Download as pdf or txt
Download as pdf or txt
You are on page 1of 82

Module-5

Organic Thin Film Transistors and


Lasing Action of Thin Films
Introduction to Transistors
• Device used to control movement of electrons in the circuit
• Functions as a switch or amplifier
• Successfully demonstrated in 1947 at Bell Laboratories
• Invention is credited to
• William Shockley
• John Bardeen
• Walter Brattain
• Components
• Three layers of semiconductors, two of the same type, and one of different kind.
• e. g. p-n-p or n-p-n heterojunctions
• Three terminals device: base, emitter, collector

• Material used:
Inorganic semiconductors, such as Si, Ge, GaAs
Other p-type and n-type semiconductors
Organic thin film transistors

Organic materials are excitonic materials. In this regard, they


would seem to be ideally suited for optoelectronic devices,
such as OLEDs, OPVs, and photodetectors.

But what about their uses as purely electronic materials?


Specifically, do organics hold promise as a platform for thin film
transistors useful solely for electronic applications? In this
chapter we address just this question.
Organic thin film transistors
“God made the bulk; the surface was invented by the devil.”
Wolfgang Pauli, Nobel Laureate in Physics

Organic thin film transistors (OTFTs) differ from the other devices that we have
considered in that they are controlled by the surface, rather than the film bulk. By this
we mean that the charge flows between source and drain contacts along the organic
semiconductor/dielectric interface.

From many decades of engineering of conventional semiconductors, it is indisputable


that surface physics, and particularly surface charge transport is far more complex
than analogous phenomena in the bulk.

The complications are inherent in the 2D nature of interfaces, introducing defects and
lattice reconstruction that occurs at surfaces. Hence, the understanding and
optimization of OTFTs necessarily implies that we examine the complex microscopic
properties of organic semiconductor/gate insulator interfaces.
Organic thin film transistors-Basics

The organic thin film transistor (OTFT) is a


metal– insulator–semiconductor field effect
transistor, or MISFET, whose basic
configuration is shown in Fig. 8.1.

It is a three terminal device comprising a


source (S) and a drain (D) contact made to a
thin film semiconducting channel.

An insulating layer (the gate insulator) is in


contact with at least one surface of the
semiconductor.
The gate contact (G) is located on the
opposite side of the insulator, bridging
the gap between S and D.
Organic thin film transistors-Basics
Figure 8.1 also defines several important
geometric factors:
The gate width (W) is typically much larger
than its length (L).
The channel thickness is d, and the insulator
thickness is ti.

Charge is introduced into the channel by


injection from the source contact due to a
potential difference from the source to the
gate, VGS.
Then conduction in the channel occurs by
applying a voltage between source and
drain, VDS.

Since the amount of charge injected into the channel depends on the gate voltage, the
source-drain current, ID, is modulated by the magnitude of VGS.

Thus, field effect transistors are essentially gate-controlled resistor.


OFET devices are a largely simple construction containing an organic
semiconductor layer, an insulating layer, and source-drain-gate
electrodes. These components allow for the evaluation of fundamental
transistor parameters including mobility, operation voltage, and driving
stability.
VD

VG
p- or n- channel device
The sign of the charge that carries the drain current determines whether the OTFT is a
p- or n-channel device.

In most organics, the channel itself is undoped, although it may have a residual
concentration of holes or electrons that determine the position of the Fermi energy at
equilibrium.

This is a different operating regime than for an inorganic MISFET where the channel
region is doped n- or p-type, and hence the semiconductor bulk, or body, supplies the
carriers at the insulator/semiconductor interface.

For example, the high hole mobility in pentacene makes this material useful in p-
channel OTFTs.

A negative voltage at the gate draws holes to the semiconductor/insulator interface,


while a negative voltage at the drain leads to conduction within the channel. These are
the same bias conditions used for an inorganic p-channel MISFET.
Transistor Configurations

The BG/TC transistor is possibly the easiest to fabricate, and hence is the most common
architecture since it allows the S and D contacts to be deposited through a shadow mask.

Another commonly employed architecture is the bottom gate/bottom contact (BG/BC)


configuration.

This places all contacts on the same surface of the semiconductor, which allows for
photolithographic pattering to achieve very short gate geometries. The gate contact is
accessed by extending it beyond and parallel to the source and drain contacts.

The selection of a particular architecture is ultimately based on the choice of materials,


the patterning method, and the demands placed by the circuit on its intended application
(e.g. high current, high bandwidth, low noise, etc.).
Transistor Configurations
The device pictured in Fig. 8.1 has a bottom gate top source-drain contact configuration,
or briefly, it is a bottom-gate/top-contact (BG/TC) transistor. However, as shown in Fig.
8.2, many other possible architectures realized by placing the gate and channel contacts
on either the same or opposing sides of the channel semiconductor.
History of organic thin film transistors (OTFT)

The history of thin fill transistor can be traced


back to a patent by Lilienfeld who proposed, but
did not demonstrate a “device for controlling
electric current” employing CuS as the channel
material, and Al2O3 as the gate insulator
(Lilienfeld, 1933).

The first realization of a thin film transistor grown


on an insulating substrate employed a vapor
deposited microcrystalline film of CdS as the
semiconductor, Au as the gate electrode, and SiO2
as the gate insulator.

This TG/BC structure was fabricated on a glass


substrate. As shown in Fig. 8.3, it bears all of the
attributes of modern OTFTs (Weimer, 1962)
OTFT- Merocyanine Channel
The door to OTFTs was opened by depositing a merocyanine channel on
interdigitated S and D contacts patterned onto a 300 nm thick SiO2 gate insulator.
This combination of materials used a 1 Ω cm p-Si substrate that served as the gate
electrode. The BG/BC configuration is shown in Fig. 8.4 (Kudo et al., 1984).

The device did not enter the saturation regime, nor


did the channel pinch off, even at +15 V, indicating
excessive background conduction of the merocyanine
dye.

A small hole field effect mobility of 10-5 cm2 /V s was


extracted from the device characteristics in the
unsaturated, or linear regime.

Hence, while this device has the elements of an OTFT,


it did not show clear transistor operation.

The device dimensions were L = 20 μm, W = 7 mm, and d = 40 nm.


OTFT- Polythiophene thin film channel
Apparently, the first report of an OTFT that showed all of the characteristics of a functioning
transistor exhibiting both linear and saturated drain current characteristics, used a
conjugated polythiophene thin film channel in a BG/BC configuration
(see Fig. 8.5) (Tsumura et al., 1986).

The polythiophene film was deposited on a SiO2 gate insulator on a 4–8 Ω-cm n-Si
substrate gate contact.

The film was electrochemically de-doped in solution to ensure a non-conducting channel


in the “off” state.

Dimensions; L = 10 μm, W = 2 mm, and d = 140 nm, ti = 300 nm


OTFT- Polythiophene thin film channel

The drain-current characteristics have three


characteristics that are expected of all practical
enhancement mode FETs:

(i) A linear onset of ID vs. VDS extending up to


VDS≃VGS.
(ii) At higher VDS, the ID becomes independent
of VDS as the transistor enters the saturation
regime at pinch-off.
(iii) The drain current, ID = 0 as VGS = 0.

A hole mobility of 10-5 cm2 /V s is inferred from


analysis of these characteristics.
α-hexathiophene (α-6T) OTFT

Within three years, a vacuum-deposited (vs. solution processed) small molecule


thin film of α-hexathiophene (α-6T) OTFT was demonstrated (Horowitz et. al
1989).)

The devices were in a BG/TC configuration, again using n-Si as the gate contact and
substrate, and a 300 nm thick SiO2 gate oxide.

The OTFT had dimensions of L = 90 μm, W =310 μm, and d = 70 nm.

The as-deposited films exhibited excessive channel conductivity, and hence required
de-doping by heating in air for 3 h at 120°C after deposition.

Following de-doping, the enhancement mode transistors exhibited a saturation-


regime mobility of 10-4 cm2 /V s, which was the highest reported up to that time.
OTFT on flexible poly(parabanic acid) (PPA) substrate
The next significant advance in OTFTs was made by replacing the Si gate coated with an
SiO2 gate insulator by a metal electrode and an organic dielectric, cyanoethylpullulan
(CYEPL) in the structure shown in Fig. 8.6, inset (Garnier et al., 1990).

The BG/TC structure starts with a 25 μm thick, flexible poly(parabanic acid) (PPA) substrate.

A 1 mm wide by 20 mm long Au gate contact stripe is then vacuum deposited, followed by


solution drop deposition of CYEPL.

Depositing an α-6T channel and Au S and D electrodes completed the L = 50 μm, W = 5 mm,
and d = 50 nm OTFT.

This structure has several characteristics that distinguished it from previous work.

A metal strip replaced the conducting substrate used previously as the gate contact. This
allows for applying different gate biases to different transistors integrated on the same
substrate, which is required in almost all electronic circuit designs.

The device characteristics is shown in Fig. 8.6.


OTFT on flexible poly(parabanic acid) (PPA) substrate
OTFT on flexible poly(parabanic acid) (PPA) substrate

The use of a plastic substrate allows for bending and twisting without changing the
device characteristics, as shown in Fig. 8.6.

This takes advantage of the inherent flexibility of van der Waals bonded solids, arguably
one of the most important characteristics of organic electronic devices. And finally, the
high quality of the organic insulator/channel interface resulted in an unprecedented
high hole field effect mobility of 0.43 cm2 /V-s.

This represents at least a 100-fold improvement over previously demonstrated devices,


and indeed made this OTFT competitive with amorphous Si thin film transistors then in
use in LCD back-planes.

The high mobility also leads to a low VGS ~ 10 V compared with previous devices where
VGS ~ 50 V was common (cf. Fig. 8.5). This early, high performance OTFT considerably
heightened interest in the technology, and subsequently led to its further rapid
development.
Metal–insulator–semiconductor contacts
The key to understanding OTFT operation is the
metal gate–insulator–semiconductor (MIS)
capacitor. In Fig. 8.8 we show the energy level
diagram for a MIS capacitor comprising a metal
contact and semiconductor with work functions,
ϕM and ϕS, respectively.
In this example, ϕM < ϕS.
When the metal, insulator and semiconductor
are brought into contact, charge flows until
equilibrium is established. Then the Fermi
energy, EF, across the device is leveled, resulting
in bending of ELUMO, EHOMO, and the vacuum level,
EVAC, near the dielectric/semiconductor interface.
This induces a dipole of energy, ∆, in the
insulator, and depletion of holes within the
semiconductor.
From inspection of this diagram:
where χ is the electron affinity of the semiconductor and Vbi is the built-in
potential. It is convenient to define the surface potential, as the band bending
at the semiconductor/insulator interface.

In the special case where ϕM = ϕS, Vbi = 0 and the bands are flat throughout the
semiconductor channel.
This is known as the flat-band condition. The device in Fig. 8.8 achieves flat-band when
the potential applied to the metal (i.e. the gate potential in a transistor) is VGS = ψS.
In MIS systems, there are several sources of trapped charge both within the insulator
and at the interfaces. Specifically, trapped charges per area due to fixed (∆QF) and
mobile (∆QM) ions in the insulator bulk, and interface traps (∆QIT), contribute to a
change in gate voltage. The presence of charge traps in equilibrium also contributes to
the energy level bending, and hence affects the surface potential in the channel.

Defining the sum of all charges trapped at defects both within the bulk and at the
surfaces of the insulator as ∆Qi = ∆QF + ∆QM + ∆QIT (each term of which can be either
positive or negative, depending on the polarity of the trap at equilibrium), the built-in
potential becomes;

where the insulator capacitance per area in the absence of traps is

Here, both Q and C are normalized to the contact area, and εi and ε0 are the relative
dielectric constant of the insulator and the permittivity of free space, respectively
Control of the channel charge and potential
Control of the channel charge and potential by the gate voltage in an MIS capacitor is shown
in Fig. 8.9, assuming ϕM = ϕS and Qi = 0. The energy level diagrams are shown at the top of
the figure, and the charge distributions are shown at the bottom. Generally, the charge
density in the organic is small, but rarely is it zero. In the MIS capacitor, we assume the
organic is infinitely thick with a background hole density, p > 0.
At VGS = 0, the device is at thermal At VGS < 0, holes accumulate at the surface at
equilibrium and the bands are flat. density +QS. This is the accumulation regime.
Under a small positive bias less than the threshold voltage, that is, at VGS < VT,
charge is depleted from the insulator/semiconductor interface, leaving only fixed
bulk charge, ∆QB, due to the residual holes in the channel.
At VGS > VT the Fermi level in the semiconductor moves above mid-gap and the minority
charge concentration at the surface surpasses the majority carrier concentration. In this
case, the semiconductor is said to be inverted, where the number of electrons, n > p.
Note that this description is general. However, due to the low background charge
concentrations in undoped organics, typically only the accumulation regime is accessed
by applying a negative gate voltage.
The inversion condition occurs when the free charge at the interface (i.e. electrons in
Fig. 8.9) equals the holes in the bulk, and therefore creates a neutrality condition.
This is the depletion-inversion transition point, and from inspection of Fig. 8.8, occurs
when qψS = 2( EF – EI) bulk where the last term refers to the values of EF and EI in the
channel bulk.
Here, EI is the intrinsic level energy. The gate voltage at the transition point is the
threshold voltage, VT, and is defined by satisfying the condition for the total surface
charge:

In the absence of channel doping, then

where ψB is the Fermi energy relative to the intrinsic level in the bulk of the
semiconductor. The charge density in the bulk of the semiconductor is given by

where NA is the density of residual acceptors in the bulk.


The total threshold voltage is thus

The expression to the far right is valid for a lightly doped organic
semiconductor of thickness d that is completely depleted at VGS = 0.

This situation is the most commonly encountered in organic


semiconductor.
The electron surface charge density for an MIS capacitor with an electron-rich
semiconductor bulk follows

where nS is the surface electron density, (with the total charge density in the
channel of nB = nS∆y where ∆y is the width of the charge layer at the
insulator/semiconductor interface), and npS0 is the electron surface charge
density at equilibrium. For an undoped semiconductor, QS increases
exponentially in both the accumulation and strong inversion regimes
A qualitative plot of the surface charge, log(|QS|), vs. the surface energy qψS
for an MIS capacitor is provided in Fig. 8.10. This plot shows how the gate
potential controls the charge density at the semiconductor/insulator
interface.

In weak inversion, the charge


density increases only slowly with
ψS
Importantly, at flat band, the
semiconductor is in equilibrium
where the number of thermally
generated holes and electrons are
equal, thus resulting in the absence
of a net charge at the interface.
OTFT Operation
Since organics have a low background charge density, charge is drawn into the
channel via injection from the S contact by applying a bias, |VGS| > |VT|.
Furthermore, the channel is “normally off” when no charge exists at VGS = 0. This
corresponds to accumulation enhancement mode operation.

For the unusual case where the channel is heavily doped, it can be conducting at
VGS = 0, in which case it is “normally on.” This is the depletion mode of operation,
which is rarely exploited in OTFTs due to the difficulty in obtaining sufficiently high
free charge concentrations via doping.

In the accumulation-enhancement mode, at |VGS| > |VT|, injected charge in the


channel diffuses along the insulator interface to the drain.
OTFT Operation
Figure 8.11 shows the OTFT along the length of
the gate at different VDS.
The illustration shows the extent of the
conduction region from VDS = 0 (Fig. 8.11a), to
VDS = VDSsat (Fig. 8.11b), that is, the point at
which the drain potential leads to a pinch-off
of the charged layer at the drain.

At small VGS, therefore, the channel current is


in the linear regime when ID vs. VDS
approximately follows Ohm’s law. In this
regime, the channel resistance is relatively
independent of VDS.
OTFT Operation

At |VDS| = | VDSsat|, the current


demanded by the drain potential can
no longer be supplied by the source
without increasing VGS needed to
increase the charge density at the
insulator interface.

This creates a depletion region


between the pinch-off point and the
source, preventing the drain current
from increasing further.

Thus, when |VDS|>|VDSsat|, the free charge recedes a distance, ∆x, from the drain
(Fig. 8.11c). The drain current is pinned at IDSsat assuming there is no series
resistance contributed by the depleted gate region.

Since the drain current is constant at higher voltages, this is known as the
saturation regime.
Organic Field-effect Transistors OFET Device Structure and
Operating Mechanism
Organic Field-effect Transistors (OFETs) Are Useful Devices for Testing the
Properties of Organic Semiconductors

OFETs Are Three Electrode Devices


Operating Mechanism of an OFET in
Hole Transporting Mode
Operating Mechanism /
working-principle of
OFET (Figure-1)

~70-80% organic semiconductors


are of p-type.

S – Source Electrode After inducing positive charges in p type organic


D – Drain Electrode
semiconductor via applying VG< 0, Apply VD< 0 to
Gate – Gate Electrode
VG – Voltage Between Source and Gate
collect the charge carriers (i.e. holes in p-type organic
VD – Voltage Between Source and Drain Semiconductor) from Source(S) to Drain (D).
L – Channel Length, Distance between Source and Drain Electrodes (Typical Value ~200 µm)
W – Channel Width, Distance of the Source and Drain Electrodes in the Direction Orthogonal to
the Channel Length Direction (Typical Value ~2,000 µm)
Operating Mechanism/ working principle of OFET
No charges are injected
when VG = 0 V

When VG >0 V applied, n-


When VG< 0 V applied, p-type
type OFET is realized.
OFET is realized.

Illustration of an OFET working principle with respect to applied VG


I-V Characteristics: two types of typical plots for p-type organic
Semiconductor

In the sub-threshold regime the device is


turning from OFF (low current) to ON
(high current)
Field Effect Mobilities
Two important performance parameters of organic TFTs are, charge carrier
mobility (μ) and Ion/IOFF ratio

p-type channel mobility is common in organic field effect transistor. However,


recently significant progress has been made in discovery of n-type semiconductors

Pentacene
n-type channel mobility
p-type channel mobility
μ= 0.18 cm2/V.s
μ= 2-3 cm2/V.s ION/IOFF = >106
ION/IOFF = >108

n-type channel semiconductors normally have electron withdrawing cyan (CN), (nitro)
amide (CO-NH), and fluoro (F) groups attached to the basic molecular skeleton
Commonly used p - and n- type organic Semiconductors

Reference: Materials Today,Volume 7, Issue 9, September 2004, Pages 20-27


Commonly used n type organic Semiconductors

NDI, naphthalene
diimide;

F16CuPc,
perfluorocopperphthalocyanine;
perylene;

PTCDA, 3,4,9,10-
perylene-tetracarboxylic dianhydrid
and its derivaties;

PDI, N,N´-dimethyl 3,4,9,10-perylene


tetracarboxylicdiimide;
C60;
PCBM, methanofullerene [6,6]-phenyl
C61-butyric acid methyl ester
Field Effect Mobilities
The charge carrier mobility's are not only determined by molecular
structure of the organic semiconductors, but also by the intermolecular
force that guides the structural order in the thin film.

The upper limits in microscopic mobilities of organic molecular crystals,


determined at 300 K by time-of-flight experiments, are falling between
1 and 10 cm2 V-1s-1.

The weak intermolecular interaction forces in organic semiconductors,


most usually van der Waals interactions with energies smaller than 10
kcal mol-1, may be responsible for this limit.
Field Effect Mobilities
In contrast, in inorganic semiconductors such as Si and Ge, the atoms
are held together with very strong covalent bonds, which for the case
of Si have energies as high as 76 kcal mol-1 In these semiconductors,
charge carriers move as highly delocalized plane waves in wide bands
and have a very high mobility.

Band transport is not applicable to disordered organic semiconductors,


in which carrier transport takes place by hopping between
localized states and carriers are scattered at every step.
Field Effect Motilities
• The boundary between band transport and hopping
is defined by materials having mobilities between
0.1 and 1 cm2 V-1 s-1.

• Most of the organic semiconductors have the


mobility's at this boundary

• Highly ordered organic semiconductors, such as


several members of the acene series including
anthracene and pentacene, have room temperature
mobilities in this intermediate range.
ION/IOFF ratio

The ION/IOFF ratio is the value of the OFF current


to the ON current, and is reported in terms of
10x. In general, it is important to minimizes IOFF
through minimizing dopants and leak and
maximizing ION through increasing the mobility
of the semiconductor system.
Applications
1. Organic field-effect transistors (OFETs) are unique devices capable of serving as testbeds to
evaluate the mobility of organic semiconductors.
2. They can be used as logic switches because they have a relatively large difference in current
between their ON and OFF states. The three terminal geometry allows for the control of charge
induction, and thus charge flow, through the gate electrode.

3. Fabrication of Organic light-emitting transistors (OLETs) by combining OLED and OFET


characteristics

4. Flexible and wearable sensors, and sweat sensors

5. BG/BC device configuration can be used as address transistor for activating display pixels
OLET test device.

Organic light-emitting transistors (OLETs) are possibly the smallest integrated optoelectronic
devices that combine the switching and amplification mechanisms of organic field-effect
transistors (OFETs) and the electroluminescent characteristic of organic light-emitting diodes
(OLEDs).
The OLETs are considered ideal for developing the next-generation flexible display technology,
electrically pumped organic lasers, and optical communication technology.
OLET test device.
Fig. 5 (a) Sensing response of the chitosan and SWNT-modified wearable sensors as a
function of glucose concentration in 0.1PBS, artificial tears, artificial sweat, and saliva; (b)
photographs of the sensors attached to an artificial eyeball and an artificial arm, the real-
time response of the glucose level in the artificial sweat from the sensor attached in (c)
artificial eye ball and (d) artificial hand
Flexible display backplanes

Organic transistor arrays are well-suited for information


display backplanes to activate voltage-driven pixels
(Gelinck et al., 2006, Yoneya et al., 2006). These displays
include LCDs and electrophoretic displays.

While a-Si TFTs can serve as the pixel driver, their


fabrication requires high temperature processing that is
incompatible with lightweight, flexible plastic substrates
that are often used for this application.

A 320x240 pixel, 4.7” quarter video graphics array


(QVGA) electrophoretic display driven by a solution-
processed pentacene OTFT backplane is shown in Fig.
8.111a (Gelinck et al., 2006).

The backplane is fabricated on a 25 μm thick PEN


substrate, with a single BG/BC address transistor per
pixel.
LASING PROCESS

LASER : Light Amplification by Stimulated Emission of Radiation


Laser
• The acronym laser refers to Light Amplification by
Stimulated Emission of Radiation.
• A laser is an optical resonator that comprises three
components: a gain medium, a low loss optical
cavity, and a pump source (electrical or optical).
• The pump intensity must be sufficient to drive the
gain medium into an excited state population
inversion, leading to stimulated emission.
What is inside a laser?

Lasers are composed of three essential elements :


(1) A gain medium
(2) A pumping source that excites the gain medium ( optical or electrical)
(3) A low loss optical cavity comprised of two mirrors that direct the light beam back and
forth through the gain medium.
An optical cavity, resonating
cavity or optical resonator is an
arrangement of mirrors or other
optical elements that forms
a cavity resonator for light waves.

Optical cavities are a major


component of lasers,
surrounding the gain medium and
providing feedback of the laser
light.

Light confined in a resonator will


reflect multiple times from the
mirrors, and due to the effects
of interference, only certain patterns
and frequencies of radiation will be
sustained by the resonator, with the
others being suppressed by
destructive interference.
Cavity and mode characteristics
The laser medium is confined to a cavity that ensures that only certain photons of
a particular frequency, direction of travel, and state of polarization are generated
abundantly.

The cavity is essentially a region between two mirrors, which reflect the light
back and forth. This arrangement can be regarded as a version of the particle in a
box, with the particle now being a photon. As in the treatment of a particle in a
box (refer module 1), the only wavelengths that can be sustained satisfy

where n is an integer and L is the length of the cavity. That is, only an integral
number of half-wavelengths fit into the cavity; all other waves undergo
destructive interference with themselves. In addition, not all wavelengths that
can be sustained by the cavity are amplified by the laser medium (many fall
outside the range of frequencies of the laser transitions), so only a few contribute
to the laser radiation. These wavelengths are the resonant modes of the laser.
Simplified process involved in laser action
Ordinary Source of light

Monochromatic and incoherent Source


Of Light

Monochromatic and coherent Source of Light

(Example:LASER)
Two level energy diagram for lasing
Process of laser light emission through population inversion is known as lasing action.
The requirement of population inversion, number of particles in higher energy state is larger
than ground state, can be understood by beginning with two-level energy diagram.

A schematic representation of a two-level energy diagram. The circles represent the


number of atoms in each state, eight in the ground state and four in the excited state.
The energy density of the light is described by two related quantities.

The radiant energy density, ρ, is defined as the radiant energy per


unit volume and has units of J. m -3

The spectral radiant energy density, ρϑ , is a measure of the radiant


energy density per unit frequency, ρϑ = dρ/dϑ, and has units of J.m-3 .s

Because the transition between states 1 and 2 occurs only if light at


ϑ= ϑ12 is provided, we will be interested in the spectral radiant energy
density at ϑ12 ρϑ( ϑ12 ), of the incident light source.

A phenomenological approach that describes the rates of the various transitions between
electronic states was proposed by Einstein.
An illustration of the absorption process. Light of energy hv12 = E2- E1 can be absorbed by
an atom, which causes the atom to make a transition from the ground state to an
electronically excited state.
The spontaneous-emission process. Light of energy hv12 = E2 - E1 is emitted by an excited
atom when the atom makes a transition from the electronically excited state to the ground
state

The rate of spontaneous emission is simply proportional to the number of atoms in the
excited state, N2(t), at time t.
The stimulated-emission process. Incident light of energy hv12 = E2 - E 1 stimulates an atom
in an excited electronic state to emit a photon of energy hv12 and thereby causes the atom
to make a transition from the excited electronic state to the ground electronic state
Note:

 the stimulated-emission process amplifies light intensity; one


photon at frequency ϑ12 stimulates an atom to emit another,
thus generating a second photon at frequency ϑ12
 In a large sample of atoms, this process can occur many times,
resulting in a substantial amplification of an incident light beam at
frequency ϑ12

 Lasers are devices that exploit the amplification of light through


stimulated emission (recall that the word, laser, stands for light
amplification by stimulated emission of radiation).
Upon exposure to light, a sample of atoms simultaneously
undergoes all three processes, absorption, spontaneous emission,
and stimulated emission. Thus, the rate of change in the population of
either the ground electronic state or the excited electronic state must
be the sum of the rates of these three individual processes :
A Two-level system can not achieve a Population Inversion
 Lasers are designed to amplify light by the stimulated emission of radiation.
 For this amplification to occur, a photon that passes through the sample of atoms must
have a greater probability of stimulating emission from an electronically excited atom than
of being absorbed by an atom in its ground state.
 This condition requires that the rate of stimulated emission be greater than the rate of
absorption:

 The stimulated emission is more probable than absorption only when N2 > N1, or when
the population of the excited state is greater than that of the lower state. Such a
situation is called population inversion.

Boltzmann distribution Equation

N2< N1
The ratio of the number of atoms in electronically excited states to the total number of
atoms, N2/ N total is plotted as a function of time for a two-level system. The number of atoms in
the excited state is always less than that in the ground state in a two-level system. Therefore, a
two-level system can never achieve a population inversion.
Three level lasers
Requirement of laser action
- The existence of a metastable excited state, an excited state with a long
enough lifetime for it to participate in stimulated emission.
- The existence of a greater population in the metastable state than in the
lower state where the transition terminates, for then there will be a net
emission of radiation

One way of achieving population inversion is


illustrated in Fig. 14.28. The molecule is excited
to an intermediate state I, which then gives up
some of its energy nonradiatively and changes
into a lower state A; the laser transition is the
return of A to the ground state X. Because
three energy levels are involved overall, this
arrangement leads to a three-level laser.
Three level lasers
• The disadvantage of this three-level
arrangement is that it is difficult to achieve
population inversion, because so many
ground-state molecules must be converted to
the excited state by the pumping action.
Four level lasers
The arrangement adopted in a four-level laser
simplifies this task by having the laser transition
terminate in a state A′ other than the ground state
(Fig. 14.29).

Because A′ is unpopulated initially, any population in


A corresponds to a population inversion, and we can
expect laser action if A is sufficiently metastable.

Moreover, this population inversion can be


maintained if the A′ ← X transitions are rapid, for
these transitions will deplete any population in A′
that stems from the laser transition, and keep the
state A′ relatively empty.
Organic Semiconductor Lasers
The four-level energy scheme for a laser based on Förster energy transfer from an optically
pumped Alq3 host to a DCM guest is shown in Fig. 6.198. Many such guest–host schemes
have been developed for optically pumped lasing based on both small molecule and
polymer active media (Hide et al., 1996, Kozlov et al., 1997a, 1998, Vasdekis et al., 2006).

Alq3

DCM
Lasers exhibit a clear threshold delineating regions of spontaneous and stimulated
emission. The threshold is identified by an abrupt increase in slope efficiency,
which is the ratio of the output intensity to the pump energy (or in the case of an
electrically pumped device, the drive current). Examples of the threshold and high
slope efficiency above threshold are illustrated for Alq3: DCM (2.5%) double and
single heterostructure, Fabry–Pérot lasers on InP substrates in Fig. 6.202. Below
threshold, the spontaneous emission efficiency is only ~ 1%, whereas it rises
abruptly once lasing is reached.
A laser used to read CDs emits red light of wavelength 700 nm. How many photons does
it emit each second if its power is 0.10 W?

You might also like