br />

8n 2 2<br />

s<br />

sp<br />

ĝ()<br />

104


Gain coefficients<br />

This represents the dependence of the population difference N on<br />

the photon-flux density .<br />

Substituting N into the expression for the gain coefficient,<br />

2<br />

c<br />

( ) N<br />

( ) N<br />

2 2<br />

8n<br />

<br />

sp<br />

gˆ(<br />

)<br />

<br />

We obtain the saturated gain coefficient<br />

() <br />

0<br />

()<br />

1 / s<br />

()<br />

Where the small-signal gain coefficient<br />

c 2<br />

0<br />

() N 0<br />

() N 0<br />

8n 2 2 sp<br />

ĝ()<br />

105<br />

105


Gain coefficients<br />

<br />

<br />

The gain coefficient is a decreasing function of the photonflux<br />

density .<br />

When equals its saturation value s () = s , the gain<br />

coefficient is reduced to half its unsaturated value.<br />

<br />

)<br />

1<br />

0.5<br />

0<br />

0.01 0.1 1 10<br />

s<br />

()<br />

106<br />

106


Amplified spontaneous emission<br />

<br />

<br />

<br />

The resonant medium that provides amplification via<br />

stimulated emission also generates spontaneous emission.<br />

The light arising from the spontaneous emission, which is<br />

independent of the input to the amplifier, represents a<br />

fundamental source of laser amplifier noise.<br />

Whereas the amplified signal has a specific frequency,<br />

direction, and polarization, the noise associated with<br />

amplified spontaneous emission (ASE) is broadband,<br />

multidirectional, and unpolarized.<br />

=> it is possible to filter out some of this noise by following the<br />

amplifier with a narrowband optical filter, a collection<br />

aperture, and a polarizer.<br />

107


Amplified spontaneous emission<br />

Spontaneous<br />

photon flux<br />

Filter and<br />

polarizer<br />

Input<br />

photon flux<br />

<br />

Spontaneous emission is a source of amplifier noise. It is<br />

relatively broadband, radiated in all directions, and<br />

unpolarized. Optics can be used at the output of the amplifier<br />

to limit the spontaneous emission noise to a narrow optical<br />

band, solid angle d and a single polarization.<br />

108


Amplifier noise<br />

<br />

<br />

<br />

<br />

The ASE of a laser amplifier is directly proportional to the optical<br />

bandwidth of the amplifier.<br />

To increase the signal-to-noise ratio (SNR) at the amplifier output, the<br />

total noise power can be reduced to a minimum by placing at the output<br />

end of an amplifier an optical filter that has a narrow bandwidth matching<br />

the bandwidth of the optical signal.<br />

Because of the spontaneous emission noise, the SNR of an optical signal<br />

always degrades after the optical signal passes through an amplifier.<br />

The degradation of the SNR of the optical signal passing through an<br />

amplifier is measured by the optical noise figure of the amplifier defined<br />

as<br />

F o<br />

SNR in<br />

SNR out<br />

where SNRin and SNRout represent the values of the optical SNR at the<br />

input and output ends of the amplifier.<br />

109

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!